Вы находитесь на странице: 1из 271

Accepted Manuscript

Fatigue and Fracture Behavior of Bulk Metallic Glasses and Their Composites

Haoling Jia, Gongyao Wang, Shuying Chen, Yanfei Gao, Weidong Li, Peter K.
Liaw

PII: S0079-6425(18)30070-7
DOI: https://doi.org/10.1016/j.pmatsci.2018.07.002
Reference: JPMS 524

To appear in: Progress in Materials Science

Received Date: 21 March 2017


Revised Date: 11 May 2018
Accepted Date: 1 July 2018

Please cite this article as: Jia, H., Wang, G., Chen, S., Gao, Y., Li, W., Liaw, P.K., Fatigue and Fracture Behavior
of Bulk Metallic Glasses and Their Composites, Progress in Materials Science (2018), doi: https://doi.org/10.1016/
j.pmatsci.2018.07.002

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Fatigue and Fracture Behavior of Bulk Metallic

Glasses and Their Composites


Haoling Jia1, a, Gongyao Wang1, a, Shuying Chen1, Yanfei Gao1, 2, *, Weidong Li1, *,

and Peter K. Liaw1, *


1
Department of Materials Science and Engineering, The University of Tennessee,

Knoxville, TN 37996, USA.


2
Materials Science and Technology Division, Oak Ridge National Laboratory,

Oak Ridge, TN 37831, USA

Abstract

A fundamental understanding of the fatigue and fracture behavior of bulk

metallic glasses (BMGs) and their composites is of critical significance for designing

new BMG systems and developing new manufacturing and processing techniques so

as to broaden the scope of applications of BMGs and their composites. However, the

fatigue and fracture studies on BMGs are limited so far, compared to other

mechanical properties. The present work reviews the fatigue and fracture behavior of

BMGs and their composites, as well as that of metallic-glass films, ribbons, and

wires. The grand challenge for the fatigue and fracture performance of BMGs is:

a
These authors contributed equally to this work.
*
Corresponding authors. Tel.: +1 865 974 6356.

E-mails: ygao7@utk.edu (Y.F. Gao), lei432378yu@gmail.com (W.D. Li), pliaw@utk.edu

(P.K. Liaw).

1
What produces a large difference among the fatigue and fracture results of BMGs? In

fact, many factors could be involved, including the composition, specimen geometry,

surface condition, temperature, cyclic frequency, etc. The present work discusses

nearly all factors that could affect the fatigue and fracture behavior of BMGs and

their composites. Furthermore, the mechanisms of fatigue-crack initiation,

propagation, and fracture of BMGs and their composites in different loading

conditions and environments will be outlined, analyzed, and discussed. Future

research directions of fatigue and fracture of BMGs and their composites are provided

for reference.

Keywords: Metallic glasses; Composites; Fatigue; Fracture; Mechanical properties

2
Table of Content


Abstract .................................................................................................................... 1
1. Introduction ......................................................................................................... 5
1.1. Bulk Metallic Glasses ..................................................................................... 5
1.2. Fatigue Failure in BMGs ................................................................................. 8
2. Deformation Mechanisms in BMGs .................................................................. 11
3. High-Cycle Fatigue Behavior ............................................................................ 16
3.1. Fatigue Behavior of Metallic Glasses (Ribbons or Wires) in Air at Room
Temperature .................................................................................................. 17
3.2. Fatigue Behavior of BMGs and Their Composites in Air at Room Temperature
20
3.2.1. Bending Fatigue ..................................................................................... 20
3.2.2. Uniaxial Fatigue .................................................................................... 27
3.2.3. Rotating Fatigue .................................................................................... 33
3.3. Fatigue Behavior in Specific Environments ................................................... 34
3.3.1. Fatigue in Vacuum ................................................................................. 34
3.3.2. Corrosion Fatigue .................................................................................. 35
4. Fatigue-Crack-Propagation Behavior............................................................... 38
4.1. Ambient Conditions ...................................................................................... 38
4.2. Other Environments ...................................................................................... 40
5. Fatigue Failure Mechanisms ............................................................................. 42
5.1. Crack Initiation ............................................................................................. 44
5.2. Crack Propagation......................................................................................... 46
5.3. Fatigue-Failure Mechanisms ......................................................................... 48
6. Size Effects on Fatigue Behavior of BMGs ....................................................... 56
6.1. Size Effects on Fatigue Behavior of BMGs ................................................... 57
6.2. Size Effects on Fatigue Properties of Fatigue-Damaged BMGs ..................... 59
6.3. Size Effects on Fracture Toughness of BMGs ............................................... 61
7. Improving Fatigue Resistance ........................................................................... 66
7.1. TFMG on BMGs substrates .......................................................................... 66
7.1.1. Preparation of TFMGs ....................................................................... 67
7.1.2. Fatigue Properties of TFMGs ............................................................. 67
7.2. Four-Point-Bending Fatigue Behavior of TFMG Materials ........................... 72
7.3. Mechanical Properties of TFMGs ................................................................. 74
7.3.1. Hardness of TFMGs ............................................................................... 74
7.3.2. Microcompression of TFMGs ................................................................. 75

3
7.3.3. Extraordinary plasticity of TFMGs at room temperature ........................ 77
7.3.4. Nano-Scale FEM on TFMG-Substrate Materials .................................... 78
8. Fracture Toughness of BMGs and MGMCs..................................................... 79
8.1. Assessment of Fracture Toughness of BMGs ................................................ 81
8.2. Fracture Toughness of Monolithic BMGs ..................................................... 82
8.3. Fracture Toughness of Metallic Glasses Composites ..................................... 85
9. Ductile vs. Brittle Fracture in BMGs ................................................................ 88
9.1. Empirical Interpretation with Elastic Constants ............................................. 90
9.2. Microstructural Rationalization with the Heterogeneity Concept ................... 91
9.3. Design Ductility in BMGs............................................................................. 96
10. Notch Effect on Fracture and Fatigue of BMGs ............................................. 97
10.1. Notch Effect on Fracture ............................................................................. 99
10.1.1. Notch Sensitivity of Strength ................................................................. 99
10.1.2. Notch Sensitivity of Ductility .............................................................. 103
10.2. Notch Effect on Fatigue ............................................................................ 105
11. Failure Modeling and Lifetime Prediction.................................................... 106
11.1. Continuum-Mechanics Models ................................................................ 107
11.1.1. Rudnicki-Rice Instability Model.......................................................... 107
11.1.2. Free-Volume-Model Simulations ........................................................ 110
11.1.3. Other Types of Continuum Models...................................................... 112
11.2. Fatigue-Damage Simulation in BMGs ..................................................... 113
11.3. Mesoscale Model: STZ Dynamics Simulations for Cyclic Indentation with
Finite-Element Modeling (FEM) ................................................................. 118
11.4. Ductile to Brittle Transition in BMGs ...................................................... 122
11.4.1. Cavitation Process as a Precursor of Failure ..................................... 123
11.4.2. Continuum Plastic Process Zone versus Cleavage Fracture ............... 124
11.4.3. Roles of Structural Heterogeneities .................................................... 125
11.5. Modeling Size Effects on Fatigue Life of BMGs ..................................... 126
11.5.1. Theoretical Model on Fatigue Life of BMGs with Different Sizes ........ 127
11.5.2. Prediction on Fatigue Life of BMGs with Different Sizes .................... 129
12. Future Directions ........................................................................................... 130
13. Concluding Remarks ..................................................................................... 133
Acknowledgements .............................................................................................. 137

4
1. Introduction

1.1. Bulk Metallic Glasses

Crystalline materials usually consist of grains with different sizes and specific

microstructures. Thus, they contain many crystalline defects, such as dislocations and

grain boundaries. The movement of dislocations under load results in the plastic

deformation of crystalline materials, which is the reason why crystalline alloys cannot

achieve the theoretical strength needed to break the atomic bonds. However, grain

boundaries are easy to promote corrosion and/or chemical reactions (for example,

oxidation and sulfidation). Therefore, mechanical properties of crystalline materials

strongly rely on their crystalline structures. The atomic origins of the strength and

ductility of crystalline materials can be explained by the well-established dislocation

theory [1]. The limitations of crystalline materials structures can be changed through

the formation of a glassy structure with randomly-packed atoms, generally called

metallic glasses. Unlike the crystalline alloys, amorphous solids, such as metallic

glasses, lack long-range order characteristics [2, 3]. The disordered structure and

metastable state make metallic glasses exhibit unusual structural properties and non-

conventional deformation mechanisms [4, 5]. In fact, some amorphous materials (i.e.,

polymers, glasses, and plastics) have been applied widely to our daily life. However,

amorphous metallic alloys represent a relatively-new class of materials, compared to

other amorphous materials [4]. Before 1960, only amorphous thin films were

successfully deposited at very low temperatures [6]. An amorphous alloy was first

synthesized in 1960 through rapidly-quenching an Au-Si alloy directly with rates up

5
to 106 K/s [6]. The significance of the work is that the process of the nucleation and

growth of crystalline phases could be kinetically bypassed in some molten alloys to

produce metallic glasses. However, the specimen geometry was severely restricted to

thin ribbons, foils, and powders, because the high rate of heat transfer was required to

prevent crystallization at that time. Since then, remarkable progress was made in

exploring alloy compositions for excellent glass formers with ever-lower critical

cooling rates. In the late 1980s, Inoue’s group discovered new multicomponent bulk-

metallic-glass (BMG) systems with lower critical cooling rates and thicknesses of

several millimeters [7, 8]. A family of multicomponent Zr-based BMGs (e.g., Zr-Cu-

Ni and Zr-Cu-Ni-Al BMGs) were also developed later [7-9]. Moreover, Peker and

Johnson developed a quinary alloy, Zr-Ti-Cu-Ni-Be, with lower critical cooling rates

down to 1 K/s in 1993 [10]. The alloy, Zr41.2Cu12.5Ni10Ti13.8Be22.5 (in atomic percent,

at.%), was the first commercial BMG and named as Vitreloy 1 (Vit 1) [3]. Since then,

a vast number of glass-forming alloy systems were greatly developed, for example,

binary, ternary, and multi-component alloy systems. These metallic-glass alloy

systems involved Al-, Cu-, Fe-, La-, Mg-, Ni-, Pd-, Ti-, and Zr-based alloys [11-28].

Because they lack microstructural features, such as grains, grain boundaries,

and dislocations, BMGs exhibit many excellent properties: high strengths (even an

ultra-high strength of over 5 GPa); high hardness; high strength to weight ratios;

superior elastic limits; low coefficients of frictions; high scratch and wear resistances;

good corrosion resistances; net-shape castability; and good soft magnetic behavior

[29-35]. BMGs have been used to produce many products, for example: sporting

goods, watch parts, electromagnetic casings, optical parts, ornamental parts, choke

6
coils, power inductors, magnetic-field-identification systems, electromagnetic-wave-

shielding sheets, micro-geared motor parts, pressure sensors, surface-coating

materials, and medical instruments [2].

Nevertheless, the excellent properties of BMGs are accompanied by their lack

of homogeneous plastic deformation without dislocation-mediated crystallographic

slip [5]. Inhomogeneous deformation, in general, happens in BMGs under the

unconstrained conditions because of the formation and propagation of highly-

localized shear bands where a high amount of plastic strains is accumulated in a very

narrow region (i.e., 10 - 20 nm) [4, 36-38]. Although plastic strains are very large at

the localized shear bands, the overall plastic deformability of BMGs at low

temperatures is disappointingly low (< 2 - 3%) [36]. Thus, this kind of brittleness

seriously confines the applications of BMGs as a potential candidate for engineering

and structural materials.

Extensive experimental investigations using a wide variety of characterization

techniques have been applied to elucidate the microstructural root cause of the

brittleness in BMGs. A few important ones include the mechanical heterogeneity

characterization with nanoindentaiton [39, 40] and dynamic force microscopy [41],

morphological analysis of fracture surfaces with scanning electron microscopy (SEM)

and atomic force microscopy (AFM) [42, 43], and systematic measurements of

Poisson’s ratio and the ratio of the elastic shear modulus to the bulk modulus [44, 45].

In the meantime, various theoretical models on the basis of atomic operations were

proposed and used for exploring deformation mechanisms in metallic glasses. One of

such models is the free volume model, which models the inhomogeneous deformation

7
in metallic glasses as a result of the biased accumulation of free volumes at certain

locations under the action of stress [46-49]. Another commonly used model is the

shear transformation zone (STZ) model, which treats clusters of atoms as the carriers

of plastic deformation in metallic glasses. The STZ model was originally proposed by

Argon [50] and later elaborated by Langer and Falk [51, 52], and has been widely

used in molecular dynamics simulations of deformation and failure processes in

metallic glasses [51-54]. For interpreting the quasi-cleavage fracture features

observed in BMGs, the tension transformation zone (TTZ) model may find more

usages [55]. TTZs are also local clusters of atoms similar to STZs in size but having

reduced relaxation timescales and are more amenable to brittle fracture when

subjected to loads. Recently, realizing that metallic glasses may contain a great

amount structural heterogeneities, a flow unit mode was proposed to study the

inhomogeneous nature of metallic glasses [49]. Flow units are essentially loosely

packed regions inside an elastic metallic glass matrix, possess low modulus and

strength, and behave like viscous liquids.

1.2. Fatigue Failure in BMGs

Based on the US national survey of service failures, the fatigue-related

failures for ground vehicles reached 75 percent [56]. In addition, an approximately 90

percent of all service failures due to mechanical causes is associated with fatigue [57].

These results indicate that fatigue is a very important characteristic for structural

applications. Therefore, fatigue studies of BMGs are necessary to be performed for

their engineering applications.

8
Fatigue is the progressive, localized, and permanent structural change in

materials, which are subjected to repeated or fluctuating strains at nominal stresses

that are less than the yield strength of the materials [58]. Fatigue may culminate into

cracks and lead to fracture after sufficient fluctuating loads. The fatigue damage is

due to the simultaneous action of the cyclic stress, tensile stress, and plastic strain,

without which a fatigue crack will not initiate and propagate. The plastic strain

resulting from the cyclic stress initiates the crack, while the tensile stress promotes

crack growth and propagation.

Under usual loading conditions, fatigue failure might occur. In the fatigue-

damage process, fatigue cracks usually initiate near the singularities that are located

just below the surface, such as scratches, sharp changes in the cross section,

inclusions, etc. Even in a flaw-free metal and no stress concentrators, fatigue cracks

may form. If the alternating stress amplitude is high enough, plastic deformation takes

place, which will lead to the slip steps on the surface. Under further loading, the

microcracks then grow up to form more macrocracks, which will grow until fracture

toughness is exceeded. Therefore, for crystalline materials, the fatigue-failure process

can be divided into five stages [58]: (1) Cyclic plastic deformation prior to fatigue-

crack initiation; (2) Initiation of one or more microcracks; (3) Propagation of

microcracks to form macrocracks; (4) Propagation of macrocracks; and (5) Final

failure. The fatigue life is generally presented by the number of loading cycles to

initiate and propagate a crack to a critical size.

Recently, mechanical properties of BMGs are being studied extensively [59-

66]. However, fatigue studies of BMGs are comparatively few. Thus, the

9
understanding of the fatigue behavior of BMGs is very limited so far. Starting from

1975, the fatigue behavior of metallic glasses was reported [67, 68]. In general, the

samples of these metallic glasses are ribbons or wires. Until 1998, Gilbert et al. [69]

first reported the fatigue results of BMGs under four-point-bending loads. From then

on, several researchers have conducted the fatigue studies on different BMGs

(including Zr-, Fe-, and Cu-based BMGs) with various loading modes, for example,

three-point-bending, four-point-bending, tension-tension, and rotating modes [70-82].

However, these results show that the fatigue limits of BMGs spanned from 8% to

50% of the fracture strength of BMGs [83, 84]. Based on these fatigue studies of

BMGs, it is very clear that BMGs exhibit a wide range of fatigue properties. Then, a

question arises: What produced such a large difference among these fatigue results of

BMGs? In fact, according to the fatigue investigation of crystalline alloys, many

factors could be involved, such as the composition, material quality, specimen

geometry, chemical environment, surface condition, temperature, cyclic frequency,

mean stress, and residual stress [83, 84]. Nevertheless, some factors might play a

more important role than others.

In this review paper, the fatigue behavior of BMGs and their composites is

summarized and discussed. In addition, it will also include the metallic-glass films,

ribbons, and wires. Moreover, the detailed discussions will be made to address the

factors affecting the fatigue behavior of BMGs and their composites. The

mechanisms of fatigue-crack initiation, propagation, and fracture of BMGs and their

composites will be outlined, analyzed, and discussed. A fundamental understanding

of the fatigue behavior of BMGs and their composites is of critical significance for

10
designing new BMG systems and developing new manufacturing and processing

techniques so as to broaden the scope of applications of BMGs and their composites.

This review paper is organized in the following ordering: starting with deformation

mechanisms in BMGs and their composites, then high-cycle fatigue studies, fatigue-

crack-growth investigations, and fracture-toughness studies, followed by fatigue-

failure mechanisms and theoretical modeling and lifetime predictions.

2. Deformation Mechanisms in BMGs

Since there are no grains and dislocations in metallic glasses, the plastic-

deformation mechanisms of metallic glasses are different from conventional

crystalline alloys, which usually involve the formation and motion of dislocations. In

general, the deformation behavior of metallic glasses can be classified as being either

homogeneous or inhomogeneous. Homogeneous deformation of a metallic glass is

the uniform deformation of the metallic-glass sample on a macroscopic scale. The

shape and size of the cross sections of the deforming sample change simultaneously

everywhere along the loading axis and remain self-similar. No macroscopic shear

localization is generally observed. In contrast to the homogeneous deformation, the

inhomogeneous deformation is the catastrophic shear failure, characterized by the

formation of localized shear bands, and major shear-band propagation [85, 86].

However, it is found that homogeneous or inhomogeneous deformation mechanisms

for BMGs depend upon the deformation conditions (i.e., applied stresses, strain rates,

and temperatures).

11
The exact nature of the local atomic motion during deformation of metallic

glasses is still being actively studied. In general, when metallic glasses deform, the

fundamental unit process must be a local rearrangement of atoms that can

accommodate shear strains [4]. Spaepen applied the ‘‘free-volume” model to simulate

the plastic flow of metallic glasses [46]. In general, the free volume is a part of the

atom’s nearest neighbor cage where the atoms can move around without an energy

change, as shown in Figure 1(a) [46]. In the metallic glass, the free volume is

distributed statistically among all atoms, since the atoms pack randomly [46]. Thus,

the deformation behavior in the metallic glasses occurs as a series of diffusion-like

local atomic jumps into vacant sites in regions of large free volumes [5]. Later Argon

[50] proposed the shear-transformation-zone (STZ) model. On the atomic scale, shear

strains in metallic glasses under an applied stress are accommodated by local

rearrangements of atoms around free-volume regions, which is different from the

dislocation motion in crystalline alloys. The strain-accommodating local

rearrangement theory involves two modes of thermally-activated shear transformation

[50]: (1) homogeneous and (2) inhomogeneous plastic flows. At high temperatures

(i.e., 0.6Tg < T < Tg, here Tg is the glass-transition temperature), shear transformation

occurs via diffuse rearrangements with small shear strains in spherical regions of 5-

atom diameters. At low temperatures (i.e., 0.6Tg), the shear-transformation behavior

produces a narrow disk-shaped volume element. In fact, the STZ includes a local

cluster of atoms that carry out an inelastic shear distortion from one relatively-low

energy configuration to a second such configuration through an activated

configuration of a higher energy and volume. The applied shear strain continues to

12
increase, when one STZ produces a localized distortion of the surrounding material.

As a result, the large planar bands of STZs form, which are called as shear bands.

Therefore, the plastic deformation of metallic glasses on the macroscopic scale is

essentially a biased accumulation of local strains incurred through the operation of

STZs and the redistribution of free volumes.

Homogeneous deformation often occurs at high temperatures where the

metallic glasses can exhibit significant plasticity [87]. A large tensile ductility was

universally obtained in BMGs for the supercooled-liquid region [87-89]. The research

revealed that the deformation behavior of metallic glasses in the supercooled liquid

region could be Newtonian or non-Newtonian, depending upon the testing

temperature and strain rate. At high temperatures and low applied stresses, the plastic

flow is Newtonian, which means that the strain rate is proportional to the stress.

However, at higher applied stress levels, the flow is non-Newtonian, which

demonstrates that the stress sensitivity of deformation decreases quickly [4]. Metallic

glasses are sensitive to structural instabilities during deformation in the non-

Newtonian regime besides conventional mechanical flow instabilities (i.e., necking)

[4]. The presence of crystallized regions may in fact contribute to the non-Newtonian

character of the rheology. For example, the characteristic transition from the

Newtonian to non-Newtonian flows at high strain rates was observed, when a

Zr52.5Al10Ti5Cu17.9Ni14.6 (at.%) glass, which included the precipitation of nanocrystals,

was deformed [89, 90].

Inhomogeneous deformation usually takes place, when a metallic glass is

deformed at low temperatures (e.g., room temperature). In general, inhomogeneous

13
deformation is characterized by forming localized shear bands, followed by the fast

propagation of these bands, and sudden fracture. These shear bands are approximately

10 to 20 nm in width [38]. Although large plastic strains form in these localized shear

regions, the entire plastic deformation of the specimen is generally very low (~ 2 - 3%)

[36]. It was assumed that the decrease of the viscosity within shear bands resulted in

the localized deformation and inhomogeneous flow. In general, two primary

hypotheses are proposed to explain the plastic flow of localized shear bands in

metallic glasses [91]: (1) the viscosity in shear bands decreases because of the

formation of free volumes [46]; and (2) The viscosity in shear bands drops due to the

generation of local adiabatic heating [92].

It was found that the local temperature in shear bands could exceed Tg, or

even possibly the melting temperature, when a BMG sample fails at room

temperature [36]. The vein patterns and melting marks were generally observed on

fracture surfaces of BMGs, which are consistent with shear softening in the bands.

The local temperature rise resulting from the local heating was estimated to be from

less than 0.1 K to a few thousand K [92, 93]. This large discrepancy in the

temperature evaluation is due to the great difficulty of directly measuring the

temperature in extremely-small distances and short-time scales [91]. Lewandowski

and Greer employed a fusible coating to estimate the temperature in shear bands [93].

During the bending experiment, metallic-glass specimens were coated with a thin

layer of tin. The melting of the coating was observed at the places where shear bands

intersected the surface, as seen in Figure 2. This fact exhibits the direct evidence of

the temperature rise of 200 K in the operating shear bands near the surface. This

14
fusible coating method has some limitation because it only detects a single

temperature per experiment due to the melting point of the coating. However, it can

obtain excellent spatial (~ 100 nm) and temporal (~ 30 ps) resolution [93]. In fact,

additional approximations are still needed to obtain more specific information about

the actual maximum temperature of the operating shear band. The measurements

suggest that remarkable temperature excursions (i.e., rises of thousands of degrees for

a few nanoseconds) occur within a shear band during its operation. Lewandowski and

Greer demonstrate that shear bands cannot be fully adiabatic through making a lower

bound estimate on the shear band propagation time [93].

Yang et al. employed a state-of-the-art infrared (IR) camera to in situ observe

the evolution of shear bands in a Zr-based BMG. They found that multiple nanoscaled

shear bands were arrested in the BMG samples with decreasing temperatures and

shear strains along the lengths prior to the final fracture [Figure 3(a)] [91, 92]. Figure

3(b) showed a three-dimensional (3D) view of the temperature distribution of the “hot

band” on the image of Figure 3(a). Yang et al. found that the temperature was the

highest at the initiation site and the lowest at the end of the shear band. It was found

that the averaged maximum and mean temperature rises were ~ 0.77 and 0.25 K,

respectively, for these observed shear bands [91]. The width of shear bands reported

in the literature is approximately 10 nm. Thus, if we assume that all the heat observed

in a single 0.40-mm “hot band” on the thermograph from the IR camera was

originally generated from a 10-nm-wide shear band, the temperature increase in the

shear band was estimated as high as 650 K, based on the observed average mean

temperature increase of 0.25 K inside the “hot band” from the IR image [91]. In

15
general, there are no work-hardening mechanisms available in amorphous structures.

It is found that the viscosity drops rapidly, and significant softening takes place in

metallic glasses, while the temperature is close to Tg [92]. This fact could cause the

rapid shear-band propagation and catastrophic failure in BMGs. Yang et al. [92]

employed a shear-transformation-zone (STZ) model to carry out a detailed theoretical

analysis on the mechanical work and heat generation within a STZ unit. They found

that a substantial temperature increase inside the shear band could occur by the

collective STZ deformation and report that the calculated shear-band temperatures for

La-based, Mg-based, Pd-based, Cu-based, Zr-based, and Fe-based BMGs were very

close to Tg [92]. This interesting discovery suggests that the catastrophic failure of a

BMG resulted from the sudden drop in the viscosity inside the shear band due to the

temperature increase close to Tg. Based on the free-volume model and a thermo-

viscoplastic model, Gao et al. [94] suggested that the heat conduction introduces a

length scale that helps explain the shear-band spacing. The above studies strongly

suggest that the softening mechanism in metallic glasses is the major factor that

resulted in the deformation of materials. More recent research on the deformation

characteristics of BMGs can be found in [1-5], and will not be reviewed here since

the focus of this paper is on the failure analysis of BMGs.

3. High-Cycle Fatigue Behavior

The safe-life design based on the infinite-life criterion reflects the classic

approach to fatigue, which was initially developed through the 1800s and early 1900s,

due to the increasing number of failures resulting from the machinery produced

16
dynamic loads in the industry revolution. The safe-life, infinite-life design philosophy

was the first to address this need. The design method is stress-life, and a general

representation would be S-N (stress vs. log number of cycles of failure). Failure in S-

N testing is typically defined by the total separation of the sample.

The general applicability of the stress-life method is restricted to

circumstances where continuum, “no crack” assumptions can be applied. The

advantages of this method are the simplicity and ease of applications, and it can

provide some initial perspective on a given situation. It is best applied in or near the

elastic range, addressing constant-amplitude loading situations in what has been

called the long-life regime. The stress-life or S-N approach is principally one of a

safe-life, infinite-life regime. It is generally categorized as a “high-cycle fatigue”

methodology, with most considerations based on maintaining the elastic behavior in

the sample/components examined. The “no cracks” requirement is in place, although

all test results inherently include the influence of the discontinuity population present

in the samples.

3.1. Fatigue Behavior of Metallic Glasses (Ribbons or Wires) in Air at Room

Temperature

Since metallic glasses were discovered, their mechanical properties were

studied extensively. However, metallic glasses could only be fabricated into small

sizes, such as wires or ribbons, before 1990s. In general, these metallic glasses were

found to have high static strengths, consistently high toughness, and good resistance

to chemical and pitting corrosion [67]. In addition, metallic glasses can be

manufactured in continuous lengths, with "finished" cross-sectional dimensions, in a

17
single step, i.e., by continuous rapid quenching from the melt [68]. Therefore,

metallic glasses demonstrated the favorable combination of the high strength,

reasonable cost of constituent elements and economy of fabrication, which indicates

the practical use for structural reinforcement applications. Thus, it becomes of interest

to further characterize the mechanics of these materials, particularly with respect to

their fracture toughness and response to cyclic loading.

Ogura et al. performed tension-tension fatigue tests (mean stress/stress

amplitude = 1) at room temperature in air on the amorphous filament (0.5 ~ 0.7 mm

in width, 20 ~ 25 m in thickness) of a Pd80Si20 alloy [67]. They found from the S-N

characteristics of the Pd80Si20 amorphous alloy: 1) The shape of its S-N curve of the

Pd80Si20 amorphous alloy is very similar to those for crystalline metals or for

polymers; 2) The S-N curve of the Pd80Si20 amorphous alloy exhibits a distinct

fatigue limit of about 400 MPa based on the stress range. The ratio of the fatigue limit

to tensile strength is approximately 0.3; 3) The critical number of cycles at which the

fracture stress become independent of stress cycle is around 4 x 104 cycles.

Davis has performed fatigue studies on Ni49 Fe29P14B6Si2 metallic glasses in a

strip form and Pd77.5Cu6Si16.5 metallic glasses in a wire form [68]. All tests were

conducted in the tension-tension mode. Samples of the Ni49Fe29P14B6Si2 alloy were

tested in "center hole" and "gauged" configurations, while the Pd77.5Cu6Si16.5

amorphous alloy was tested as uniform cross-section filaments. The fatigue limits for

Ni49Fe29P14B6Si2 and Pd77.5Cu6Si16.5 metallic glasses are approximately 320 and 71

MPa, respectively, based on the stress range. Davis found that for the

Ni49Fe29P14B6Si2 strips, failure occurred under partial or full plane-strain crack

18
propagation conditions. However, the final failure by general yielding occurred only

for peak cyclic stresses within a percent or so of the yield stress. On the other hand,

the failure of the Pd77.5Cu6Si16.5 wires always happened due to general yielding.

Verduzco et al. have performed cyclic compressive/tensile bending on Fe-Cr-

Si-B metallic-glass wires using a controlled-strain double pulley testing machine [95].

Two series of glassy-alloy wires, Fe78-xCrxSi10B12 (at.%, 0  x  8) and Fe77.5-

xCrxSi7.5B15 (at.%, 0  x  8) (diameters from 90 to 130 m), were fabricated, using

the rotating water-bath melt-spinning process. They found that the fatigue

performance of the amorphous alloys increases with increasing the Cr content at the

low stress-amplitude level. However, the trend is not so distinct in the range of high-

stress amplitudes. Particularly, the Fe73.5Cr4Si7.5 B15 alloy exhibited better fatigue

resistance than the higher Cr content of the Fe79.5Cr8Si7.5B15 alloy did under high-

stress conditions in the case of the Fe77.5-xCrxSi7.5B15 alloy series. Verduzco et al. [95]

thought that two main reasons could explain why the fatigue performance of the

glassy alloy wires improved with the partial substitution of Fe by Cr: (1) Cr improved

the wire castability, and, therefore, the surface uniformity became better, which

reduced and eventually eliminated periodic necks formed along the wire during

casting. Finally, the consequential stress concentrations were reduced and (2) Cr

resulted in a protective chromic oxide film formed on the surface, which helped

enhance the resistance against the atmospheric corrosion attack. In addition, the

fatigue performance was superior for the Fe77.5-xCrxSi7.5B15 series of alloy wires than

for the Fe78-xCrxSi10B12 series, which probably resulted from the higher intrinsic

strength due to the higher B content.

19
Compared with Pd- and Ni-based metallic glasses, Fe-based metallic glasses

demonstrated long fatigue lifetimes and high fatigue limits, as exhibited in the after-

mentioned Figure 4(a). This fact could be due to that Fe-based metallic glasses

generally exhibited the higher intrinsic strength than Pd- and Ni-based metallic

glasses.

3.2. Fatigue Behavior of BMGs and Their Composites in Air at Room Temperature

After 1990, new types of BMGs were synthesized. Their mechanical

properties were studied extensively. Moreover, many studies of the high-cycle fatigue

behavior have been conducted with Zr-based BMGs and involved various test

configurations: bending, uniaxial, and rotating-fatigue conditions.

3.2.1. Bending Fatigue

Due to convenient sample mounting and dismounting, relatively-simple

sample preparation, and a uniform maximum tensile stress on the specimen surface,

which makes it possible to test brittle materials in tension without special sample

gripping [96], the bending or flexural fatigue setup has widely been employed for

assessing the fatigue resistance of engineering materials. For four- and three-point

bending setups, as shown in [97], the nominal maximum tensile stress, , within the

bending sample can be calculated from the beam theory, using the following

equation:

, (3-1)

20
where P is the applied load, b is the specimen width, t is the specimen thickness, S1 is

the inner span (zero for the three-point bending setup), and S2 is the outer span.

Since BMGs are brittle alloys, and their sizes are limited so far, three- and

four-point bending-fatigue studies are employed widely to perform fatigue

investigations on BMGs. The first high-cycle fatigue results of BMGs was reported

based on 4-point-bending experiments [69]. Gilbert et al. employed a setup with the

inner span of 10.2 mm and outer span of 20.3 mm to conduct fatigue studies on the

Vitreloy 1 alloy. However, the fatigue limit reported was as low as approximately 150

MPa, based on the stress range. In addition, Menzel and Dauskardt [71] conducted

four-point-bending fatigue on Vitreloy 1 alloys using a different setup that increased

the outer span from 20.3 mm to 30 mm. They reported similar fatigue results, and the

fatigue limit was about 190 MPa. However, Launey et al. performed fatigue

experiments using a four-point-bending setup with the inner span of 30 mm and outer

span of 60 mm on annealed Vitreloy 1 specimens [98]. The specimens that were

annealed for 73 min. at 610 K exhibited higher fatigue limits (550 MPa) than those

that were annealed for 2 min. at 573 K (390 MPa), which was due to the reduction in

the free volume, which made the plastic deformation more difficult [98]. Moreover,

they carried out 3-point bending experiments on as-cast and anneal-relaxed Vitreloy 1

samples and reported that the relaxed alloys exhibited higher endurance limits than

the as-cast alloys [99].

Morrison et al. studied the four-point-bending fatigue on the

Zr52.5Cu17.9Ni14.6 Al10.0Ti5.0 (at.%) BMG alloy, which has a high fatigue limit of 850

MPa [100]. They showed that the assumption of a constant stress between the inner

21
span is only accurate for a particular range of fixture and specimen geometries. For

example, an outer-span/inner-span ratio is between 4 and 5 [96]. In addition, Wang et

al. investigated four-point-bending fatigue on Zr50Cu40Al10, Zr50Cu30Al10Ni10, and

Zr50Cu37Al10Pd3 (at.%), which possess the fatigue limits ranging from 540 MPa to

640 MPa. Moreover, the fatigue behavior of the Zr 50Cu37Al10Pd3 BMG was studied,

based on a three-point-bending setup, which exhibited a fatigue-endurance limit of

631 MPa [101]. Recently, El-Shabasya and Lewandowski performed fatigue coaxing

experiments on a Zr-based bulk-metallic glass (Vitreloy 1) under three-point bending

at room temperature [102]. The coaxing effect in fatigue refers to the improvement of

the high-cycle-fatigue strength by the application of a gradually-increasing stress

amplitude, usually starting below the fatigue limit. They reported that the fatigue limit

is around 200 MPa, based on the stress range. They found that a fatigue coaxing cycle

often increased the fatigue limit (at least double the value of the fatigue limit

according to the conventional fatigue test), as shown in Figure 5 [102]. However,

annealed (but not crystallized) samples, which were tested at the same condition, also

showed some benefits to the fatigue behavior. This fact suggested that the prevention

of fatigue-crack nucleation via shear banding may provide some benefits to the S-N

behavior, whether this trend arises due to fatigue coaxing or via annealing. In order to

determine the source(s) of the improved fatigue limit in the coaxed samples, more

study is needed.

Moreover, four-point-bending fatigue has been employed to study an

amorphous steel, Fe48Cr15Mo14Er2C15B6 (at.%) [81] and Cu-based BMG,

(Cu60Zr30Ti10)99Sn1 (at.%) [103]. A setup of an inner span of 10 mm and outer span of

22
20 mm was used for the four-point-bending fatigue experiments. As a result,

Fe48Cr15Mo14Er2C15B6 showed a fatigue-endurance limit of 682 MPa.

Fe48Cr15Mo14Er2C15B6 demonstrated a brittle fracture mode, which shows no any

characteristic vein patterns visible on the fracture surface. Moreover, the stress-

fatigue life curve revealed that the lifetime of the amorphous steel decreased abruptly

with increasing the applied stress [81]. Freels et al. [103] reported that the

(Cu60Zr30Ti10)99Sn1 BMG had a fatigue limit of 350 MPa, which is lower than the

amorphous steel.

The three- or four-point-bending-fatigue results of BMGs from the literature

are summarized in Tables 1-3, and Figure 6. Figure 6(a) and (b) display the S-N

curves of the Zr-based BMGs and other-based BMG alloys, respectively. Based on

the three- and four-point-bending fatigue experiments, the fatigue limits of Zr-based

BMGs range from 150 to 850 MPa, while the fatigue-limit range of other-based

BMGs vary from 200 to 1,250 MPa. Here, the fatigue limit is based on the applied

stress range (maximum stress - minimum stress). In Figure 6(a), the S-N curves of Zr-

based BMGs can be divided into two regions, named as the low fatigue-endurance

limit and high fatigue-endurance limit regions, respectively. In the present work, the

critical fatigue-endurance limit of BMGs is ~ 600 MPa, as marked in Figure 6(a),

above which the BMG alloys will be defined as high fatigue-endurance limit

materials. There are four BMG alloys located in the low fatigue-endurance limit

region, including Zr41.2Ti13.8Cu12.5Ni10Be22.5 [69], Zr44Ti11Ni10Cu10Be25 [98],

Zr60.14Cu22.3Fe4.85Al9.7 Ag3 [104], and Zr52.5Cu17.9Ni14.6Al10Ti5 [105], while in the high

fatigue-endurance limit region, three alloys are involved: Zr50Cu40Al10 [106],

23
Zr50Cu37Al10Pd3 [106], and Zr50Cu30Al10Ni10 [106]. Moreover, it is interesting to

notice that in the low fatigue-endurance limit region, all alloy systems have five

elements. However, in the high fatigue-endurance limit region, the alloy systems have

four or three elements.

In Figure 6(b), the S-N curves of other-based BMGs can be divided into three

regions: low, high, and extra high fatigue-endurance limit regions. Note that BMGs

here exclude Zr-based BMG alloys. The critical fatigue-endurance limits of BMGs

are 600 and 1,000 MPa, respectively, as marked in Figure 6(b). For the alloys with

the fatigue-endurance limits above 1,000 MPa, the BMG alloys will be defined as an

extra high fatigue-endurance limit materials; below 600 MPa, the BMG alloys will be

defined as low fatigue-endurance limit materials; while for the alloys with fatigue-

endurance limits between 600 and 1,000 MPa, they are classified as high fatigue-

endurance limit materials. In Figure 6(b), four BMG alloys are located in the extra

low fatigue-endurance limit region, including Cu47.5Zr47.5Al5 [82], (Cu60Zr30Ti10)99Sn1

[103], Cu60Zr30Ti10 [107], Cu47.5Zr38Hf9.5 Al5 [82], and Cu45Zr45Ag7Al3 (at.%) [108].

Three alloys are located in the high fatigue-endurance limit region, including

Fe48Cr15Mo14Er2C15B6 (amorphous steel) [81], Ti41.5Zr2.5Hf5Cu42.5Ni7.5Si1 [107], and

Ni60Zr20Nb15Al5 (at.%) [107]. Two alloys are located in the high fatigue-endurance

limit region, including (Fe50Co50)72B20Si4Nb4 [107], [(Co60Fe40)75B20Si5]96Nb4 (at.%)

[107]. It can be observed that all the Cu-based BMG alloys are located in the low

fatigue-endurance limit region. However, the amorphous steel, Ti-based, and Ni-

based BMG alloys are located in the high fatigue-endurance limit region, while the

Co-based and Fe-based BMGs are in the extra high fatigue-endurance limit region.

