† Corresponding author. E-mail:
Amorphous materials are ubiquitous and widely used in human society, yet their structures are far from being fully understood. Metallic glasses, a new class of amorphous materials, have attracted a great deal of interests due to their exceptional properties. In recent years, our understanding of metallic glasses increases dramatically, thanks to the development of advanced instrumentation, such as in situ x-ray and neutron scattering. In this article, we provide a brief review of recent progress in study of the structure of metallic glasses. In particular, we will emphasize, from the scattering perspective, the multiscale structures of metallic glasses, i.e., short-to-medium range atomic packing, and phase transitions in the supercooled liquid region, e.g., crystallization and liquid-to-liquid phase transition. We will also discuss, based on the understanding of their structures and phase stability, the mechanical and magnetic properties of metallic glasses.
Metallic glasses (MGs) constitute a new class of metallic materials with unique properties, including high strength, high elastic limit, high wear and corrosion resistance, and excellent soft magnetic behavior, for structural and functional applications.[1,2] MGs can be formed when a liquid is frozen very quickly. The liquid–glass (LG) transition is thus featured with drastic slowing down of relaxation time (increasing of viscosity) upon cooling. L–G transition is considered as the consequence of the cooperative development of short-range ordering (SRO) and medium-range ordering (MRO),[3] which have incompatible local symmetry compared to the long-range density ordering toward crystallization. The L–G transition is thus a process accompanied not only by structure change but also by dynamic evolution. The physical origin of L–G transition has been one of the most challenging, unresolved issues in condense matter physics.[4–6] Studying the atomic packing schemes of liquid and glassy states for MGs (Section
As cutting-edge techniques, scattering techniques, including neutron and high-energy synchrotron x-ray scattering, are very convenient for in situ study of the structure and dynamics of both liquids and glass under different experimental conditions, such as pressure- or temperature-dependent experiments. Non-destructive, high-penetration scattering experiments equipped with large area 2D detectors can probe structure information in reciprocal space over a wide momentum transfer Q (
Atomic packing has been a long-standing issue for amorphous solids, liquids, and soft matters. Unlike crystalline materials, disordered materials do not have periodic unit cells that can perfectly fill in three-dimensional space and pack to achieve long-range ordering. The most efficient atomic packing configurations in crystalline materials are the face-centered cubic (FCC) and the hexagonal closest packed (HCP). Metallic glass (MG) can be formed by quenching molten liquid from a high temperature to room temperature. Although free volumes can be cast from loose packing liquid into dense packing glass, the packing coefficient for MGs is still close to that of their crystalline counterparts.[9] Bernal’s dense random hard-sphere packing model[10,11] was proposed about a half-century ago and has been widely cited for monoatomic liquid or glass. However, it fails to describe the atomic arrangement of MGs that have more than two components; these have been found to have a pronounced topological/chemical short-range order (SRO).[12–14] The global packing scheme for MGs remains mysterious.
The studies of atomic packing from the short-to-medium range can aid in understanding GFA and the properties of MGs. As shown in Fig.
Recently, fractal packing schemes at medium-range scale have been experimentally observed in 37 MGs by Ma et al.,[21] which show a higher packing efficient than that of cubic packing and icosahedral order packing as demonstrated by simulations. Compared to cubic packing, fractal means self-similarity and scale invariance, and this has been clearly demonstrated in a variety of systems including polymers, colloids, and nanoparticles in suspension, etc.[22] The primary evidence for fractal packing, as shown in Fig.
An early MD simulation study by Barmin et al.[24] analyzed the atomic structure of pure amorphous Re and Re–Tb amorphous alloys in the framework of percolation theory, and demonstrated that the dense areas in the amorphous matrix form the fractal skeleton. Later, Chen et al.[25] proposed a dimensionality crossover between fractal short-range (
Today, the capabilities of electron microscopes (EMs) are being used to directly observe the structure of MGs. Recently, Hirata et al.[13] demonstrated that a single distorted icosahedron with partial FCC cubic symmetry could be directly observed for the Zr80Pt20 MG by using angstrom-beam electron diffraction in a double Cs-corrected EM, and this single distorted icosahedron was later proposed to be the frustration element contributing to glass-formation. Researchers have used 3D EM to observe the packing mechanism of advanced materials.[26–29] Now, combining and utilizing both 3D imaging and angstrom-beam diffraction, the fractal packing for MGs can be directly observed. These new observational abilities will help to unravel the mystery of atomic packing in MGs.