24
Besides monolithic BMG alloys, the fatigue behavior of metallic-glass-matrix

composites (MGMCs) are also reported, mainly employing the four-point-bending

fatigue tests, as summarized in Figure 4(b). It can be observed that the fatigue limits

of MGMCs vary from 100 to 1,200 MPa. Flores et al. first reported the four-point-

bending-fatigue results regarding a Zr-based BMG composite, Zr56.2Ti13.8Nb5.0Cu6.9

Ni5.6Be12.5 (at.%) [109], based on a setup of an inner span of 10.3 mm and an outer

span of 20 mm. The studied composite was based on the Vitreloy 1 alloy and

contained a relatively-fine dispersion of dendrites within the glass matrix [109]. The

BMG composite exhibited a higher fatigue-endurance limit (296 MPa) than the

Vitreloy 1 alloy [69]. In two separate works of applying four-point-bending fatigue

tests, Qiao et al. reported that a nano-particle (~ 5 nm) dispersed MGMC with the

composition of (Zr58Ni13.6Cu18Al10.4)99Nb1 [110] and a Zr58.5Ti14.3Nb5.2Cu6.1Ni4.9Be11

[111] MGMC with randomly-distributed crystalline dendrites have a fatigue-

endurance limit of 559 MPa and 567 MPa, respectively. El-Shabasy et al. reported the

high-cycle-fatigue behavior of a nanostructured composite produced via the extrusion

of amorphous Al89Gd7Ni3 Fe1 alloy powders at different extrusion ratios (ER = 5:1,

10:1, and 20:1) under three-point bending at a stress ratio R = 0.1 [112]. They found

that fatigue behavior of the Al89Gd7Ni3 Fe1 composites, as well as hardness and bend

strength, could be improved with increasing the extrusion ratio. The four-point-

bending fatigue tests were also conducted on a dendritic-inclusion ( phase)-

dispersed MGMC (Zr39.6Ti33.9Nb7.6Cu6.4Be12.5 in the composition and termed DH3),

and an enhanced fatigue-endurance limit was noticed [113]. Figure 7(a) compares the

stress-life (S-N) fatigue data of the DH3 alloy with another MGMC termed LM2

25
(Zr56.2Ti13.8Nb5.0Cu6.9Ni5.6Be12.5), monolithic metallic glasses , and traditional alloys

[113]. It is seen that both the fatigue lives and fatigue limit of the DH3 alloy are

remarkably higher than those of the LM2 alloy [109], the monolithic Vitreloy1 [69,

71], and the metallic glass ribbon [67, 68, 114]. Indeed, its fatigue strength is

sufficiently high to be comparable to that of the 300-M high-strength steel [115]and

the 2090-T81 aluminum alloy [116]. The remarkably excellent fatigue properties of

the DH3 alloy stems from an exquisite microstructural design, i.e., utilizing the

semisolid processing to optimize the distribution of second-phase dendrites so that the

microstructural length scale (interdendritic spacing) can match the crack length scale.

By doing so, shear bands are confined within interdendritic regions (Figure 7(c)) and

a wide distribution of damage around the crack tip is promoted (Figure 7(b)), thereby

boosting the alloy’s fatigue lifetime and fatigue strength extraordinarily. Recently, the

fatigue behavior of the MGMC, Zr48Cu47.5Co0.5Al4, is found to be much improved

than the reported results on the MGMCs, as shown in Figure 4(b). It can be observed

that the endurance limit of Zr48Cu47.5Co0.5Al4 is ~ 1,200 MPa located in the extra high

fatigue-endurance limit zone, while the other reported MGMCs are all in the low

fatigue-endurance limit zone [84], which should be attributed to the stress-induced

martensitic transformation of crystalline inclusions from B2-CuZr to B19’-ZrCu

[117]. Moreover, the fatigue results of other kinds of BMGs are reported, including

ribbons, wires, and foams, as summarized in Figure 4(a), which can also be separated

into two regions: low and high fatigue-endurance limit regions, with the critical stress

of 600 MPa. It can be observed that the Fe-based wire and Pd-based foam have high

26
fatigue-endurance limits, while the Fe-based ribbon and Zr-based shot-peened BMG

alloys have low fatigue-endurance limits.

Note that the test volume in the four-point-bending condition is generally

larger, compared to the three-point-bending condition. Because the larger test volume

could contain more defects, stress raisers, and free volumes, which will increase the

possibilities for the shear-band formation and crack initiation, three-point-bending

fatigue usually produces higher fatigue limits than four-point-bending fatigue in the

same condition [103]. Based on these bending-fatigue results from literature, it is

found that different compositions, testing conditions, and fabrication procedures

could affect the fatigue behavior of BMGs.

3.2.2. Uniaxial Fatigue

Uniaxial-fatigue experiments usually involve the tension-tension, tension-

compression, and compression-compression fatigue. The fatigue-loading mode could

influence the fatigue behavior of BMGs and their composites. Under compression–

compression loading, BMGs generally demonstrate better fatigue resistance than

them under tension-tension loading. The tensile stresses could be more effective in

growing the crack of BMGs than the compressive stresses. In general, the frequencies

from 0.13 Hz to 20 Hz were employed in the uniaxial fatigue experiments of BMGs

and their composites so far. Cylindrical samples were generally used for uniaxial

fatigue testing of BMGs and their composites. The S-N curves from the uniaxial

fatigue of various BMGs and their composite are summarized and plotted together for

comparison in Figure 8, and the fatigue results are listed in Table 3.

27
Based on the literature results summarized in Figure 8, the fatigue limits of

BMGs under the uniaxial loading vary from approximately 140 to 1,000 MPa. Here,

the uniaxial-fatigue data of BMGs can be divided into two regions: low fatigue-

endurance limit and high fatigue-endurance-limit regions. Above the fatigue-

endurance limit of 600 MPa in the S-N curves, the BMG alloys are defined as the

high fatigue-endurance limit materials. It can be observed that, in Figure 8, the

Zr50Cu40Al10 and Zr50Cu37Al10Pd3 are in the high fatigue-endurance-limit region,

while the BMG alloys, Zr76.6Cu12.3Al3.5Ni7.6, Zr56.2Ti13.8Nb5.0Cu6.9Ni5.6Be12.5,

Zr55Cu30Ni5 Al10, and Zr50Cu39Al10Pd1, are in the low fatigue-endurance-limit region.

Therefore, the fatigue lifetimes of BMGs under a uniaxial-loading condition exhibit a

very wide range, which can be significantly dependent on the specimen composition,

fabrication process, and loading condition, etc.

Fatigue studies under uniaxial loading have been extensively performed on the

Vitreloy 1 BMGs alloy in the literature [76, 77, 79, 118, 119]. Both notched and

tapered Vitreloy 1 samples were employed to carry out tension-tension fatigue

investigations with a similar specimen geometry, which revealed that the fatigue

lifetimes were inversely related to the oxygen content, and fatigue cracks generally

initiated from inclusions or porosity in the specimens [76, 77]. Moreover, tension-

compression and compression-compression uniaxial fatigue results of Vitreloy 1 were

also reported [118], which displayed longer fatigue lifetimes and higher endurance

limits for the compression-compression fatigue, compared with tension-tension and

tension-compression fatigue. Although the tension-compression fatigue showed a

lower endurance limit, as compared to tension-tension fatigue, both of these testing

28
configurations had comparable fatigue lifetimes at high stress levels. Based upon

these studies, it is found that the tensile stresses are the driving force of the most

damaging fatigue processes in BMGs [118]. Compressive stresses can cause the

increased stress concentration at the initiation site during the tension-compression

fatigue loading. Therefore, the fatigue crack in the tension-compression fatigue

initiates more easily than that in tension-tension fatigue and thus, a low endurance

limit in the tension-compression fatigue could occur.

Moreover, tension-tension and compression-compression fatigue studies were

reported on Zr50Cu37Al10Pd3 BMGs, as listed in Table 3 [79, 119]. The BMG alloy,

Zr50Cu37Al10Pd3, under compression-compression loading exhibited longer lifetimes

than those under tension-tension loading, which is similar to the reported results of

the Vitreloy 1 alloys [118]. However, the endurance limits of the Zr50Cu37Al10Pd3

BMGs under both loading conditions are comparable, while Vitreloy 1 showed a

higher endurance limit under compression-compression loading than that under

tension-tension loading. In general, the failure mode of BMGs under compression-

compression loading is similar to the failure mode under monotonic-compressive

loading, which usually proceeds by the unstable fracture along a single shear band

[118]. However, the fracture surface under tension-tension fatigue is perpendicular to

the loading direction [72-79], while compression-compression fatigue exhibits the

fracture surface with an angle of approximately 45 to the loading direction [118,

119]. No fatigue striations are generally found on the fractured surface after the

compression-compression fatigue [118, 119]. However, fatigue striations are clearly

observed on the fracture surface after tension-tension fatigue, which indicates that

29
stable fatigue-crack growth occurs [74-79]. Therefore, tension loading is the essential

driving force to produce striations in BMGs.

Tension-tension fatigue studies on Zr52.5Cu17.9Al10Ni14.6Ti5 have been carried

out on BMGs in different labs [72, 120]. Peter et al. [72] performed tension-tension

fatigue experiments using notched cylindrical samples with a sinusoidal wave and a

frequency of 10 Hz. Zhang et al. [120] conducted tension-tension fatigue experiments

on plate samples using a triangle wave with a frequency of 1 Hz. The fatigue results

reported by Peter et al. exhibited much longer lifetimes than those reported by Zhang

et al. Moreover, the tension-tension-fatigue behavior was investigated on notched

cylindrical samples of Zr50Cu40Al10 and Zr50Cu30Al10Ni10 BMGs using a sinusoidal

wave and a frequency of 10 Hz. However, these alloys exhibited better fatigue

resistance than Vitreloy 1 [74, 75]. In addition, tension-tension fatigue experiments

were conducted on plate samples of Zr55Cu30Al10 Ni5 [121] and Zr766Cu12.3Al3.5Ni7.6

[122] BMGs, and their fatigue results indicated poor fatigue behavior for these

BMGs, as illustrated in Figure 8.

Moreover, Freels et al. has investigated the cyclic-compression behavior of a

Cu45Zr45Al5 Ag5 BMG at frequencies of 10 and 40 Hz [123]. They used cylindrical

samples with a diameter of 2.95 mm and an aspect ratio of 2. Here, the aspect ratio is

defined as the values between the length and diameter of a specimen. A relatively-

high fatigue-endurance limit (1,418 MPa) and fatigue ratio (0.77) were reported,

where the fatigue ratio is defined as the fatigue-endurance limit to the static tensile

strength of a material. In general, it is found that the fracture under cyclic

compression occurred in a pure shear mode, and the fracture surface forms an angle

30
of 41 with respect to the loading axis, which was similar to the monotonic-

compressive fracture angle for the Cu45Zr45Al5Ag5 (at.%) BMG [123]. The fracture

surface under cyclic-compression loading demonstrates a morphology nearly identical

to the monotonic-compression fracture surface, which involved three distinct patterns

[123]: (1) vein-like regions, (2) river-like regions, and (3) intermittent smooth

regions. They thought that each pattern is developed in a similar manner to that under

monotonic compression. Besides many shear bands and cracks, areas of “chipping”

were generally observed on the outside surfaces of the fatigue specimens, which were

not found on the outside surfaces of the monotonic-compression samples.

Compression-compression fatigue experiments have been conducted on Ca-

based BMG alloys with a geometry of 4 × 4 × 4 mm3 [124]. The Ca-based BMGs

exhibit the unique properties, such as low density, low Young’s modulus, which is

comparable to the modulus of human bones, low glass-transition temperature, and a

wide super-cooled liquid temperature range [124]. The Ca-Mg-Zn-based alloys have

great potential for use in biomedical applications because Ca, Mg, and Zn are

biocompatible. The fatigue lifetime of the Ca65Mg15Zn20 (at.%) BMG alloy generally

increased from 104 to 106 cycles, when the maximum applied stress decreased from

about 240 to 140 MPa, and its fatigue limit is about 140 MPa at 106 cycles [124].

Compared to the Zr-based BMGs, the Ca-based BMGs exhibit shorter lifetimes and

lower endurance limits. Moreover, different from the Zr-based BMGs, the

Ca65Mg15Zn20 BMG samples shattered into very small fragments after the final

fracture. Fatigue–fracture surfaces demonstrated a mixture of fracture features similar

to compression experiments. These results indicate that the fatigue-fracture behavior

31
of the Ca-based BMGs under compression-compression loading is more complex

than that of the Zr-based BMGs where the fracture occurs only along specific shear

planes. Besides BMGs, the uniaxial-fatigue behavior of BMG composites has been

studied as well. Fujita et al. [125] investigated tension-tension fatigue on

nanocrystalline Zr55Cu30Al10Ni5 BMGs. In addition, Wang et al. [77] carried out

tension-tension-fatigue studies on the Zr56.2Ti13.8Nb5.0Cu6.9Ni5.6 Be12.5 BMG

composite, and found that the fatigue resistance of the composite was not better than

that of BMGs. Wei et al. [126] conducted compression-compression fatigue tests on

the Cu47.5Zr48Al4Co0.5 BMG composite, and reported that the composite exhibited a

lower fatigue limit but higher fatigue ratio in comparison with the Cu 46.5Zr46.5Al7

(at.%) BMG. The composite’s inferior fatigue limit was attributed to its lower yield

strength, and its higher fatigue ratio was thought to be induced by the “blocking

effect” and “martensitic transformation effect” of the B2 crystalline phase, which

could effective impede the propagation of fatigue cracks in the BMG composite

[126].

Based on the comparison of the uniaxial fatigue behavior of the Zr-based

BMGs, it is found that the fatigue-loading mode, sample geometry, material quality,

BMGs composition, and microstructure could have important influence on the fatigue

behavior. Usually BMGs under compression-compression loading have excellent

fatigue resistance. The fatigue lifetimes under the Mode-II fracture are much longer

than those under the Mode-I fracture. Therefore, the crack growth in BMGs could be

driven by tensile stresses. Compared to BMGs, BMG composites generally exhibit

poor fatigue behavior.

32
3.2.3. Rotating Fatigue

The rotating-beam-fatigue experiments are similar to four-point-bending

fatigue tests, which have a constant-bending moment between two bending points.

However, the sample is rotating, and the dead weight is fixed. Yokoyama et al. [80,

127] studied the fatigue behavior of several Zr-based BMGs using rotating-fatigue

experiments with hour-glass-shaped-bar specimens. The applied frequency of the

rotating fatigue is 50 Hz. They found that the Zr50Cu37Al10Pd3 alloy exhibited the

highest endurance limit (shown in Figure 9 and Table 4), and the Zr50Cu40Al10 alloy

exhibited the lowest endurance limit. In general, Zr50Cu30Al10Ni10, Zr50Cu39Al10Pd1,

Zr50Cu38Al10Pd2, Zr50Cu35Al10Pd5, and Zr50Cu37Al10Pd7 (at.%) showed comparable,

middle-range endurance limits (Figure 9). The improvement in the fatigue

performance of the Zr-based BMGs as the compositions changed was interpreted

through a linear relationship between the endurance limit and the volume change

from the as-cast state to the structurally-relaxed state at Tg, as shown in Figure 9(b)

[80]. Specifically, since a large volume change in a BMG sample usually implies a

large amount of free volumes contained in the sample, Figure 9(b) tells us that the

magnitude of the endurance limit of a given composition of the Zr-based BMGs is

governed by the amount of free volumes it holds. A large amount of free volumes in

BMGs usually creates many structural heterogeneities and results in better ductility.

The fatigue properties, therefore, are enhanced [80].

Moreover, high-cycle fatigue of the Pd40Cu30Ni10 P20 BMG was investigated,

using the rotating-fatigue experiments with an applied frequency of 60 Hz [128].

Yokoyama et al. claimed that no fatigue limit was found in the S-N curve for the

33
Pd40Cu30Ni10P20 BMG [128]. They thought that the fatigue is sensitive to the

structural defects of inclusions and micro pores. The inclusions act as the fatigue-

crack-initiation rate, and the micro-pores cause the embrittlement of the

Pd40Cu30Ni10P20 (at.%) BMG. The fatigue lifetime of the Pd40Cu30Ni10P20 BMG is

comparable to that of the Zr50Cu37Al10Pd7 BMG, as presented in Figure 9 [80, 127].

3.3. Fatigue Behavior in Specific Environments

Since BMGs exhibit good corrosion resistance and bio compatibility, they can

be considered as potential structural and biomedical materials [129-131]. Therefore,

BMGs can be used not only in air but also in other environments. Thus, the fatigue

behavior in various environments is important to be characterized. Some researchers

have performed fatigue studies on BMGs in vacuum and corrosive environments [72,

78, 122, 132]. Moreover, a smaller number of fatigue studies have been performed in

biological environments.

3.3.1. Fatigue in Vacuum

The tension-tension fatigue of the Zr52.5Cu17.9Al10Ni14.6Ti5 (at.%) BMG has

been investigated in vacuum [72]. It was reported that the fatigue lifetimes were

lower in vacuum than in air, which could be due to the hydrogen embrittlement. An

ionization gauge with a hot tungsten filament was on during the fatigue tests in

vacuum. Therefore, the dissociation of the residual water vapor at the hot tungsten

filament of the ionization gauge could result in the hydrogen embrittlement of the

Zr52.5Cu17.9Al10Ni14.6Ti5 (at.%) BMG. Therefore, the authors indicated that an

environmental effect resulted in the lower fatigue lifetimes in vacuum than in air [72].

34
Moreover, Wang et al. [78] performed the fatigue-behavior studies on Zr50Cu40Al10,

Zr50Cu30Al10Ni10, and Zr50Cu37Al10Pd3 BMGs in vacuum and reported that no

difference of the fatigue behavior in vacuum and in air for these BMG alloys was

found [78]. Therefore, these results may suggest that the relative humidity in the lab

is generally not detrimental to the fatigue-lifetime performance of BMGs.

3.3.2. Corrosion Fatigue

Corrosion fatigue is the damage or failure process, when a material is

subjected to the interaction of the corrosion environment and fluctuating stresses. The

corrosion-fatigue behavior is of critical importance for structural materials used in

corrosive environments. In the literature, the corrosion fatigue has been studied in the

phosphate-buffered saline (PBS) electrolyte and NaCl solutions.

Maruyama et al. [122] reported S-N curves based on the uniaxial-fatigue tests

with plate specimens of the Zr65Cu15Ni10Al10 BMG alloy in air and in a phosphate-

buffered saline (PBS) electrolyte with a physiologically-relevant oxygen content.

They did not observe any difference between the S-N curves in the air and PBS

environments. Moreover, they found that the fatigue-fracture surfaces in both cases

were similar, and the fatigue crack generally initiated from defects or impurity

particles near the sample surface. In addition, Huang et al. [132] studied fatigue

behavior of the (Zr0.55Al0.10Cu0.30Ni0.05)99Y1 (at.%) BMG in the PBS solution at 37 ºC.

They conducted four-point-bending corrosion-fatigue experiments in a

physiologically-relevant environment, and the results were compared with those

obtained in air at room temperature. It was found that the corrosive environment did

not significantly affect the fatigue lifetime at high stress levels, while the corrosive

35
environment had a detrimental effect on the fatigue resistance at low stress levels.

Moreover, the fatigue limit was decreased by 40% in the physiologically-relevant

environment, compared with those in air environments. In general, at low stress

levels, environmental degradation influenced crack initiation, with the assistance of

the cyclic force and/or the abrasion of loading pins. Furthermore, crack propagation

was affected even more by the environment. The corrosion-fatigue degradation

mechanism for the (Zr0.55Al0.10Ni0.05Cu0.30)99Y1 BMG in phosphate-buffered saline

(PBS) (37 °C) was determined to be anodic dissolution [132].

The Zr52.5Cu17.9Ni14.6 Al10Ti5 BMG generally has good fatigue resistance in

both uniaxial [72, 73] and four-point-bending configurations [100]. Moreover, it also

exhibits excellent corrosion resistance in both 0.6 M NaCl (3.5 wt.% NaCl) and PBS

environments [33]. Morrison et al. [33] conducted the four-point-bending-fatigue

experiments on the Zr52.5Cu17.9Ni14.6Al10Ti5 BMG alloy in a 0.6 M NaCl electrolyte.

Comparing the results with those in air, the environment had an increasingly-

deleterious effect on the fatigue lives, as the applied stress decreased due to the

increasing exposure time to the degradative environment. It was found that the

fatigue-endurance limit in the 0.6 M NaCl electrolyte reduced approximately 88%

than that tested in air (Figure 10) [133]. Three kinds of fracture morphologies on the

Zr52.5Cu17.9Ni14.6 Al10Ti5 (at.%) BMG fracture surface were observed: (1) mixed

morphologies of small areas with fatigue striations and large areas with the typical

vein pattern; (2) fracture morphologies with alternating areas of smooth steps

separated by abrupt changes; and (3) typical tension-tension fatigue-fracture

morphologies containing a crack-growth region with striations, a transition region,

36
and a fast-fracture region [133]. Although multiple pits were found near the fracture

surfaces, there was no clear evidence, which showed that the crack initiated from

these pits. Morrison et al. thought that the degradation mechanism for the

Zr52.5Cu17.9Ni14.6 Al10Ti5 (at.%) BMG alloy was the anodic dissolution of the alloy

instead of the hydrogen embrittlement [33]. In general, the anodic dissolution (stress-

assisted dissolution) happens in materials that form passive films in a corrosive

environment. The localized plastic deformation at the crack-initiation site or crack tip

resulting from the cyclic stresses causes the rupture of the passive film. As a result,

the newly-exposed bare metal because of the rupture of the passive film serves as the

anode. Nevertheless, the unbroken passive film will serve as the cathode in an

electrochemical circuit. Therefore, the crack propagates in a corrosive condition due

to the continuous anodic dissolution of the newly-exposed bare metal.

The Vitreloy-type BMG compositions with low atomic fractions of late

transition metals (LTMs) were found to exhibit a combination of exceptionally large

supercooled liquid region and good glass-forming ability. Due to the low LTM

atomic fractions, these compositions were thought to exhibit good corrosion

characteristics [134]. The corrosion and corrosion-fatigue behavior of Zr35Ti30Be35

and Zr35Ti30Be29Co6 (at.%) has been investigated and compared to the traditional

Vitreloy glass, Zr52.5Cu17.9Ni14.6 Al10Ti5 (at.%), and to other crystalline engineering

alloys used widely in saline environments, such as the 18/8 stainless steel, Alclad

24S-T, and annealed Monel [134]. It was found that the low-LTM Vitreloy glasses

exhibited corrosion rates of less than 1 m/year, which are lower by more than one

order of magnitude, compared to the traditional Vitreloy glass and the conventional

37
engineering metals. Wiest et al. suggested that the high corrosion resistance of

Zr35Ti30Be35 and Zr35Ti30Be29Co6 (at.%) was attributed to the low fraction or

complete absence of LTM elements, facilitating the formation of a chemically-

homogeneous passive layer without “weak spots” [134]. However, the corrosion-

fatigue performance of Zr35Ti30Be29Co6 and Zr35Ti30Be35 (at.%) was poor, as less than

10% of their yield strength is retained at 107 cycles, which was comparable to

traditional Vitreloy glasses but significantly lower than conventional crystalline

alloys. Wiest et al. thought that the poor corrosion performance is probably due to a

retarded reformation of the passive layer at the extending crack tip, possibly caused

by their low pitting potential [134].

Based on these fatigue studies in specific environments, it could be concluded

that the alloy composition and the environment can result in drastically-different

corrosion-fatigue behaviors of BMGs.

4. Fatigue-Crack-Propagation Behavior

The fatigue failure usually includes two processes: fatigue-crack initiation and

fatigue-crack growth. Therefore, the fatigue life is generally determined by the time

to initiate a crack and the time to grow the crack till the critical size. Thus, the

fatigue-crack-propagation behavior of BMGs is very important to predict and

evaluate the fatigue behavior of BMGs. Some literature results regarding the fatigue-

crack-propagation behavior of BMGs are summarized here.

4.1. Ambient Conditions

38
The studies on the fatigue-crack-growth behavior of BMGs have focused on

the Vitreloy 1 BMG at the beginning. Gilbert et al. [133] first conducted the fatigue-

crack-propagation experiments on the 7-mm-thick, 38-mm-wide compact-tension

[C(T)] Vitreloy 1 BMG specimens and reported its fatigue-crack-growth-rate results.

It was found that the fatigue-crack-growth rates and the fatigue-threshold-stress-

intensity-factor range (K, where K = Kmax. - Kmin., Kmax. is the maximum stress

intensity, and Kmin. is the minimum stress intensity) in the Vitreloy 1 alloys were

comparable to ductile crystalline alloys [70]. Moreover, Schroeder et al. [135] and

Flores et al. [109] employed C(T) specimens to investigate the fatigue-crack-

propagation behavior of Vitreloy 1 as well. Later, Zhang et al. [136] performed the

fatigue-crack-propagation tests on Vitreloy 1 alloys using single-edge-notched-beam

samples. Schroeder et al., Flores et al., and Zhang et al. reported the similar fatigue-

crack-propagation results to those by Gilbert et al. However, the fatigue-crack-growth

rates that these researchers reported were slightly faster than those reported by Gilbert

et al. [133]

Moreover, Launey et al. conducted the fatigue-crack-propagation experiments

on the Zr44Ti11Ni10Cu10Be25 (at.%) samples [98]. Nakai et al. [137] studied the

fatigue-crack-growth behavior in the Zr55Cu30Al10Ni5 (at.%) BMG. Philo et al.

recently performed the fatigue-crack-growth tests on Zr58.5Cu15.6Ni12.8Al10.3Nb2.8

(at.%) BMG [138]. Launey et al., Nakai et al., and Philo et al. reported similar results

to those of the Vitreloy 1 alloy. In addition, fatigue-crack-propagation behavior in a

nano-crystalline Zr55Cu30Al10Ni5 (at.%) BMG [139] and a

Zr56.2Ti13.8Nb5.0Cu6.9Ni5.6Be12.5 (at.%) BMG composite [109] were studied. These

39
results indicated that the studied composites exhibit similar fatigue-crack-growth

behaviors to the fully-amorphous BMGs. The variation in fatigue-crack-propagation

rates (da/dN) of several Zr-based BMGs as a function of the stress-intensity-factor

range (K) is shown in Figure 11. The fatigue-crack-growth results of a metallic

glass, Ni78Si10B12 (at.%), [140] are also plotted for comparison. In general, metallic

glasses exhibited a small fatigue threshold and a large steady crack-growth range.

Some BMG composites with the delicately-designed microstructure, however, may

exhibit considerably enhanced fatigue-crack-growth resistance than monolithic

BMGs. Such examples include the Zr36.6Ti31.4Nb7Cu5.9Be19.1, Zr25Ti43Nb7Cu6Be19,

and Zr43Ti25Nb7Cu6Be19 composites with dendritic second-phases [141]. In these

composites, the microstructural length scale (the dendrite arm spacing and

interdendrite length) was carefully designed through semisolid processing so that

shear bands and propagating cracks could be arrested within interdendritic regions

during deformation. As a result, these composites exhibited a fatigue threshold stress-

intensity-factor range (K0) of 5.0 – 5.7 MPa m1/2 [141], which is more than three

times higher than the Vitreloy 1 (~ 1.5 MPa m1/2) [136]. In addition, these composites

possess Paris exponent values about twice higher than that of the Vitreloy 1, i.e., 2.4 –

3.5 vs. 1.5 [136].

4.2. Other Environments

Hess et al. [142] investigated the fatigue-crack-growth behavior of the

Vitreloy 1 alloy at elevated temperatures below Tg. The authors reported that there

were no differences in the fatigue-crack-growth rates in the temperature range of 25

40
to 220C. However, the value of Kth was found to increase as the testing temperature

was increased from 100 to 220C, as seen in Table 5 [142]. In addition, the values of

the effective-threshold-stress-intensity-factor range, (= Kmax - Kcl at the

fatigue threshold, where Kcl is the stress intensity determined by the crack-closure

load), for this BMG at elevated temperatures are slightly less than those Kth, and

they have the same trend with the increase of temperature.

Schroeder et al. [143] examined the fatigue-crack-growth behavior of the

Zr41.2Cu12.5Ni10Ti13.8Be22.5 (at.%) alloy at room temperature in several environments:

aerated, deionized water; aerated 0.5 M NaClO 4; aerated 0.5 M Na2SO4; aerated

0.005 M NaCl; 0.05 M NaCl; and the 0.5 M NaCl. It was found that the deionized

water resulted in a slight increase in the fatigue-crack-growth rates, compared to the

behavior in air (Figure 12). On the other hand, the fatigue-crack-growth rates in the

aqueous 0.5 M NaCl solution were as much as three orders of magnitude faster than

those found in air [143]. Moreover, comparing these experiments in 0.5, 0.05, and

0.005 M NaCl solutions, the steady-state fatigue-crack-growth rates were

proportional to the concentration of NaCl in the solution, and the Kth values

remained similar. Finally, the fatigue-crack-propagation rates in 0.5 M NaClO4 and

0.5 M Na2SO4 solutions were about an order of magnitude slower than those reported

in 0.5 M NaCl [143]. These fatigue results showed that aqueous sodium-chloride

solutions had a critical effect on the acceleration of the fatigue-crack-propagation

rates and the reduction of the Kth values in this Zr-based BMG. The authors

theorized that the marked environmental effect on fatigue-crack propagation in

aqueous sodium-chloride solutions could be attributed to a stress-corrosion crack-

41
growth mechanism, which involved the stress-assisted, anodic dissolution at the crack

tip.

5. Fatigue Failure Mechanisms

To compare the fatigue behaviors of BMGs, the broad ranges of S-N behavior

for Zr-based BMGs under bending, uniaxial, and rotating loading are presented in

Figure 13(a) [83]. An examination of these plots demonstrates that the fatigue

lifetimes of Zr-based BMGs are comparable under bending, uniaxial tension, and

rotating loading. However, the fatigue lifetimes under uniaxial compression appear to

be longer than those derived through bending, uniaxial tension, and rotating fatigue

experiments. The S-N curves of some typical crystalline alloys are also plotted in

Figure 13(a) for comparison. The ultrahigh-strength steel (300-M) exhibited higher

endurance limits than the IN 718 superalloy, Ti-6Al-4V, and Zirconium alloys [144].

The IN 718 superalloy had a greater fatigue limit than Ti-6Al-4V and Zr alloys while

Ti-6Al-4V presented a higher fatigue limit than Zr alloys. It is obvious that in Figure

13(a), the fatigue behavior of Zr-based BMGs spans the wide range (i.e., the value of

the fatigue-endurance limits varies from 150 to 1,050 MPa). These fatigue limits of

BMGs are comparable with the conventional, crystalline materials depending upon

the exact BMG alloy tested and the test methods. It is possible to obtain Zr-based

BMGs with different fatigue behaviors through changes in the alloy composition or

the fabrication procedure. Figure 13(b) and (c) display the relationship of fatigue-

endurance limit and fatigue ratio vs. UTS, respectively. It can be observed that BMGs

(the area in red color) have the highest fatigue strengths and UTS among all the listed

42
materials, including steel, high entropy alloys (HEAs), Zr-based, Al-based, Cu-based,

Ti-based, and Ni-based alloys. In Figure 13(b), with the increase of material UTS, the

endurance limit will increase in a linear fashion, approximately equal to 0.5 for most

materials. However, BMGs have a wider fatigue ratio range from 0.1 – 0.6.

Recently, the new concepts, HEAs are widely studied because of their

excellent strength, ductility, and resistance to corrosion [145-149]. The fatigue

properties of Al0.5CoCrCuFeNi (at.%) HEA have been investigated by Tang and

Hemphill et al. [147, 148] using four-point bending, which is also presented in the

Figure 13(a). Compared with other crystal alloys and most of the BMGs,

Al0.5CoCrCuFeNi HEAs exhibit a larger fatigue limit, indicating that HEAs have the

potential to be the candidates in the fatigue application. So more attention will be paid

on the fatigue studies and further work needs to be performed in HEAs in depth.

The endurance limits are plotted as a function of the tensile/compressive

strengths of these various Zr-based BMGs in [83]. There was no distinct correlation

found between the endurance limit and strength in Zr-based BMGs, which suggests

that the mechanism of the tensile/compressive deformation is different from that of

the fatigue fracture in the Zr-based BMGs.

In general, the fatigue-crack-growth data of Zr-based BMGs, as shown in

Figure 11, follow the simple Paris power-law equation in the steady state:

, (5-1)

where m is the crack-growth exponent, and C is a scaling constant. The values of C,

m, and Kth of Zr-based BMGs and composites are summarized in Table 6 [83]. As

shown in Figure 11, the fatigue-crack-growth rates of the Zr-based BMGs and

43
composites are comparable to those observed in traditional crystalline alloys, such as

a ultrahigh-strength steel (300-M) and an age-hardened aluminum alloy (2090-T81)

[70]. For Zr-based BMGs and composites, the exponent, m, is generally in the range

of 1 to 3, which is typical of ductile crystalline alloys in the Paris regime [70, 109,

133, 135, 136]. However, the Kth values of Zr-based BMGs are usually lower than

those of crystalline alloys. The fatigue fracture toughness that starts the unstable

fatigue-crack propagation in the Zr-based BMGs is also much lower than in these

crystalline alloys. This feature could be the reason why BMGs fail after a limited

number of cycles once crack initiation has occurred.

5.1. Crack Initiation

It was generally found that the fatigue-crack-propagation behaviors of Zr-

based BMGs were similar to some crystalline alloys [70]. However, the stress-life

fatigue results demonstrated that the total lifetimes of Zr-based BMGs were generally

shorter than those of the crystalline alloys [70]. This significant difference could be

associated with crack initiation. Since there are no microstructural barriers like grain

boundaries, which could serve as a local arrest point to newly-initiated or pre-existing

cracks, a fatigue crack in BMGs could initiate easily [70]. Obviously, the crack-

initiation mechanisms are very important in the study of the fatigue behavior in

BMGs. It was noted that fractures in BMGs have been observed to initiate from

casting defects, such as porosities, inclusions, or surface flaws like polishing

scratches, during cyclic loading [76]. In other studies, fatigue cracks reportedly

initiated from shear bands [71, 74, 110]. The formation and propagation of shear

bands are usually explained by the free-volume theory under monotonic loading.
44
Some researchers also found that the fatigue behavior of BMGs is related to the free

volume [80, 99]. Moreover, the molecular-dynamics calculations for metallic glasses

suggests that the accumulation of the free volume and the resultant damage could

happen at much lower cyclic loads, compared to the loads that are required for the

same processes to happen under monotonic loading [144]. It was also found that

damage initiation was controlled by local atomic arrangements and fluctuations in the

local free volume [144]. This trend suggests that the shear bands could form at much

lower loads during fatigue of BMGs. Using a thermodynamic modeling, Wright et al.

[150] showed that nanometer-scale voids could form due to the free-volume

coalescence in a shear band during deformation. The size to which these voids grow

is limited by the growth kinetics. In a tensile-stress state, the void growth and linkage

are promoted by the tensile forces. In a compressive-stress state, the void growth is

retarded, which may promote the shear-band multiplication and, thus, a larger strain

to failure. High-resolution-transmission-electron-microscopy (HRTEM) studies have

observed a large concentration of nanometer-scale voids in shear bands, which are the

result of the coalescence of the excessive free volume in the shear band [151, 152].

These voids could result in the localized stress concentration during cyclic loading.

Thus, a fatigue crack could initiate from the shear band where the void defects

concentrated.

Although very little is known about the initiation mechanisms of BMGs in the

corrosion solutions, Morrison et al. [100] found that multiple pits were observed near

the fracture surfaces. However, there was no clear evidence that a crack initiated from

45
pits [100]. This fact suggests that the initiation mechanisms of BMGs in air and other

specific environment could be similar.

5.2. Crack Propagation

As discussed above, the fatigue-crack-growth data for Zr-based BMGs

generally conforms to the Paris power-law, Eq. (5-1). The fatigue-crack-growth rates

in the Paris-law region are close to those observed in some traditional crystalline

alloys, [e.g., an ultrahigh-strength steel (300-M) and an age-hardened aluminum alloy

(2090-T81)] [70]. Moreover, after fatigue experiments, the fatigue-crack-growth

region on the fracture surface of Zr-based BMGs generally demonstrated the distinct

striations [70, 74, 75, 80]. For crystalline alloys, the striation formation is due to the

irreversible blunting and re-sharpening of the crack tip during cyclic loading.

Similarly, the striations observed in fatigue-fractured BMGs could be associated with

the process of blunting and re-sharpening. Using this model for the striation

formation, the fatigue-crack-growth rate is directly proportional to the range of the

crack-tip-opening displacement () [70]. The following equation can be obtained

using continuum-mechanics arguments:

, (5-2)

where is the yield stress, E = E in plane stress or E/(1 - 2) in plane strain ( is the

Poisson’s ratio), and  is a scaling constant (~ 0.01 to 0.1 for the mode-I crack

growth), which is a function of the degree of the slip reversibility and elastic-plastic

properties of the material [70]. During the monotonic and cyclic experiments of Zr-

based BMGs, the shear bands are observed readily [62, 70, 71, 74]. Moreover, Gilbert

46
et al. employed Eq. (5-2) to predict the fatigue-crack-growth rates, and the calculated

results showed a good agreement with the experimentally-measured crack-growth

rates when  = 0.01 [70]. These facts suggest that the mechanism for cyclic crack

propagation in Zr-based BMGs could be related to the blunting and re-sharpening of

the crack tip.

In general, the striation spacing was reported to be larger than the fatigue-

crack-growth rate, da/dN [70, 118]. The fracture surface showed that striations did

not extend by the same width, which meant that the entire crack front did not

propagate consistently with one loading cycle. This inconsistency could be associated

with the non-uniform crack extension along the crack front, which could result in the

larger striation spacing [70]. In addition, Hess et al. suggested that this disagreement

between the striation spacing and the crack-propagation rate was due to the necessity

of the damage accumulation at the crack tip prior to the crack advancement in BMGs

[118]. These explanations are based on an assumption that one cycle forms one

striation. For crystalline alloys, the assumption is generally accepted since the

striation spacing typically agrees with the da/dN during the fatigue-crack-growth

experiments. However, for BMGs, it is still unknown whether this assumption is

valid. In fact, two kinds of striations (coarse and fine striations) have been reported in

the fatigue-crack-propagation region [153, 154]. The coarse striations are easy to be

observed, even at a low magnification. However, if the coarse striations are carefully

observed at high magnifications, some fine striations can be found on the coarse

striations [153]. It is suggested that the striation spacing of the fine striations could

match the fatigue-crack-growth rate, da/dN [154].

47
Although the crack-initiation mechanism of BMGs in a NaCl solution could

be similar to that in air, the fatigue-crack-growth behavior in the NaCl solution is

obviously different from that in air, and the fracture surfaces in the NaCl solution (no

striation was observed) are also markedly different. This difference suggests that the

fatigue-crack-growth mechanisms varied in the aqueous solution and in the air.

Schroeder et al. [143] found that the environmentally-assisted fatigue-crack-growth

behavior of the Zr-based BMGs in the NaCl solution could be attributed to a stress-

corrosion-cracking (SCC) process, which is observed under constant loads in the

same solution [143]. It was generally found that SCC and corrosion-fatigue

experiments in the NaCl solutions yield the same stress-corrosion fatigue behavior,

and the mechanistic relationship between-SCC and corrosion fatigue is consistent

with the experimental results [143].

Many BMGs have demonstrated excellent corrosion resistance [155], which

could be due to their structural and chemical homogeneity. However, amorphous

alloys do not always exhibit superior corrosion resistance. Potential dynamic-

polarization studies suggest that the amorphous phase is not necessary for alloys to

improve their corrosion resistance [143]. Indeed, the good corrosion resistance in

amorphous alloys was because of the presence of a highly-stable passive film,

concentrated in certain alloying elements known to provide the good corrosion

resistance in crystalline metals [143]. This fact suggests that the composition of the

alloy could be more important than whether the structure is crystalline or amorphous

for the stress-corrosion resistance.

5.3. Fatigue-Failure Mechanisms

48
According to Wang’s work [156], the fatigue characteristics and mechanisms

of MGs can be characterized through the real-time, high-resolution TEM. The sample

preparation and experimental setup for the TEM characterization are displayed in

Figure 14 with detailed description in Ref. [156]. Figure 15(a) [156] shows three

high-magnification images of the notch-tip region for the untested, cyclically-strained,

and monotonically-strained specimens, respectively. However, for the specimen

under monotonic loading with a strain of ∼ 12%, no obvious surface roughening was

visible at the free surface of the notch (Figure 15(a) [156], Right). It is, thus, seen in

Figure 15(b) [156] that cyclic loading leads to surface damage on the atomic-scale

roughening after several hundred cycles. Figure 15(b) [156] exhibits three bright-field

TEM images of the notch-tip region after 980, 1,470, and 1,960 cycles, respectively,

from which the following mechanistic processes can be proposed. Local cyclic

deformation on the roughened surface of the notch tip leads to the initiation of a

fatigue crack. Many tiny nanograins are nucleated in this highly-strained region from

the high-resolution video photography in Ref. [156]. After 1,960 cycles, the

nanocrystal was grown to a diameter of ∼ 27 nm. In Figure 15(c) [156], Left and

Center, the formation and growth of the nanocrystal is further elucidated by high-

resolution TEM images after 1,470 and 1,960 displacement cycles, respectively.

Figure 15(c) [156], Left and Center show a larger grain (G1) neighboring a smaller

grain (G2) with a fast Fourier transformation (FFT) analysis of this area in Figure

15(c), Left Inset. After an additional 490 cycles, G1 grew larger by consuming G2, as

was supported by the corresponding inverse FFT image in Figure 15(c) [156], Right.

49
The temperature-driven structural disorder-to-order transitions, such as the

crystals nucleation in liquid upon cooling, have been extensively studied [157-159].

However, stress-driven disorder-to-order transitions are seldom investigated [160,

161]. Although the ability of MGs to plastically deformation under the stress by STZs

is well known, how stress enhances “atomic diffusivity” in MGs? To quantify and

understand the atomic-structure evolution, molecular dynamics (MD) simulations of

the cyclic-stress–induced glass–to-crystal transition can be carried out on a binary

MG Al50Fe50 with the details of MD simulation setup in Ref. [156]. The evolution of

atomic geometries is analyzed by the following “deformation–diffusion”

decomposition [156],

d ji  d 0ji J i  s ji , j  Ni , (5-3)

where i is the analyzed atom, j ∈ Ni are i’s original neighbors in a reference

configuration, d 0ji is the original distance between atoms, j and i, in the reference

configuration, and d ji is their current distance vector. The first term in the above

equation reflects local deformation, while the second term represents additional

atomic movements or shuffling beyond a mere shape change. The extent of local

diffusion is defined by

1 1
|s min  | d 0ji Ji  d ji | ,
2 2
Di2  ji |  (5-4)
Ni jNi Ni Ji jNi

where the local deformation gradient, Ji, needs to be numerically optimized to

minimize Di2 . Although Di2 might be small, it accumulates during cyclic

deformation. Figure 16(a) shows the atomic configuration of the notch area after 275

cycles of strain, in which the long-range periodic packing of atoms clearly

50
corresponds to crystallization in the active zone. Statistics of the atomistic Di2 after

different cycles are plotted in Figure 16(c). It reveals that the variation of the atom

fraction vs. Di2 for the first half cycle showed a sharp decrease. Then it gradually

increased to reach a peak value at the large diffusional displacement ( Di2 ) as the

cycle number increased, e.g., the corresponding Di2 equals 5.8 Å2 for the peak after

275 cycles. Although the peak Di2 becomes sharper with increasing numbers of

cycles, it occurs at an approximately constant square displacement value of about DC2

= 5.5 Å2, which characterizes the initiation of crystallization. Moreover, the averaged

showed a nonlinear increase with increasing strain cycles, as shown in the black

solid line of Figure 16(d), in which the average grain size ( d  A ), where A is the

grain area, can be classified into three regimes according to the cycle number. For

Regime I (cycles from 1 to ∼ 110), the average grain size remains subnanometer

(Figure 16(d), the curve with circles). For Regime II (cycles from 110 to ∼ 175),

rapid grain growth begins, following the incubation period in Regime I. After the

cycle number exceeds 175 (regime III), the grain-growth rate decreases significantly.

The atomistic simulation of monotonic deformation (at a strain of ∼ 8%) of

the same sample geometry is shown in Figure 16(b) to compare with cyclic

deformation. The active zone was along the plane with the maximum shear stress.

Neither obvious localization nor long-period packing of atoms (of a glassy structure)

was observed in these zones.

Based on the above discussions, the fatigue-failure process of BMGs under a

tensile-stress state can be described, as illustrated in [83]. First, when a BMG sample

51
is cyclically loaded, a shear band will form. In the shear band, voids can develop

during deformation due to the free-volume coalescence. The growth and linkage of

voids will be assisted by a tensile-stress state, perhaps leading to a large stress

concentration near these voids. A fatigue crack can initiate from these voids due to

the resultant-stress concentration. Then, under a cyclic-tensile load, the fatigue crack

will open and grow. At the crack tip, a small plastic zone will form. The plastic zone

can blunt the main crack tip. However, many shear bands or crack branches will

develop near the crack tip. This crack-branching phenomenon was observed by Flores

et al. when the crack-growth behavior of the Zr41.25Ti13.75Ni10Cu12.5Be22.5 (at.%) BMG

was examined [162]. Hence, the fatigue crack will propagate along another favorable

direction. Therefore, the process of blunting and re-sharpening will form the striated

crack-propagation region when the fatigue experiments of BMG samples are

conducted under a tensile-stress state. In case that BMGs include inclusions and

porosities, a crack will initiate from these casting defects easily. The fatigue crack

will propagate following the process of blunting and re-sharpening.

Regarding the compression-compression fatigue, the fatigue-fracture surface

is similar to that observed in the monotonic-compression test, which forms by the

unstable fracture along a primary shear band [118]. There are no fatigue striations

observed on the fractured surface after the compression-compression fatigue [118,

119]. This trend suggests that the mechanism of the fatigue facture under the

compression-compression stress seems to be similar to that under a compression-

stress state. The free volume in a shear band is expected to increase during the

deformation of BMGs. Any free volume created in a shear band during the

52
deformation is highly unstable and tends to form nanometer-scale voids [150]. Since

a compressive stress will retard the void growth and linkage, the fatigue crack is

difficult to originate under a compressive-stress state. But it is anticipated that the

shear stress can also lead to void growth, similar to the shear fault in geomaterials.

We also note that the formation of voids will decrease the density of the material and

its resistance to deformation in the shear band. After several compression-

compression fatigue cycles, one primary shear band gradually becomes weak and

cannot sustain the compressive stress. Finally, the BMG sample fails along one

primary shear band, which forms a shear-fracture angle with respect to the stress axis.

This fact could also explain why the fatigue lifetime under a compression-

compression cyclic loading was longer than that under a tension-tension cyclic

loading. With a tensile-stress state, the growth of voids would be promoted by the

tensile stress, while a compressive-stress state would retard the void growth.

Recently, the shear-band mechanism for fatigue-crack propagation in BMGs

is modeled from the framework of linear elastic-fracture mechanics (LEFM) [163].

The LEFM simulation has been successfully applied in many disciplines, such as

geophysics and tribology [164]. However, Wang et al.’s work is for the first time

integrated with a localized shear-band model. In the present study, the detailed

algorithm for calculating the stress-intensity factor (SIF) can be found in Ref. [163].

In this simulation work, shear bands are assumed to be uniformly spaced

along the 45° direction, which usually develop at an early stage of the fatigue-crack

growth, as indicated by the dashed lines in Figure 17(a). The fatigue crack advances

and step-by-step crack path can be calculated using the maximum energy-release-rate

53
theory (G) [165]. In the simulation, a fatigue crack is assumed to grow in the

direction of the maximum value of G. In Figure 17(a), the possible crack-growth

directions at point, P, can be PA, PB, or PC. In the simulation, the stress intensity

factor (SIF) analysis is performed on all three tentative branches, PA, PB, or PC. By

comparing the resultant G, the preferred crack-growth direction is determined along

which attains the maximum G. After this calculation, the geometry of the fatigue

crack is updated for the succeeding simulation step. Finally, the crack-growth path

can be plotted by repeating the above simulation steps.

In general, the atomic density and flow viscosity are low within shear bands,

since many free volumes are introduced, which makes shear bands to be easily

deformed in BMGs [4]. A fatigue crack initiates from the opened shear band, and

then, the crack preferentially propagates along the shear bands than outside the shear

bands [80]. Subsequently, the fatigue crack will advance every fatigue cycle in a

direction perpendicular to the tensile-stress direction. At the crack tip, a plastic zone

is covered by many formed shear bands, as schematically displayed in Figure 18(a)

[163, 166]. Then, the crack propagates one fine striation spacing every fatigue cycle,

as schematically shown in Figure 18(b). In general, the crack may propagate for

several cycles along one shear band and, then, change to another shear band.

Therefore, the coarse striation forms by crack propagation for several cycles, as

schematically presented in Figure 18(c). With the increase in the crack length, the

maximum SIF, Kmax, becomes larger and larger. When Kmax reaches the fracture

toughness, KIc, the BMG sample fails. Note that for BMGs, two kinds of striations

can be observed in the crack-propagation region [163]. The fine striation spacing is

54
related to the advance of the fatigue crack by one fatigue cycle, while the formation

of coarse striations is due to the presence of the discontinuous shear bands ahead of

the fatigue crack. The LEFM simulation results are shown in Figure 17(b) [163],

where the two adjacent segments represent the crack propagation by one coarse

striation [Figures 17(a) and (b)], which is progressively determined at each simulation

step. The present numerical simulation demonstrates that the incipient crack kinking

from the free surface may transverse several coarse striations and propagates along

the direction roughly perpendicular to the applied stress, which is in agreement with

the experimental results [163].

Moreover, the LEFM model can also be used to investigate the crack-growth

behavior of BMGs under compression loading [167]. It has been found that tensile

microcracks may be generated from the pores under a compressive load. Computation

of SIFs for short radial cracks emanating from a circular hole shows that KII is

relatively smaller than KI, which, therefore, verifies a Mode-I dominant theory for

fatigue-crack initiation from pores. Here, the Mode-I and Mode-II SIFs, i.e., KI and

KII, respectively, are normalized by a factor,  1  r0 [167]:

K I  1  r0 FI and K II  1  r0 FII , (5-5)

where the non-dimensionalized SIFs, FI and FII, are functions of λ, a/r0, θ, and φ,

which are schematically shown in Figure 19(a). The origin of the coordinate system is

placed at the center of the circular hole of a radius, r0, in an infinitely extended solid.

A crack of length, a, emanates from the perimeter of the circular hole. The orientation

angles, θ and φ, can be an arbitrary combination. The remote compressive load, σy =

55
σ1, is applied along the y-axis direction. The lateral confinement is simulated by

applying a stress, σx = λσ1, along the x-axis direction Figure 19(a).

Consider first a radial crack emanating from a circular hole subjected to

uniaxial compression. Computations on two short cracks (a/r0 = 0.01 or 0.1) are

performed to investigate the properties of crack initiation. Figure 19(b) shows the

variation of FI and FII for different inclined angles, θ. For both cases, the Mode-I SIF

vanishes for θ smaller than 60, since the crack faces are closed. A maximum F I

occurs at θ = 90, in which case, a pure mode I is evidenced due to symmetry. For the

shorter crack (a/r0 = 0.01), the Mode-II SIF tends to be much smaller. This trend

demonstrates that a radial crack is much easier to be initiated by the mode-I splitting

instead of the mode II shearing. However, for a pre-existing crack whose length is

comparable to one tenth of the radius of the circular hole, the Mode-II effect may not

be negligible, and it becomes paramount when crack closure is involved (0< θ <

60). Their analysis results also show that the behavior of crack growth is extremely

sensitive to the effect of lateral confinement. Axial splitting can be caused by the

unstable growth of macroscopic cracks, resulting from such lateral-confinement

effects as misalignment, off-axis loading, or over-constraining the sample ends [167].

6. Size Effects on Fatigue Behavior of BMGs

In general, fatigue accounts for approximately 90% of all service failures

caused by mechanical damages [57], which can be affected by many factors, such as

the material composition, specimen geometry, chemical environment, temperature,

mean stress, residual stress, and surface condition, etc. In these factors, the geometry

56
effect is a critical one in affecting the fatigue behavior of metallic glasses, as reported

in Refs. [168-170]. Thin wires or ribbons of metallic glasses under bending show

large plasticity, but thicker plates fail catastrophically with very limited plasticity

under bending [168, 171-173], which suggests that the significant bend ductility can

be obtained only when the sample size is below a critical value. These results suggest

a size effect for the bend ductility of metallic glasses. Therefore, an interesting topic

arises: How does the sample size influence the fatigue behavior of BMGs? Moreover,

the improved plasticity of BMGs is achieved through the generation of multiple shear

bands, especially when the sample size becomes smaller [168, 169]. Nevertheless, the

effect of these multiple shear bands on the fatigue behavior in BMGs is still not clear.

Therefore, it is of critical importance in understanding the size effects on the fatigue

behavior of ductile metallic glasses.

6.1. Size Effects on Fatigue Behavior of BMGs

Recent studies on the fatigue behavior of metallic-glass nanowires by

molecular-dynamic (MD) simulations showed that MGs will not fatigue under strain-

controlled compression-compression tests [174]. It found that irreversible

deformation occurs during all fatigue simulations. However, the MG nanowire does

not suffer from structural damage, and no softening occurs during cyclic loading.

Jang et al. [175, 176] reported a significant strength increase and highly-localized-to-

homogeneous deformation mode change, when the size of BMG specimens decreases

to a nano-meter scale. The strength of the BMG pillar increases with decreasing the

specimen diameter on a micro-meter scale, and reaches the maximum value of ~ 2.6

GPa at a specimen diameter of 800 nm, while the bulk (millimeter) scale specimens

57
have a yield strength of 1.7 GPa. Jang et al. [177] further studied the fatigue behavior

of micron-sized BMG pillars and found that the fatigue-endurance limit is very close

to the yield strength of BMGs, which is consistent with the MD simulation results

that metallic glasses do not fatigue at submicron scales [174]. The above studies

presented that the size of BMG specimens does influence the fatigue behavior of

metallic glasses under fatigue.

For studying the size effects on the fatigue behavior of monolithic BMGs, Zr-

based alloys are usually employed [178]. In the current work, two compositions

[Zr50Cu40Al10, and Zr50Cu30Al10Ni10, at.%] with different sizes are used under four-

point-bending fatigue loading. In Figure 20 [178], it shows the S-N curves of metallic

glasses [(Zr50Cu40Al10 and Zr50Cu30Al10Ni10 (at.%)] with two different sizes (marked

as small and large samples in the graph). It can be observed that for the same size,

Zr50Cu30Al10Ni10 (at.%) BMGs (705 MPa for large samples and 397 MPa for small

samples) exhibited higher endurance limits than Zr50Cu40Al10 (at.%) BMGs (602 MPa

for large samples and 232 MPa for small samples). Under four-point-bending, large-

sized BMG samples show longer fatigue lifetimes and higher endurance limits than

those with small sizes. However, above the endurance limit, the lifetimes of small

samples decreased slower than those of large samples with increasing applied stress.

The fractography of the small-size samples after four-point-bending tests can

be characterized by SEM, in which two types of failure modes can be found, flexural

and fracture failure, during fatigue experiments. Here, the flexural failure means that

metallic-glass specimens become bent with multiple uniform shear bands, while the

fracture failure denotes that a BMG specimen produces one or more primary fatigue

58
cracks and breaks into two or more pieces along these primary fatigue cracks. Figure

21(a) [178] presents the flexural failure of one small fatigue sample, in which many

shear bands can be observed on both compressive- and tensile-stress surfaces. Figure

21(b) [178] demonstrates primary and secondary shear bands on the tensile surface.

The primary shear bands exhibited an angle of ~ 55º with respect to the tension stress

direction (Figure 21(a)). After failure, some shear steps are observed near the corner

(Figure 21(c) [178]). From the lateral side, it can be found that shear bands appear

from the surface and propagate toward the center of the specimen from both the

tensile and compressive sides, as shown in Figure 22 [178]. It can be observed that

the shear bands from the tensile side (~ 590 µm) tend to be longer than those from the

compression side (~ 410 µm). The average shear-band spacing on the tensile and

compressive sides are approximately 192 and 155 µm, respectively.

6.2. Size Effects on Fatigue Properties of Fatigue-Damaged BMGs

The fatigue properties influenced by different sample sizes have been

discussed in Section 6.1. However, the fatigue property of BMG specimens with

different sizes after fatigue-induced damage has not been studied extensively, which

will be an interesting topic [179].

The detailed schematic diagram of the experimental plan on this fatigue-

induced damage work was presented in Figure 23. The BMG specimens used are

cylindrical rods with the composition of (Zr 55Cu30Ni5Al10)98Er2 (at.%), in a diameter

(D) of 6 mm and a length (L) of 25 mm, which are loaded under a constant-load

compression-compression fatigue test. For the first run of fatigue tests, denoted as the

1st run, an as-cast BMG specimen is used. The repeated loading/unloading fatigue

59
process continues until the specimen fails. Then, the damaged region is removed, as

illustrated in this figure, and the leftover is used for the 2 nd-run tests. The

experimental condition of the 2 nd-run test is identical to the 1st-run tests, except that

the sample is shorter due to the removal of the fractured region. The series of tests

was continued until the L/D ratio of the leftover was less than 0.8.

Figure 24(a) shows the S-N curves of the as-cast and after-fatigue BMG alloy

(Zr55Cu30Ni5Al10)98Er2. The open circle is the fatigue life of BMG specimens with a

fixed L/D ratio (1.67) at various stress ranges (typical fatigue tests). The colored

symbols are results from this fatigue-induced-damage study, which have a larger L/D

ratio. The red color represents the first-cycle tests (as-cast samples), and the blue

color denotes the proceeding tests using the after-1st-round leftover. It can be

observed that the first-cycle tests of the specimens show comparable on shorter

fatigue life than the second – cycle tests [red symbols in Figure 24(a) and (b)].

As discussed in Section 6.1, the geometry, here is the changes of L/D ratios,

would affect the fatigue life of a BMG specimen. The fatigue life of each test is

plotted as a function of L/D ratio in Figure 24(c). The results do not show strong

correlation between the L/D ratio and the fatigue life within the L/D range from ~ 0.8

to 4. Therefore, the dominant factor that affects the fatigue life of a specimen in this

experiment could not be the L/D ratio. In the following of this section, statistical

analyses are employed to examine the effect of sample size on the fatigue life.

Therefore, the fatigue-failure mechanism of the large (6-mm in-diameter) Zr-

based BMGs under cyclic compression-compression loading is suggested and

illustrated in Figure 25 [179]. The crack initiates at the weakest points (such as

60
extrinsic defect sites) in the sample, then, propagates slowly at the beginning, and

leaves the striation on the crack surface. The crack continues to grow until the

specimen cannot sustain the stress, and then the fast-shearing process starts, which

leads to the catastrophic failure. During the crack-propagation process, the fatigue

damage is localized. When one crack starts to grow, the rest of the sample still

undergoes elastic deformation. The microstructure away from the crack region

generally remains unchanged. Therefore, after cutting off the damaged part, the

remaining part performs as an as-cast specimen. The above mechanism implies that

the fatigue-endurance limit of BMGs under cyclic compression-compression stresses

can be greatly improved by reducing the number of defects (crack-initiation sites,

such as inclusions) in the material. It should be pointed out that even elastic

deformation at elevated temperature (but still below Tg) can result in structural

disordering or rejuvenation [180]. The corresponding fatigue tests have not been

reported, which can be an interesting line of future work.

6.3. Size Effects on Fracture Toughness of BMGs

Similar to the fatigue behavior of metallic glasses, the fracture toughness of

BMGs is also affected by their sample sizes. For large sample sizes, C(T) specimens

are prepared in Figure 26(a) [181], while single-edge notched-bend [SE(B)]

specimens are fabricated for small sample dimensions, as exhibited in Figure 26(c)

[181].

The American Society of Testing and Materials (ASTM) standard E399 [182]

distinguishes between three types of failure in a fracture-toughness test based on the

load – (crack mouth) displacement (P-V) curve, as shown in Figure 26(b). Type I

61
describes a material, which fails after significant plastic deformation; here, the

intersection of the load with the 95% secant line (Figure 26(b)) is used to calculate a

conditional fracture toughness, KQ. Type II can be seen for a material that shows

crack propagation (“pop-in”), followed by a further increase in the load before the

catastrophic failure (Figure 26(b)). Here, the peak load before crack propagation is

used to calculate KQ. Type III shows a material, which behaves in a linear-elastic

manner, with just minor plasticity before the catastrophic failure (Figure 26(b)). In

this case, the maximum load is used to calculate KQ. Further details can be found in

the standard [182].

The results of all valid fracture-toughness tests [KIc (plane-strain fracture

toughness), and JIc] are shown in Figure 27 [181] as a function of the uncracked

ligament width, b = W (sample width) – a (crack length), which is illustrated in

Figure 26. For comparison purposes, JIc-toughness values were represented as a stress

intensity, referred to as KJIc toughness values, using the standard mode-I, linear-

elastic, K-J equivalence relationship:

EJ Ic
K JIc  , (6-1)
1  v2

with a Young’s modulus of E = 85.6 GPa and a Poisson’s ratio of ν = 0.375.

Additionally, to show the data for all valid and invalid tests, KIc, KJIc, and KQ

values are plotted as a function of the ligament size, b, in Figure 28(a). Finally, a

summary of all data is shown in Table 7 [181], together with details, such as the

sample dimensions, and the critical sample thickness, Bcrit, and the ligament width,

bcrit, values needed to achieve a valid test by ASTM E399. Additionally shown are the

62
types of failure based on the load–displacement curve categories of the ASTM

standard E399 [182] and the pre-cracking method, as in Figure 28(b).

Samples with the largest ligament size, four C(T) samples with b = ~ 10 mm,

showed a fracture toughness KIc of ~ 25.3 MPa m1/2. These results represent the

linear-elastic, plane-strain fracture toughness for the Zr52.5Cu17.9Ni14.6Al10Ti5 (at.%)

BMG measured using C(T) samples. All other tests were performed, employing SE(B)

samples with ligament sizes of b = ~ 1 – 2 mm. Twelve samples failed in a nearly-

linear-elastic manner with very minor plastic deformation, conforming to Type-III

failure by the ASTM standard E399 [182]. Five of them failed at relatively low loads,

leading to KQ values between 26.2 and 44.9 MPa m1/2. Thus, the sample dimensions

are sufficient for the measured numbers to qualify as valid KIc. One of those samples,

however, had an a/W of 0.58 and hence cannot be strictly included as a valid KIc

value, as indicated by the equations below for the size requirements of KIC testing:

KQ
a, b, B  2.5( ) 2 and 0.45  a / W  0.55 . (6-2)
 YS

The remaining four values led to an average KIc of ~ 35.7 MPa m1/2 (Table 7).

The data point with a/W = 0.58 is invalid as KIc according to ASTM E399 [182].

However, it is valid in terms of JIc measurements denoted as KJIc by ASTM E1820

[315], which allows for a/W ratios between 0.45 and 0.7. Three more samples

allowed for valid JIc measurements leading to KJIc = ~ 39.0 MPa m1/2 (Table 7). The

other four samples that showed Type-III failure, as well as seven samples that failed

as Type-I with significant amounts of plasticity, did not meet the requirements to

qualify for either KIc or JIc. For Type-III samples, the failure load was used to

calculate KQ, whereas for Type-I samples, the load intersecting with the 95% secant

63
line was used [Figure 26(b)]; the resulting KQ for those samples was found to be 58.2

± 16.1 MPa m1/2 (Table 7). Results of the statistical analysis of all valid data revealed

a statistically significant (p < 0.05) correlation between the valid fracture toughness

(KIc, KJIc) and the ligament size, b, assuming a linear regression. However, when

considering only the valid KIc data, the correlation just barely missed the criterion for

significance (p = 0.054).

The three SE(B) samples that were only notched but not pre-cracked had b = ~

2 mm and failed catastrophically in a nearly linear-elastic manner, showing Type-III

failure by ASTM standard E399 [182]; their (apparent) toughness values were far

higher, at KQ = ~ 93.9 MPa m1/2. Both specimens with cracks smaller than a/W = 0.45

and larger than a/W = 0.55 did not seem to show a clear trend toward larger or smaller

fracture-toughness values. Additionally, no obvious difference between pre-cracking

in tension or compression could be discerned. Scanning-electron-microscopy

examinations of the fracture surfaces showed no obvious defects associated with the

different locations in the castings for the various samples. Both the C(T) and SE(B)

samples, which yielded valid toughness values, showed a relatively-even crack

propagation front and just minor shear banding ahead of the pre-crack [Figure 29(a)].

In comparison, the SE(B) samples, which did not meet the requirements for either KIc

or KJIc failed, and showed significant plastic deformation, multiple shear-band

formation and blunting at the tip of the pre-crack (Figure 29(b)). These samples also

showed a clear trend to rougher fracture surfaces, bifurcations, and significant

deviations from a mode I crack path.

64
Based on the experimental results of size effects on the fracture toughness of

bulk metallic glass (Vitreloy 105), a set of phenomenon were noticed: 1) The small-

sized samples expected to be valid will give a wider scatter in KQ values, a smaller

yield of valid tests, and possibly somewhat elevated toughness values. Such behavior

is distinct from polycrystalline metals, where the size requirements of ASTM

standard E399 have been shown to be quite conservative; 2) Measured conditional KQ

toughness values are found to increase, with the increased scatter, with progressively

decreasing uncracked ligament widths. The greater scatter is likely related to the size-

dependent bending ductility of BMGs, as samples below a certain critical size are

able to prevent catastrophic failure in bending by the formation of multiple shear

bands throughout the extent of the uncracked ligament; 3) Samples smaller than that

required by ASTM standard E399 for valid tests are allowed by the J-integral-based

analysis of ASTM standard E1820. However, according to the experimental study,

such tests were found to give either significantly higher KJIc toughness values relative

to KIc or invalid results according to the standard. Such behavior is quite different

from crystalline metals and is most likely related to the distinct local strain-softening

behavior found in metallic glasses, which gives rise to the non-unique crack-tip stress

and strain fields; 4) Toughness values measured using samples smaller than those

required by ASTM E399 for valid fracture-toughness values should be considered as

size dependent, even when considered valid by JIc measurements. When only small-

sized BMG samples are permitted by processing limitations or other factors, it would

be advisable to compare new BMG compositions to other compositions using exactly

the same sample geometry and testing configuration to allow fair comparisons.

65
7. Improving Fatigue Resistance

Since the fatigue properties are very important for the application of materials,

how to improve the materials’ fatigue performance can be an interesting topic? From

the literature on fatigue improvements, thin films can be good candidates, which

attracted more and more attentions recently, especially the thin-film metallic glass

(TFMG) [97, 183-185]. Similar to BMGs, TFMGs have unique properties, compared

with the conventional crystalline films, such as high strength, high toughness, large

elastic limits, and high-corrosion resistance, etc. In our previous work [97], there is

the plot of yield strength vs. plasticity of ceramic, metallic, and TFMG, as shown in

Figure 30. It can be observed that TFMGs have much higher yield strengths than

ceramic and metallic coatings, while still keeping good ductility. According to the

work in Refs. [184-188], the TFMGs have been successful in improving the fatigue

properties, such as the fatigue life and fatigue endurance limit of different kinds of

substrates, including steels, Ni-based, Al-based, and Ti-based alloys, as shown in

Figure 31 [187-190]. The fatigue performance of these TFMG-substrate material

systems have been reviewed in detail in Ref. [97], which will not be included in the

present work. However, recently we found that the fatigue properties of the BMG

could be improved by the TFMGs, which will be reviewed in the current paper.

Beyond that, the following contents will also be reviewed and discussed in this

section: 1) the TFMG on other crystalline alloy substrates and mechanical properties

of TFMG.

7.1. TFMG on BMGs substrates

66
7.1.1. Preparation of TFMGs

TFMGs are fabricated on the p-type (100) silicon (Si) substrate via radio-

frequency (RF) magnetron sputtering using a single target, as illustrated in Figure

32(a) [191]. This system consists of four rectangular targets and a rotating substrate

holder in the center of the chamber. Pure element targets of designed compositions

were used, such as the Zr53Cu29Al12Ni6 (at.%) TFMG. The Zr-Cu-Al-Ni alloy target

was placed at the bottom side of the chamber, while the substrate was located on the

top at a 10-cm working distance. In order to prevent undesirable chemical reactions,

the base pressure of the chamber was maintained below 5 x 10-7 Torr before

deposition. The working gas of Ar was, subsequently, introduced into the chamber at

a flow rate of 20 Standard Cubic Centimeters per Minute (sccm) with the working

pressure being controlled at 10 mTorr. During sputtering, it was detected that the

chamber temperature could rise up to ~ 80 ºC. After deposition, the TFMGs were

cooled to RT and taken out of the chamber for subsequent testing.

7.1.2. Fatigue Properties of TFMGs

The composition of TFMG was measured to be Zr60.2Cu24.4Al10.6Ni4.8 (at.%)

by EDS, while the BMG substrate has a composition of Zr50Cu30Al10Ni10 (at.%).

Figure 32(b) shows the stress-life (S-N) curve of the BMG substrate coated with a 200

nm-thick Zr-based thin-film metallic glass. It can be observed that the 200 nm-thick

TFMG can improve the fatigue life of the BMG substrate. Based on the applied

maximum stress, the fatigue-endurance limits of the BMG substrate and the substrate

with a 200 nm-thick TFMG are 300 MPa and 400 MPa, respectively. It can be

67
noticed that at both high and low stress levels, the improvement of the fatigue lives is

very pronounced by the Zr-based TFMGs, which is a value slightly different from the

reported work on the case of conventional metallic materials, such as steel and Ni-

based alloys [187, 188]. For the conventional metallic alloys, the fatigue behavior of

TFMG-coated specimens is usually improved more pronouncedly at the low stress

levels, compared with their performance at high stress levels. What are the

deformation mechanisms for the improved fatigue behavior of TFMG-BMG? This

can be investigated using the SEM and TEM to characterize the fractography of the

fatigue-fractured specimens at different stress levels in this section.

Figure 33 shows the SEM fractographs of the TFMG-BMG with a 200 nm-

thick Zr-based metallic-glass film after fatigue-fractured at σmax = 450 MPa. As

marked in Figure 33(a), three regions can be observed on the fracture surface,

including fatigue-crack-initiation (Stage I), crack-propagation (Stage II), and fast

failure (Stage III) regions. A fatigue crack initiates from the corner of the tensile

surface exhibiting a striation-type structure (close to the TFMG, marked with a red

circle), and, then, propagates inside with crack-growth direction indicated by the red

arrow on the micrograph. There is a distinct boundary between the fatigue-crack-

growth region and the fast-fracture region [Figure 33(b)], which presents very

different microstructures in various regions. The final fracture (Stage III) surface

shows a porous structure surface, which is usually a vein-like structure in the reported

work of the BMG fatigue deformation [36, 76]. A tilted-angle image [Figure 33(c)]

shows the interface region of the fatigue-fractured specimens after loading at the

maximum stress of 450 MPa, which clearly indicates that after the severe plastic

68
deformation and final fatigue fracture, only slight film delamination appears in the

area close to the fracture surface, while the TFMG remained well adhered with the

substrate in the other region. Moreover, no observable cracks are found on the

TFMG, which is a very good indication of the good adhesion between the TFMG and

BMG substrate. This is due to the excellent film ductility, which can accommodate

the fatigue deformation of substrates during cyclic loading.

Fatigue-damage mechanisms of crystalline materials have been well studied

and understood, and many theories are available. In general, slip bands (SBs),

twinning, deformation bands (DBs), and grain boundaries (GBs) are the preferential

sites for the nucleation of fatigue cracks in single-phase materials. Since BMGs are

non-crystalline materials, their fatigue-crack initiation and growth mechanisms could

be different from those of crystalline alloys. The basic fatigue-damage mechanism of

BMGs is still unclear. However, the deformation mechanism of metallic glasses is

usually attributed to the presence of shear bands and plastic flows. Because the plastic

flow is confined to narrow regions (shear bands) of BMGs, the presence of a shear

band in a metallic-glass specimen reduces its strength by providing a site for further

plastic flows [48, 74].

In general, a shear band or crack usually initiates from the casting defect and

porosity in a BMG specimen introduced during alloy casting and leads to a

preliminary failure due to the stress concentration in these sites [192]. These defects

could have a significant effect on the mechanical properties of a BMG specimen,

especially, on the fatigue properties [76, 154]. A Weibull-size effect appears, and the

shear-band embryo heterogeneously in a millimetre-sized MG specimen due to a

69
higher population of casting flaws in the specimen [193]. When the scale of the MG

specimen down to a submicron- or nano-scale, the deformation of a thin MG

specimen would require the homogeneous nucleation of shear bands, which would

lead to an increase of yield strength of the specimen [194].

The mechanism behind the improved fatigue resistance in the TFMG-coated

BMG substrate [Figure 32(b)] is rationalized as the TFMG shielding the substrate

from early shear-band initiation at surface defects. In a bare BMG, shear bands can

easily initiate at highly stress-concentrated defects on the sample surface, shear off

the surface, and promptly evolve into microcracks. Firmly binding a layer of TFMG

to the surface of the BMG, on the other hand, plays a role in protecting the substrate

and retarding the initiation of shear bands and fatigue cracks, thereby boosting the

fatigue resistance. This operation takes place at stage I in a three-stage fatigue-crack

initiation process in the TFMG/BMG system schematically shown in Figure 34 [195].

Also, the delayed shear-band initiation is partially attributed to the reduced overall

stress level on the surface of the substrate through the stress redistribution by the

TFMG, which has been testified by comparative indentation simulations [195].

Nevertheless, the merit of the TFMG coating in proving the fatigue resistance of the

BMG substrate ceases once the shear bands transit into shear steps at stage II.

Thereafter, the TFMG loses its effect in protecting the substrate as the severe stress

concentration at the film/substrate interface tends to promote the swift growth of the

shear steps and formation of cracks (stage III).

Recently, Wang, et al. [196] discovered that the nucleation length for shear

banding in metallic glasses can reach as high as 500 nm in the case of constrained

70
deformation, such as indentation. Therefore, a 260 nm-thickness TFMG is expected

to deform homogeneously without embryonic shear banding in this study.

Nanoindentation tests were carried out on the BMG sample and TFMG on the Si

wafer at a maximum applied loading of 1,200 μN. In order to avoid the substrate

effect, the film was prepared with a thickness of 500 nm. Nanoindentation curves for

the BMG substrate and TFMG were obtained. The pop-in or serration events can be

observed in the loading curve of BMGs, associated with the coalescence of local

STZs, which lead to an individual shear band underneath the indent area [4, 194,

197]. However, no pop-in event can be observed in the TFMG curve, which validate

the result that the shear-band nucleation lengths are as high as 500 nm reported by

Wang et al. [196].

The TEM EDS line-scan was used to analyze the composition at the interface

between the Zr-based TFMG and BMG substrate, as shown in Figure 35 [198, 199].

An apparent dark line presents at the interface between the film and substrate, while

no significant contrast can be observed in the shear-band offset region in Figure

35(c). Based on the line-scan results in Figure 35(e), obvious drops in the X-ray

intensity for all elements are observed at the corresponding position for the interface

between the film and substrate. This trend suggests that the less dense region at the

interface might have resulted from undesirable preferential etching during the FIB

TEM sample preparation. However, no apparent intensity difference is shown at the

interface in the shear-band offset region, as presented in Figure 35(d). It is assumed

that the energy released during the shear-band offset formation, which led to a

localized temperature rise in the offset region. As a result, the diffusion and fusion

71
occurred in the offset region between the MG film and substrate. It has been reported

by Chu et al. [183] that the TFMG coating can cover the surface weak points and

decrease shear-band initiating point at the surface of coated samples as well as absorb

the deformation energy, when the coated BMG substrate undergoes plastic

deformation.

7.2. Four-Point-Bending Fatigue Behavior of TFMG Materials

According to the reported data [183-185, 190, 200-202], TFMGs have been

used extensively to improve the mechanical behavior of various substrate materials,

especially in the field of wear resistance and hardness, without adversely affecting

their desirable properties [201]. Numerical studies have reported that the fatigue

behavior of the film/substrate system is closely dependent on film properties, such as

the structure, composition, thickness, hardness, ductility, and film/substrate interfacial

adhesion, et al. [97, 203, 204]. Therefore, the fatigue behavior of various TFMG-

substrate systems is summarized in Figure 31, which shows the S-N curves of the

316L stainless steels coated with Zr-based TFMGs (Zr47Cu31Al13Ni9, at.%) of two

different thicknesses (200 nm and 1 μm). It can be observed that both TFMG-

substrate systems could improve the fatigue life of the steel substrate, while the 1-μm-

thick-TFMG case has a better performance in improving the fatigue life and strength

of the substrate. The fatigue-endurance limits of the bare-steel substrate, and

substrates coated with 200-nm- and 1-μm-thick films are 700 MPa, 750 MPa, and 775

MPa, respectively. Note that the fatigue-lifetime improvement is generally much

more significant at lower stress levels in Figure 31. However, only limited beneficial

effects of the TFMG on the fatigue-life improvement can be observed at high stress

72
levels [97]. Similar fatigue improvement can also be observed on the TFMG-coated

Ni-based, Zr-based, Al-based, and Ti-based alloys. For example, in the S-N curves of

Ti-based (Ti-6Al-4V) alloy substrates with and without coatings, three kinds of

coating materials with the thicknesses of 200 nm are used: TiN, single-layer

Zr50Cu27Al16Ni7 (at.%) TFMG, and bilayer TFMG/Ti coatings. The S-N results

indicate that the fatigue life of Ti-based alloys is improved for all three cases, among

which the TFMG/Ti bilayer case exhibits the best fatigue lifetime and endurance limit.

The deformation mechanisms and morphology of these TFMG-substrate materials is

quite similar to those of TFMG-BMG materials, which can be found in Ref. [97].

Finally, the fatigue lifetimes and endurance limits of the reported TFMG-

substrate material systems, together with two monolithic Zr- and Cu-based BMGs,

are plotted in Figure 31. For the four-point-bending fatigue data (A, B, and C), the

results can be divided into high (C, larger than 700 MPa), medium (B, 300 – 700

MPa), and low (A, smaller than 300 MPa) fatigue-endurance limit regions in the

TFMG-substrate material systems. The tension-tension fatigue data of TFMG-

substrate material systems are in Region D, while the monolithic BMG

[Zr55Cu30Ni5Al10 and (Cu60Zr30Ti10)99Sn1, at.%] four-point-bending fatigue data is in

region E. Note that the fatigue lifetime and endurance limit of the coated 316L

stainless steel (in Region C) cannot be improved by the Fe-based TFMG, TiN, and

pure-Cu film adhesion, due to the intrinsic brittle nature of films and poor

film/substrate. For the good adhesion cases in Region C, such as Zr- and Cu-based

TFMGs on the Ni-based alloy, the lifetime and fatigue-endurance limit can only be

improved more or less. The detailed discussion for the fatigue performance of these

73
TFMG-substrate materials can be found in Ref. [97]. Furthermore, the film properties

can affect the film/substrate adhesion in the film/substrate system, which will result in

the dramatic difference in the fatigue behavior of the film/substrate system.

7.3. Mechanical Properties of TFMGs

The basic nanomechanical properties of TFMGs, including hardness and

Young’s modulus, were obtained, employing the classic Berkovich nanoindentation

approach. One of the reported composition of TFMGs is Zr 53Cu29Al12Ni6 (at.%),

which were deposited on a (100) silicon (Si) substrate with four different thicknesses

(400, 600, 800, and 1,000 nm) [191]. Moreover, compression tests assisted with FIB

were reported to reveal the possible size-and-substrate effect on the mechanical

behavior of the TFMG-based micropillars [191].