The study of MG structure stability can aid in exploring the development of new alloys with excellent GFA and the creation of the structure-property correlation for MGs. The studies of structure stability include two aspects: relaxation and crystallization. When heating an MG to the supercooled liquid region between glass transition
Amorphous phase separation in metallic glasses thus has been a controversial issue in the past several decades because of the large negative heat of mixing among its main constituent elements. Thermodynamics predicts that amorphous phase separation will never occur in an alloy system of the negative heat of mixing but also predicts that it is possible in an alloy system of the positive heat of mixing.[39,40] As shown in Tables
To find trustworthy direct evidence is the key issue for resolving this long-standing issue. In 2003, Wang et al.[32] investigated the amorphous-to-crystalline transformation in BAM 11, i.e., Zr
The existence of amorphous phase separation is a long-standing issue for Pd–Ni–P alloys. In 1976, Chen et al. proposed the possibility of amorphous phase separation in Pd-based alloys.[60] Later, researchers countered this possibility, finding that efforts were made in the samples using cooling paths as shown by path (i) and (ii) in Fig.
Another interesting aspect is how the decomposition in an amorphous state affects subsequent crystallization. The occurrence of nano-scale amorphous phase separation may enhance the rearrangement of chemical species prior to crystallization in BAM-11 and help the system to achieve a new metastable equilibrium. Initially, the chemical composition of the alloy is homogeneous due to the fast quenching rate. During annealing, like-clusters (solute-centered SRO) are prone to stay together and to connect to be a network, perhaps by following these possible proposed mechanisms: a single atom jump[77] of neighboring mobile atoms between clusters, e.g., solvents, the fast diffusion of small interstitial atoms,[78] and the creation/annihilation of “defects” like free volume.[79–81] These mobile atoms play the role of “connecting atoms.” “Connecting atoms”[82] are easier to be obtained for patching the “imperfect” or “weaker” region (e.g., “holes” and “cavities”). Having begun patching a region, then larger scale clusters (∼nm scale) can be formed, thus further enhancing the connectivity of the network.[83] Finally, different networks characterized by different building blocks, i.e., solute-centered SROs, would intertwine with each other at the nanometer scale and then fill in the whole space. One of the networks is mainly formed by Cu and Ni centered clusters, and another network is rich in Al and Ti centered clusters. Based on the framework of amorphous phase separation, we propose two stages for phase transformations at the early stage of crystallization of bulk MGs, such as BAM-11, as follows: the formation of a metastable phase; a stabilized crystalline phase with complex structure.
At the first stage, a metastable phase would be incubated, which was proposed as a large cube structure[84] packed by solute-centered clusters such as icosahedral SRO. As pointed by Sheng et al.,[20] the ‘full’ icosahedra with complete five-fold symmetry formed to frustrate crystallization ordering due to the size effect of constituent elements. In the beginning, there should be a lot of full icosahedra for the multi-component alloy with a variation of atomic size. The final state of BAM 11 was identified as t-Zr2Ni phase.[31] There are large discrepancies between the initial state and the final state of alloy either in chemical and topological. The amorphous phase separation may pave the way for the transformation of metastable phases by reducing the chemical disordering. To some degree, most of the full icosahedra would be destroyed to be distorted forms because of the reducing size effect by phase separation during annealing. It is possible to reconstruct the distorted icosahedra with incomplete five-fold symmetry to transform it into a more stable phase with higher crystalline ordering.[85,86] So the formation of metastable phase such as a big cube phase with distorted icosahedra may lower the nucleation barrier and further favor the crystallization process.
Nanoscale solute partition happens at the second stage due to the difference of the heat of mixing (as shown in Table
Because the structure of supercooled liquid is completely different from that of crystals, it is difficult to form a complex crystalline phase directly, especially for those MGs that have good GFA. Recently, the MD simulation by Tang et al.[89] of interfacial structure for two alloys, Ni–Al and Zr–Cu, with difference GFA showed that the crystalline ordering in the Zr–Cu with better GFA just extended to limited atomic layers. For Zr–Cu alloy, there is no pre-crystalline ordering in its liquid compared to that of the Ni–Al alloy. It is reasonable that a metastable phase may be formed for Zr–Cu prior to the final crystalline phase to lower crystal/liquid interfacial energy. Recently, Lan et al.[90] observed a ‘suspicious’ metastable phase for Zr–Cu–Al MG. As shown in Fig.