7.3.1. Hardness of TFMGs

To measure the TFMGs’ hardness (Hf) and elastic modulus (Ef), a sequential

nanoindentation approach was employed, which uses the load function consisting of

numerous loading/holding/unloading cycles, as exhibited in Figure 36(a) [191].

Following the Oliver-Pharr’s (O-P) method [205], the apparent elastic modulus of the

thin film was obtained after each load cycle, which increases with the indentation

depth (h) [185]. By fitting the curves of Ef vs. h to the modified King’s model [206],

with the Poisson’s ratios of the TFMG and Si being taken as 0.36 and 0.28,

respectively, the intrinsic moduli of both TFMG and Si can be extracted

simultaneously, as shown in Figure 36(b) [191]. It can be observed that the Young’s

moduli of the TFMGs remain almost constant at ~ 117 GPa, while the Young’s

74
modulus of the Si substrate was ~ 183 GPa, which agrees with the reported data [207-

209].

Note that the hardness (~ 5.5 GPa) of the Zr-based TFMGs in Figure 36(c)

[191] is unexpectedly low, about ~ 20% smaller than that (~ 6.2 GPa) of the

corresponding Zr-based BMGs. The interesting point is that this apparent film

“softening” phenomenon occurs not only to Zr-based TFMGs but also to other based

TFMGs. To highlight this trend, the hardness-versus-modulus data of different types

of Zr-based TFMGs can be plotted, as illustrated in Figure 36(d) [191], which shows

that the TFMG hardness is all below that of the BMGs with a similar modulus. As

analogous to the known relation between the strength (σy) and Young’ modulus (E) of

metallic glasses [210], i.e., σy = 0.02E [Figure 36(d)].

7.3.2. Microcompression of TFMGs

Based on our nanoindentation results, the hardness of the Zr-based TFMG (~

5.5 GPa) is about half of that of the Si substrate (~ 12.5 GPa). This trend implies that

the microcompression could be used to supplement nanoindentation for further

studying the mechanical behavior of TFMGs.

(a) Yield strengths of TFMGs

Figure 37 gives the SEM micrographs of the TFMG/Si composite micropillars

with different sizes and shapes for microcompression tests [191]. From

microcompression tests, the elastic moduli and yield strengths of the TFMGs can be

measured. Figure 38(a)-(d) display the typical load-displacement curves obtained

from the micropillars of different sizes and shapes [191]. Using the linear portion of

these curves, the elastic modulus of the TFMG-substrate micropillar could be

75
extracted after accounting for the tapering of the micropillar and the substrate

compliance [211]. Then, the elastic modulus of the TFMG top layer can be obtained

by subtracting the stiffness of the Si substrate out of the total stiffness of the

micropillar. Following the already established method [185], the Young’s moduli of

the TFMGs were obtained at ~ 108 GPa, agreeing with the previous nanoindentation

result.

Since the Si under-layer is much stronger than the TFMG top-layer [Figure

36(c)], the yield strengths, σy, of the TFMGs can be extracted directly from the load-

displacement curves of the micropillars. Figure 39(a) displays the yield strengths vs.

the tf/D ratio (tf and D are the film thickness and pillar diameter, respectively) [191],

from which it can be observed that yield strengths of TFMGs can be viewed as a

constant (~ 2.6 GPa) for tf/D > 0.5, regardless of the film thickness. However, σy

increases with decreasing the tf/D ratio, when tf/D < 0.5. Along the same line for the

indentation hardness, the constant value of ~ 2.6 GPa was taken to be the intrinsic

yield strength of the TFMG. In Figure 39(b) [191], the data of σy vs. E obtained for

TFMGs are plotted together with those of the monolithic metallic glasses, from which

it can be observed that all experimental data collapse onto the linear curve of σy =

E/50. The current finding strongly implies that the previous film softening, as

revealed by the nanoindentation experiments, should stem from a tensile residual

stress in TFMGs, which facilitates yielding in the TFMGs-Si system, but is released

after FIB milling and plays no role in microcompression.

(b) Post-yielding behavior of TFMGs

76
The post-yielding behavior of metallic glasses under compression is usually

characterized by a serrated load-displacement or pop-in response. For a fixed film

thickness (such as 600 nm), pop-in can be observed at the tf/D ratio of 2.0, as shown

in Figure 38(a). However, no evident pop-in can be found for other cases with the

corresponding tf/D ratio of 1.2, 0.3, and 0.1, respectively, as shown in Figure 38(b)-

(d). After the microcompression tests, the deformation morphologies of the

micropillars are examined, using the high-resolution SEM, as exhibited by the insets

of Figure 38(a)-(d), in which discernible shear banding can be found on the deformed

micropillars even though there is no existing prominent displacement pop-in.

Moreover, significant work hardening could be observed in the load-displacement

curves of micropillars with a low tf/D ratio, or when the plastic flow in the micropillar

with a large tf/D ratio has progressed to a later stage.

7.3.3. Extraordinary plasticity of TFMGs at room temperature

Four-point bending fatigue tests on Zr50Cu30Al10Ni10 (at.%) BMGs coated

with a layer of 260 nm thick Zr60Cu24Al11Ni5 TFMG carried out at room temperature

and a strain rate of ~ 10 s-1 show that the TFMG can experience ~ 4000% shear strain

[199]. Cross-sectional transmission electron microscopy (XTEM) analyses reveal no

formation of any shear bands, cracks, and delamination, indicative of homogeneous

plastic deformation. Likewise, conducting the Berkovich nanoindentatoin tests on the

same TFMG/BMG system results in a uniform thickness reduction of 61% without

any sign of the inhomogeneous plastic flow, as confirmed by XTEM analyses and the

fact that almost no pop-in events but the first yielding one are noticed on the

measured load-depth (P-h) curve [199]. On the top of the extraordinary plasticity,

77
another direct result of the homogeneous deformation in the TFMG is its higher

hardness in comparison to the BMG with shear-banding softening.

The absence of shear bands and presence of homogeneous plastic flow in the

TFMG after undergoing drastic plastic deformation in both the fatigue cycle and

nanoindentation tests are presumed to be a consequence of confining the thickness of

TFMG below the critical shear-band nucleation size, which may range from ~ 100 –

500 nm [199]. Although this reasoning seems plausible, the explicit determination of

the critical shear-band nucleation length scale in the TFMG under investigation is

lacking. Conclusive evidences should, therefore, be sought in future efforts.

7.3.4. Nano-Scale FEM on TFMG-Substrate Materials

At the macroscopic scale, a geometric constraint hinders shear-band

propagation in BMGs, thus improving their plasticity through shear-band

multiplication [97, 212]. Here, the existence of a geometric constraint at the nano-

scale is investigated on the bonded TFMG/Si interface [191].

As TFMGs are pressed against their underlying substrate, a hydrostatic

pressure builds up with its magnitude varying with the tf/D ratio. To gain a

quantitative insight into the variation of such a hydrostatic pressure, the FEM of the

elastically-deformed TFMGs is performed. Assuming an intact TFMG/Si interface,

we can simplify the FEM by having the upper surface of the TFMG as a free end,

while the lower surface as a fixed end. Moreover, the diamond punch was rigid when

it is in contact with the TFMG. Figures 40(a)–(d) [191] display the simulated

distribution of the hydrostatic pressure in the TFMGs at four different tf/D ratios (1,

0.5, 0.25, and 0.125) at the same elastic strain of ~ 2%, in which a large high

78
hydrostatic pressure is observed at the TFMG/Si interface, spreading out to the bulk

of the TFMG. With the tf/D ratio decreasing from 1 to 0.125, the average hydrostatic

pressure is increased from ~ 0.5 GPa to ~ 2 GPa. To a certain extent, these numerical

results explain the rising trend of the yield strengths measured in microcompression,

as seen in Figure 39(a). Using the Drucker-Prager (DP) model, the average

hydrostatic pressure can be extracted in TFMGs, which was done by taking σy = 2.6 ±

0.3 GPa as the intrinsic yield strength free of the substrate effect. In such a case, the

hydrostatic pressure, p, exerted by the Si substrate can be simply expressed as

1 f
p ( l   y ) , (7-1)
f

where denotes the substrate-affected yield strength. As seen in Figure 40(e) [191],

the estimated hydrostatic pressure, p, emerging at the yielding point of the TFMG

increases from ~ 0.5 to ~ 1 GPa, as the tf/D ratio of the micropillar decreases from ~

0.2 to ~ 0.1.

8. Fracture Toughness of BMGs and MGMCs

Compared with amorphous alloys, crystalline materials exhibit ordered

structures with morphological features, such as grains. Under deformation, the defects

associated with these structural features (dislocations) become mobile, resulting in the

high fracture toughness. However, the elastic-energy threshold to activate these

defects is often low, leading to low yield strengths. For example, Ni-based alloys

(ductile metals) have the high fracture toughness (> 100 MPa m1/2), but a fairly-low

plastic yield strength (< 500 MPa), as shown in Figure 41 [213, 214]. By contrast,

amorphous materials could potentially yield at much higher strengths, due to the lack

79
of microstructural defects, but with the relatively-limited plasticity. In this respect, the

toughness and strength are invariably mutually exclusive in many classes of materials

[215]. Therefore, achieving the combination of the good strength and toughness of

materials remains a great challenge.

Unlike brittle oxide glasses, metallic glasses are generally likely to yield

plastically under an opening stress, leading to substantial fracture toughness for most

metallic glasses. The reported toughness-strength data for metallic glasses bridge the

gap between brittle ceramics and marginally-tough metals, as shown in Refs. [45, 213,

214, 216]. To be specific, the reported fracture-toughness values vary from ~ 1 MPa

m1/2 (for brittle rare-earth and ferrous BMGs) [217, 218] to ~ 100 MPa m1/2 (for

tougher noble BMGs) [219, 220], while the reported strengths range from ~ 0.5 GPa

(for weak rare-earth BMGs) [217] to ~ 5 GPa (for strong ferrous BMGs) [35], as

shown in Figure 41.

From the mechanistic point of view, when an opening stress of the order of

the material yield strength is applied, plastic shear sliding appears, confined within

shear bands at the nano-scale, usually oriented along planes of the maximum resolved

shear stress. Such shear bands propagate by slip, which can evolve into opening

cracks at some critical shear strain. The shearing can be sustained until low-density

configurations develop, and critical cavities eventually appear. It is usually believed

that large fracture-toughness values are theoretically possible for BMGs with a good

capacity of undergoing multiple configurational shear rearrangements before forming

critical cavities [214]. To this end, the BMGs seem to exhibit such a capacity, since it

demonstrates an unusual propensity for shear flows without cavitation, which leads to

80
very high fracture toughness of BMGs and will be discussed in depth in the following

sections.

8.1. Assessment of Fracture Toughness of BMGs

The measurement of the fracture toughness of brittle BMGs is straightforward.

One can use either fatigue pre-cracked specimens to measure values or notched

specimens without pre-cracking to give notch toughness [221]. If hoping to avoid

intricate sample preparation and fracture mechanics testing, indentation tests are

available for providing a rapid evaluation of fracture toughness. The indentation

fracture-toughness scales with the length of cracks emitting from the corners of

indents through , where is the indentation load, is the half

crack length, is the Young’s modulus, and is the hardness [218, 221]. Yet, the

evaluation of the fracture toughness of BMGs with good plasticity can be extremely

challenging, because meeting the fracture-mechanics requirements for the linear-

elastic K-field dominance and the development of plane-strain conditions have a

demand on specimen sizes. For example, to measure a linear-elastic fracture

toughness KIc value of ~ 200 MPa m1/2 (KIc is the critical value of the stress intensity,

K, at crack initiation, or fracture instability), sample dimensions larger than 45 mm

will be considered to be valid, which is in excess of the critical casting thickness of

most BMGs. To satisfy the small-scale yielding conditions while still properly

characterizing the extension of the crack, a crack-tip opening displacement (CTOD)

approach can be employed. Specifically, the relationship between the J-integral (the

nonlinear strain-energy release rate) and  t (CTOD) is used, given by

81
J  dn 0 t , (8-1)

where  0 is the flow stress, and d n is a constant tabulated from the strain-hardening

exponent, n, of the materials.

The K-based fracture-toughness values were back-calculated from the J

measurements, using the standard Mode-I J-K equivalence

K J  [ JE / (1  v 2 )] . (8-2)

Then, the extensive crack-growth resistance of MGMCs can be determined by

the equivalent stress intensity, KJ, instead of R curves.

The CTOD approach has been verified with a ductile-phase-reinforced

metallic-glass composite, also exhibiting extensive plasticity [214]. Good agreement

was obtained between these two measurement techniques. Accordingly, the CTOD

approach will be used to determine the fracture toughness of a metallic glass with

critical thickness below the width required for direct J-integral toughness

measurements.

8.2. Fracture Toughness of Monolithic BMGs

The metallic glass, a Pd-based (Pd79Ag3.5P6Si9.5Ge2, at.%) alloy, is used for an

example, showing a small degree of apparent hardening in bending. The Mode-I

(tensile-opening) fracture toughness of the glassy Pd79Ag3.5P6Si9.5Ge2 was determined

in the single-edge/notched-beam bending geometry using micro-notched beams.

Evaluation of toughness requires the nonlinear elastic fracture mechanics to

characterize the resistance curve (R curve) analysis, which is experimented by the  t -

R curve [Figure 42(a)] [214]. Results for the stress intensity, KJ, back-calculated from

82
the J measurements are shown in Figure 42(b), which is characterized by the KJ-R

curve. The BMG demonstrates the extensive rising R-curve behavior, indicative of

stable crack growth over hundreds of micrometers. A near-steady-state fracture

toughness measured in terms of the stress intensity is ~ 200 MPa m1/2, which is an

exceptionally high value for any material.

The mechanisms of toughening in the metallic glasses can be examined by

carrying out the fracture-toughness tests in the SEM, which can monitor the evolution

of damage ahead of the crack tip. The high toughness is obtained by stabilizing the

plastic-flow processes at the opening crack tip to form a distributed damage zone

accompanied by significant plastic shielding. The specific mechanisms contributing

to the good toughness of the glassy Pd79Ag3.5P6Si9.5Ge2 (at.%) can be described in

terms of a three-step process. First, shear bands form along the fan-shaped (Prandtl-

field) slip lines that bend back toward the crack plane [Figure 42(c)-(e)] [214]. With

the development of the Prandtl-field, more localized shear sliding appears, leading to

very large shear offsets [Figure 42(f)]. Furthermore, when a critical sliding strain is

reached with increasing the load, an extended shear band evolves to a crack at the

notch tip [Figure 42(g)]. Extensive shear bands can be noticed to persist ahead of the

crack tip, which leads to evident crack-tip blunting [Figure 42(h)]. Finally, as the slip

bands bend back to the crack plane, enabling substantial shear sliding, the crack

remains stable on its plane, which promotes stable crack extension during fracture

[Figure 42(f)-(h)].

Some other metallic glasses, e.g., Mg-based BMG, may fracture in a nearly

ideal brittle manner, exhibiting a fracture toughness of ~ 2 MPa m1/2 [217]. As the

83
fracture toughness of metallic glasses varies from very small to very large values, it

was recognized that the vein pattern manifested on the fracture surfaces of all samples,

tough in different length scales [217]. By measuring the wavelength of the vein

features (i.e., the average width of the sampled dimples) as the plastic-zone size, w,

Xi et al. correlated the plastic zone size with the fracture toughness, KIC, by

, where is the material yield strength [217]. On the other hand, some

other authors believed that the area fraction of the river-like pattern, rather than the

vein pattern, on the fracture surface of BMGs is proportional to the macroscopic

plasticity, even though a detailed quantification was not provided [222]. The rationale

behind is that the river-like pattern is closely related with the secondary or tertiary

shear bands generated on the fracture surface [222]. In another line of work of

characterizing the fracture behavior of the Zr 57Ti5 Cu20Ni8Al10 (at.%) metallic glass at

the temperatures ranging from 128 K to 298 K, Hufnagel et al. showed that the stress

intensity factor at fracture (not true fracture toughness) and the plastic-zone size more

or less followed a standard relation [223]. In their work, the plastic

zone size was measured by virtue of in situ high-energy X-ray scattering during the

course of fracture testing. The approach starts with mapping out the elastic strain field

around the crack tip by computing the peak shift in the recorded X-ray scattered

intensity, from which the stress field and different stress invariants are calculated,

using the elastic constants. The plastic-zone size is estimated to be the extent of the

region over which the von Mises stress reaches a plateau value, due to the fact the X-

ray scattering merely maps the elastic-stress field. The shape of the plastic zone

estimated in such a way exhibits a pair of butterfly-shaped lobes, whose extent is in a

84
rough agreement with that from the constitutive prediction [224]. Quantitatively, it is

calculated as the radius of gyration of the region over which the von Mises stress

reaches the plateau value, i.e., , where N is the number of

pixels, is the position vector of the ith pixel at which the von Mises stress has

reached the plateau, and is the average position vector for all N pixels [223].

8.3. Fracture Toughness of Metallic Glasses Composites

Compared to BMGs, the toughness of metallic glasses matrix composites

(MGMCs) can be significantly improved, since the essential element to develop high

toughness metallic glasses is to prevent single shear-band failures upon loading. By

introducing a second phase in the form of crystalline dendrites, crystalline phases in

the MGMCs act as barriers to arrest shear-band propagation over appropriate size

scales, leading to the increased ductility, toughness, and fatigue resistance [113, 225-

228]. In the literature, Launey et al. [229] employed a nonlinear-elastic fracture-

mechanics (J-integral) characterization, including the resistance-curve (R-curve)

determination, to estimate the fracture resistance of MGMCs and to interpret behavior

according to the prevalent mechanisms of microstructural damage and toughening.

In the current review, two compositions of MGMCs [Zr 36.6Ti31.4Nb7Cu5.9Be19.1

(DH1, at.%) and Zr39.6Ti33.9Nb7.6Cu6.4Be12.5 (DH3, at.%)] are employed for the

toughness studies with the volume percentages of second-phase dendrites (-phase)

of 42% and 67%, respectively [225]. R-curves were measured in terms of the J-

integral to study the fracture resistance as a function of crack extension, Δa, under

monotonical loading. Specimens were fatigue precracked, and then loaded under a

85
displacement control in three-point bending. After the onset of cracking, crack

extension was monitored from the elastic unloading load-line compliance [230].

As shown in Figure 43, the MGMCs exhibit evident rising R curves,

compared to monolithic Vitreloy 1. It can be noticed that the immediate unstable

fracture occurs at a stress intensity of 54 MPa m1/2 in Vitreloy 1, while for DH1 and

DH3 alloys, R-curves are initially steep (from crack-tip blunting), followed by stable

cracking until the steady “plateau” is reached at ~ 155 and 200 MPa m1/2, respectively.

In these measures, the composites are ~ 3 - 4 times tougher than the metallic-glass

matrixes. For comparison with metallic glasses and MGMCs, stainless steel is also

listed in Figure 43. However, these data of metallic glasses are limited by their

specimen-size requirements for the J dominance. According to the American Society

for Testing and Materials (ASTM) Standards [231], the maximum J capacity for a

specimen is given by the smaller one of Jmax = bσy/10 or Bσy/10 [b is the uncracked

ligament, B is the specimen thickness, and σy is the flow stress] and crack extension

cannot exceed Δamax = 0.25b, beyond which the R-curve data are invalid [229]. As the

samples size of both DH1 and DH3 met this criteria, the fracture toughness of the

DH1 alloy can be accurately ascribed to be JQ = JIc = 96 kJ/m2, equivalent to a KJc of

97 MPa m1/2, and the DH3 alloy to be JQ = JIc = 282 kJ/m2, equivalent to a KJc of 157

MPa m1/2. Consequently, in comparison to monolithic BMGs, where KIc can be as

low as 20 MPa m1/2 [70, 162, 232], these MGMCs exhibit toughnesses that are

significantly higher than monolithic BMGs, and further are comparable to the

toughest crystalline metallic alloys (stainless steel) at these high strength levels.

86
During the steady-state crack growth, the increased fracture resistance is

associated with the extensive plastic zone [Figure 44(a)] and a wide damage

distribution around the crack tip [Figure 44(b)). It can be noticed that in Figure 44(c),

plastic deformation occurs through the development of highly-organized shear bands

distributed uniformly along the crack path until they are blocked by the crystalline

dendrites [Figure 44(c)].

In order to develop the high toughness of metallic glasses, the essential is to

prevent single shear-band failures, which can be achieved through the confinement of

the shear bands as a result of the mismatch in the plastic response. As reported in Ref

[229], the characteristic microstructural length scale has to match the critical

mechanical length scale, the crack size for fracture. Thus, single shear bands can be

arrested before failure, which leads to the formation of multiple shear bands, and

therefore enhanced plasticity and toughness. Take the DH3 alloy for an example, by

constraining the initial deformation band, the efficient multiple shear-band formation

results in extensive plastic shielding and improved fracture toughness (~ 200 MPa

m1/2), which is almost one order of magnitude higher than the monolithic BMG

(Vitreloy 1).

According to the above results, the values of the fracture energy and

toughness of BMGs and MGMCs are comparable to those for the toughest

engineering metals (e.g., low-carbon steels). Considering that the metallic glass lacks

microstructural defects, such as dislocations, which can rearrange to shield the stress

and suppress crack opening, it is quite remarkable to achieving such high fracture

resistance. Thus, an unusual combination of very high strength and toughness is

87
possible - a feature perhaps unparalleled by most monolithic materials. In Figure 41

[213, 214], an Ashby map showing toughness-versus-strength ranges for oxide

glasses, engineering ceramics, engineering polymers, and engineering metals, as well

as for monolithic metallic glasses and ductile-phase-reinforced MGMCs. As shown in

this toughness-versus-strength map, the metallic glasses lie outside the benchmarks of

the strongest and toughest steels. In summary, the combination of toughness and

strength, which is the level of damage tolerance, can be potentially accessible for

metallic glasses, which extends beyond the limiting ranges towards the levels

previously impossible for many traditional materials.

9. Ductile vs. Brittle Fracture in BMGs

One of the most mysterious questions in BMGs is their widely-varied fracture

behavior, i.e., some may exhibit a great deal of plastic flow locally or globally upon

failure [45, 219] whilst others fracture in a perfectly-brittle manner with the fracture

toughness as low as 2 MPa m1/2 [44, 45, 217, 233]. The remarkable fluctuation in the

fracture mode is not only found in compositionally-different BMGs but also those

with identical compositions but subjected to dissimilar processing histories. Figure 45

gives the SEM images of the fracture surfaces of two compositionally-different Cu-

based BMGs failed by the uniaxial compression tests [234]. In Figure 45(a), the

Cu48Zr48Al2 Ag2 (at.%) exhibits typical a vein-like fracture pattern alongside plenty of

melted liquid beads, which signals that the sample is fractured in a ductile manner via

the extensive shear strain and localized melting. On the other hand, the fracture

surface of the Cu40Zr40Al10Ag10 (at.%) BMG in Figure 45(b) manifests the

88
characteristics of the cleavage fracture, indicative of brittle failure. It is easily seen

that a slight change in the chemical composition can transform BMGs from the

ductile to brittle failures.

Most of BMGs possess a certain degree of ductility in the as-cast state, but as

they are annealed at elevated temperatures, the ductility gradually loses, and the

fracture becomes favoring brittleness. The longer the annealing time is, the more

brittle the samples turn. When the annealing temperature is in excess of the glass-

transition temperature, Tg, the embrittlement usually resultes from partial

crystallization, which can be easily detected and confirmed through X-ray diffraction

or TEM scanning [233, 235]. Nevertheless, even when the crystallization is

suppressed, namely annealed BMGs at the temperatures below Tg, the ductile-to-

brittle transition can still be observed. Figure 46 shows the bending fractographs of

the Zr52.5Cu17.9Ni14.6Al10Ti5 (at.%) metallic glass (Tg ≈ 400 oC) in the as-cast state and

annealed at 300 oC for various durations [40]. The fracture surface of the as-cast

sample in Figure 46(a) exhibits a typical vein pattern and localized melting, which is

characteristic of ductile fracture. Extensive dimples are noticed on the fracture surface

of the 21-hour annealed sample in Figure 46(c), and a mixture of vein pattern and

dimples feature the 9-hour annealed sample in Figure 46(b). In the case of 168 hour-

annealed sample in Figure 46(d), a mirror-like featureless fracture surface presents.

Since it is known that in BMGs, the vein pattern is a result of the fluid meniscus

instability in the ductile fracture while dimples are caused by the nano-scaled

cavitation in the brittle fracture [236, 237], the tendency of the fracture characteristic

changing from the vein pattern to dimples and ultimately to the mirror-like surface is

89
a clear indication of the fracture mode switching from the ductile manner gradually to

brittle manner. The embrittlement of BMGs upon annealing below Tg is believed to

be associated with the structural relaxation [233], which will be elucidated in the

following sections.

9.1. Empirical Interpretation with Elastic Constants

In polycrystalline metals, the ratio of the shear modulus, µ, to the bulk

modulus, B, is suggested to be a good indicator of the intrinsic plasticity or brittleness

[238, 239]. The high µ/B favors brittleness and vice versa. And the mechanistic

reasoning is relevant to the competition of the dislocation-mediated plastic

deformation resisted by µ and the brittle transgranular or intergranular fracture

through the tensile separation of non-close-packed atomic planes, which is resisted by

B. Though a very different deformation mechanism appears in BMGs, as opposed to

their crystalline counterparts (i.e., shear softening mediated by the operation of shear

bands vs. working hardening regulated by dislocation motion), their brittleness or

plasticity is advised to still be able to be clued from µ/B, or equivalently the Poisson’s

ratio, . Lewandowski et al. [45] complied the elastic constants and fracture data of a

large amount of as-cast bulk metallic glasses with dissimilar compositions, and

compositioanlly identical but differently-annealed ones. For both categories, by

plotting the fracture energy as a function of the µ/B and , they found a clear

separation of the ductile and brittle regime, as shown in Figure 47. Figure 47 suggests

that BMGs with µ/B > 0.41 – 0.43 or < 0.31 – 0.32 are brittle and vice versa. This

rule is sort of empirical since it is a loaned concept from crystalline materials and lack

of persuasive mechanistic explanations in spite of attempts [240, 241].


90
While the above rule holds for many BMGs [44, 240, 242], some exceptions

have been reported. Li et al. reported a sharp ductile-to-brittle transition in the as-cast

and variously annealed Zr52.5Cu17.9Ni14.6Al10Ti5 (at.%) BMGs even when µ/B varies

slightly from 0.277 to 0.296 or 0.373 to 0.365 [40]. Similarly, Kumar et al.

observed the severe embrittlement of Vit-1 and Vit-4 BMGs after annealing at the

temperature below Tg, but the change in µ/B is marginal and remains smaller (0.293 –

0.337) than the proposed critical value [233]. Raghavan et al. found that the annealing

can bring Zr41.2Ti13.8Cu12.5Ti10Be22.5 (at.%) metallic glasses into brittleness but the

in all samples still remain greater (0.346 - 0.358) than the proposed critical value

(0.31 – 0.32) [243]. Also, both the brittle Pd43Cu27Ni10P20 (at.%) metallic glasses

rendered from low cooling rates during casting, and the ductile Pt 57.5Cu14.7Ni5.3P22.5

metallic glasses retain similar µ/B (~ 0.2), which is well below the critical value of

0.41 – 0.43 [244]. All these findings put the universal applicability of the µ/B or

criterion into suspicion. Given that it is an empirical rule without the solid physical

basis, its practical applications would be confined, and the physics-based criteria that

can mechanistically explain the ductility/brittleness of metallic glasses is in need.

9.2. Microstructural Rationalization with the Heterogeneity Concept

In polycrystalline metals, the ductile-to-brittle transition is straightforward to

understand from the microstructural point of view. Ductility or brittleness in these

materials is associated with the collective mobility of dislocations, namely great

dislocation mobility leads to considerable plastic flows, and less mobile dislocations

end in the cleavage fracture. However, this uncomplicated task becomes very intricate

in their amorphous counterparts, metallic glasses. Due to the amorphous nature and
91
lack of obvious microstructural deformation mediation mechanism as in crystalline

materials, it is extremely challenging in BMGs to directly measure any

microstructural features that are linked with their macroscopic ductility or brittleness.

Researchers have been strived long to identify feasible microstructural features that

are applicable for determining the intrinsic ductility or brittleness of BMGs.

Fortunately, through numerous studies, it is now widely acknowledged that the

microstructure of metallic glasses is intrinsically heterogeneous on the scale from

nano and micro, and these heterogeneities are nailed down as the structural reason

dictating the ductile or brittle fracture behavior of metallic glasses [245-247]. The

structural heterogeneity controls the fracture mode of BMGs through the scenarios

depicted in Figure 48 [40]. The metallic glass is thought of consisting of a glass-type

hard matrix and randomly distributed heterogeneities, which are relatively soft. The

crack propagation in the glass-type matrix can advance swiftly with little impedance,

inclining to lead to the catastrophic brittle failure. Conversely, the heterogeneities are

of ductile nature and have a great capability of blunting a sharp crack and makes the

fracture ductile. If a BMG would fail in a ductile or brittle manner relies on the

competition of these two conflicting deformation mechanisms. When a metallic glass

sample is in the as-cast state, a relatively-large amount of heterogeneities disperse in

the glass-type matrix, and the deformation is primary commanded by the ductile

heterogeneities, and the macroscopic ductility will be present, as depicted in Figure

48(a) [40]. Annealing the as-cast metallic glass tends to gradually annihilate the

structural heterogeneities. As the annealing temperature or time increases, the

heterogeneities will eventually become depleted, and the sample transforms to a pure

92
glass state in which the fracture turns into be brittle with near-zero macro-plasticity.

This trend is the scenario described in Figure 48(c). For those intermediately

structurally relaxed, the heterogeneities still exist but are not as dense as those in the

as-cast state. Therefore, their effectiveness for preventing the brittle failure of

metallic glasses is not as good as that in the as-cast state, resulting in the limited

ductility. The widely scattered ductile or brittle behavior in the as-cast metallic

glasses of different compositions can be rationalized by this model. Different

compositions and cooling rates during casting will incur varied densities of structural

heterogeneities, which thereafter lead to either ductile or brittle fracture in a similar

fashion as delineated in Figure 48.

Due to the difficulty of the direct quantitative characterization of nano or

micro-scale structural heterogeneities, they can be firstly probed in the form of

mechanical heterogeneity using statistical nanoindentation tests, followed by linking

the mechanical heterogeneity back to the structural heterogeneity with a stochastic

structural model developed by Li et al. [39, 40]. The structural model unifies both the

homogeneous shear band nucleation process (thermally activated) and heterogeneous

shear band formation process (defect-assisted), and is given in the form of

ftotal  1  exp  A0 Ei  v* max


pop in
kBT   def Vs  , (9-1)

*
where v is the activation volume, k B is the Boltzmann constant, T is the absolute

13
 16 Ppop in Er2 
temperature,  max  0.445 
 9 3 R 2 
is the maximum shear stress at the first pop-
 

in of nanoindentation tests, Ei  x    t 1et dt


x
is the exponential integral,


93
  
A0  3n0V
P
exp    with n0 being a pre-factor,  the intrinsic nucleation
P  B 
k T

energy barrier, the constant nanoindentation loading rate, and V the stressed

material volume, Vs is the sampling volume with Vs a  Vˆs  def  max


3
  ,

13
 3PR 
a  is the contact radius,  def is the strength of defects (i.e., structural
 4 Er 

heterogeneities), and  def is the defect density.

The mechanical heterogeneities in the as-cast and variously annealed

Zr52.5Cu17.9Ni14.6 Al10Ti5 (at.%) BMGs measured with statistical nanoindentation tests

equipped with a 3.80 μm radius spherical indenter [248] are shown as symbols in

Figure 49(a). By fitting the Eq. (9-1) to the measurement data, one can quantitatively

extract the density of heterogeneities,  def . The extracted heterogeneity density,  def

, as a function of the annealing temperature and time is plotted in Figure 49(b). It is

clearly seen that increasing the annealing temperature and time are two equivalent

ways to facilitate the embrittlement of metallic glasses, which is in accord with

experimental observations [233, 234, 243]. The extracted heterogeneity density can

then be correlated back to the global fracture-energy density measured by notched

three-point bending tests or any other tests on the same samples. Figure 50 is a plot of

such kind. It is found that the global fracture energy density is related to the

 
heterogeneity density via an exponential relationship Etotal  c1 exp c2 def , where c1

and c2 are fitting parameters. This exponential relation allows one to associate the

macroscopic material property (the fracture-energy density) with the microstructural

94
parameter (the heterogeneity density) of metallic glasses, and can serve as an

effective way to predict the failure nature of a metallic glass (ductile for brittle) once

knowing its heterogeneity density.

In addition to the nanoindentation, the structural heterogeneities in metallic

glasses can be also probed with many other techniques, either qualitatively or

quantitatively. Firstly, the color contrast in TEM images can imply the existence of

heterogeneities, and these heterogeneities may either possess chemical-composition

fluctuations or not, relative to the matrix depending on the processing histories [235,

249]. Figure 51(a) show a micro-scale heterogeneity observed in such a way [249].

The heterogeneities can be also measured by utilizing the amplitude-modulation

dynamic atomic force microscopy (AM-AFM). Figure 51(b) presents the AM-AFM

mapped phase shift image of the hype-quenched Zr53Cu36Al11 metallic glass. The

heterogeneous intensity distribution in the phase-shift image corresponds to the varied

viscoelasticity in the metallic glass, and is a direct signal of the spatial heterogeneities

caused by structure variation [41, 250]. Note that the surface roughness variation in

the inhomogeneous phase shift observed is ruled out [41, 250]. Besides, molecular

dynamics simulations are also found to be useful for gaining insights into the

characteristics of structural heterogeneities [251-254]. For instance, it is shown in

Figure 51(c) that in the Cu50Zr50 BMG, the heterogeneities appear as approximately

elliptical shapes, and different heterogeneities tend to interconnect in a way similar to

“grain boundaries” in crystalline materials [251]. Synchrotron radiation x-ray

scattering is another technique that can be utilized to investigate the structural

heterogeneities [255]. Figure 51(d) shows the reduced pair distribution of the

95
Al88Ce8Co4 (at.%) BMG mapped on two distinct local regions (discs of the diameter

of 100 nm), C and IC regions. Clearly, the C-region has more loosely-packed atomic

structure, compared to the IC-region, in correspondence to the structural

heterogeneity. Since the β-relaxation in BMGs stems from localized atomic motions

confined to isolated regions, the structural heterogeneities can be inquired by

measuring the β-relaxation with dynamic mechanical analysis or others [250, 255].

Figure 51(e) gives the internal friction detected by the dynamical mechanical

analysis, in which the reduced cooling rate leads to the more pronounced β-relaxation

moving towards lower temperatures, indicating an increased amount of

heterogeneities [255]. Moreover, Sun et al. developed a micromechanical model on

the basis of Eshelby’s theory used for quantitatively-calculating the volume fraction

of heterogeneities in BMGs by referencing their measured elastic constants. The

example given in Figure 51(f) is the volume fraction of the structural heterogeneities

calculated for the Vit105 BMGs annealed for various times, based on the measured

shear modulus [256].

9.3. Design Ductility in BMGs

BMGs, though exceptional in strength and elastic limit, find very limited

structural applications even after a few decades of unremitting efforts from academia

and industries. One of the major reasons is that most of them exhibit very limited

ductility. Designing BMGs with considerable ductility has been long pursued for the

sake of prompting their sophisticated engineering applications. On one hand, the

ductility can be enhanced by extrinsic means, such as introducing geometric

constraints or second phases [85, 86, 97, 228, 257-259]. Nonetheless, a more

96
promising and universal approach is through intrinsic toughening. Growing endeavors

on revealing the factors governing the intrinsic ductility of metallic glasses have shed

light on practically selecting and designing ductile ones. The elastic constant ratio,

µ/B, criterion (or the Poisson’s ratio, , criterion), though having limited preciseness,

can still be used as an empirical principle for selecting and designing intrinsically-

ductile BMGs. As a matter of fact, successful applications of this principle have been

reported in the literature. For example, Schroers and Johnson reported [219] that a

plastic strain of 20% was measured in Pt-based BMGs with a high Poisson’s ratio of

0.42. Furthermore, since it is already widely accepted that the density of structural

heterogeneities is the microstructural reason governing the macroscopically brittle or

ductile fracture behavior of BMGs, designing heterogeneities in a controllable

manner is a more robust way to produce ductile metallic glasses. Li et al.

demonstrated the success of this path by mechanically-imprinting BMGs to produce a

heterogeneous structure of hard and soft regions [260]. Overall speaking, the

application of this principle is still immature, given that a comprehensive picture of

how the structural heterogeneities affect the ductility is still incomplete. More

investments on fundamentally understanding and verifying the interconnection

between the heterogeneities and ductility in metallic glasses is in need prior to large-

scale applications of this principle.

10. Notch Effect on Fracture and Fatigue of BMGs

Engineering materials and structures, when present in a bulk form, oftentimes

contain flaws of various sizes and shapes. In the majority of circumstances, these

97
flaws are where the stress becomes highly concentrated and the material failure

initiates. From the design point of view, it is desirable that the materials with a high

level of flaw tolerance are designed or selected for engineering applications. Notch,

which is a source of stress concentration artificially created, plays important roles in

assessing the flaw tolerance of a material prior to putting it into practical usages. In

laboratory experiments, notches are usually shaped on material samples, followed by

conducting standard mechanical tests (e.g., tensile, compressive, bending, or fatigue

tests), thus permitting the study of the notch sensitivity of materials. Usually, a lower

notch sensitivity implies that a material has a greater tolerance to notches, holes,

grooves, or other geometric discontinuities and vice versa.

For most existing engineering materials, it is generally accepted that brittle

materials will notch weaken, and highly ductile material will notch strengthen [261].

For instance, the strength of notched brittle ceramics (e.g., Ti3SiC2 ceramics) is

substantially reduced, relative to their un-notched counterparts [262]. In contrast,

ductile crystalline metals, such as coppers, show the notch-strengthening effect, i.e.,

notched samples exhibit higher strength than un-notched ones [263]. Apart from the

failure strength, the plasticity of a material could also be altered by notches. For

example, the tensile plasticity becomes lowered in notched copper samples despite

their strength are enhanced [263]. As a matter of fact, the notch effect in either

crystalline brittle or ductile materials is associated with the dislocation-mediated

deformation mechanism [263]. Different from traditional crystalline materials,

metallic glasses, as a class of new engineering materials, possess amorphous

structures and lack dislocation-mediated deformation, which makes their strength and

98
plasticity theory very distinct. The notch effect in these material turns out to be much

more challenging to understand. In the following, we attempt to give a compressive

summary of what has majorly been done in this particular area.

10.1. Notch Effect on Fracture

10.1.1. Notch Sensitivity of Strength

The notch effect on the strength of BMGs have been considerably studied in

recent years but no consensus has been reached till now. In the literature, notch

strengthening, weakening, and insensitivity have all been reported. Notch

strengthening is documented in circumferentially-notched specimens of Pd78Cu6Si16

[264], Zr64.13Cu15.75Ni10.12Al10 [265], and Zr-Ti-Ni-Cu-Be (at.%) [266, 267] BMGs.

The fracture strength of notched samples grows with the increasing ratio of the notch

depth to notch width. Besides, in the thin plate specimens of the Cr 50Zr50 model

metallic glass with symmetric notches, it is noticed that the strength escalates not only

with the increasing ratio of the notch depth to width but also the sharpness of the

notches [268]. Notch weakening is reported in circumferentially-notched specimens

of the Zr41.25Ti13.75Ni10Cu12.5Be22.5 (at.%) metallic glass [269], thin plate specimens of

Zr52.5Ni14.6Al10Cu17.9Ti5 and Ti40Zr25Ni3Cu12Be20 (at.%) BMGs with single-edge

notch, central notch, antisymmetric double-edge notches, and a symmetric double-

edge notches [270, 271], and nano-sized Ni75P25 metallic glass with single-edge notch

[272]. In the plate sample of BMGs with antisymmetrically-inclined edge notches, the

notch strengthening and weakening may switch from one to another as the notch-

inclination angle varies. On the other hand, Qu et al. and Sha et al. reported the notch

99
insensitivity of the strength in the plate specimen of the Zr 52.5Cu17.9Ni14.6 Al10Ti5 and

Zr54.5Cu20Al10Ni8Ti7.5 (at.%) BMGs with double-edge notches [263], and the Cu50Zr50

model metallic glass with varied notch sizes and shapes [273]. Some important results

are summarized in Figure 52.

Excluding the compositional effect, it is noticed that the reported

controversies in the notch effect on BMGs are majorly originated from the notch

configuration and sample geometry. The introduction of notches will make the stress

in a sample deviate from the uniaxial state, and varying the notch depth, width, angle

or sharpness will continue to alter the stress state. The change in the stress state may

in turn leads to a switch of the failure mode. By considering the stress state at the

center of the notched section of the BMG sample, Pan et al. rationalizes the notch

strengthening and weakening with the following model [265]

, (10-1)

where is the hydrostatic stress, is the von Mises equivalent stress, and is a

geometric ratio representing the ratio of the notch depth to notch width. A small

represents a deep notch and vice versa. Since the hydrostatic stress, , and von

Mises stress, , govern the cavitation-failure and shear banding mechanism,

respectively, Eq. (10-1) provides a quantitative description of the competition

between these two failure mechanisms. Then they introduced a material parameter, ,

where and are the critical stresses for cavitation and shear localization,

respectively. If is very large (e.g., > 2), the increase of will never result in a

hydrostatic stress that reaches . Instead, the shear localization will be first reached,

100
and the deformation in such a scenario is featured as shear banding, manifesting notch

weakening. For a small (e.g., ~ 1.4 in Pan et al. [265]), the increase of leads to a

transition from shear localization to void growth and coalescence, the material

exhibiting the notch-strengthening effect. The fracture-surface morphology in Figure

53(a) shows that in the notched BMG sample with a below the critical value, the

failure is characteristic of the shear-banding-governed brittle failure. On the contrary,

as becomes in excess of the critical value, the deformation transforms to the

Mode-I fracture in Figure 53(c). Detailed SEM imaging reveals a rough fracture

surface with numerous dimples and micro-cracks in this fractured sample [265],

which is an implication of the nucleation, growth, and coalescence of voids during

deformation. The transition of the failure mode from the shear banding to cavitation

as the stress state, , varies with the notch geometric factor, , is further confirmed

with molecular dynamic simulations. As shown in Figure 53(b) and (d), plastic

deformation localizes into an inclined shear band in the widely-notched sample (a

large ) but a void forms at the center of the deeply-notched sample (a small ).

Although the notch strengthening and weakening are properly explained under this

framework [265], the notch insensitivity is not touched. Li et al. [85] explores this

trend with the numerically modeling, i.e., the normal fracture and shear-banding-

induced fracture in BMGs are represented with a cohesive interface model and a

weak-zone model, respectively, and their synergistic effects on the notch sensitivity is

investigated [85]. With such an approach, it is pointed out that the large-scale-

bridging (LSB) behavior is responsible for the notch insensitivity in BMGs with the

101
normal fracture along the notch plane being the dominant mechanism. This is

evidenced by the circle- and star-marked curves at ≤ 1 in Figure 54, where

the notch depth, , is comparable or less than the cohesive zone size, ( , ,

and is the Young’s modulus, peak stress, and corresponding separation in the

cohesive-interface model). They also suggest that an additional weak zone of low

strength from the notch roots (i.e., formed shear bands accompanying the normal

fracture) will further reduce the stress concentration and extend the notch-

insensitivity regime, as shown by the diamond-marked curve in Figure 54.

Alternatively, Qu et al. [262] characterized the notch geometry with the stress

concentration factor, , and justify the notch sensitivity of BMGs and other materials.

The data summarized in Figure 55 depicts the relation of the notch strength ratio,

( and are the strengths of notched and un-notched samples,

respectively), and the stress concentration factor, , for BMGs alongside ductile

metals and brittle ceramics for comparison. NSR > 1 indicates notch strengthening,

NSR < 1 implies notch weakening, and NSR ~ 1 means notch insensitivity. It is seen

from Figure 55 that the NSRs of BMGs decrease from 1.25 to 0.75, as the stress

concentration factor, , ramps from 1.0 to 5.5. Notch geometries with small (< 3)

in BMGs result in NSR > 1 and, thus, notch strengthening. On the other hand,

Notches with large (> 3) cause NSR < 1 and notch weakening. At the threshold

value of , NSR = 1 and the strength of BMGs are insensitive to the notches

[262, 274]. Since the NSR of BMGs runs across the NSR = 1 threshold line as

changes, notch strengthening, notch weakening, and notch insensitivity could all be

present in BMGs depending on the notch configuration in test samples, essentially

102
consistent with the scattered data reported in the literature. Furthermore, Figure 55

also confirms the universally-observed notch strengthening in ductile crystalline

metals and notch weakening in brittle ceramics (AZ80A Mg alloys behaves like

BMGs due to their relatively low ductility).

10.1.2. Notch Sensitivity of Ductility

The notch effect on the ductility of BMGs is less controversial, and it is

generally acknowledged that the ductility can be improved with the introduction of

symmetric double-edge notches [85, 86, 270, 271, 275]. An un-notched BMG sample,

under either tensile or compressive loading, usually fails along a dominant shear band,

exhibiting very limited or near-zero plasticity, as shown in Figures 56(a) and (b)

[271]. Introducing symmetric double-edge notches, on the other hand, can improve

the ductility by ~ 10%, as indicated in Figure 56(a). The underlying mechanism is

that notches create a large-scale stress gradient that facilitates the initiation of

multiple shear bands at notch roots as well as impedes that fast propagation of the

generated shear bands. On top of that, if notches are sufficiently deep, shear bands

generated and propagated from two notches may interact with each other, causing

deflection and further deferral of shear-band propagation. Such a transformed

deformation mechanism can be glimpsed in Figure 56(c). Notch toughening in BMGs

is intimately related to the shear-banding mediated deformation. In crystalline

materials in which dislocations are the deformation carrier, the similar notch

toughening is unlikely [271].

Given that the interaction of shear bands is one of the major mechanism

responsible for the improved ductility in symmetrically-notched BMG samples, it is

103
anticipated that the depth of the notches has an impact on the effectiveness of the

ductility enhancement. Li et al. parametrically investigated the effect of the notch

depth on the shear-band arrangement and, therefore, the ductility enhancement. Their

results are shown in Figure 57. Figures 57(a)-(c) are shear-band patterns predicted

from the Rudnicki-Rice model based on the elastic stress field [85, 86, 97, 257], while

Figures 57(d)-(f) are from the finite-element modeling based on the free volume

constitutive law [85, 86, 97, 228, 257, 258]. When the notch depth is small (a = 2

mm), symmetric shear bands are found to radiate from the notch roots in Figures 57(a)

and (d). Since the two notches are far apart, the shear bands radiating from two

notches have no chance to meet and interact each other. As a result, the shear-band

interaction mechanism is absent in this case, and the ductility enhancement is

attributed solely to the generation of multiple shear bands at notch roots and the

blocking effect of shear-band propagation by the large-scale stress gradient. Due to

the absence of the shear-band interaction, the effectiveness of the ductility

enhancement is not optimal. As the notches become deeper, shear bands propagating

from two notches are made to be possibly interacting with each other, and the

ductility of BMGs is greatly enhanced. Such an effect can be seen from Figures 57(b)

and (e) but more obvious from (c) and (f). As two notches become closer, it is also

noted that additional curved shear bands that directly connect two notches appear

from the Rudnicki-Rice instability theory prediction in Figures 57(b) and (c). These

shear bands are caused by the out-of-plane shear according to the stress field analysis

[85] and confirmed by experimental observations [276].

104
It is worth noting that not all notches can help enhance the ductility of BMGs.

For example, a single notch, asymmetric double-edge notch, and central hole are

found to be ineffective in incurring multiple shear bands, and the failure in samples

with such notches still happens catastrophically along a dominant shear band as in un-

notched ones [271]. However, one may intentionally design artificial notches (e.g.,

antisymmetric double-edge notches with different inclination angle [275]) to gain

ductility in brittle BMGs. In general, there is a plenty of room for exploration.

10.2. Notch Effect on Fatigue

For years, it is widely accepted the tensile strength of material is positively

correlated with their fatigue strength [124, 147, 277]. Recent experiments, however,

reported a breakdown of this correlation, finding that the fatigue strength remains

invariable or reduced as the tensile strength of the material is increasing, particularly

at high-stress levels [278]. When notches are present, the fatigue strength of a

material might become even more different from its uniaxial strength. This is to say

that the notch effect on the fracture strength of BMGs summarized in the previous

section is not directly applicable to their fatigue strength. The notch effect in the

fatigue strength of BMGs needs separate treatments. Figures 58(a) and (b) compile

the uniaxial fatigue data for Zr41.2Ti13.8Cu12.5Ni10Be22.5 (at.%) BMGs and Zr-based

BMGs of different compositions without a notch and with different types of notches

[279]. It is noted that for all Zr-based BMGs, the notched samples exhibit higher

fatigue strength and fatigue endurance limit than the un-notched ones, indicative of

the notch-strengthening effect. In addition, taper-notched samples show slightly

superior fatigue performance relative to the sharp-notched ones.

105
In the first place, the notch-strengthening effect of the fatigue of BMGs is due

to the proliferation of shear bands at the notch roots via the large stress gradient,

analogous to the mechanism in the uniaxial fracture of notched BMGs. Besides, after

shear bands evolve into cracks, their growth in the notched sample is also very

different from that in the un-notched sample. Figures 59(a) and (b) give the SEM

observations of the fatigue-crack propagation paths for the un-notched and notched

BMGs, respectively. In the un-notched sample, the fatigue crack-growth trajectory is

flat and without the formation of shear bands, implying brittle failure without any

obstacle, as expected in an amorphous structure. In contrast, the crack growth in the

notched sample displays a staircase-like path with shear bands formed at each step.

Clearly, in a highly zig-zag fatigue-crack propagation path, more energy will be

consumed, and the fast fracture of BMGs under cyclic loading will be inhibited. From

the microstructural viewpoint, when cracks develop from shear bands at the notch

roots and grow into the sample, a high concentration of free volumes is biased

towards the crack tip by cyclic loading, which promotes the initiation of shear bands

at the subcritical stress. The stress gradient is enlarged by the notches, cracks, and

shear bands and the propagation of shear bands or cracks is suppressed accordingly.

These mechanisms collectively regard the occurrence of the rapid fracture and extend

the fatigue life of notched BMGs. This trend is similar to the observed high fatigue

endurance limit in BMG composites, with second phases promoting the formation of

multiple shear bands and deflecting crack propagation [113, 259].

11. Failure Modeling and Lifetime Prediction

106
In this section, we first review continuum constitutive models that predict the

shear-band initiation and spatial arrangements, which may help explain the ductility

enhancement in a number of applications where geometric constraints are introduced.

Molecular dynamics and mesoscale STZ simulations will be discussed in Sections

11.2 and 11.3, respectively. The ductile to brittle transition in BMGs is discussed in

Section 11.4 from the viewpoints of the cavitation, structural heterogeneity, and

crack-tip process zone. Lastly, fatigue modeling on size effects is discussed in Section

11.5.

11.1. Continuum-Mechanics Models

Geometric constraints are a typical approach to block the shear-band

propagation, and, thus, increase the shear-band number, reduce the strain on

individual shear bands, and delay the failure of the sample. Examples can be found in

the thin-film MGs, BMG composites, and samples with variable height-to-diameter

ratios [258, 280]. For example, Li et al. [281] studied the effect of nanocrystalline Ni-

15%Fe (weight percent, wt.%) coatings on Zr-based BMGs, which results in the

increased shear-band density and enhanced ductility. In order to better understand

these geometric effects, we need to have a predictive theory that elucidates the shear-

band arrangements. The Rudnicki-Rice instability theory and the free-volume model,

which have been described in detail in Refs. [46, 97, 282-284], are powerful tools

along this line of study.

11.1.1. Rudnicki-Rice Instability Model

107
Shear bands and strain localization are the consequence of the constitutive

instability, i.e., a bifurcation phenomenon from a homogeneous deformation field to

an inhomogeneous one, when the constitutive equation has a loss of the elliptical

stability [282]. Since MGs demonstrate pressure sensitivity, the Rudnicki-Rice model

is adopted, giving rise to the shear-band angle,

  N min ,
0  tan 1 (11-1)
N max  

where  0 is the angle between the shear-band plane and the largest principal stress

axis in the plane of  I ,  III  , where  I ,  II , and  III are principal stresses,  I ,

1
 II , and  III are principal deviatoric stresses,   1        N 1   ,  is
3

Poisson’s ratio, Nmax   I  , N   II  , Nmin   III


  , and    mises 3 . For a

pressure-sensitive material, one can define the coefficient of internal friction,  , and

the dilatancy factor,  . When    , the deformation is associative. When both are

zero, the material obeys the Mises-plasticity model. The procedure of obtaining the

shear-band arrangements is specified as follows. The elastic simulation can be first

performed to obtain the stress fields, and then the above equation is used to find out

the potential shear-band directions in the entire sample. Streamline plots can be

constructed from the direction fields, which correspond to candidate shear bands.

One practical example is given in Figure 60(a) [257], where a three-

dimensional (3D) half-symmetric ABAQUS model under Rockwell indentation was

developed with a film/substrate thickness ratio of 1/20. In this model, both the MG

substrate and film are treated as a pure elastic body, which means that the

108
deformation behavior of the substrate and coating is dominated by their Young’s

moduli (E) and Poisson’s ratios (  ), with E = 88.6 GPa (substrate) and 122 (film)

GPa, and  = 0.3 (substrate) and 0.34 (film), respectively. The Rudnicki-Rice

instability theory is employed to predict the directions of shear bands for both the

monolithic [Figure 60(b)] [97] and coated BMGs [Figure 60(c)] [97], where µ + β = 0

indicates that the materials deformation is pressure insensitive and associative, with

the detailed explanation of µ and β in Ref. [284]. The prediction, from the instability

theory, gives typical radial shear-band patterns under the indentation for both the bare

substrate and film/substrate cases, which is consistent with many reported

experimental results [257, 284, 285]. In Figure 60(b), the blue solid curves indicate

the predicted shear-band directions, while the red dashed curves are along principal

shear-stress directions. Since there are no shear-band constraint conditions on the

surface of the monolithic BMG specimen, only major shear bands appear and

propagate in the MG substrate. However, in the coating/substrate material system,

more shear bands appear, as shown in Figure 60(c). It can be observed that many

short solid black and dashed green curves, which are the corresponding shear-band

directions and principal shear-stress directions, respectively, occur at the

coating/substrate interface, together with some larger shear bands (blue solid curves).

This phenomenon suggests that the local strain of the BMG produced in the

deformation process can be dispersed by more shear bands, which reduce the shear

strain in each shear band. Therefore, the plasticity of the BMG substrate is increased

by a surface coating.

109
Moreover, in the TFMG-substrate material systems, shear bands are

“reflected”, resulting in the occurrence of more short and minor shear bands [solid

curves, Figure 60(c)], when major shear bands propagate and arrive at the

film/substrate interface during deformation, as shown in Figure 60(c). It should be

noted that in our simulations, shear bands start from the substrate material, while

shear bands can be initiated at the interface in the real case. The term “reflection”

means that the shear-band directions change, since two families (before and after

reflection) of shear bands may be initiated simultaneously. This trend causes the

formation of multiple shear bands at the interface. Thus, each shear band will not

endure a large amount of shear strains. Hence, the enhanced ductility can be achieved

in the coated MG during experiments.

11.1.2. Free-Volume-Model Simulations

The initiation and propagation of shear bands can be modeled by the free-

volume model. In the classic Spaepen model, the stress-driven free-volume increase

reduces the viscosity, thus leading to the strain-softening behavior. The plastic-strain

rate is given by

 p   v*   G m    
 2 f exp    exp    sinh  , (11-2)
t  vf   k T   2 k T 
  B B

where f is the frequency of the atomic vibration,  is a geometric factor of order 1,

v * is the hard-sphere atomic volume, v f is the average free volume per atom, G m is

the activation energy,  is the atomic volume, kB is the Boltzmann constant, and T is

110
the absolute temperature. The free-volume evolution is also coupled with a

diffusional process, given by

v f   v*   G m  
 2 kBT      1 
 v* f exp    exp     cosh    1   , (11-3)
t  vf   k BT    2k BT   nD 
   v f Ceff  

*
where nD is the number of atomic jumps to annihilate a free volume of v , and Ceff

is an effective elastic constant. The above constitutive equations can be generalized to

multiaxial stress states using the effective Mises stress, and the corresponding finite-

element implementation can be used to predict the shear-band evolution [283].

As an example of this method, we investigate the effects of film/substrate

adhesion and film thickness on the mechanical properties of BMG specimens. A two-

dimensional (2D) ABAQUS model is constructed under indentation loading, with an

MG substrate and a Ti-based alloy coating. To explore the effects of coating thickness

and adhesion on the enhanced plasticity, films with varied configurations and

constitutive laws are employed. For all substrate materials, the constitutive

parameters are v f /  v*  0.05 , E / 2kBT  240 , ν = 0.33, nD = 3, α = 0.15, and

v* /  = 1. The normalized loading rate is shown below:

h G m
 exp( )  2.3 106 s 1 . (11-4)
Rf kBT

To investigate the effects of coating adhesion on the ductility and fatigue

enhancement of BMGs, different indentation contours were simulated in [97]. The

coating material used in the present work is Ti, which is treated as a purely-elastic

body, with a Young’s modulus of E = 122 GPa and a Poisson ratio of ν = 0.34.

Experimental studies [183] reveal that the film/substrate adhesion plays a

111
crucial role in enhancing the ductility of BMGs. Consistent with the experimental

observations, poor film/substrate adhesion tends to bring about less shear-band

reflection and branching, which is attributed to film delamination occurring easily in

the poor adhesion case. As a result, only relatively-limited enhanced plasticity is

obtained in the poorly-bonded coating material, thus leading to poor fatigue

improvement. For the perfect film-adhesion condition, the fatigue-crack initiation

stage is significantly elongated, thus prolonging the overall fatigue life of this coating

material. This trend is consistent with our fatigue experimental results that good

adhesion usually leads to the longer fatigue life of coated specimens [183].

Based on the experimental and simulation analysis above, a summary of the

fatigue-crack initiation and propagation mechanisms in the TFMG-substrate material

systems is illustrated in [97]. It can be observed that fatigue-crack initiation in the

TFMG-substrate system results from the dislocation pileup in crystalline substrates

under cyclic loading, with the factors affecting fatigue-crack-initiation: (i)

film/substrate interface adhesion; (ii) surface roughness; (iii) coating properties; and

(iv) coating thickness. After the fatigue crack initiates in the TFMG/substrate

systems, some minor shear bands will be formed in the TFMG. Thereafter, the

fatigue-crack-propagation process will not be affected significantly by the TFMG,

which will result in the rapid failure of the TFMG/substrate system. Beneficial-effect

factors on fatigue behavior of the TFMG/substrate systems are: (i) medium-strength

substrate materials; (ii) good TFMG/substrate adhesion; (iii) TFMGs with good

ductility; and (iv) proper film thickness.

11.1.3. Other Types of Continuum Models

112
The constitutive models, as long as incorporating a strain-softening

mechanism, can predict strain localization. Obviously, the ones with predictive

capabilities need to have a direct connection to mesoscale shear-banding behavior. A

vast number of experiments have shown the pressure sensitivity of the deformation

behavior of BMGs. Hence, the incorporation of the Mohr-Coulomb yield criterion

[286] or other types of methods, such as the ones used in the Rudnicki-Rice model,

can be adopted. The deformation behavior is found to be sensitive to the

environmental temperature and loading rate [287]. Thus, the thermally-activated

processes (e.g., the one used in the free-volume model) has to be calibrated with

respect to these two variables. The onset of shear-band initiation corresponds to a

large rate of the plastic work and, thus, heat generation. Hence, the thermal transport

analysis should be employed since the material also demonstrates thermal softening

[94, 288]. We also note that the glass structure on atomic- and meso-scales may not

be homogeneous. Thus, the role of structural heterogeneity is of critical importance,

which will be discussed in details shortly.

11.2. Fatigue-Damage Simulation in BMGs

There was an attempt to apply the free-volume based continuum constitutive

model to elucidate the damage mechanism in BMGs subjected to tension-

compression fatigue [289]. The simulations found that the fatigue life of the BMGs

was closely related to the free-volume accumulation and shear-band evolution. When

the applied cyclic strain was in the elastic range, no stress drops were observed on the

stress-loops curves, and the samples would never fail due to the invariable free

volume density [289]. In the plastic range, however, a stress drop was always

113
observed. The stress drop was believed to correspond to a main shear band running

through the sample, thereby causing fatigue failure [289]. Both the applied strain

amplitude and the cycling frequency could notably affect the cycle number at which a

stress drop occurred, i.e., fatigue life. Although the continuum-constitutive models

can predict the shear-band initiation and propagation, but they lack the capability of

predicting failures or damage accumulations. Atomistic models have been used in the

fatigue-damage study. An example is given here that investigates the effects of cyclic

loading on the deformation behavior of a 3-dimensional, four-component amorphous

Lennard–Jones solid. The potentials were chosen with a wide range of atomic sizes in

order to create a stable amorphous state. All atomic interactions are represented by

the standard Lennard–Jones potential (Uσ) given in the equation below [290]

  12   6
U  4 [( ) ( ) ], (11-5)
r r

where ε and r are the pair equilibrium well depth and separation, respectively. The

subscripts, α and , represent different atom types. The glass-transition temperature

for this system is ~ 0.63T0, where kBT0 = ε11. All of the simulations have been

conducted at 0.2T0, which is roughly the RT and much below the glass transition

temperature, Tg.

Shear loading can be simulated by imposing a specified load on the areas

within five atoms of the deformation boundary. The atomic structure during and after

deformation can be characterized by two different methods: 1) the first one is to

calculate the free volume of the structure, following the work of Sietsma and Thijsse

[291]; 2) The other way is to locate deformation regions and keep track of the local

excess strain of all atoms, following the work of Falk and Langer [51].
114
To study the effects of cyclic loading on the nonuniformity of the shear strain,

a structure that developed a shear band under monotonic loading was cycled under a

stress between zero and the onset of the localization level. Figure 61 shows the

comparison between cycling the system and just holding the same structure at that

load for the same period of time [144]. Note that the indicated atomic positions were

initially horizontal before loading. Therefore, it can be noticed that the atoms in the

cyclic-loading model have larger displacements than these in the just-holding model.

In this model, the total free volume and excess strain can be tracked during

cyclic loading, as shown in Figure 62 [144], respectively, in which the free volume

and excess strain also gradually increase as cyclic loading progresses (cycles one, two,

and five). Figure 62 reveals that the plastic-deformation events occur precisely in

those regions with higher levels of free volumes. It can be observed that the

deformation in the cycle one occurs in the region of the highest free volume in the

initial structure. As the shear transformations occur, more free volumes are created,

which in turn lead to more shear-transformation events to occur in this region upon

loading.

From the modeling efforts, the shear-localization process is not limited to

deform within a single applied load, which means that cyclic loading can promote the

formation of more shear bands. The free volume created by a shear transformation

can precipitate localization over several cycles, while the elastic unloading of other

atoms can create even more potential free volumes. However, it should be noted that

when the load is applied in the opposite direction, shear transformation will occur in

115
different locations and, therefore, interrupt the accumulation of free volumes in

particular locations.

In summary, the current MD method can be used to determine the evolutions

in free volumes and excess strains during monotonic or fatigue deformation of MGs,

which is simulated, using a four-component amorphous Lennard-Jones solid. In the

model, it demonstrates how the stress state affects the distribution of free volumes

and how regions with excess free volumes preferentially deform, which is helpful in

understanding the mechanism for the rapid initiation of fatigue damage and shear-

band formation in MGs under fatigue loading.

To understand the fatigue deformation of BMGs on the atomic scale, the

atomic bond rotation angle (ABRA) method can be employed to demonstrate where

the local shear strains appear under cyclic loading [292]. The ABRA method is

quantified by calculating the change of the angle, φij, between mutual orientations of

the nearest-neighbor bond vectors, which is given by

 ij ij'
ij  cos 1 ( ), (11-6)
|  ij ||  ij' |

where  ij and  ij' are the nearest-neighbor bond vectors before and after deformation.

Since all atomic bonds can rotate during shear deformation, the angle, φij, is closely

related to the local atomic shear strain.

The model used in Ref. [293] consisted of 47,424 atoms, in a mixture of 50%

(at.%) Zr and 50% (at.%) Cu in a cubic size of 9.5 nm  9.5 nm  9.5 nm, as shown

in Figure 63(a) [293]. After 10, 50, and 100 cycles of tensile loading at σmax = 2 GPa,

Figures 63(b)-(d) [293] show the excessive strain distribution of a slice parallel to the

116
xz plane of the Zr-Cu amorphous alloy, which is extracted from the 3D-simulation

results for better view. In these contours, the green-color sphere indicates the higher

strain levels, while the blue-color sphere indicates smaller strains. During cyclic

loading, two kinds of deformation mechanisms on the atomic scale work

simultaneously to dominate the deformation process of MGs: 1) One is the elastic

deformation with the reversible atomic rearrangement; and 2) The other is the plastic

deformation with the irreversible atomic rearrangement. From Figure 63, it can be

observed that most excessive strains would accumulate with increasing cycles. As

reported in Ref. [247], an individual STZ has been suggested to be ~ 15 Å in diameter,

corresponding to ~ 120 atoms for the CuTi BMG model. In the simulation results,

these irreversible-deformation atoms constructed the independent STZs distributed

randomly in the MGs at the early stages of the simulation, as shown in Figure 63(b),

in which the STZs with a higher strain level (green color) have a size of ~ 10 - 20 Å

[247]. Hence, it can be inferred that the irreversible events dominate the majority of

the deformation under cyclic loading, leading to the development of STZs. With

increasing loading cycles, the larger plastic-strain zone appeared from the connection

of individual STZs, as exhibited in Figure 63(c). However, the distribution of shear

strains does not seem to lie along a specific direction even after 100 loading cycles,

which is different from the case that the linear distribution of shear strains is localized

along a specific direction (~ 40 - 60º) with respect to the loading axis [294].

Moreover, the progress of excessive strains tends to connect and form a network

existing in the current 3D space model, which is presented as the green color in

Figures 63(c) and (d).

117
From Figures 63(b)-(d), it can be noticed that under cyclic loading, the growth

rate of shear deformation appears to be slower in the middle and final stages,

compared with that in the early stage. To better understand the growth behavior of

flow defects during cyclic loading, the statistics of accumulating deformations is

employed with the numbers of deformed atoms collected in Figure 64 (green color in

Figure 63) [293]. It reveals that the deformed atom number increases fast in the early

stage, but becomes gradually saturated in later stages. This tendency corresponds well

to the observation in the sliced plots in Figures 63(b)-(d), in which apparent

difference in STZ numbers are observed between Figures 63(b) and (c). However,

only minor difference can be noticed between Figures 63(c) and (d). The gradual

saturation tendency of the plastic-flow growth rate suggests a resistance against the

further development of STZs, after the accumulating deformation reaches a specific

level under fatigue loading.

11.3. Mesoscale Model: STZ Dynamics Simulations for Cyclic Indentation with

Finite-Element Modeling (FEM)

It should also be noted that the atomistic processes are better modeled by the

STZ model than the free-volume model. In the STZ model, the deformation process

is proposed as a collective activity of a few atoms that rearrange to achieve a

characteristic shear strain under an applied shear stress, as shown in Figure 1(b) [50,

295]. The STZ can be viewed as a stress-biased, thermally-activated event, which

permits the STZ-activation laws to be written in terms of state variables, including

stress, temperature, and local structural parameters (free volume), et al. [50]. By

appropriately modeling the dynamics of these events, the local information of

118
individual atomic motions can be revealed, as well as capture the fundamental

physics of deformation [51]. In the STZ model, an amorphous material can be treated

as an elastic continuum, consisting of an ensemble of potential STZs defined on a

mesh. In practice, a continuum mesh is substituted for collections of atoms, as shown

schematically in Figure 1(c). The shearing of an STZ can be treated as an Eshelby

inclusion problem [296], in which there is an activation energy barrier for shearing of

an STZ [50]. The activation-energy barrier is determined by Argon [50] in the

equation below

7  5 2(1   ) 2 ˆ
F  [    ] (T ) 02  0 , (11-7)
30(1   ) 9(1   ) 2 0  (T )

where the first term represents the strain energy of an STZ sheared by the

characteristic shear strain, γ0. The second term is the strain energy for a temporary

dilatation to allow the atoms to rearrange into the sheared position, in which β

represents the ratio of the STZ dilatation during the transformation to the shear strain,

γ0. The third term represents the energy required to freely shear an STZ, with  equal

to the peak interatomic shear resistance between atoms. Beyond that, ν is Poisson’s

ratio, μ(T) is the temperature-dependent shear modulus, and Ω0 is the STZ volume.

The predictive capability of the STZ model requires the knowledge of structural

heterogeneity.

Beyond simulating the fatigue behavior of metallic glasses, as shown in the

above sections, the STZ dynamics simulation method can also be employed to

calculate the case of cyclic indentation loading [295], in which the STZ can be treated

as the basic unit process for shape change using a coarse-grained method. The

metallic glass is simulated by partitioning the volume into an ensemble of potential

119
STZs, which are mapped onto the mesh of FEM. The nanoindentation simulation is

carried out on a two-dimensional (2D) simulation cell comprised of plane-strain

elements, with the indenter modeled as a rigid surface, which is loaded along the y-

direction. More information on the simulation details can be referred in Ref. [297].

Note that the present STZ model does not specifically incorporate the structural state

variable of MGs, such as free volumes, which makes it very difficult to explore the

intrinsic structural hardening or softening. However, the current model can be used to

validate the concept of microplasticity upon cyclic loading.

Before studying the response of cyclic nanoindentation, monotonic-

nanoindentation tests are firstly investigated. As a baseline for the subsequent

analysis, an ideal elastic indentation is conducted, with the simulated load–

displacement (P-h) curve (solid black line) shown in Figure 65 [297]. When STZs are

allowed to activate, the P-h result is exhibited in Figure 65 (the blue curve). The

simulation results can be observed that the P-h response of the model with the STZ

activation initially follows the elastic curve exactly, but, begins to evidently deviate

from the elastic curve as the plastic flow starts at a depth of ~ 2 nm.

Corresponding to the positions, ‘A’, ‘B’ and ‘C’, on the simulated P–h curve,

corresponding snapshots during the indentation are provided in the bottom of Figure

65, also marked as A, B, and C, respectively [297]. In these snapshots, the red solid

line denotes the outer envelope of the material, in which the local von Mises stress

exceeds the nominal yield stress of the metallic glass. However, the high von Mises

stress is not necessarily related to the activation of STZs, since there is a significant

region at loads where some part of the material reaches the yield stress without the

120
STZ activity. In this region, the material is still under elastic deformation, leading to a

symmetric yield envelope, as exhibited in the panel ‘A’ of Figure 65 [297]. With

increasing load (point ‘B’), the stress and strain field is perturbed and redistributed

due to the activation of STZs, resulting in the asymmetric and irregular envelope.

Note that at point ‘B’, the global P-h curve is still in excellent agreement with the

elastic curve, despite of the activity of numerous STZs and the considerable portion

of materials larger than the nominal yield stress. Far beyond point ‘B’ (such as, point

‘C’), the departure of P-h curve from the ideal elastic curve is evident, corresponding

to significant plastic deformation, is large, as illustrated in panel ‘C’ of Figure 65

[297].

Regarding the cyclic-indentation simulation, four displacement amplitudes are

conducted at 1.2, 1.6, 2.0, 2.4, and 2.8 nm, respectively, with the simulated P-h

curves illustrated in Figure 66 [297]. For the low amplitude case (1.2 and 1.6 nm), the

STZ is not activated, leading to a perfectly-elastic response. On the other side, for the

largest displacement amplitude of 2.8 nm, the abundant STZ activity appears beneath

the indenter, and the measurable dissipation occurs after the first cycle. At

intermediate displacement amplitudes of 2.0 and 2.4 nm, the significant STZ activity

beneath the indenter can be observed, but the relatively-little permanent deflection

appears in the P–h curves. Despite the appearance of elastic conditions, the significant

STZ activity beneath the indenter can be observed in [297] for these two intermediate

displacement cases.

Qualitatively, the current simulations match well with those from the

experimental work, and validate the hypothesis of microplastic structural

121
rearrangements. That is the cyclic loading can lead to the undetected microplasticity,

and the structural change via the STZ activation can occur progressively during the

process of several load cycles. However, it should be noted that the present model

does not consider the hardening mechanism in the microplasticity-affected regions.

11.4. Ductile to Brittle Transition in BMGs

One way to characterize the ductile versus brittle behavior in a given material

is to investigate its crack-tip response. From the top-down point of view, a crack tip

will be surrounded by a set of characteristic fields that depend on the fracture modes

and the stress-intensity factors. Approaching the exact crack tip, one will see the

continuum plastic process-zone and, then, “messy” process zone that depends on

damage processes. From the macroscopic point of view, the ductile versus brittle

behavior results from the competition between the crack-tip blunting due to plasticity

and the cleavage that results in brittle fracture. If the cleavage stress is low, a

considerable degree of plastic deformation will take place, which leads to a noticeable

resistance curve. The deformation behavior is, thus, ductile, and the fracture is by the

cavitation initiation, growth, and coalescence. From the atomistic point of view, the

ductile versus brittle behavior is governed by the competition between dislocation

activities and cleavage. For example, if a dislocation can be easily emitted from the

crack tip than cleavage process, the crack tip will be blunted. For another example, if

the dislocation motion is sluggish, a large stress is, thus, required to drive the

nucleated dislocations away from the crack tip. Thus, the material may fail via

cleavage fracture.

122
When using the above views to study the ductile versus brittle behavior in

MGs, we obviously have the following questions: what are the plastic-zone

characteristics when deformation is realized by individual shear bands, and what is

the cleavage-fracture process, when there are no crystalline weak planes?

11.4.1. Cavitation Process as a Precursor of Failure

As opposed to the commonly-seen cavitation-induced ductile failure in

crystalline alloys, it is found that atomic scale fluctuations in metallic glasses will

reduce the hydrostatic stress needed for cavity nucleation [236, 237, 298, 299].

Murali et al. [236] studied two types of metallic glasses, Fe80P20 and Cu50Zr50, which

are brittle and ductile, respectively. Their molecular simulations in Figure 67

demonstrate the brittle fracture by nanoscale voids ahead of the crack tip, as opposed

to the ductile fracture where shear localization blunts the crack tip. The fundamental

difference of these two types of MGs is that the density fluctuations in the FeP MG

are higher than those in the CuZr MGs. In other words, there are considerable

atomistic heterogeneities in the FeP MG that leads to a wide range of fluctuations in

the cavitation stress. Thus, the material is prone to the cavitation process. Additional

continuum mechanics simulations also suggest the elevation of the hydrostatic stress

on the emanated shear band from the crack tip [298, 299]. Consequently, it is

anticipated that for brittle metallic glasses, cavities may directly nucleate from the

crack tip due to the low cavitation stress, or follow the initiation of the shear band that

is accompanied with an increased hydrostatic stress and then take place on the shear-

band plane.

123
In the free-volume model, Wright et al. [150] suggested that the excess free

volume generated in shear bands upon deformation have increased free energy.

Consequently, voids may potentially nucleate from the coalescence of the excess free

volume, and the subsequent growth and linkage may depend on the stress state and

strain history. This feature provides an alternative view of cavity or void nucleation,

in contrast to the model in Figure 68.

11.4.2. Continuum Plastic Process Zone versus Cleavage Fracture

In analogy to the crack-tip process-zone studies in Tvergaard and Hutchinson

[300], here we present an investigation of the competition between the cleavage

fracture and crack-tip blunting in the metallic glass by numerically determining the

response of a crack subjected to a history of the applied stress-intensity factor. The

brittle fracture is modeled by a traction-separation law in the crack plane [301], and

the constitutive law of the metallic glass employs the free-volume model [283]. The

line of thought is similar to that in Steif [47], i.e., the delayed response of the free-

volume evolution changes the rate of the stress redistribution ahead of a potential

flow. In Steif [47], this flaw is a hole, while here in Figure 68, it is a crack tip.

Because of the strain softening and strain rate hardening behavior, it is found that at

high loading rates, the fracture behavior is of brittle nature, while multiple shear

bands nucleate and reduce the stress concentration at low loading rates, as shown in

Figure 68.

The crack extension, a , can be represented by a dimensional analysis as

a  K K  E 
 a  , t0 , max , , , v fi ,  ,  , nD  , (11-8)
R0  K0 K0 0 0 

124
where K 0 is the toughness of the cohesive interface,  max is the interface strength,

2
1 K 
R0   0  is the characteristic length, and other parameters are from the free-
2   max 

volume model. The ductile-versus-brittle response map in Figure 68 suggests that

when the loading rate is large, there is no sufficient time for strain localization to take

place, thus leading to brittle fracture.

11.4.3. Roles of Structural Heterogeneities

A metallic-glass sample cannot have the perfect glassy packing of the

constituent atoms everywhere, and multiscale structural heterogeneities may exist.

The roles of atomic heterogeneities [236, 302] may lead to fluctuations of the atomic-

scale cavitation stress or yield strength. The comparison between two types of

metallic glasses in Figure 68 suggests one way of tuning the ductile versus brittle

behavior by the atomic heterogeneity. Even for the same metallic glass but with

various thermal treatments, the mechanical responses can vary significantly. It has

been shown in Li et al. [39, 86] that a longer annealing time and higher annealing

temperature will make the same metallic glass brittle. Although studies, such as

synchrotron x-ray and differential scanning calorimetry, do not show noticeable

difference among these annealed samples, other microstructural analyses indicate the

importance of structural heterogeneities. Nanoindentation pop-in studies with

properly corrected machine stiffness effect by Li et al. [39, 86, 248] can be used to

probe the distribution of weak zones (or structural heterogeneities) in these samples,

and it is found that annealing leads to increased spacing of these weak zones. Digital

125
image correlation studies [303] show the evolution of homogeneous into

inhomogeneous deformation fields by the correspondence between tiny stress drops

on the stress-time curve and many small shear-bands as determined by the DIC

technique in Figure 69. All these studies reveal the importance of various length

scales of structural heterogeneities in understanding the failure modes of metallic

glasses.

11.5. Modeling Size Effects on Fatigue Life of BMGs

In the reported experimental works, large samples tend to have longer fatigue

lives than small samples under the same bending-stress condition, as shown in Figure

19(b) and Figure 70 [178]. Figure 70 shows the S-N curves based on the fatigue-life

data of Zr50Cu30Al10Ni10 (at.%) BMG samples with two different thicknesses, 2 and 3

mm. The significantly-degraded fatigue lifetime and endurance limit of the 2 mm

BMG sample were attributed to the extensive interaction of abundant primary and

secondary shear bands [178]. On one hand, the high density of shear bands facilitated

the formation of fatigue cracks under cyclic loading as they served as the embryos of

cracks. On the other hand, the high density of shear bands and later formed micro-

cracks could substantially reduce the overall strength of the sample. The combined

effect of these two factors leaded to an early failure of the 2 mm BMG sample. By

contrast, the larger 3 mm sample formed very few shear bands during the fatigue test.

It was, thus, unable to form as many fatigue cracks as the 2 mm sample. Also, its

strength was maintained at a much higher level in comparison to the 2 mm sample.

Consequently, the 3 mm sample exhibited a much higher fatigue resistance than the 2

mm one [178]. However, compared with the experimental research, the modeling

126
work of the size effects on the fatigue life of BMGs under bending has not been

extensively investigated, which is usually studied based on the Weibull statistics

below [304, 305]. Since the failure of BMGs under bending is largely controlled by

the shear-band processes [178], the Weibull statistics cannot adequately describe the

dependence of the fatigue-life of BMGs on the sample size. Therefore, new statistical

models for the fatigue-life prediction are proposed after considering the size

dependence of the bending-fatigue life.

11.5.1. Theoretical Model on Fatigue Life of BMGs with Different Sizes

A commonly-used analytical representation of the S–N curves is given by

 c ( S  S0 )  d , S  S0
N  , (11-9)
  , S  S 0

where c and d are positive material parameters, N, S, and S0 are the cycles to failure,

applied stress, and fatigue-endurance limit, respectively. Taking the natural logarithm

on the above equation, resulting in

   1 log( S  S0 ), S  S0
log( N )   0 , (11-10)
 , S  S0

where γ0 = log(c) and γ1 = −d.

Then the fatigue life at a given stress level, S, follows the Weibull distribution

with the cumulative distribution function (CDF) given by

N 
F (N |  , ( S ))  1  exp( ( ) ), (11-11)
 (S )

and the probability density function (PDF) given by

127
dF ( N |  , ( S ))  N  1 N 
f ( N |  , ( S ))   ( ) exp( ( ) ), (11-12)
dN  (S )  (S )  (S)

where  is the Weibull shape parameter, and θ(S) is the Weibull-scale parameter

depending on the stress, S, according to

   1 log( S  S0 ), S  S0
log( ( S ))   0 , (11-13)
 , S  S0

The fatigue-life model given by the above equation is called Model I, which is

actually an extension of the widely-applied Weibull accelerated life-test model in

reliability engineering with the addition of the fatigue-endurance limit parameter and

can be used to analyze and predict the fatigue life of metallic glasses with identical

size. In order to incorporate the size effects in the fatigue-life model, two size-effect

fatigue-life models are proposed.

The first size-effect fatigue-life model (Model II) is an empirical extension of

Model I by including the thickness dependence to the Weibull shape and scale

parameters. The CDF of Model II is given by

N  (h)
F ( N |  (h), ( S , h))  1  exp( ( ) ), (11-14)
 ( S , h)

where the scale and shape parameters are assumed to be

log( ( S , h))  0  1 log( S  S0 (h))  2h, S  S0 (h) , (11-15)

and

log(  (h))  0  1h , (11-16)

, respectively, where h represents the thickness of the material under bending, and

S0(h) is the fatigue-endurance limit of the material with the thickness, h.

128
The second size-effect fatigue-life model (Model III) incorporates mechanistic

understanding of the fatigue failure of BMGs. The CDF of Model III is assumed to be

N 
F ( N |  , 0 ( S ))  1  exp( h 1 ( ) ), (11-17)
 (S )

where h0(S) is given by

log( ( S ))  0  1 log( S  S0 (h)), S  S0 (h) , (11-18)

Model III is proposed, based on the following two assumptions. First, Model

III assumes that the failure of BMGs under bending is a weakest-link process. At RT,

the catastrophic failure in BMGs is mainly caused by the inhomogeneous deformation,

resulting from the localized shear band [163, 178]. The fatigue life of BMGs under

bending may be mainly controlled by the number of shear bands formed during

bending. Second, the number of shear bands in a sample depends on the shear-band

spacing. According to the reported data [168, 173, 178], the shear-band spacing scales

linearly with the sample thickness for a wide range of thickness values. The results

exhibit consistent trends regardless of BMGs and test loading.

11.5.2. Prediction on Fatigue Life of BMGs with Different Sizes

Figure 71(a)-(c) plot the median fatigue lives and the 95%-predictive intervals

predicted by Model I, II, and III, respectively [306]. The median fatigue life can be

used to describe the relationship between the applied stress and the average fatigue-

life response. The 2.5 and 97.5 quantiles can be used to construct a 95%-predictive

interval for the fatigue life, and to quantify the scatter in the fatigue-life cycles. Model

I is capable of describing the S–N behavior, capturing the variability in the observed

129
fatigue-life data, and providing reasonable estimation of the endurance limit. It,

therefore, provides a general approach to analyzing the S–N data of BMGs.

Figure 71(d) compares the median lives predicted by the three models [306].

Overall, the three models produce consistent predictions. The discrepancy among the

results may be caused by (1) different assumptions used by the models, (2) possible

higher-order effects of the thickness on the logarithms of the Weibull parameters in

Model II, and (3) significant effects from factors other than the shear-band density in

Model III.

The shear-band-based fatigue-life model (Model III) can reasonably capture

the size effects on the fatigue life of the BMG samples under bending. The modeling

results, therefore, provide the evidence that the number of shear bands formed during

the bending experiments may have a significant influence on the fatigue life. During

bending-fatigue experiments, small-sized BMG samples are easier to form shear

bands and, consequently, to initiate fatigue cracks from shear bands than large-sized

BMG samples, as illustrated in Figure 70 [178]. Therefore, a small-sized BMG

sample tends to exhibit a shorter fatigue life than a large-sized BMG sample.

Overall, the current study explored possible modeling approaches to including

the size effects in the statistical fatigue-life models under bending fatigue. Two

models (Models II and III) were developed, based on Model I. Model II is an

empirical model, and Model III incorporates the mechanistic knowledge of the

fracture and failure mechanisms, which provide promising results on the fatigue-life

prediction.

12. Future Directions

130
According to the numerous studies reviewed in this article, the fatigue and

fracture behavior of BMGs and MGMCs has been investigated in the last two

decades. However, there are still many questions needed to be further understood, as

described below.

1) Although the true nature of the deformation and fracture mechanisms in

BMGs could be attributed to shear bands, the characteristic for the formation of shear

bands during the cyclic deformation of BMGs is still unclear. Thus, how a fatigue

crack initiates in BMGs is the primary question that needs to be addressed.

2) Compared with the reported fatigue results on the Zr-based BMGs, the

fatigue behavior of other-based BMGs are still far from well investigated, as shown in

Figure 6. It can be noticed that Co-based and Fe-based BMGs even have a much

better fatigue-endurance limit than the Zr-based metallic glasses because of high

strength in Co-based and Fe – based BMGs than Zr-based BMGs.

3) Schuh et al. [4] suggested that the difference in the fatigue results of BMGs

from different groups could result from the effect of residual stresses and structural

relaxation on fatigue-crack propagation. There are generally residual stresses in BMG

samples since the cooling rates between the interior and exterior of BMG samples are

different during the casting process. Then, how residual stresses affecting the fatigue

life of BMGs can be an interesting topic.

4) According to Yokoyama et al.’s and Launey et al.’s works [80, 99], there

seems a conflicting result concerning the effect of free volumes on fatigue behavior

of BMGs. Yokoyama et al. thought that the free volume could improve the fatigue

properties of BMGs, while Launey suggested that the free volume could decrease the

131
fatigue-endurance limits of BMGs. Therefore, free-volume effects on the fatigue

behavior of BMGs must be further investigated.

5) Insufficient studies have been reported on the mechanical behavior of BMG

substrates coated with TFMGs, especially fatigue properties, which can be a potential

research direction to increase the fatigue behavior of BMGs, thus advancing the

application of BMGs, considering that TFMG-coated BMGs have shown a greatly-

improved bending ductility to 13.7% from almost zero ductility [183].

6) The fracture toughness of BMGs can vary from ~ 1 MPa m1/2 (for brittle

rare-earth and ferrous BMGs) [217, 218] to ~ 100 MPa m1/2 (for tougher noble BMGs)

[219, 220]. Many researchers have attempted to unveil the root cause of the

substantially scattered fracture toughness values. In spite of some degrees of success

along this line of work, e.g., a positive correlation between the Poisson’s ratio and the

fracture toughness of many BMGs were suggested [44, 45], the structural origin of

the scattered fracture toughness values remain unsettled, and requires more dedicated

endeavors.

7) The thermal history could alter the fracture behavior of some BMGs

dramatically, for instance, BAM11 could switch from somewhat ductile fracture to

perfectly brittle fracture when subjected to long-time annealing [40]. Currently, it is

widely believed that the ductile-to-brittle transition in BMGs is caused by the drastic

change in the density of structural heterogeneities they contain. Nevertheless, due to

the amorphous structure, it is challenging to characterize these heterogeneities in a

definite and reliable manner even though various experimental and modeling

techniques have been applied for this purpose. Determining an explicit relationship

132
between the ductile-to-brittle transition and the structural heterogeneities is, therefore,

still one the primary future directions of efforts.

8) Traditional brittle materials, e.g., glasses and ceramics, generally exhibit

notch weakening effect, i.e., the strength of a material decreases with the increasing

notch depth. However, the notch effect in brittle BMGs is rather complicated.

Depending on the compositions and processing histories, BMGs could exhibit notch

weakening, notch strengthening, or notch insensitivity [265]. Although some

researches attempted to explain the abnormal notch effect observed in BMGs from

the standpoints of shear-band operations [85] and failure-mode change [265], a

rigorous interpretation is still lacking.

9) The simulation work on the fatigue and fracture behavior of BMGs is still

quite limited, compared with the reported simulation results on the other mechanical

behavior of BMGs, due to the difficulty of modeling on the fatigue and fracture

behavior of materials [163, 167].

10) The deformation mechanisms of TFMG materials can be investigated with

the methods of FEM and MD simulations. But the actual failure mechanisms might

not be faithfully captured in the FEM model because of the limitation in the

constitutive model and in MD simulations because of limited spatio-temporal scales.

13. Concluding Remarks

Based on the current review of the fatigue and fracture behavior of BMGs and

MGMCs, the conclusions can be summarized as below:

133
1) The fatigue-loading mode, sample geometry, BMG composition, material

quality, fabrication procedure, and microstructure could influence the fatigue

behavior of Zr-based BMGs. BMGs under compression-compression loading

demonstrate excellent fatigue resistance. BMGs exhibit good tension-tension-fatigue

behavior. The tensile stress could be the driving force in affecting the crack-growth

behaviors in BMGs. Compared to fully-amorphous BMGs, BMG composites

generally exhibit poor fatigue resistance. The fatigue behavior of Zr-based BMGs has

been shown to be dependent upon the free volume in BMGs and the test environment.

2) The fatigue-endurance limits of Zr-based BMGs vary from 150 to 1,050

MPa. Although the fatigue behavior spans a wide range, these fatigue limits of BMGs

are comparable with the conventional, crystalline materials, such as ultrahigh-strength

steel (300-M), IN 718 superalloy, Ti-6Al-4V, and Zirconium alloys, depending upon

the exact BMG alloy tested and the experimental methods. The Zr-based BMGs with

the excellent fatigue resistance could be developed through changing the alloy

composition or the fabrication procedure.

3) The Zr-based BMGs and composites exhibit similar fatigue-crack-growth

behavior in air. Fatigue–crack-growth rates were found to be faster in aqueous

solutions than in air. The environmentally-assisted fatigue-crack-propagation

behavior in an aqueous sodium-chloride solution results from a stress-corrosion

crack-growth mechanism, which involves the stress-assisted, anodic dissolution at the

crack tip.

4) The fatigue crack initiates from the casting defects or shear bands.

Nanometer-scale voids could form due to the free-volume coalescence in a shear

134
band. In a tensile-stress state, the void growth and coalescence are promoted by the

tensile stress and can initiate fatigue cracks. The fatigue striations in the crack-

propagation region are formed due to the process of blunting and re-sharpening of the

crack tip. In a compressive-stress state, the formation of voids will decrease the

density of the material and its resistance to deformation in the shear band. The BMG

fails along one primary shear band, which forms a shear fracture angle with respect to

the stress axis under compression-compression fatigue loading, while the fracture

surface of the tension-tension is generally perpendicular to the stress axis. Moreover,

there are no striations on the compression-shear-fracture plane, while distinct

striations are found on the tension-tension fatigue-fracture surface.

5) TFMGs, as well as composited TFMGs, provide a possible solution to

utilize their unique properties to increase the plasticity and fatigue resistance of

substrates, without weakening the substrate strength. The fatigue behavior and

underlying micro-mechanisms of TFMGs were studied with both experimental and

FEM-simulation methods.

6) Zr-based and Cu-based TFMGs, as well as annealed Zr-based TFMGs, are


employed to improve the four-point-bending fatigue resistance of the 316L stainless-

steel, Ni-, Zr-, Al-, and Ti-based alloy substrates, while TiN, pure Cu, and Fe-based

TFMGs cannot extend the four-point-bending fatigue life of steel substrates. The

enhancement of fatigue life is mainly attributed to the good ductility and strength of

the films, leading to the multiplication of shear bands in TFMGs, and, thus,

prolonging the fatigue-crack-initiation state of film/substrate materials.

135
7) The fracture toughness of BMGs fluctuates in a wide range of ~ 1 – 100
MPa m1/2, depending on the compositions and processing histories. Rare-earth, Mg-

based, and ferrous BMGs normally manifest very low fracture toughness values,

fracturing in a brittle manner. On the other hand, some Ti- and Pt-based BMGs can

exhibit rather large fracture toughness, and their fracture is also fairly ductile.

Annealing an individual BMG below for a long time may drastically embrittle it.

The fracture toughness of metallic glasses matrix composites (MGMCs) can be

significantly improved by introducing second phases to retard the propagation of

shear bands.

8) For many BMG systems, a critical Poisson’s ratio, , or a

critical ratio of shear modulus to bulk modulus, , are found to

be effective in deciding the occurrence of the ductile-to-brittle transition. When >

or < , the BMGs are tough. In contrary, BMGs are brittle when <

or > . However, this criterion is not universally applicable and some

exceptions have been found in some Zr-, Pt-, and Pd-based BMGs. More rigorous

criteria for the ductile-to-brittle transition in BMGs are being actively searched.

9) Unlike traditional brittle materials whose strengths tend to be brought


down by introducing notches, notch strengthening, notch weakening and notch

insensitivity could all be present in BMGs depending on the notch configuration in

test samples. The strengths of some BMGs can be enhanced by the introduction of

notches, because the introduced notches change the stress state in the samples,

136
therefore, lead to a switch of the failure mode, i.e., from shear failure to cavitation

failure.

Acknowledgements

We would like to acknowledge the financial support of the National Science

Foundation: (1) the Division of the Design, Manufacture, and Industrial Innovation

Program, under Grant No. DMI-9724476, (2) the Combined Research-Curriculum

Development (CRCD) Programs, under EEC-9527527 and EEC-0203415, (3) the

Integrative Graduate Education and Research Training (IGERT) Program, under

DGE-9987548, (4) the International Materials Institutes (IMI) Program, under DMR-

0231320, (5) the Major Research Instrumentation (MRI) Program, under DMR-

0421219, (6) the Division of Materials Research, under DMR-0909037, and DMR-

1611180, and (7) The Division of Civil, Mechanical, and Manufacturing Innovations,

CMMI 1300223 to the University of Tennessee, Knoxville, with Dr. D. Durham, Ms.

M. Poats, Dr. C. J. Van Hartesveldt, Dr. J. Giordan, Dr. D. Dutta, Dr. W. Jennings,

Dr. L. Goldberg, Dr. C. Huber, Dr. C. R. Bouldin, Dr. A. Ardell, Dr. Thomas

Siegmund, Dr. G. Shiflet, and Dr. D Faikas as Program Directors, respectively. PKL

would like to acknowledge the Department of Energy (DOE), Office of Fossil

Energy, National Energy Technology Laboratory (DE-FE-0011194) with the program

manager, Dr. J. Mullen. PKL very much apprecipates the support of the U.S. Army

Research Office project (W911NF-13-1-0438) with the program manager, Dr. M. P.

Bakas and Dr. D. M. Stepp. YFG also acknowledges the support from Materials and

137
Engineering Division, Basic Energy Sciences (BES), the U.S. Department of Energy

(DOE).

138
References

[1] Chen M. Mechanical behavior of metallic glasses: microscopic understanding of

strength and ductility. Annu Rev Mater Res. 2008;38:445-69.

[2] Inoue A, Takeuchi A. Recent development and application products of bulk glassy

alloys. Acta Mater. 2011;59:2243-67.

[3] Johnson WL. Bulk glass-forming metallic alloys: science and technology. MRS

Bull. 1999;24:42-56.

[4] Schuh CA, Hufnagel TC, Ramamurty U. Mechanical behavior of amorphous

alloys. Acta Mater. 2007;55:4067-109.

[5] Trexler MM, Thadhani NN. Mechanical properties of bulk metallic glasses. Prog

Mater Sci. 2010;55:759-839.

[6] Klement W, Willens RH, Duwez P. Non-crystalline structure in solidified gold–

silicon alloys. Nature. 1960;187:869-70.

[7] Inoue A, Zhang T, Masumoto T. Zr-Al-Ni amorphous alloys with high glass

transition temperature and significant supercooled liquid region. Mater Trans, JIM.

1990;31:177-83.

[8] Inoue A, Zhang T, Nishiyama N, Ohba K, Masumoto T. Preparation of 16 mm

diameter rod of amorphous Zr65Al7.5Ni10Cu17.5 alloy. Mater Trans, JIM.

1993;34:1234-7.

[9] Inoue A. High strength bulk amorphous alloys with low critical cooling rates

(overview). Mater Trans, JIM. 1995;36:866-75.

[10] Peker A, Johnson WL. A highly processable metallic glass:

Zr41.2Ti13.8Cu12.5Ni10.0Be22.5. Appl Phys Lett. 1993;63:2342-4.

139
[11] Zhang QS, Deng YF, Zhang HF, Ding BZ, Hu ZQ. Cyclic softening of

Zr55Al10Ni5Cu30 bulk amorphous alloy. J Mater Sci Lett. 2003;22:1731-4.

[12] Zhang W, Inoue A. Formation and mechanical strength of new Cu-based bulk

glassy alloys with large supercooled liquid region. Mater Trans. 2004;45:1210-3.

[13] Ponnambalam V, Poon SJ, Shiflet GJ. Fe–Mn–Cr–Mo–(Y,Ln)–C–B (Ln =

Lanthanides) bulk metallic glasses as formable amorphous steel alloys. J Mater Res.

2004;19:3046-52.

[14] Lu ZP, Liu CT, Thompson JR, Porter WD. Structural amorphous steels. Phys

Rev Lett. 2004;92:245503.

[15] Inoue A, Nishiyama N, Matsuda T. Preparation of bulk glassy

Pd40Ni10Cu30P20 alloy of 40 mm in diameter by water quenching. Mater Trans,

JIM. 1996;37:181-4.

[16] He Y, Schwarz RB, Archuleta JI. Bulk glass formation in the Pd–Ni–P system.

Appl Phys Lett. 1996;69:1861-3.

[17] He Y, Dougherty GM, Shiflet GJ, Poon SJ. Unique metallic glass formability

and ultra-high tensile strength in Al-Ni-Fe-Gd alloys. Acta Metall Mater.

1993;41:337-43.

[18] Inoue A, Sobu S, Louzguine DV, Kimura H, Sasamori K. Ultrahigh strength Al-

based amorphous alloys containing Sc. J Mater Res. 2004;19:1539-43.

[19] Yi S, Park TG, Kim DH. Ni-based bulk amorphous alloys in the Ni–Ti–Zr–(Si,

Sn) system. J Mater Res. 2000;15:2425-30.

140
[20] Choi-Yim H, Xu D, Johnson WL. Ni-based bulk metallic glass formation in the

Ni–Nb–Sn and Ni–Nb–Sn–X (X=B,Fe,Cu) alloy systems. Appl Phys Lett.

2003;82:1030-2.

[21] Zhang T, Inoue A. New bulk glassy Ni-based alloys with high strength of 3000

MPa. Mater Trans. 2002;43:708-11.

[22] Louzguine DV, Inoue A. Nanocrystallization of Cu–(Zr or Hf)–Ti metallic

glasses. J Mater Res. 2002;17:2112-20.

[23] Inoue A, Nakamura T, Nishiyama N, Masumoto T. Mg-Cu-Y bulk amorphous

alloys with high tensile strength produced by a high-pressure die casting method.

Mater Trans, JIM. 1992;33:937-45.

[24] Inoue A, Masumoto T. Mg-based amorphous alloys. Mater Sci Eng A.

1993;173:1-8.

[25] Lin XH, Johnson WL. Formation of Ti–Zr–Cu–Ni bulk metallic glasses. J Appl

Phys. 1995;78:6514-9.

[26] He G, Eckert J, Hagiwara M. Glass-forming ability and crystallization behavior

of Ti–Cu–Ni–Sn–M (M=Zr, Mo, and Ta) metallic glasses. J Appl Phys.

2004;95:1816-21.

[27] Li R, Pang S, Ma C, Zhang T. Influence of similar atom substitution on glass

formation in (La–Ce)–Al–Co bulk metallic glasses. Acta Mater. 2007;55:3719-26.

[28] Jiang QK, Zhang GQ, Chen LY, Wu JZ, Zhang HG, Jiang JZ. Glass formability,

thermal stability and mechanical properties of La-based bulk metallic glasses. J

Alloys Compd. 2006;424:183-6.

141
[29] Inoue A. Stabilization of metallic supercooled liquid and bulk amorphous alloys.

Acta Mater. 2000;48:279-306.

[30] Löffler JF. Bulk metallic glasses. Intermetallics. 2003;11:529-40.

[31] Inoue A, Shen B, Takeuchi A. Developments and applications of bulk glassy

alloys in late transition metal base system. Mater Trans, JIM. 2006;47:1275-85.

[32] Peter WH, Buchanan RA, Liu CT, Liaw PK, Morrison ML, Horton JA, et al.

Localized corrosion behavior of a zirconium-based bulk metallic glass relative to its

crystalline state. Intermetallics. 2002;10:1157-62.

[33] Morrison ML, Buchanan RA, Leon RV, Liu CT, Green BA, Liaw PK, et al. The

electrochemical evaluation of a Zr-based bulk metallic glass in a phosphate-buffered

saline electrolyte. J Biomed Mater Res A. 2005;74A:430-8.

[34] Miller M, Liaw PK. Bulk metallic glasses: an overview. New York, NY:

Springer; 2010.

[35] Inoue A, Shen B, Koshiba H, Kato H, Yavari AR. Cobalt-based bulk glassy alloy

with ultrahigh strength and soft magnetic properties. Nat Mater. 2003;2:661-3.

[36] Liu CT, Heatherly L, Horton JA, Easton DS, Carmichael CA, Wright JL, et al.

Test environments and mechanical properties of Zr-base bulk amorphous alloys.

Metall Mater Trans A. 1998;29:1811-20.

[37] Donovan PE, Stobbs WM. The structure of shear bands in metallic glasses. Acta

Metall. 1981;29:1419-36.

[38] Pekarskaya E, Kim CP, Johnson WL. In situ transmission electron microscopy

studies of shear bands in a bulk metallic glass based composite. J Mater Res.

2001;16:2513-8.

142
[39] Li W, Bei H, Tong Y, Dmowski W, Gao YF. Structural heterogeneity induced

plasticity in bulk metallic glasses: from well-relaxed fragile glass to metal-like

behavior. Appl Phys Lett. 2013;103:171910.

[40] Li W, Gao Y, Bei H. On the correlation between microscopic structural

heterogeneity and embrittlement behavior in metallic glasses. Sci Rep. 2015;5:14786.

[41] Liu YH, Wang D, Nakajima K, Zhang W, Hirata A, Nishi T, et al.

Characterization of nanoscale mechanical heterogeneity in a metallic glass by

dynamic force microscopy. Phys Rev Lett. 2011;106:125504.

[42] Wang G, Wang YT, Liu YH, Pan MX, Zhao DQ, Wang WH. Evolution of

nanoscale morphology on fracture surface of brittle metallic glass. Appl Phys Lett.

2006;89:121909.

[43] Wang G, Zhao DQ, Bai HY, Pan MX, Xia AL, Han BS, et al. Nanoscale

periodic morphologies on the fracture surface of brittle metallic glasses. Phys Rev

Lett. 2007;98:235501.

[44] Lewandowski JJ, Shazly M, Shamimi Nouri A. Intrinsic and extrinsic

toughening of metallic glasses. Scripta Mater. 2006;54:337-41.

[45] Lewandowski JJ, Wang WH, Greer AL. Intrinsic plasticity or brittleness of

metallic glasses. Philos Mag Lett. 2005;85:77-87.

[46] Spaepen F. A microscopic mechanism for steady state inhomogeneous flow in

metallic glasses. Acta Metall. 1977;25:407-15.

[47] Steif PS. Ductile versus brittle behavior of amorphous metals. J Mech Phys

Solids. 1983;31:359-88.

143
[48] Steif PS, Spaepen F, Hutchinson JW. Strain localization in amorphous metals.

Acta Metall. 1982;30:447-55.

[49] Wang Z, Wang WH. Flow unit model in metallic glasses. Acta Phys Sin.

2017;66:176103.

[50] Argon AS. Plastic deformation in metallic glasses. Acta Metall. 1979;27:47-58.

[51] Falk ML, Langer JS. Dynamics of viscoplastic deformation in amorphous solids.

Phys Rev E. 1998;57:7192-205.

[52] Langer JS. Shear-transformation-zone theory of deformation in metallic glasses.

Scripta Mater. 2006;54:375-9.

[53] Shi Y, Falk ML. Strain localization and percolation of stable structure in

amorphous solids. Phys Rev Lett. 2005;95:095502.

[54] Shi Y, Falk ML. Atomic-scale simulations of strain localization in three-

dimensional model amorphous solids. Phys Rev B. 2006;73:214201.

[55] Jiang MQ, Ling Z, Meng JX, Dai LH. Energy dissipation in fracture of bulk

metallic glasses via inherent competition between local softening and quasi-cleavage.

Philos Mag. 2008;88:407-26.

[56] Sharp ML, Nordmark GE, Menzemer CC. Fatigue design of aluminium

components and structures. New York, NY: McGraw-Hill; 1996.

[57] Suresh S. Fatigue of materials. Cambridge, UK: Cambridge University Press;

1998.

[58] ASM International Handbook Committee. ASM handbook. Volume 19, Fatigue

and fracture. Materials Park, Ohio: ASM International; 1996.

144
[59] Choi-Yim H, Conner RD, Szuecs F, Johnson WL. Processing, microstructure

and properties of ductile metal particulate reinforced Zr57Nb5Al10Cu15.4Ni12.6

bulk metallic glass composites. Acta Mater. 2002;50:2737-45.

[60] Mukai T, Nieh TG, Kawamura Y, Inoue A, Higashi K. Effect of strain rate on

compressive behavior of a Pd40Ni40P20 bulk metallic glass. Intermetallics.

2002;10:1071-7.

[61] He G, Löser W, Eckert J, Schultz L. Phase transformation and mechanical

properties of Zr-base bulk glass-forming alloys. Mater Sci Eng A. 2003;352:179-85.

[62] Li H, Fan C, Tao K, Choo H, Liaw PK. Compressive behavior of a Zr-based

metallic glass at cryogenic temperatures. Adv Mater. 2006;18:752-4.

[63] Fan C, Li H, Kecskes LJ, Tao K, Choo H, Liaw PK, et al. Mechanical behavior

of bulk amorphous alloys reinforced by ductile particles at cryogenic temperatures.

Phys Rev Lett. 2006;96:145506.

[64] Bharathula A, Lee S-W, Wright WJ, Flores KM. Compression testing of metallic

glass at small length scales: Effects on deformation mode and stability. Acta Mater.

2010;58:5789-96.

[65] Pan J, Chen Q, Liu L, Li Y. Softening and dilatation in a single shear band. Acta

Mater. 2011;59:5146-58.

[66] Sun BA, Pauly S, Tan J, Stoica M, Wang WH, Kühn U, et al. Serrated flow and

stick–slip deformation dynamics in the presence of shear-band interactions for a Zr-

based metallic glass. Acta Mater. 2012;60:4160-71.

[67] Ogura T, Masumoto T, Fukushima K. Fatigue fracture of amorphous Pd-

20at.%Si alloy. Scripta Metall. 1975;9:109-13.

145
[68] Davis LA. Fatigue of metallic glasses. J Mater Sci. 1976;11:711-7.

[69] Gilbert CJ, Lippmann JM, Ritchie RO. Fatigue of a Zr-Ti-Cu-Ni-Be bulk

amorphous metal: stress/life and crack-growth behavior. Scripta Mater. 1998;38:537-

42.

[70] Gilbert CJ, Schroeder V, Ritchie RO. Mechanisms for fracture and fatigue-crack

propagation in a bulk metallic glass. Metall Mater Trans A. 1999;30:1739-53.

[71] Menzel BC, Dauskardt RH. Stress-life fatigue behavior of a Zr-based bulk

metallic glass. Acta Mater. 2006;54:935-43.

[72] Peter WH, Liaw PK, Buchanan RA, Liu CT, Brooks CR, Horton JA, et al.

Fatigue behavior of Zr52.5Al10Ti5Cu17.9Ni14.6 bulk metallic glass. Intermetallics.

2002;10:1125-9.

[73] Peter WH, Buchanan RA, Liu CT, Liaw PK. The fatigue behavior of a

zirconium-based bulk metallic glass in vacuum and air. J Non·Cryst Solids.

2003;317:187-92.

[74] Wang GY, Liaw PK, Peter WH, Yang B, Yokoyama Y, Benson ML, et al.

Fatigue behavior of bulk-metallic glasses. Intermetallics. 2004;12:885-92.

[75] Wang GY, Liaw PK, Peter WH, Yang B, Freels M, Yokoyama Y, et al. Fatigue

behavior and fracture morphology of Zr50Al10Cu40 and Zr50Al10Cu30Ni10 bulk-

metallic glasses. Intermetallics. 2004;12:1219-27.

[76] Wang GY, Liaw PK, Peker A, Yang B, Benson ML, Yuan W, et al. Fatigue

behavior of Zr–Ti–Ni–Cu–Be bulk-metallic glasses. Intermetallics. 2005;13:429-35.

146
[77] Wang GY, Liaw PK, Peker A, Freels M, Peter WH, Buchanan RA, et al.

Comparison of fatigue behavior of a bulk metallic glass and its composite.

Intermetallics. 2006;14:1091-7.

[78] Wang GY, Liaw PK, Yokoyama Y, Peter WH, Yang B, Freels M, et al.

Influence of air and vacuum environment on fatigue behavior of Zr-based bulk

metallic glasses. J Alloys Compd. 2007;434-435:68-70.

[79] Wang GY, Liaw PK, Yokoyama Y, Peker A, Peter WH, Yang B, et al. Studying

fatigue behavior and Poisson's ratio of bulk-metallic glasses. Intermetallics.

2007;15:663-7.

[80] Yokoyama Y, Liaw PK, Nishijima M, Hiraga K, Buchanan RA, Inoue A.

Fatigue-strength enhancement of cast Zr50Cu40Al10 glassy alloys. Mater Trans, JIM.

2006;47:1286-93.

[81] Qiao DC, Wang GY, Liaw PK, Ponnambalam V, Poon SJ, Shiflet GJ. Fatigue

behavior of an Fe48Cr15Mo14Er2C15B6 amorphous steel. J Mater Res.

2007;22:544-50.

[82] Qiao DC, Fan GJ, Liaw PK, Choo H. Fatigue behaviors of the Cu47.5Zr47.5Al5

bulk-metallic glass (BMG) and Cu47.5Zr38Hf9.5Al5 BMG composite. Int J Fatigue.

2007;29:2149-54.

[83] Wang GY, Liaw PK, Morrison ML. Progress in studying the fatigue behavior of

Zr-based bulk-metallic glasses and their composites. Intermetallics. 2009;17:579-90.

[84] Wang G, Liaw PK. Bending-fatigue behavior of bulk metallic glasses and their

composites. JOM. 2010;62:25-33.

147
[85] Li W, Bei H, Gao Y. Effects of geometric factors and shear band patterns on

notch sensitivity in bulk metallic glasses. Intermetallics. 2016;79:12-9.

[86] Li W, Gao Y, Bei H. Instability analysis and free volume simulations of shear

band directions and arrangements in notched metallic glasses. Sci Rep. 2016;6:34878.

[87] Nieh T, Mukai T, Liu C, Wadsworth J. Superplastic behavior of a Zr-10Al-5Ti-

17.9Cu-14.6 Ni metallic glass in the supercooled liquid region. Scripta Mater.

1999;40:1021-7.

[88] Kawamura Y, Nakamura T, Kim KB. Superplasticity in Pd40Ni40P20 metallic

glass. Mater Sci Forum. 1999;304:349-54.

[89] Nieh TG, Schuh C, Wadsworth J, Li Y. Strain rate-dependent deformation in

bulk metallic glasses. Intermetallics. 2002;10:1177-82.

[90] Nieh TG, Wadsworth J, Liu CT, Ohkubo T, Hirotsu Y. Plasticity and structural

instability in a bulk metallic glass deformed in the supercooled liquid region. Acta

Mater. 2001;49:2887-96.

[91] Yang B, Liu CT, Nieh TG, Morrison ML, Liaw PK, Buchanan RA. Localized

heating and fracture criterion for bulk metallic glasses. J Mater Res. 2006;21:915-22.

[92] Yang B, Morrison ML, Liaw PK, Buchanan RA, Wang G, Liu CT, et al.

Dynamic evolution of nanoscale shear bands in a bulk-metallic glass. Appl Phys Lett.

2005;86:141904.

[93] Lewandowski JJ, Greer AL. Temperature rise at shear bands in metallic glasses.

Nat Mater. 2005;5:15.

[94] Gao YF, Yang B, Nieh TG. Thermomechanical instability analysis of

inhomogeneous deformation in amorphous alloys. Acta Mater. 2007;55:2319-27.

148
[95] Verduzco JA, Hand RJ, Davies HA. Fatigue behaviour of Fe–Cr–Si–B metallic

glass wires. Int J Fatigue. 2002;24:1089-94.

[96] Zhai T, Xu YG, Martin JW, Wilkinson AJ, Briggs GAD. A self-aligning four-

point bend testing rig and sample geometry effect in four-point bend fatigue. Int J

Fatigue. 1999;21:889-94.

[97] Jia H, Liu F, An Z, Li W, Wang G, Chu JP, et al. Thin-film metallic glasses for

substrate fatigue-property improvements. Thin Solid Films. 2014;561:2-27.

[98] Launey ME, Busch R, Kruzic JJ. Effects of free volume changes and residual

stresses on the fatigue and fracture behavior of a Zr–Ti–Ni–Cu–Be bulk metallic

glass. Acta Mater. 2008;56:500-10.

[99] Launey ME, Busch R, Kruzic JJ. Influence of structural relaxation on the fatigue

behavior of a Zr41.25Ti13.75Ni10Cu12.5Be22.5 bulk amorphous alloy. Scripta

Mater. 2006;54:483-7.

[100] Morrison ML, Buchanan RA, Liaw PK, Green BA, Wang GY, Liu CT, et al.

Four-point-bending-fatigue behavior of the Zr-based Vitreloy 105 bulk metallic glass.

Mater Sci Eng A. 2007;467:190-7.

[101] Wang GY, Qiao DC, Yokoyama Y, Freels M, Inoue A, Liaw PK. Effects of

loading modes on the fatigue behavior of Zr-based bulk-metallic glasses. J Alloys

Compd. 2009;483:143-5.

[102] El-Shabasy AB, Lewandowski JJ. Fatigue coaxing experiments on a Zr-based

bulk-metallic glass. Scripta Mater. 2010;62:481-4.

149
[103] Freels M, Liaw PK, Wang GY, Zhang QS, Hu ZQ. Stress-life fatigue behavior

and fracture-surface morphology of a Cu-based bulk-metallic glass. J Mater Res.

2007;22:374-81.

[104] Wang Y, Liu Y, Liu L. Fatigue behaviors of a Ni-free ZrCuFeAlAg bulk

metallic glass in simulated body fluid. J Mater Sci Technol. 2014;30:622-6.

[105] Naleway SE, Greene RB, Gludovatz B, Dave NKN, Ritchie RO, Kruzic JJ. A

highly fatigue-resistant Zr-based bulk metallic glass. Metall Mater Trans A.

2013;44:5688-93.

[106] Wang GY, Liaw PK, Yokoyama Y, Freels M, Buchanan RA, Inoue A, et al.

Effects of partial crystallization on compression and fatigue behavior of Zr-based

bulk metallic glasses. J Mater Res. 2007;22:493-500.

[107] Fujita K, Zhang W, Shen B, Amiya K, Ma CL, Nishiyama N. Fatigue properties

in high strength bulk metallic glasses. Intermetallics. 2012;30:12-8.

[108] Zhang LK, Chen ZH, Chen D, Zhao XY, Zheng Q. Four-point-bending-fatigue

behavior of the Cu45Zr45Ag7Al3 bulk metallic glass. J Non·Cryst Solids.

2013;370:31-6.

[109] Flores KM, Johnson WL, Dauskardt RH. Fracture and fatigue behavior of a Zr–

Ti–Nb ductile phase reinforced bulk metallic glass matrix composite. Scripta Mater.

2003;49:1181-7.

[110] Qiao DC, Liaw PK, Fan C, Lin YH, Wang GY, Choo H, et al. Fatigue and

fracture behavior of (Zr58Ni13.6Cu18Al10.4)99Nb1 bulk-amorphous alloy.

Intermetallics. 2006;14:1043-50.

150
[111] Qiao JW, Huang EW, Wang GY, Yang HJ, Liang W, Zhang Y, et al.

Characteristic of improved fatigue performance for Zr-based bulk metallic glass

matrix composites. Mater Sci Eng A. 2013;563:101-5.

[112] El-Shabasy AB, Hassan HA, Liu Y, Li D, Lewandowski JJ. High cycle fatigue

behavior of a nanostructured composite produced via extrusion of amorphous

Al89Gd7Ni3Fe1 alloy powders. Mater Sci Eng A. 2009;513-514:202-7.

[113] Launey ME, Hofmann DC, Johnson WL, Ritchie RO. Solution to the problem

of the poor cyclic fatigue resistance of bulk metallic glasses. P Natl Acad Sci.

2009;106:4986-91.

[114] Alpas AT, Edwards L, Reid CN. Fracture and fatigue crack propagation in a

nickel-base metallic glass. Metall Trans A. 1989;20:1395-409.

[115] Brown WF. Aerospace structural metals handbook. Syracuse, NY: Metals and

ceramics information center 1989.

[116] Rao KTV, Ritchie RO. Fatigue of aluminium—lithium alloys. Int Mater Rev.

1992;37:153-86.

[117] Wu Y, Xiao Y, Chen G, Liu CT, Lu Z. Bulk metallic glass composites with

transformation-mediated work-hardening and ductility. Adv Mater. 2010;22:2770-3.

[118] Hess PA, Menzel BC, Dauskardt RH. Fatigue damage in bulk metallic glass II:

experiments. Scripta Mater. 2006;54:355-61.

[119] Qiao D, Wang G, Jiang W, Yokoyama Y, Liaw PK, Choo H. Compression-

compression fatigue and fracture behaviors of Zr50Al10Cu37Pd3 bulk-metallic glass.

Mater Trans, JIM. 2007;48:1828-33.

151
[120] Zhang ZF, Eckert J, Schultz L. Fatigue and fracture behavior of bulk metallic

glass. Metall Mater Trans A. 2004;35:3489-98.

[121] Yokoyama Y, Fukaura K, Sunada H. Fatigue properties and microstructures of

Zr55Cu30Al10Ni5 bulk glassy alloys. Mater Trans, JIM. 2000;41:675-80.

[122] Maruyama N, Nakazawa K, Hanawa T. Fatigue properties of Zr-based bulk

amorphous alloy in phosphate buffered saline solution. Mater Trans, JIM.

2002;43:3118-21.

[123] Freels M, Wang GY, Zhang W, Liaw PK, Inoue A. Cyclic compression

behavior of a Cu–Zr–Al–Ag bulk metallic glass. Intermetallics. 2011;19:1174-83.

[124] Wang G, Liaw PK, Senkov ON, Miracle DB, Morrison ML. Mechanical and

fatigue behavior of Ca65Mg15Zn20 bulk-metallic glass. Adv Eng Mater. 2009;11:27-

34.

[125] Fujita K, Takayama T, Inoue A, Zhang T, Kimura H. Effect of nanocrystalline

dispersion on fatigue in a Zr-based bulk metallic glass. J Jpn Inst Met. 2003;67:79-84.

[126] Wei R, Chen LB, Li HP, Li FS. Compression-compression fatigue behavior of

CuZr-based bulk metallic glass composite containing B2 phase. Intermetallics.

2017;85:54-8.

[127] Yokoyama Y, Fukaura K, Inoue A. Effect of Ni addition on fatigue properties

of bulk glassy Zr50Cu40Al10 alloys. Mater Trans, JIM. 2004;45:1672-8.

[128] Yokoyama Y, Nishiyama N, Fukaura K, Sunada H, Inoue A. Rotating-beam

fatigue strength of Pd40Cu30Ni10P20 bulk amorphous alloy. Mater Trans, JIM.

1999;40:696-9.

152
[129] Jia H, Muntele CI, Huang L, Li X, Li G, Zhang T, et al. A study on the surface

structures and properties of Ni-free Zr-based bulk metallic glasses after Ar and Ca ion

implantation. Intermetallics. 2013;41:35-43.

[130] Huang L, Cao Z, Meyer HM, Liaw PK, Garlea E, Dunlap JR, et al. Responses

of bone-forming cells on pre-immersed Zr-based bulk metallic glasses: Effects of

composition and roughness. Acta Biomater. 2011;7:395-405.

[131] Huang L, Pu C, Fisher RK, Mountain DJH, Gao Y, Liaw PK, et al. A Zr-based

bulk metallic glass for future stent applications: Materials properties, finite element

modeling, and in vitro human vascular cell response. Acta Biomater. 2015;25:356-68.

[132] Huang L, Wang G, Qiao D, Liaw PK, Pang S, Wang J, et al. Corrosion-fatigue

study of a Zr-based bulk-metallic glass in a physiologically relevant environment. J

Alloys Compd. 2010;504:S159-S62.

[133] Gilbert CJ, Ritchie RO, Johnson WL. Fracture toughness and fatigue-crack

propagation in a Zr–Ti–Ni–Cu–Be bulk metallic glass. Appl Phys Lett. 1997;71:476-

8.

[134] Wiest A, Wang G, Huang L, Roberts S, Demetriou MD, Liaw PK, et al.

Corrosion and corrosion fatigue of Vitreloy glasses containing low fractions of late

transition metals. Scripta Mater. 2010;62:540-3.

[135] Schroeder V, Gilbert CJ, Ritchie RO. A comparison of the mechanisms of

fatigue-crack propagation behavior in a Zr-based bulk amorphous metal in air and an

aqueous chloride solution. Mater Sci Eng A. 2001;317:145-52.

153
[136] Zhang H, Wang ZG, Qiu KQ, Zang QS, Zhang HF. Cyclic deformation and

fatigue crack propagation of a Zr-based bulk amorphous metal. Mater Sci Eng A.

2003;356:173-80.

[137] Nakai Y, Yoshioka Y. Stress corrosion and corrosion fatigue crack growth of

Zr-based bulk metallic glass in aqueous solutions. Metall Mater Trans A.

2010;41:1792-8.

[138] Philo SL, Heinrich J, Gallino I, Busch R, Kruzic JJ. Fatigue crack growth

behavior of a Zr58.5Cu15.6Ni12.8Al10.3Nb2.8 bulk metallic glass-forming alloy.

Scripta Mater. 2011;64:359-62.

[139] Fujita K, Inoue A, Zhang T. Fatigue crack propagation in a nanocrystalline Zr-

based bulk metallic glass. Mater Trans, JIM. 2000;41:1448-53.

[140] Alpas AT, Edwards L, Reid CN. Fatigue crack propagation in an Ni78Si10B12

metallic glass with amorphous and partially crystalline structures. Mater Sci Eng.

1988;98:501-4.

[141] Boopathy K, Hofmann DC, Johnson WL, Ramamurty U. Near-threshold fatigue

crack growth in bulk metallic glass composites. J Mater Res. 2009;24:3611-9.

[142] Hess PA, Dauskardt RH. Mechanisms of elevated temperature fatigue crack

growth in Zr–Ti–Cu–Ni–Be bulk metallic glass. Acta Mater. 2004;52:3525-33.

[143] Schroeder V, Ritchie RO. Stress-corrosion fatigue–crack growth in a Zr-based

bulk amorphous metal. Acta Mater. 2006;54:1785-94.

[144] Cameron KK, Dauskardt RH. Fatigue damage in bulk metallic glass I:

simulation. Scripta Mater. 2006;54:349-53.

154
[145] Yeh JW, Chen SK, Lin SJ, Gan JY, Chin TS, Shun TT, et al. Nanostructured

high-entropy alloys with multiple principal elements: Novel alloy design concepts

and outcomes. Adv Eng Mater. 2004;6:299-303.

[146] Zhang Y, Zuo TT, Tang Z, Gao MC, Dahmen KA, Liaw PK, et al.

Microstructures and properties of high-entropy alloys. Prog Mater Sci. 2014;61:1-93.

[147] Hemphill MA, Yuan T, Wang GY, Yeh JW, Tsai CW, Chuang A, et al. Fatigue

behavior of Al0.5CoCrCuFeNi high entropy alloys. Acta Mater. 2012;60:5723-34.

[148] Tang Z, Yuan T, Tsai C-W, Yeh J-W, Lundin CD, Liaw PK. Fatigue behavior

of a wrought Al0.5CoCrCuFeNi two-phase high-entropy alloy. Acta Mater.

2015;99:247-58.

[149] Seifi M, Li D, Yong Z, Liaw PK, Lewandowski JJ. Fracture toughness and

fatigue crack growth behavior of as-cast high-entropy alloys. JOM. 2015;67:2288-95.

[150] Wright WJ, Hufnagel TC, Nix WD. Free volume coalescence and void

formation in shear bands in metallic glass. J Appl Phys. 2003;93:1432-7.

[151] Li J, Wang ZL, Hufnagel TC. Characterization of nanometer-scale defects in

metallic glasses by quantitative high-resolution transmission electron microscopy.

Phys Rev B. 2002;65:144201.

[152] Li J, Spaepen F, Hufnagel TC. Nanometre-scale defects in shear bands in a

metallic glass. Philos Mag A. 2002;82:2623-30.

[153] Wang G, Liaw PK, Yoshihiko Y, Peker A, Freels M, Fielden D, et al. Direct

comparisons of the fatigue behavior of bulk-metallic glasses and crystalline alloys.

Key Eng Mater. 2008;378-379:329-38.

155
[154] Wang GY, Liaw PK, Yokoyama Y, Inoue A, Liu CT. Fatigue behavior of Zr-

based bulk-metallic glasses. Mater Sci Eng A. 2008;494:314-23.

[155] Morrison ML, Buchanan RA, Liaw PK, Green BA, Wang GY, Liu CT, et al.

Corrosion–fatigue studies of the Zr-based Vitreloy 105 bulk metallic glass. Mater Sci

Eng A. 2007;467:198-206.

[156] Wang C-C, Mao Y-W, Shan Z-W, Dao M, Li J, Sun J, et al. Real-time, high-

resolution study of nanocrystallization and fatigue cracking in a cyclically strained

metallic glass. P Natl Acad Sci. 2013;110:19725-30.

[157] Turnbull D. Formation of crystal nuclei in liquid metals. J Appl Phys.

1950;21:1022-8.

[158] Debenedetti PG. Metastable liquids: concepts and principles. Princeton, NJ:

Princeton University Press; 1996.

[159] Kawasaki T, Tanaka H. Formation of a crystal nucleus from liquid. P Natl Acad

Sci. 2010;107:14036-41.

[160] Kim JJ, Choi Y, Suresh S, Argon AS. Nanocrystallization during

nanoindentation of a bulk amorphous metal alloy at room temperature. Science.

2002;295:654.

[161] Jiang WH, Atzmon M. The effect of compression and tension on shear-band

structure and nanocrystallization in amorphous Al90Fe5Gd5: a high-resolution

transmission electron microscopy study. Acta Mater. 2003;51:4095-105.

[162] Flores KM, Dauskardt RH. Enhanced toughness due to stable crack tip damage

zones in bulk metallic glass. Scripta Mater. 1999;41:937-43.

156
[163] Wang G, Liaw PK, Jin X, Yokoyama Y, Huang EW, Jiang F, et al. Fatigue

initiation and propagation behavior in bulk-metallic glasses under a bending load. J

Appl Phys. 2010;108:113512.

[164] Jin X, Keer LM, Chez EL. Numerical simulation of growth pattern of a fluid-

filled subsurface crack under moving hertzian loading. Int J Fract. 2007;142:219.

[165] Palaniswamy K, Knauss W. On the problem of crack extension in brittle solids

under general loading. Mechanics Today, Volume 4. Oxford, UK: Pergamon Press;

1978.

[166] Ogura T, Fukushima K, Masumoto T. Propagation of fatigue cracks in

amorphous metals. Scripta Metall. 1975;9:979-83.

[167] Jin X, Wang G, Keer LM, Liaw PK, Wang QJ. Modeling crack growth from

pores under compressive loading with application to metallic glasses. J Alloys

Compd. 2011;509:S115-S8.

[168] Conner RD, Li Y, Nix WD, Johnson WL. Shear band spacing under bending of

Zr-based metallic glass plates. Acta Mater. 2004;52:2429-34.

[169] Guo H, Yan PF, Wang YB, Tan J, Zhang ZF, Sui ML, et al. Tensile ductility

and necking of metallic glass. Nat Mater. 2007;6:735-9.

[170] Greer JR, De Hosson JTM. Plasticity in small-sized metallic systems: Intrinsic

versus extrinsic size effect. Prog Mater Sci. 2011;56:654-724.

[171] Inoue A, Wang XM. Bulk amorphous FC20 (Fe–C–Si) alloys with small

amounts of B and their crystallized structure and mechanical properties. Acta Mater.

2000;48:1383-95.

157
[172] Inoue A, Katsuya A, Amiya K, Masumoto T. Preparation of amorphous Fe-Si-

B and Co-Si-B alloy wires by a melt extraction method and their mechanical and

magnetic properties. Mater Trans, JIM. 1995;36:802-9.

[173] Conner RD, Johnson WL, Paton NE, Nix WD. Shear bands and cracking of

metallic glass plates in bending. J Appl Phys. 2003;94:904-11.

[174] Shi Y, Louca D, Wang G, Liaw PK. Compression-compression fatigue study on

model metallic glass nanowires by molecular dynamics simulations. J Appl Phys.

2011;110:023523.

[175] Jang D, Greer JR. Transition from a strong-yet-brittle to a stronger-and-ductile

state by size reduction of metallic glasses. Nat Mater. 2010;9:215-9.

[176] Jang D, Gross CT, Greer JR. Effects of size on the strength and deformation

mechanism in Zr-based metallic glasses. Int J Plast. 2011;27:858-67.

[177] Jang D, Maaß R, Wang G, Liaw PK, Greer JR. Fatigue deformation of

microsized metallic glasses. Scripta Mater. 2013;68:773-6.

[178] Wang GY, Liaw PK, Yokoyama Y, Inoue A. Size effects on the fatigue

behavior of bulk metallic glasses. J Appl Phys. 2011;110:113507.

[179] Chuang C-P, Yuan T, Dmowski W, Wang G-Y, Freels M, Liaw PK, et al.

Fatigue-induced damage in Zr-based bulk metallic glasses. Sci Rep. 2013;3:2578.

[180] Tong Y, Iwashita T, Dmowski W, Bei H, Yokoyama Y, Egami T. Structural

rejuvenation in bulk metallic glasses. Acta Mater. 2015;86:240-6.

[181] Gludovatz B, Naleway SE, Ritchie RO, Kruzic JJ. Size-dependent fracture

toughness of bulk metallic glasses. Acta Mater. 2014;70:198-207.

158
[182] ASTM E399-12e3. Standard test method for linear-elastic plane-strain fracture

toughness KIc of metallic materials. West Conshohocken, PA: ASTM International;

2012.

[183] Chu JP, Greene JE, Jang JSC, Huang JC, Shen Y-L, Liaw PK, et al. Bendable

bulk metallic glass: effects of a thin, adhesive, strong, and ductile coating. Acta

Mater. 2012;60:3226-38.

[184] Chu JP, Huang JC, Jang JSC, Wang YC, Liaw PK. Thin film metallic glasses:

preparations, properties, and applications. JOM. 2010;62:19-24.

[185] Chu JP, Jang JSC, Huang JC, Chou HS, Yang Y, Ye JC, et al. Thin film

metallic glasses: Unique properties and potential applications. Thin Solid Films.

2012;520:5097-122.

[186] Chu JP, Lee CM, Huang RT, Liaw PK. Zr-based glass-forming film for fatigue-

property improvements of 316L stainless steel: annealing effects. Surf Coat Technol.

2011;205:4030-4.

[187] Liu FX, Liaw PK, Jiang WH, Chiang CL, Gao YF, Guan YF, et al. Fatigue-

resistance enhancements by glass-forming metallic films. Mater Sci Eng A.

2007;468-470:246-52.

[188] Liu FX, Chiang CL, Chu JP, Gao YF, Liaw PK. Effects of glass-forming

metallic film on the fatigue behavior of C-2000 Ni-based alloy. Mater Res Soc Symp

Proc: Cambridge University Press; 2006. p. 1-6.

[189] Liu FX. Microstructures and mechanical behavior of multicomponent metallic-

glass thin films [Doctoral thesis]. Knoxville, TN: The University of Tennessee; 2009.

159
[190] Chang YZ, Tsai PH, Li JB, Lin HC, Jang JSC, Li C, et al. Zr-based metallic

glass thin film coating for fatigue-properties improvement of 7075-T6 aluminum

alloy. Thin Solid Films. 2013;544:331-4.

[191] Ye JC, Chu JP, Chen YC, Wang Q, Yang Y. Hardness, yield strength, and

plastic flow in thin film metallic-glass. J Appl Phys. 2012;112:053516.

[192] Cheng YQ, Ma E. Intrinsic shear strength of metallic glass. Acta Mater.

2011;59:1800-7.

[193] Lee CJ, Lai YH, Huang JC, Du XH, Wang L, Nieh TG. Strength variation and

cast defect distribution in metallic glasses. Scripta Mater. 2010;63:105-8.

[194] Greer AL, Cheng YQ, Ma E. Shear bands in metallic glasses. Mater Sci Eng R

Rep. 2013;74:71-132.

[195] Yu C-C, Chu JP, Jia H, Shen Y-L, Gao Y, Liaw PK, et al. Influence of thin-film

metallic glass coating on fatigue behavior of bulk metallic glass: Experiments and

finite element modeling. Mater Sci Eng A. 2017;692:146-55.

[196] Wang S, Ye YF, Sun BA, Liu CT, Shi SQ, Yang Y. Softening-induced plastic

flow instability and indentation size effect in metallic glass. J Mech Phys Solids.

2015;77:70-85.

[197] Schuh CA, Nieh TG. A nanoindentation study of serrated flow in bulk metallic

glasses. Acta Mater. 2003;51:87-99.

[198] Jia HL. Improving mechanical properties of bulk metallic glasses by

approaches of in-situ composites and thin films [Doctoral thesis]. Knoxville, TN: The

University of Tennessee; 2015.

160
[199] Yu C-C, Lee CM, Chu JP, Greene JE, Liaw PK. Fracture-resistant thin-film

metallic glass: Ultra-high plasticity at room temperature. APL Mater. 2016;4:116101.

[200] Stoudt MR, Cammarata R, Ricker RE. Suppression of fatigue cracking with

nanometer-scale multilayered coatings. Scripta Mater. 2000;43:491-6.

[201] Kim KR, Suh CM, Murakami RI, Chung CW. Effect of intrinsic properties of

ceramic coatings on fatigue behavior of Cr–Mo–V steels. Surf Coat Technol.

2003;171:15-23.

[202] Cairney JM, Tsukano R, Hoffman MJ, Yang M. Degradation of TiN coatings

under cyclic loading. Acta Mater. 2004;52:3229-37.

[203] Jaeger G, Endler I, Bartsch K, Heilmaier M, Leonhardt A. Fatigue behavior of

duplex treated TiCxN1−x- and Ti1−xAlxN-hard coating steel compounds. Surf Coat

Technol. 2002;150:282-9.

[204] Jaroslav M. Coatings and surface modification technologies: a finite element

bibliography (1995–2005). Modell Simul Mater Sci Eng. 2005;13:935.

[205] Oliver WC, Pharr GM. An improved technique for determining hardness and

elastic modulus using load and displacement sensing indentation experiments. J

Mater Res. 1992;7:1564-83.

[206] Saha R, Nix WD. Effects of the substrate on the determination of thin film

mechanical properties by nanoindentation. Acta Mater. 2002;50:23-38.

[207] Chou HS, Huang JC, Chang LW, Nieh TG. Structural relaxation and

nanoindentation response in Zr–Cu–Ti amorphous thin films. Appl Phys Lett.

2008;93:191901.

161
[208] Liu MC, Lee CJ, Lai YH, Huang JC. Microscale deformation behavior of

amorphous/nanocrystalline multilayered pillars. Thin Solid Films. 2010;518:7295-9.

[209] Chou HS, Huang JC, Chang LW. Mechanical properties of ZrCuTi thin film

metallic glass with high content of immiscible tantalum. Surf Coat Technol.

2010;205:587-90.

[210] Johnson WL, Samwer K. A universal criterion for plastic yielding of metallic

glasses with a (T/Tg)2/3 temperature dependence. Phys Rev Lett. 2005;95:195501.

[211] Yang Y, Ye JC, Lu J, Liu FX, Liaw PK. Effects of specimen geometry and base

material on the mechanical behavior of focused-ion-beam-fabricated metallic-glass

micropillars. Acta Mater. 2009;57:1613-23.

[212] Zhang ZF, Zhang H, Pan XF, Das J, Eckert J. Effect of aspect ratio on the

compressive deformation and fracture behaviour of Zr-based bulk metallic glass.

Philos Mag Lett. 2005;85:513-21.

[213] Ritchie RO. The conflicts between strength and toughness. Nat Mater.

2011;10:817.

[214] Demetriou MD, Launey ME, Garrett G, Schramm JP, Hofmann DC, Johnson

WL, et al. A damage-tolerant glass. Nat Mater. 2011;10:123.

[215] Launey ME, Ritchie RO. On the Fracture Toughness of Advanced Materials.

Adv Mater. 2009;21:2103-10.

[216] Ashby MF, Greer AL. Metallic glasses as structural materials. Scripta Mater.

2006;54:321-6.

[217] Xi XK, Zhao DQ, Pan MX, Wang WH, Wu Y, Lewandowski JJ. Fracture of

brittle metallic glasses: brittleness or plasticity. Phys Rev Lett. 2005;94:125510.

162
[218] Hess PA, Poon SJ, Shiflet GJ, Dauskardt RH. Indentation fracture toughness of

amorphous steel. J Mater Res. 2005;20:783-6.

[219] Schroers J, Johnson WL. Ductile bulk metallic glass. Phys Rev Lett.

2004;93:255506.

[220] Gu XJ, Poon SJ, Shiflet GJ, Lewandowski JJ. Compressive plasticity and

toughness of a Ti-based bulk metallic glass. Acta Mater. 2010;58:1708-20.

[221] Madge S. Toughness of bulk metallic glasses. Metals. 2015;5:1279.

[222] Jiang F, Zhang DH, Zhang LC, Zhang ZB, He L, Sun J, et al. Microstructure

evolution and mechanical properties of Cu46Zr47Al7 bulk metallic glass composite

containing CuZr crystallizing phases. Mater Sci Eng A. 2007;467:139-45.

[223] Hufnagel TC, Vempati UK, Almer JD. Crack-tip strain field mapping and the

toughness of metallic glasses. PLOS ONE. 2013;8:e83289.

[224] Tandaiya P, Ramamurty U, Ravichandran G, Narasimhan R. Effect of

Poisson’s ratio on crack tip fields and fracture behavior of metallic glasses. Acta

Mater. 2008;56:6077-86.

[225] Hays CC, Kim CP, Johnson WL. Microstructure controlled shear band pattern

formation and enhanced plasticity of bulk metallic glasses containing in situ formed

ductile phase dendrite dispersions. Phys Rev Lett. 2000;84:2901-4.

[226] Qiao JW, Sun AC, Huang EW, Zhang Y, Liaw PK, Chuang CP. Tensile

deformation micromechanisms for bulk metallic glass matrix composites: from work-

hardening to softening. Acta Mater. 2011;59:4126-37.

163
[227] Hofmann DC, Suh J-Y, Wiest A, Duan G, Lind M-L, Demetriou MD, et al.

Designing metallic glass matrix composites with high toughness and tensile ductility.

Nature. 2008;451:1085-9.

[228] Jia HL, Zheng LL, Li WD, Li N, Qiao JW, Wang GY, et al. Insights from the

lattice-strain evolution on deformation mechanisms in metallic-glass-matrix

composites. Metall Mater Trans A. 2015;46:2431-42.

[229] Launey ME, Hofmann DC, Suh JY, Kozachkov H, Johnson WL, Ritchie RO.

Fracture toughness and crack-resistance curve behavior in metallic glass-matrix

composites. Appl Phys Lett. 2009;94:241910.

[230] Haggag FM, Underwood JH. Compliance of a three-point bend specimen at

load line. Int J Fract. 1984;26:R63-R5.

[231] ASTM. Annual book of ASTM standards: Metal test methods and analytical

procedures. Metals - mechanical testing; elevated and low-temperature tests;

metallography. Section 3. Volume 3.1. West Conshohocken, PA: ASTM; 2007.

[232] Lowhaphandu P, Lewandowski JJ. Fracture toughness and notched toughness

of bulk amorphous alloy: Zr-Ti-Ni-Cu-Be. Scripta Mater. 1998;38:1811-7.

[233] Kumar G, Rector D, Conner RD, Schroers J. Embrittlement of Zr-based bulk

metallic glasses. Acta Mater. 2009;57:3572-83.

[234] Zhang Q, Zhang W, Inoue A. Transition from plasticity to brittleness in Cu-Zr-

based bulk metallic glasses. Mater Trans, JIM. 2007;48:1272-5.

[235] Kim KB, Das J, Baier F, Tang MB, Wang WH, Eckert J. Heterogeneity of a

Cu47.5Zr47.5Al5 bulk metallic glass. Appl Phys Lett. 2006;88:051911.

164
[236] Murali P, Guo TF, Zhang YW, Narasimhan R, Li Y, Gao HJ. Atomic scale

fluctuations govern brittle fracture and cavitation behavior in metallic glasses. Phys

Rev Lett. 2011;107:215501.

[237] Murali P, Narasimhan R, Guo TF, Zhang YW, Gao HJ. Shear bands mediate

cavitation in brittle metallic glasses. Scripta Mater. 2013;68:567-70.

[238] Kelly A, Tyson WR, Cottrell AH. Ductile and brittle crystals. Philos Mag.

1967;15:567-86.

[239] Rice JR, Thomson R. Ductile versus brittle behaviour of crystals. Philos Mag.

1974;29:73-97.

[240] Liu Y, Wu H, Liu CT, Zhang Z, Keppens V. Physical factors controlling the

ductility of bulk metallic glasses. Appl Phys Lett. 2008;93:151915.

[241] Lewandowski JJ, Gu XJ, Shamimi Nouri A, Poon SJ, Shiflet GJ. Tough Fe-

based bulk metallic glasses. Appl Phys Lett. 2008;92:091918.

[242] Poon SJ, Zhu A, Shiflet GJ. Poisson’s ratio and intrinsic plasticity of metallic

glasses. Appl Phys Lett. 2008;92:261902.

[243] Raghavan R, Murali P, Ramamurty U. On factors influencing the ductile-to-

brittle transition in a bulk metallic glass. Acta Mater. 2009;57:3332-40.

[244] Kumar G, Prades-Rodel S, Blatter A, Schroers J. Unusual brittle behavior of

Pd-based bulk metallic glass. Scripta Mater. 2011;65:585-7.

[245] Cheng YQ, Ma E. Atomic-level structure and structure–property relationship in

metallic glasses. Prog Mater Sci. 2011;56:379-473.

[246] Mayr SG. Activation energy of shear transformation zones: a key for

understanding rheology of glasses and liquids. Phys Rev Lett. 2006;97:195501.

165
[247] Zink M, Samwer K, Johnson WL, Mayr SG. Plastic deformation of metallic

glasses: Size of shear transformation zones from molecular dynamics simulations.

Phys Rev B. 2006;73:172203.

[248] Li W, Bei H, Qu J, Gao Y. Effects of machine stiffness on the loading–

displacement curve during spherical nano-indentation. J Mater Res. 2013;28:1903-11.

[249] Lee MH, Lee JK, Kim KT, Thomas J, Das J, Kühn U, et al. Deformation-

induced microstructural heterogeneity in monolithic Zr44Ti11Cu9.8Ni10.2Be25 bulk

metallic glass. Phys Status Solidi-R. 2009;3:46-8.

[250] Zhu F, Nguyen HK, Song SX, Aji DPB, Hirata A, Wang H, et al. Intrinsic

correlation between β-relaxation and spatial heterogeneity in a metallic glass. Nat

Commun. 2016;7:11516.

[251] Hu YC, Guan PF, Li MZ, Liu CT, Yang Y, Bai HY, et al. Unveiling atomic-

scale features of inherent heterogeneity in metallic glass by molecular dynamics

simulations. Phys Rev B. 2016;93:214202.

[252] Ding J, Patinet S, Falk ML, Cheng Y, Ma E. Soft spots and their structural

signature in a metallic glass. P Natl Acad Sci. 2014;111:14052-6.

[253] Peng C-X, Song K-K, Wang L, Şopu D, Pauly S, Eckert J. Correlation between

structural heterogeneity and plastic deformation for phase separating FeCu metallic

glasses. Sci Rep. 2016;6:34340.

[254] Huang X, Ling Z, Wang YJ, Dai LH. Intrinsic structural defects on medium

range in metallic glasses. Intermetallics. 2016;75:36-41.

[255] Gu B, Liu F. Characterization of structural inhomogeneity in Al88Ce8Co4

metallic glass. Acta Mater. 2016;112:94-104.

166
[256] Sun BA, Hu YC, Wang DP, Zhu ZG, Wen P, Wang WH, et al. Correlation

between local elastic heterogeneities and overall elastic properties in metallic glasses.

Acta Mater. 2016;121:266-76.

[257] An ZN, Li WD, Liu FX, Liaw PK, Gao YF. Interface constraints on shear band

patterns in bonded metallic gass films under microindentation. Metall Mater Trans A.

2012;43:2729-41.

[258] Liu MC, Huang JC, Chen KW, Lin JF, Li WD, Gao YF, et al. Is the

compression of tapered micro- and nanopillar samples a legitimate technique for the

identification of deformation mode change in metallic glasses? Scripta Mater.

2012;66:817-20.

[259] Qiao J, Jia H, Liaw PK. Metallic glass matrix composites. Mater Sci Eng R

Rep. 2016;100:1-69.

[260] Li BS, Shakur Shahabi H, Scudino S, Eckert J, Kruzic JJ. Designed

heterogeneities improve the fracture reliability of a Zr-based bulk metallic glass.

Mater Sci Eng A. 2015;646:242-8.

[261] Hertzberg RW. Deformation and fracture mechanics of engineering materials.

3rd ed. New York, NY: Wiley; 1989.

[262] Qu R, Zhang P, Zhang Z. Notch effect of materials: strengthening or

weakening? J Mater Sci Technol. 2014;30:599-608.

[263] Qu RT, Calin M, Eckert J, Zhang ZF. Metallic glasses: notch-insensitive

materials. Scripta Mater. 2012;66:733-6.

[264] Kimura H, Masumoto T. Plastic constraint and ductility in tensile notched

specimens of amorphous Pd78Cu6Si16. Metall Trans A. 1983;14:709-16.

167
[265] Pan J, Zhou HF, Wang ZT, Li Y, Gao HJ. Origin of anomalous inverse notch

effect in bulk metallic glasses. J Mech Phys Solids. 2015;84:85-94.

[266] Lei X, Li C, Shi X, Xu X, Wei Y. Notch strengthening or weakening governed

by transition of shear failure to normal mode fracture. Sci Rep. 2015;5:10537.

[267] Varadarajan R, Lewandowski JJ. Stress-state effects on the fracture of a Zr-Ti-

Ni-Cu-Be bulk amorphous alloy. Metall Mater Trans A. 2010;41:1758-66.

[268] Sha ZD, Pei QX, Liu ZS, Zhang YW, Wang TJ. Necking and notch

strengthening in metallic glass with symmetric sharp-and-deep notches. Sci Rep.

2015;5:10797.

[269] Flores KM, Dauskardt RH. Mean stress effects on flow localization and failure

in a bulk metallic glass. Acta Mater. 2001;49:2527-37.

[270] Zhao JX, Zhang ZF. Comparison of compressive deformation and fracture

behaviors of Zr- and Ti-based metallic glasses with notches. Mater Sci Eng A.

2011;528:2967-73.

[271] Zhao JX, Wu FF, Qu RT, Li SX, Zhang ZF. Plastic deformability of metallic

glass by artificial macroscopic notches. Acta Mater. 2010;58:5420-32.

[272] Gu XW, Jafary-Zadeh M, Chen DZ, Wu Z, Zhang Y-W, Srolovitz DJ, et al.

Mechanisms of failure in nanoscale metallic glass. Nano Lett. 2014;14:5858-64.

[273] Sha ZD, He LC, Pei QX, Pan H, Liu ZS, Zhang YW, et al. On the notch

sensitivity of CuZr nanoglass. J Appl Phys. 2014;115:163507.

[274] Zhang ZF, Qu RT, Liu ZQ. Advances in fracture behavior and strength theory

of metallic glasses. Acta Metall Sin. 2016;52:1171-82.

168
[275] Qu RT, Zhao JX, Stoica M, Eckert J, Zhang ZF. Macroscopic tensile plasticity

of bulk metallic glass through designed artificial defects. Mater Sci Eng A.

2012;534:365-73.

[276] Sarac B, Schroers J. Designing tensile ductility in metallic glasses. Nat

Commun. 2013;4:2158.

[277] Tóth L, Yarema SY. Formation of the science of fatigue of metals. Part 1.

1825–1870. Mater Sci. 2006;42:673-80.

[278] Furuya Y, Matsuoka S. Improvement of gigacycle fatigue properties by

modified ausforming in 1600 and 2000 MPA-class low-alloy steels. Metall Mater

Trans A. 2002;33:3421-31.

[279] Wang XD, Qu RT, Wu SJ, Duan QQ, Liu ZQ, Zhu ZW, et al. Notch fatigue

behavior: metallic glass versus ultra-high strength steel. Sci Rep. 2016;6:35557.

[280] Liu FX, Gao YF, Liaw PK. Rate-dependent deformation behavior of Zr-based

metallic-glass coatings examined by nanoindentation. Metall Mater Trans A.

2008;39:1862-7.

[281] Li H, Li L, Fan C, Choo H, Liaw PK. Nanocrystalline coating enhanced

ductility in a Zr-based bulk metallic glass. J Mater Res. 2007;22:508-13.

[282] Rudnicki JW, Rice JR. Conditions for the localization of deformation in

pressure-sensitive dilatant materials. J Mech Phys Solids. 1975;23:371-94.

[283] Gao YF. An implicit finite element method for simulating inhomogeneous

deformation and shear bands of amorphous alloys based on the free-volume model.

Modell Simul Mater Sci Eng. 2006;14:1329-45.

169
[284] Gao YF, Wang L, Bei H, Nieh TG. On the shear-band direction in metallic

glasses. Acta Mater. 2011;59:4159-67.

[285] Xie S, George EP. Hardness and shear band evolution in bulk metallic glasses

after plastic deformation and annealing. Acta Mater. 2008;56:5202-13.

[286] Zhao M, Li M. Interpreting the change in shear band inclination angle in

metallic glasses. Appl Phys Lett. 2008;93:241906.

[287] Su C, LaManna JA, Gao Y, Oliver WC, Pharr GM. Plastic instability in

amorphous selenium near its glass transition temperature. J Mater Res. 2010;25:1015-

9.

[288] Zhao M, Li M. Local heating in shear banding of bulk metallic glasses. Scripta

Mater. 2011;65:493-6.

[289] Jiang Y. Numerical modeling of cyclic deformation in bulk metallic glasses.

Metals. 2016;6:217.

[290] Wales DJ, Doye JPK. Global optimization by basin-hopping and the lowest

energy structures of Lennard-Jones clusters containing up to 110 atoms. J Phys Chem

A. 1997;101:5111-6.

[291] Sietsma J, Thijsse BJ. Characterization of free volume in atomic models of

metallic glasses. Phys Rev B. 1995;52:3248-55.

[292] Li QK, Li M. Atomic scale characterization of shear bands in an amorphous

metal. Appl Phys Lett. 2006;88:241903.

[293] Lo YC, Chou HS, Cheng YT, Huang JC, Morris JR, Liaw PK. Structural

relaxation and self-repair behavior in nano-scaled Zr–Cu metallic glass under cyclic

loading: molecular dynamics simulations. Intermetallics. 2010;18:954-60.

170
[294] Shi Y. Size-independent shear band formation in amorphous nanowires made

from simulated casting. Appl Phys Lett. 2010;96:121909.

[295] Homer ER, Schuh CA. Mesoscale modeling of amorphous metals by shear

transformation zone dynamics. Acta Mater. 2009;57:2823-33.

[296] Eshelby JD. The determination of the elastic field of an ellipsoidal inclusion,

and related problems. Proc R Soc London, Ser A. 1957;241:376-96.

[297] Packard CE, Homer ER, Al-Aqeeli N, Schuh CA. Cyclic hardening of metallic

glasses under Hertzian contacts: experiments and STZ dynamics simulations. Philos

Mag. 2010;90:1373-90.

[298] Singh I, Guo TF, Murali P, Narasimhan R, Zhang YW, Gao HJ. Cavitation in

materials with distributed weak zones: Implications on the origin of brittle fracture in

metallic glasses. J Mech Phys Solids. 2013;61:1047-64.

[299] Singh I, Guo TF, Narasimhan R, Zhang YW. Cavitation in brittle metallic

glasses – Effects of stress state and distributed weak zones. Int J Solids Struct.

2014;51:4373-85.

[300] Tvergaard V, Hutchinson JW. The relation between crack growth resistance

and fracture process parameters in elastic-plastic solids. J Mech Phys Solids.

1992;40:1377-97.

[301] Gao YF, Bower AF. A simple technique for avoiding convergence problems in

finite element simulations of crack nucleation and growth on cohesive interfaces.

Modell Simul Mater Sci Eng. 2004;12:453.

171
[302] Fan C, Liaw PK, Wilson TW, Choo H, Gao YF, Liu CT, et al. Pair distribution

function study and mechanical behavior of as-cast and structurally relaxed Zr-based

bulk metallic glasses. Appl Phys Lett. 2006;89:231920.

[303] Wu Y, Bei H, Wang YL, Lu ZP, George EP, Gao YF. Deformation-induced

spatiotemporal fluctuation, evolution and localization of strain fields in a bulk

metallic glass. Int J Plast. 2015;71:136-45.

[304] Pugno NM, Ruoff RS. Nanoscale Weibull statistics. J Appl Phys.

2006;99:024301.

[305] Zhu YT, Blumenthal WR, Taylor ST, Lowe TC, Zhou B. Analysis of size

dependence of ceramic fiber and whisker strength. J Am Ceram Soc. 1997;80:1447-

52.

[306] Yuan T, Wang G, Feng Q, Liaw PK, Yokoyama Y, Inoue A. Modeling size

effects on fatigue life of a zirconium-based bulk metallic glass under bending. Acta

Mater. 2013;61:273-9.

[307] Wang G, Liaw PK, Yokoyama Y, Freels M, Inoue A. Investigations of the

factors that affected fatigue behavior of Zr-based bulk-metallic glasses. Adv Eng

Mater. 2008;10:1030-3.

[308] Nakai Y, Hosomi S. Fatigue crack initiation and small-crack propagation in Zr-

based bulk metallic glass. Mater Trans, JIM. 2007;48:1770-3.

[309] Chaussumier M, Mabru C, Shahzad M, Chieragatti R, Rezai-Aria F. A

predictive fatigue life model for anodized 7050 aluminium alloy. Int J Fatigue.

2013;48:205-13.

172
[310] Zhang ZJ, An XH, Zhang P, Yang MX, Yang G, Wu SD, et al. Effects of

dislocation slip mode on high-cycle fatigue behaviors of ultrafine-grained Cu–Zn

alloy processed by equal-channel angular pressing. Scripta Mater. 2013;68:389-92.

[311] He Z, Peng L, Fu P, Wang Y, Hu X, Ding W. High cycle fatigue improvement

by heat-treatment for semi-continuous casting Mg96.34Gd2.5Zn1Zr0.16 alloy. Mater

Sci Eng A. 2014;604:78-85.

[312] Peng L, Fu P, Li Z, Wang Y, Jiang H. High cycle fatigue properties of cast Mg–

xNd–0.2Zn–Zr alloys. J Mater Sci. 2014;49:7105-15.

[313] Li H, Liu Y, Pang S, Liaw PK, Zhang T. Corrosion fatigue behavior of a Mg-

based bulk metallic glass in a simulated physiological environment. Intermetallics.

2016;73:31-9.

[314] Liu W, Jiang L, Cao L, Mei J, Wu G, Zhang S, et al. Fatigue behavior and

plane-strain fracture toughness of sand-cast Mg-10Gd-3Y-0.5Zr magnesium alloy.

Mater Design. 2014;59:466-74.

[315] ASTM E1820-15. Standard test method for measurement of fracture

toughness. West Conshohocken, PA: ASTM International; 2015.

[316] Gludovatz B, Hohenwarter A, Catoor D, Chang EH, George EP, Ritchie RO. A

fracture-resistant high-entropy alloy for cryogenic applications. Science.

2014;345:1153.

173
Table Captions

Table 1. Fatigue limits and fatigue ratios based on the stress ranges of BMGs and their composites under bending loads.

Table 2. Fatigue limits and fatigue ratios based on the stress ranges of BMGs and their composites (different sizes) under

bending loads.

Table 3. Fatigue limits and fatigue ratios based on the stress ranges of BMGs under uniaxial loads.

Table 4. Fatigue limits and fatigue ratios based on the stress ranges of BMGs under rotating loads.

Table 5. Fatigue-crack-growth results for Vitreloy 1 at different temperatures [142].

Table 6. Fatigue-crack-growth properties of Zr-based BMGs.

Table 7. Summary of sample dimensions and pre-cracking conditions as well as fracture-toughness results, failure types and

required sample thickness, Bcrit, and ligament size, bcrit, for plane-strain and small-scale yielding conditions [181].

174
Table 1

Fracture Inner Outer Fatigue


Geometry Frequency R- Fatigue
Material strength span span limit
(mm) (Hz) ratio ratio*
(MPa) (mm) (mm) (MPa)
Zr41.2Cu12.5Ni10Ti13.8Be22.5 [99]a 1,900 2x2x60 0 48.4 20 0.1 768 0.404
Zr41.2Cu12.5Ni10Ti13.8Be22.5 [99]b 1,900 2x2x60 0 48.4 20 0.1 359 0.189
Zr50Cu37Al10Pd3 [101] 1,899 3.5x3.5x25 0 20 10 0.1 631 0.332
(Cu60Zr30Ti10)99Sn1 [103] ~ 1,800 2.85x2.85x25 0 20 10 0.1 475 0.264
Zr41.2Cu12.5Ni10Ti13.8Be22.5 [69] 1,900 3x3x50 10.2 20.3 25 0.1  152 0.080
Zr41.2Cu12.5Ni10Ti13.8Be22.5 [71] 1,900 3x3x40 10 30 5 0.1  190 0.100
Zr44Cu10Ni10Ti11Be25 [98]c 1,900 2.3x2.0x85 30 60 5 - 20 0.3 550 0.289
Zr44Cu10Ni10Ti11Be25 [98]d 1,900 2.3x2.0x85 30 60 5 - 20 0.3 390 0.205
Zr52.5Cu17.9Al10Ni14.6Ti5 [100] 1,700 3.5x3.5x30 5 20 10 0.1 850 0.500
Zr50Cu40Al10 [307] 1,821 3x3x26 10 20 10 0.1 542 0.298
Zr50Cu30Al10Ni10 [307] 1,900 3x3x26 10 20 10 0.1 635 0.334
Zr50Cu37Al10Pd3 [101] 1,899 3.5x3.5x25 10 20 10 0.1 630 0.332
Fe48Cr15Mo14Er2C15B6 [81] 4,000 2.85x2.85x25 10 20 10 0.1 682 0.162
(Cu60Zr30Ti10)99Sn1 [103] ~ 1,800 2.85x2.85x25 10 20 10 0.1 350 0.194
(Zr58Cu18Al10.4Ni13.6)99Nb1 with
1,700 2x2x25 10 20 10 0.1  559 0.329
nano particles [110]
Zr56.2Cu6.9Ni5.6Ti13.8Nb5.0Be12.5
1,480 3x3x30 10.3 20 25 0.1  296 `0.200
composites [109]
Zr39.6Cu6.4Ti33.9Nb7.6Be12.5
1,210 3x3x50 15 30 25 0.1  678 0.560
composites [113]
*: fatigue limit/fracture strength; a: samples from Howmet; b: samples from Liquid Metal; c: partially relaxed; d: stress relieved.

175
Table 2

Fracture Fatigue
Geometry Frequency R- Fatigue
Material strength Loading mode * limit
(mm) (Hz) ratio ratio **
(MPa) (MPa)
Zr56.2Cu6.9Ni5.6Ti13.8Nb5.0Be12.5
1,480 3x3x30 4PB 25 0.1  296 0.200
Composites [82]
Zr41.2Cu12.5Ni10Ti13.8Be22.5 [70] 1,900 3x3x50 4PB 25 0.1  152 0.080
Zr41.2Cu12.5Ni10Ti13.8Be22.5 [99] 1,900 2x2x60 3PB 10 0.1 768 0.404
Zr41.2Cu12.5Ni10Ti13.8Be22.5 [99] 1,900 2x2x60 3PB 10 0.1 359 0.189
Zr44Ti11Ni10Cu10Be25 [98] 1,900 2.3x2.0x85 4PB 5 - 20 0.3 550 0.289
Zr44Ti11Ni10Cu10Be25 [98] 1,900 2.3x2.0x85 4PB 5 - 20 0.3 390 0.205
Zr52.5Cu17.9Al10Ni14.6Ti5 [100] 1,700 3.5x3.5x30 4PB 10 0.1 850 0.500
(Zr58Ni13.6Cu18Al10.4)99Nb1
1,700 2x2x25 4PB 10 0.1 559 0.329
[110]
Zr55Cu30Ni5 Al10 [308] 1,560 2x20x50 Plate Bend 40 0.1 410 0.263
Fe48Cr15Mo14Er2C15B6 [81] 4,000 2.85x2.85x25 4PB 10 0.1 682 0.17
Al7050-T7451 [309] 505 20x60x200 Prismatic bend 10 0.1 200 0.4
Cu-5at.%Zn [310] 433 14x4x5 - 20 - 140 0.32
Mg96.34Gd2.5Zn1Zr0.16 [311] 367 5x4x10 Plate 20 - 130 0.35
Hour glass
Mg-3Nd-0.2Zn-0.45Zr [312] 284 - 100 - 98 0.35
Shaped
Mg66Zn30Ca3Sr1 [313] 785 3x6 10 0.1 370 0.47
Hour glass
Mg-10Gd-3Y-0.5Zr [314] 333 - 100 - 120 0.36
shaped
Al89Gd1Ni3 Fe1(5:1) [112] 803 5x10x30 3PB 15 0.1 469 0.58
Al89Gd1Ni3 Fe1(10:1) [112] 1019 5x10x30 3PB 15 0.1 490 0.48
Al89Gd1Ni3 Fe1(20:1) [112] 1064 5x10x30 3PB 15 0.1 500 0.47
Al0.5CoCrCuFeNi [147] 1344 3x3x25 4PB 10 0.1 472 0.35
* 4PB: 4-point bending, 3PB: 3-point bending; ** fatigue limit/fracture strength

176
Table 3

Fracture
Geometry Loading Frequency R- Fatigue limit Fatigue
Material strength
(mm) mode * (Hz) ratio (MPa) ratio
(MPa)
Zr56.2Cu6.9Ni5.6Ti13.8Nb5.0Be12.5
1,480 2.98 TT 10 0.1 239 0.161
composites [77]
Zr55Cu30Al10Ni5 Nano [125] 1,700 2x4x70 TT 10 0.1  340 0.200
Zr41.2Cu12.5Ni10Ti13.8Be22.5 [76] 1,850 2.98 TT 10 0.1 703 0.380
Zr41.2Cu12.5Ni10Ti13.8Be22.5 [76] 1,850 2.98 TT 10 0.1 615 0.332
Zr41.2Cu12.5Ni10Ti13.8Be22.5 [77] 1,850 2.98 TT 10 0.1 567 0.306
Zr41.2Cu12.5Ni10Ti13.8Be22.5 [118] 1,900 - CC 5 0.1 ~ 1,050 0.553
Zr41.2Cu12.5Ni10Ti13.8Be22.5 [118] 1,900 - TC 5 -1 ~ 150 0.079
Zr50Cu40Al10 [74] 1,821 2.98 TT 10 0.1 752 0.413
Zr50Cu30Al10Ni10 [74] 1,900 2.98 TT 10 0.1 865 0.455
Zr50Cu37Al10Pd3 [77] 1,899 2.98 TT 10 0.1 983 0.518
Zr50Cu37Al10Pd3 [119] 1,899 5.33 TT 10 0.1 ~ 900 0.474
Zr52.5Cu17.9Al10Ni14.6Ti5 [120] 1,660 6x3x1.5 TT 1 0.1 - -
Zr52.5Cu17.9Al10Ni14.6Ti5 [72] 1,700 2.98 TT 10 0.1 907 0.534
Zr59Cu20Al10Ni8Ti3 [120] 1,580 6x3x1.5 TT 1 0.1 - -
Zr65Cu15Al10Ni10 [122] 1,300 3x4x16 TT 20 0.1 ~ 280 0.215
Zr55Cu30Al10Ni5 [121] 1,560 1x2x5 TT 0.13 0.5 - -

* TT: tension-tension, TC: tension-compression, CC: compression-compression.

177
Table 4

Fracture
Geometry Frequency R- Fatigue limit Fatigue
Material strength
(mm) (Hz) ratio (MPa) ratio
(MPa)
Zr50Cu40Al10 [127] 1,821 4.0 50 -1 250 0.137
Zr50Cu30Al10Ni10 [127] 1,900 4.0 50 -1 500 0.263
Zr50Cu39Al10Pd1 [80] 1,909 4.0 50 -1 689 0.361
Zr50Cu38Al10Pd2 [80] 1,911 4.0 50 -1 820 0.429
Zr50Cu37Al10Pd3 [80] 1,899 4.0 50 -1 1,055 0.556
Zr50Cu35Al10Pd5 [80] 1,929 4.0 50 -1 650 0.337
Zr50Cu33Al10Pd7 [80] 1,952 4.0 50 -1 540 0.277

178
Table 5

C Kth Kth, eff


Test temperature (C) m
[(m/cycle) ( 10-10 MPam1/2)-m] (MPam1/2) (MPam1/2)
25 6.6 1.4 1.25 -
100 7.1 1.4 1.09 1.07
140 5.9 1.4 1.14 1.11
180 7.4 1.45 1.24 1.21
220 6.8 1.45 1.40 1.38

179
Table 6

C Kth
Alloys m
[(m/cycle) (MPam1/2)-m] (MPam1/2)
Zr41.2Ti13.8Cu12.5Ni10Be22.5 [72] 1.6  10-11 2.7 1-3
Zr41.2Ti13.8Cu12.5Ni10Be22.5 [82] - 1.7 1.39
Zr41.2Ti13.8Cu12.5Ni10Be22.5 [135] 2.3  10-10 1.6 1.4
Zr41.2Ti13.8Cu12.5Ni10Be22.5 [136] - 1.43-1.66 ~ 0.7
Zr44Ti11Ni10Cu10Be25 [98] 1.8  10-10 2.03 1.79
Zr55Cu30Al10Ni5 [83] - 1.4 1.8
Zr55Cu30Al10Ni5 (Nano) [139] - ~2 0.9
Zr56.2Ti13.8Nb5.0Cu6.9Ni5.6Be12.5
- 1.8 1.24
(composite) [109]

180
Table 7

KIc, KJIc, KQ Bcrit, bcrit Failure type


B (mm) W (mm) a/W b (mm) Pre-cracking
(MPa m1/2) (mm) (ASTM E399)
C(T) KIc
1.99 21.45 0.50 10.8 18.7 0.3 III Tension
2.03 21.42 0.50 10.77 28 0.68 III Tension
2.26 19.85 0.50 9.95 28.8 0.72 III Tension
2.34 19.83 0.50 9.91 25.7 0.57 III Tension
Mean: 25.3 MPa m1/2; standard deviation: 4.6 MPa m1/2

SE(B) KIc

2.01 4.23 0.55 1.91 36.5 1.15 III Tension


2.05 4.15 0.49 2.1 26.2 0.59 III Tension
2.3 3.94 0.54 1.8 35.3 1.08 III Tension
2.34 4.08 0.47 2.18 44.9 1.74 III Tension
Mean: 35.7 MPa m1/2; standard deviation: 7.7 MPa m1/2

SE(B) KJIc
2.04 4 0.58* 1.66 31.1*** 0.84 III Tension
2.33 3.95 0.45 2.18 46.5 1.87 III Tension
1.28 2.34 0.56* 1.03 35.2 1.07 III Tension
2 2.07 0.45 1.14 43 1.6 III Tension
Mean: 39 MPa m1/2; standard deviation: 7.0 MPa m1/2

SE(B) KQ sub-sized
1.99 4.22 0.51 2.07 63.5 3.49 III Tension
2.2 4.1 0.48 2.12 66.4 3.81 III Tension
1.33 2.25 0.63* 0.84 39.9 1.38 III Tension
1.1 2.13 0.7* 0.65 66.6 3.84 III Tension
1.92 1.93 0.44** 1.08 45.5 1.79 I Tension
1.87 4.03 0.52 1.93 72.6 4.56 I Tension
2.04 2.03 0.58* 0.86 37.2 1.2 I Tension
2.23 2.03 0.54 0.94 33.9 0.99 I Tension
2.07 3.97 0.48 2.06 81.6 5.76 I Tension
1.93 2.06 0.39** 1.26 66.6 3.84 I Compression
2.02 2.14 0.37** 1.35 66.3 3.8 I Compression
Mean: 58.2 MPa m1/2; standard deviation: 16.1 MPa m1/2

SE(B) KQ notched only


2.16 4.61 0.45 2.54 91.6 7.26 III -
2.15 4.08 0.45 2.23 95.9 7.96 III -
2.72 4.14 0.49 2.09 94.1 7.66 III -

181
Mean: 93.9 MPa m1/2; standard deviation: 2.2 MPa m1/2

* Samples with a/W  0.55 – 0.7.


** Samples with a/W < 0.45.
*** Not a valid KIc value due to a/W > 0.55.

182
Figure Caption

Figure 1. (a) Schematic illustration of an individual atomic jump in the free-volume

model, where an atom experiences a reduced activation energy barrier when

jumping to the nearest free-volume site [46]. (b) In the shear-transformation-

zone (STZ) model, dozens of atoms shear collectively under an applied shear

stress [50]. (c) Idealized illustration of the STZ process when atoms are

projected on a continuum mesh.

Figure 2. Tin coating on a BMG surface melting into spherical beads at shear bands,

providing an indirect evidence of local heating due to the inelastic

deformation in the shear band [93].

Figure 3. A shear band of a BMG specimen during tensile testing observed by

thermography with an IR-camera frame rate of 725 Hz [92]: (a) a typical

thermographic image of a shear band; (b) three-dimensional (3D)

temperature profile of the shear band on the third image of (a).

Figure 4. The mean stress (S) versus the number of cycles at failure (N) curves from

bending-fatigue tests for (a) metallic-glass ribbons, wires, and foams, and (b)

metallic-glass composites.

Figure 5. Mean stress versus number of cycles for Vitreloy 1 BMG based on the single

fatigue tests and fatigue-coaxing step tests. Solid arrows indicate that the

samples were not broken [102].

Figure 6. The mean stress (S) versus the number of cycles at failure (N) curves from

bending fatigue tests for (a) Zr-based and (b) other-types of BMGs.

183
Figure 7. S-N fatigue data for the Zr39.6Ti33.9Nb7.6Cu6.4Be12.5 metallic glass matrix

composite with dendritic second phases (DH3), along with some other

materials for comparison. (b) SEM image of the fatigue crack on the tensile

side showing a wide distribution of damage around the crack tip. (c) SEM

image presenting that shear bands are trapped between two dendrites. [113]

Figure 8. The S-N curves for uniaxial tension-tension fatigue tests of Zr-based BMGs

in air.

Figure 9. (a) The S-N curves for rotating fatigue of Zr-based BMGs with a frequency

of 50 Hz in air. (b) The fatigue limit-volume changes for the Zr-Cu-Al-Pd

system. [80, 127]

Figure 10. The S-N curves for 4-point-bending fatigue tests of the Zr-based BMGs with

a frequency of 10 Hz in the air and 0.6 M NaCl solution.

Figure 11. Fatigue-crack-growth rates (da/dN) vs. the stress-intensity-factor range (K)

for Zr-based BMGs and BMG composites (CT: compact-tension, SE: single-

edge notch, C: composite, CCT: center-cracked tension, NC: nano-crystalline

phase) [133].

Figure 12. The fatigue-crack-growth rate (da/dN) vs. stress-intensity-factor-range (K)

curves for Vitreloy 1 alloys in different environments [143].

Figure 13. The S-N curves of BMGs and some typical crystalline alloys in air [81, 83,

112, 147, 309-314]. (a) Stress range vs. cycles to failure. (b) Fatigue ratio vs.

ultimate tensile strength. (c) Endurance limit vs. ultimate tensile strength.

Figure 14. In-situ fatigue test of the metallic glass inside a TEM [156]. (a) Schematic of

the experimental setup. (b) SEM image showing a specimen prepared with

184
the FIB technique. (c) Bright-field TEM image of the notch tip before testing,

showing a fully-amorphous structure of specimen, as indicated by the

selected area diffraction pattern (Inset).

Figure 15. Microstructural evolution of the Al88 Fe7Gd5 metallic glass in response to

cyclic loading [156]. (a) HRTEM images of the notch-tip region for the

untested, cyclically-strained (the maximum strain of ~ 2.1%), and

monotonically-strained (the maximum strain > 10%) specimens. (b) Fatigue-

crack morphology after 980, 1,470, and 1,960 cycles, respectively. SADP

(Inset) revealed the existence of the crystalline phase ahead of the crack tip.

(c) HRTEM observation of two grains (G1 and G2) ahead of the crack tip

(Left) and their coalescence after 490 additional cycles (Center).

Figure 16. Maximum/minimum loads under a fixed value of the imposed displacement

amplitude and crack-length evolution as a function of fatigue-cycle numbers

[156]. (a) Load vs. cycle number for the first set of tests under 490 cycles. (b)

Load vs. cycle number for the third set from 981 to 1,470 cycles. (c) Crack-

length evolution with increasing cycle number.

Figure 17. (a) A fatigue crack propagating along shear bands. At point, P, there are three

possible growth directions: PA, PB, or PC and (b) Modeling predictions

based on LEFM, showing the agreement with the experimental results. [163]

Figure 18. Schematic illustration of the formation of (a) the plastic wake and plastic

zone; (b) the fine striation; and (c) the coarse striation when a fatigue crack

advances. [163]

185
Figure 19. (a) A two-dimensional linear-elastic fracture-mechanics-based modeling; (b)

The stress-intensity factor (SIF) results of short radial cracks emanating from

a circular hole. [167]

Figure 20. The applied maximum stress vs. cycles to failure curves for Zr50Cu40Al10 and

Zr50Cu30Al10Ni10 BMGs with two different size samples under four-point-

bend loading [178].

Figure 21. (a) Flexural deformation and (b) shear bands on the tensile surface of the

BMG sample with a small size after four-point-bending fatigue. (c) View

from the corner of a BMG sample with a small size after four-point-bending

fatigue [area A in (a)]. [178]

Figure 22. Lateral view of the BMG sample with a small size after four-point-bending

fatigue (a) near the compressive side [area B in Figure 21(a)] and (b) near the

tensile side [area C in Figure 21(a)]. [178]

Figure 23. A schematic diagram of the experimental design. The as-cast sample

undergoes compression-compression fatigue until it fails. Then, the damaged

part is cut off, and the leftover is used for the next fatigue test as long as the

rest part has a L/D ratio greater than 0.8 [179].

Figure 24. (a) The stress-range versus number of cycles to failure (S-N curve) data of

(Zr55Cu30Ni5Al10)98Er2 BMGs. The open circle is the fatigue life of the same

material with a fixed L/D ratio (L/D = 1.67), and a frequency of 10 Hz at

various stress ranges. The colored symbols are the data from this study. (b)

The fatigue life of samples, A, B, and C. The 1st-run (in the as-cast condition)

data is shown in red, and the 2nd run data is presented in blue. The number of

186
cycles-to-failure of the 2nd-run test starts from zero, and does not include the

cycles made in the 1st-run test. The fatigue life of the ‘‘pre-fatigued’’ sample

(leftover) is equal or longer, compared to the as-cast sample. The results

suggest that the fatigue damage in the specimen is mainly localized, and

cyclic loading has no global effect on the sample. When the fractured part is

removed, the rest of the material acts just like an as-cast material. (c) The

number of cycles to failure as a function of the L/D ratio of the specimen.

The results showed no correlation between the L/D ratio and the fatigue life.

[179]

Figure 25. Proposed fatigue-failure mechanisms of the large (6-mm in diameter)

(Zr55Cu30Ni5Al10)98Er2 (at.%) BMGs. (a) The as-cast sample contains weak

points (defect sites, such as microvoids or nanocrystalline particles formed

during the fabrication process). (b) The crack initiates at the weakest point in

the sample. The crack propagates slowly at the beginning, and leaves the

striation on the crack surface. (c) The crack continues to grow until the

sample cannot sustain the stress and then starts fast shearing of the sample

(fast fracture). The fatigue damage is found to be localized. When one crack

starts to grow, the rest of the sample still undergoes elastic deformation. The

microstructure away from the crack region generally remains unchanged. (d)

Therefore, after cutting off the damaged part, the remaining material

performs like an as-cast material with fewer defects. [179]

Figure 26. Tested specimen geometries, showing the nomenclature of their dimensions

and failure types by the ASTM standard E399 [182]. (a) Compact-tension

187
[C(T)] specimens as well as (c) Single-edge notch-bend [SE(B)] specimens

of the Zr52.5Cu17.9Ni14.6Al10Ti5 BMG, tested for fracture-toughness

measurements [181]. (b) All samples failed catastrophically after either major

plastic deformation – Type-I failure by the ASTM standard E399 – or minor

plastic deformation, thereby showing Type-III failure. Failure Type II – a

load-drop due to subcritical crack growth [181, 182].

Figure 27. Fracture toughness, KIc and KJIc, data of the Zr52.5Cu17.9Ni14.6Al10Ti5 BMG as

a function of the uncracked ligament width, b. KIc values measured in

accordance with the ASTM standard E399 [182] using both larger-sized C(T)

samples and SE(B) samples, whereas KJIc values were measured in

accordance with the ASTM standard E1820 [315] using only for SE(B)

samples [181].

Figure 28. Conditional fracture toughness, KQ, data for the Zr52.5Cu17.9Ni14.6Al10Ti5

BMG as a function of the uncracked ligament width, b. The metallic glass

showing a clear trend towards increasing KQ results with decreasing the

ligament size [181].

Figure 29. SEM images of fracture-toughness samples of the Zr52.5Cu17.9Ni14.6 Al10Ti5

BMG. (a) Larger samples that failed at lower fracture-toughness values

showing just minor shear banding from the pre-crack and a relatively-flat

fracture surface. (b) Samples with the large ductility and high KQ values

showing significant shear banding and blunting at the pre-crack tip, leaving a

large shear offset step on the fracture surface close to the pre-crack along

with a rough fracture surface. [181]

188
Figure 30. A review of the ductility and yield strength of different kinds of coatings,

including TFMGs, metallic coatings, and ceramic coatings applied to

improve mechanical properties of substrate materials [97]. For the references

from which the material data were extracted, please refer to Ref. [97].

Figure 31. S–N curves for all the reported TFMGs on different kinds of substrates,

showing the improvement of fatigue-endurance limit vs. UTS for all the

reported substrate materials, including the 316L steel, Ni-based alloy, Zr-

based alloy, Ti-based alloy, and Al-based alloy [97]. For the references from

which the fatigue data were extracted, please refer to Ref. [97].

Figure 32. (a) The schematic plot for the PEM control system and sputtering system

[191]. (b) S-N curves of the bare BMG and BMG substrate with Zr-based

TFMG [195, 198].

Figure 33. (a) SEM fractography micrograph of the BMG substrate with a 200 nm-thick

Zr-based TFMG after fatigue fractured at σmax = 450 MPa, (b) transition

regions between Stages II and III, and (c) partial film delamination due to the

severe deformation of the BMG substrate. [198]

Figure 34. Schematic description of the three-stage fatigue-crack initiation process in a

TFMG/BMG system, wherein the TFMG coating is effective in improving

the fatigue resistance of the BMG substrate at Stage I [195].

Figure 35. TEM EDS line-scans at the interface of the TFMG and BMG substrate after

fatigue fractured at σmax = 500 MPa. (a) STEM image; (b) Enlarged TEM

image of the shear offset area in (a); (c) EDS line-scan directions marked

189
with the red arrows; (d) and (e) EDS line-scan profile at the interface of the

film and substrate. [198, 199]

Figure 36. (a) The typical load-displacement curve obtained from the TFMG/Si

composite system in a sequential nanoindentation test, (b) the indentation

elastic moduli of the TFMGs (Ef) and their underlying Si substrates (Esi)

extrapolated using the modified King’s model, (c) the contact-depth-

insensitive hardness of the TFMGs (Hf) and the pure Si substrate (Hsi), and (d)

the comparison of the hardness-versus-modulus relation obtained from

BMGs and TFMGs via indentation testing. [191]

Figure 37. (a)-(e) The SEM micrographs showing the TFMG/Si composite micropillars

with different sizes and shapes. (Note that a sketched profile of the

micropillar is superimposed on the corresponding SEM micrograph for

clarity) [191].

Figure 38. (a)-(d) The typical load-displacement curves obtained from the

microcompression of TFMG/Si composite micropillars with the same film

thickness of 600 nm but different tf/D ratios, where tf and D are the film

thickness and pillar diameter, respectively. Insets showing the SEM

micrographs of the corresponding micropillars after microcompression [191].

Figure 39. (a) The variation of the yield strengths of the TFMGs obtained from

microcompression with their tf/D ratios (tf and D are the film thickness and

pillar diameter, respectively), and (b) the same trend of “strength-versus-

modulus” correlation exhibited by the TFMG and BMGs [191]. The data of

BMGs are taken from Ref. [210].

190
Figure 40. The contour plots for the distribution of the hydrostatic pressure

corresponding to an elastic strain of 2% in the TFMGs with the tf /D ratios of

(a) 1.0, (b) 0.5, (c) 0.25, and (d) 0.125, and (e) the average hydrostatic

pressure estimated in the TFMGs with the tf/D ratio less than 0.3 [191]. tf and

D are the film thickness and pillar diameter, respectively

Figure 41. Ashby map of the damage tolerance (toughness versus strength) of materials

[316].

Figure 42. Fracture-toughness measurements of the Pd79Ag3.5P6Si9.5Ge2 glass [214]. (a)

The crack-tip opening displacement (δt) is plotted against the crack extension

(Δa). (b) Fracture toughness (KJ) is plotted against the crack extension (Δa).

(c) – (k) SEM images taken during an in situ R-curve measurement of an

SE(B) specimen. The corresponding fracture toughness, KJ, values are 0 MP

m1/2 (c), 25 MP m1/2 (d), 63 MP m1/2 (e), 115 MP m1/2 (f), 144 MP m1/2 (g),

and 196 MP m1/2 (h).

Figure 43. R-curves showing resistance to fracture in terms of the stress intensity, KJ, as

a function of crack extension, Δa, for the DH1 and DH3 composite BMG,

monolithic Vitreloy 1 BMG, and a 17–7 PH stainless steel [229].

Figure 44. (a) Differential interference contrast micrograph of the DH3 composite alloy

showing extensive plasticity around the crack tip of the order of several

millimeters. (b) SEM backscattered electron image showing a wide

distribution of damage ahead of the crack tip in the DH1 alloy. The arrow

indicates the direction of crack propagation. (c) Deformation in DH3

occurring through the development of highly-organized patterns of regularly-

191
spaced shear bands distributed uniformly along the crack path. (d)

Microcracks in DH3 nucleated along the shear bands or at the

matrix/dendrite interface. [229]

Figure 45. SEM micrographs of the compressive fracture surface of the (a)

Cu48Zr48Al2 Ag2 and (b) Cu40Zr40Al10 Ag10 (at.%) BMGs. [234]

Figure 46. SEM micrographs of the bending fracture surfaces of (a) the as-cast

Zr52.5Cu17.9Ni14.6 Al10Ti5 (at.%) sample, and the same samples annealed at 300
o
C for (b) 9 hours, (c) 21 hours, and (d) 168 hours. [40]

Figure 47. (a) The correlation of the fracture energy, G, with the ratio of the shear

modulus to the bulk modulus, µ/B, for various as-cast and annealed metallic

glasses as well as oxide glasses. µ/B=0.41 - 0.43 is where the tough and

brittle regimes are divided. (b) The correlation of the fracture energy, G, with

the Poisson’s ratio, , for various as-cast and annealed metallic glasses as

well as oxide glasses. = 0.31 - 0.32 is where the tough and brittle regimes

are divided. [45]

Figure 48. A model describing the ductile vs. brittle fracture behavior of metallic glasses.

(a) Dense heterogeneities postpone crack propagation and lead to ductile

fracture. (b) Reduced heterogeneities lesson the hindrance to crack

propagation and result in diminished ductility. (c) Completely depleted

heterogeneity cause brittle fracture through catastrophic crack propagation.

[40]

Figure 49. (a) Cumulative probability of the maximum shear stress at the first pop-in for

the as-cast and annealed Zr52.5Cu17.9Ni14.6Al10Ti5 metallic glasses, measured

192
from statistical nanoindentation tests on (symbols) and predicted from the

unified structural model (solid lines). (b) The contour of the calculated

heterogeneity density for the Zr52.5Cu17.9Ni14.6 Al10Ti5 metallic glasses, as a

function of the annealing temperature and time. All indentation tests are

carried out, using a spherical indenter with a 3.80 μm radius. [40]

Figure 50. Variation of the total fracture energy with respect to the heterogeneity

density in Zr52.5Cu17.9Ni14.6Al10Ti5 (at.%) BMGs. R = 1.8 and 3.8 μm

represent the data from nanoindentation tests with two different radii of

spherical indenters. Dots points are measured data and the dashed line is the


fitting by Etotal  c1 exp c2 def  . [40]
Figure 51. Various techniques used for characterizing structural heterogeneities (a) TEM

imaging, in which the heterogeneity is displayed as color contrast in a

Zr44Ti11Cu9.8Ni10.2Be25 (at.%) BMG sample [249]. (b) Amplitude-modulation

dynamic atomic-force microscopy (AM-AFM), in which the inhomogeneous

intensity distribution on the phase-shift image of Zr53Cu36Al11 (at.%) BMG is

a signal of structural heterogeneities [250]. (c) Molecular dynamic simulation,

showing that the structural heterogeneities are approximately in elliptical

shapes in Cu50Zr50 (at.%) model MGs [251]. (d) Synchrotron radiation x-ray

scattering, in which the C-region in Al88Ce8Co4 (at.%) BMGs has more

loosely-packed atomic structures and is corresponding to the structural

heterogeneity region [255]. (e) β-relaxation, detected from the internal friction

in Al88Ce8Co4 BMGs by the dynamical-mechanical analysis [255]. (f) A

193
micromechanical model that derive the volume fraction of heterogeneities for

Vit105 metallic glasses from their measured shear modulus [256].

Figure 52. Experimental data in the literature summarizing varied notch effects found in

different metallic glasses [265], including notch strengthening [264, 265],

notch weakening [269], and notch insensitivity [262, 263, 265]. The

normalized fracture stress is the ratio of the fracture strength of the notched

samples to the un-notched ones. The notch dimension, d, represents the width

of the notched section and h the notch height.

Figure 53. (a) Fracture morphology of the notched Zr 64.13Cu15.75Ni10.12Al10 (at.%) BMG

sample with the notch dimension of d = h = 2.7 mm, showing the typical

shear fracture along a 50 o major shear band with respect to the loading axis.

(b) Failure pattern of the Cu64Zr36 (at.%) MG with the same notch dimension

ratio, d/h, as in (a), predicted from the MD simulation, in which the plastic

deformation is localized into an inclined shear band. (c) Fracture morphology

of the notched Zr64.13Cu15.75Ni10.12 Al10 (at.%) BMG sample with the notch

dimension of d = 2.7 mm and h = 0.45 mm, showing the model-I fracture at


o
an angle of 90 with respect to the loading axis. (d) Failure pattern of the

Cu64Zr36 MG with the same notch dimension ratio, d/h, as in (c), predicted

from the MD simulation, in which a void forms at the center of the sample as

it fails in Mode-I. [265]

Figure 54. The fracture stress, , normalized by the normal cohesive strength, , and

the geometric factor, , where L represents the width of the un-

notched section of a double-edge notched sample, as a function of the ratio of

194
the notch length, , to the cohesive zone size, , where and

represent the Young’s modulus and critical normal cohesive separation at ,

respectively. The data are produced from numerical modeling, which

accounts for the synergistic effect of the normal fracture and shear-banding-

induced fracture in metallic glasses simulated by a cohesive interface model

and a weak-zone model, respectively. Curves with different ratios of the

cohesive shear strength, , to the normal strength, , and the critical shear

separation, , to normal separation, , are displayed [85].

Figure 55. The notch strength ratio (NSR), which is calculated as the ratio of the

strength of the notched samples to that of un-notched ones, as a function of

the stress-concentration factor for a variety of materials, including metallic

glasses [262].

Figure 56. (a) Nominal compressive stress-strain curves for Zr52.5Ni14.6Al10Cu17.9Ti5

(at.%) BMG samples without notches and with double-edge symmetric

notches, showing the significant ductility improvement in the notched sample.

(b) Failure mode of the un-notched sample, in which the fracture is along a
o
41 major shear band with respect to the loading axis. (c) Failure mode of

the notched sample, in which multiple shear bands initiate from the roots of

the two notches, propagate radially, and interact with each other [271].

Figure 57. Effect of the notch depth, a, on the shear-band arrangement in metallic

glasses with two symmetric edge notches. (a)-(c) Prediction from the

Rudnicki-Rice instability theory based on the elastic-stress field. (d)-(e)

Predictions from the finite-element simulation using the free-volume

195
constitutive law. The notch root radius, r, and notch width, w, are held

constant for all cases [85].

Figure 58. Uniaxial fatigue data for (a) Zr41.2Ti13.8Cu12.5Ni10Be22.5 (at.%) BMGs and (b)

Zr-based BMGs of different compositions, without a notch and with different

types of notches. The data is present in the form of the stress amplitude, ,

as a function of the number of cycles to failure, [279]. For the references

from which the fatigue data were extracted, please refer to Ref. [279].

Figure 59. SEM observations of the fatigue-crack-propagation paths for (a) the un-

notched and (b) notched Ti32.8Zr30.2Ni5.3Cu9Be22.7 (at.%) BMGs loaded

cyclically. A straight crack path without the formation of shear bands

features the un-notched sample, while a staircase-like path forms in the

notched sample with the formation of multiple shear bands [279].

Figure 60. (a) To clearly visualize the three-dimensional conical shear band, free-

volume contours were plotted. Finite-element simulations were performed in

a quarter of the substrate under the spherical indentation with symmetry-

boundary conditions prescribed on the two side surfaces. Shear-band patterns

on the cross-sectional interface of the (b) bonded-interface metallic-glass

substrate and (c) bonded-interface metallic-glass film on a steel substrate

under the spherical indentation. [257]

Figure 61. Comparison of the structure that has been held at a load to the structure that

has been cycled at the same load for the same amount of time. The cyclic

loading has increased the level of the strain [144, 297].

196
Figure 62. The evolution of free volumes and excess strains under cyclic loading. The

free volume is given in units of Å3 [144].

Figure 63. (a) Schematic illustration of the geometry used in MD-simulations under

cyclic-loading conditions of Zr-Cu metallic glass (the arrows indicate the

direction of applied loading). 2-D sliced plots extracting from the 3D-

simulated results for the local-strain distribution of the stress-control fatigue

at different cycles: (b) cycle ten, (c) cycle fifty, and (d) cycle one hundred.

The color scheme represents the degree of the atomic-bond rotation, or shear

strain. The numbers of the horizontal and vertical axes are in the unit of

angstrom. [293]

Figure 64. The accumulation numbers of deformation atoms as a function of loading

cycles corresponding to the results of a stress-control mode. The change of

growth rate of plastic flows implying that a resistance against the sustained

development of STZ groups occurs when accumulation deformations reach a

specific amount during cyclic loading [293].

Figure 65. Simulated nanoindentation results for monotonic loading [297]. The graph

showing the load–displacement curve for a single monotonic indentation test,

with results for a purely-elastic contact for comparison. Snapshots of the

system during the simulation provided below the graph, as marked by ‘A’,

‘B’, and ‘C’. The red contour on the snapshots showing the region of

material that has exceeded the yield stress, while the gray regions denote the

operation of STZs.

197
Figure 66. Load–depth curves for the simulated cyclic nanoindentation where the

limiting amplitude of the cycling depth is marked by each curve, ranging

from 1.2 to 2.8 nm [297]. The elastic reference is plotted with each cycling

simulation for comparison. Cycling at depths where the load does not reach

the minimum load for STZ activation, 1.2 and 1.6 nm, results in a perfectly-

elastic material response. Cycling above the minimum load for the STZ

activation leads to plasticity through the STZ activity in all cases, 2.0–2.8 nm,

although the hysteresis in the load–displacement curve is not immediately

apparent in all cases. Snapshots of a portion of the 2.0 and 2.4 nm cycled

systems after each of the five cycles, illustrating the progressive nature of the

structural change.

Figure 67. Two types of fracture modes in brittle Fe80P20 and ductile Cu50Zr50 metallic

glasses [236]. The upper row presenting the brittle failure by nanoscale voids

while the lower one showing the crack-tip blunting by the shear band.

Figure 68. (a) Crack-tip ductile versus brittle behavior in metallic glasses, as a function

of the applied rate of the stress-intensity factor and interface strength. (b)

Representative case of crack-tip blunting with  max  0  10 and

Kt0 K0  1.1105 . (c) Representative case of a cohesive crack with

 max  0  6 and Kt0 K0  1.23 106 .

Figure 69. Strain heterogeneities, as revealed by the digital-image correlation (DIC)

technique [303]. (a) Surface of a Zr-based BMG. (b) The stress versus

loading time curve showing multiple stress drops, each of which corresponds

to a strain localization in (c). (c) Strain fields displaying localizations into

198
narrow bands. Eventually major shear bands dominating the deformation

response.

Figure 70. Fatigue-life cycles versus maximum applied stress data of Zr 50Cu30Al10Ni10

BMG samples tested under four-point-bending fatigue experiments [306].

Figure 71. Predictive results of Models I, II, and III in (a), (b), and (c), respectively.

(solid line: median fatigue life when h = 3 mm; dotted line: 95% predictive

interval for fatigue life when h = 3 mm; dashed line: median fatigue life

when h = 2 mm; dashed–dotted line: 95% predictive interval for fatigue life

when h = 2 mm). (d) Comparison of the median fatigue lives predicted by the

three models (solid line: Model I; dashed line: Model II; dashed–dotted line:

Model III). [306]

199
FIGURE 1

200
FIGURE 2

201
FIGURE 3

202
FIGURE 4

203
FIGURE 5

204
FIGURE 6

205
FIGURE 7

206
FIGURE 8

207
FIGURE 9

208
FIGURE 10

209
FIGURE 11

210
FIGURE 12

211
FIGURE 13

212
FIGURE 14

213
FIGURE 15

214
FIGURE 16

215
FIGURE 17

216
FIGURE 18

217
FIGURE 19

218
FIGURE 20

219
FIGURE 21

220
FIGURE 22

221
FIGURE 23

222
FIGURE 24

223
FIGURE 25

224
FIGURE 26

225
FIGURE 27

226
FIGURE 28

227
FIGURE 29

228
FIGURE 30

229
FIGURE 31

230
FIGURE 32

231
FIGURE 33

232
FIGURE 34

233
FIGURE 35

234
FIGURE 36

235
FIGURE 37

236
FIGURE 38

237
FIGURE 39

238
FIGURE 40

239
FIGURE 41

240
FIGURE 42

241
FIGURE 43

242
FIGURE 44

243
FIGURE 45

244
FIGURE 46

245
FIGURE 47

246
FIGURE 48

247
FIGURE 49

248
FIGURE 50

249
FIGURE 51

250
FIGURE 52

251
FIGURE 53

252
FIGURE 54

253
FIGURE 55

254
FIGURE 56

255
FIGURE 57

256
FIGURE 58

257
FIGURE 59

258
FIGURE 60

259
FIGURE 61

260
FIGURE 62

261
FIGURE 63

262
FIGURE 64

263
FIGURE 65

264
FIGURE 66

265
FIGURE 67

266
FIGURE 68

267
FIGURE 69

268
FIGURE 70

269
FIGURE 71

270

Вам также может понравиться