LLPT is defined as the transition between two liquid states with different density and entropy. LLPT has been proposed, in principle, to be a general phenomenon in all kinds of liquid.[92,93] The existence of LLPT was reported in a variety of liquids from atomic to molecular liquids, including P,[94] C,[95] Si,[96,97] Ge,[98,99] SiO2,[100] GeO2,[101] H2O,[93,102–104] Al2O3-Y2O3,[105] triphenyl phosphite (TPP),[106,107] etc. LLPT is usually accompanied by the poly-amorphous transition, i.e., liquid polymorphism. Experiments and simulations together showed that there are structure changes[94–96,98,100–102,104–106,108–110] during the occurrence of LLPT, sometimes accompanied by the changes of physical properties such as density and specific heat. Katayama et al.[94] directly observed the coexistence of two forms of liquids, i.e., molecular and polymeric phosphorus liquids, with totally different structures using in situ x-ray diffraction. Although numerical supporting evidence was reported, there were still many reports contradicting such evidence, especially in Si,[111] H2O,[104,112] Al2O3-Y2O3,[113] TPP,[114] etc. LLPT and the liquid polymorphism are still controversial issues because consistently direct observations are few and the nature of the occurrence of LLPT is unclear.
Many theories, including frustration,[18,19] a two-order-parameter model[115–119] and a random first-order transition (RFOT),[120] etc., have been suggested to explain the structure and dynamic evolution of liquid when LLPT occurs. According to frustration theory,[18,19] there exist dynamic, locally preferred structure (LPS), such as icosahedra, which differ from that of crystalline phases. LPS cannot perfectly tile to the whole space and thus give rise to an abstract reference system in which the effect of frustration has been turned off. Therefore, there exists an avoided critical point for the occurrence of a weakly first-order phase transition in a liquid state. One of the structural origins of LLPT proposed by a two-order parameters theory via Tanaka et al.[92] is the cooperative development of short- to medium-range ordering (SRO and MRO) and the frustration/competition of these so-called locally/energetically favored structures with density ordering. The two-order-parameter model has been used successfully to describe water’s anomalies.[121] The lack of theory connecting structure and dynamic makes the explanation of the LLPT difficult. RFOT[120] was developed for predicting relationships connecting the α relaxation time
The poly-amorphous phase transition in MGs was first reported by Sheng et al.[123] in a Ce55Al45 alloy induced by high pressure using in situ x-ray diffraction. Abrupt shifts of the first sharp diffraction peak position Q 1 as well as a crossover in calculated specific volume have been observed as a function of hydrostatic pressure. The structure and density changes together indicated the occurrence of polyamorphous phase transition. The possible origin revealed by ab initial MD simulation is f-electron delocalization of Ce. Later, the “polyamorphism” in a Ce75Al25 MG has also been observed by Zeng et al.[124] The direct experimental evidence of 4f-electronic delocalization was elaborated. If the f-electron delocalization gives rise to the compressibility for MGs under pressure, is the LLPT or polyamorphous transition possible in MGs without f-electron delocalization property? What is the possible origin for an LLPT induced by temperature instead of induced by pressure?
F–S transition has been reported in numerical references based on viscosity studies,[108,110,125–130] which implies the occurrence of a dynamic transition in MGs. Until recently, Li et al.[131] demonstrated that LLPT is possible in MGs during heating/cooling based on volumetric measurements results using electro-levitation for three Zr-based alloys (Zr
Wei et al.[132] reported an LLPT in the Vit 1 MGFL, and correlated the structure changes with a specific heat
Wu et al.[134] identified evidence of an LLPT accompanied by larger changes in atomic diffusion behavior using high-temperature NMR in the La50Ni35Al15 MGFL above its
We studied the structure evolution and specific volume changes for Vit 106 MGFL using electrostatic levitator.[135] Combining with the MD simulation for the Zr57Cu28Al15 alloy, approximating the Vit 106 alloys as a pseudo-ternary alloy Zr57(Cu
The nature of LLPT has also been proposed to be the enhancement of connectivity of solute-centered clusters in medium-range scale. As shown in Fig.
The integrations of intensity for coordination shells illustrate that there is an abrupt increase of intensity for the shoulder of the second neighbor shell as shown in Fig.
There are excellent reviews on the mechanical properties of MGs in recent years. A comprehensive discussion of elastic property is provided in Ref. [137]. References [138] and [139] emphasized the deformation and fracture mechanisms. The authors also constructed a deformation map.[138] Reference [140] summarized the progress in understanding shear bands. In this part, we will offer a brief summary of the overall mechanical behavior of MGs. For more detailed information, the readers can go to the review articles shown above.
One of the most impressive mechanical properties of MGs is their high yield strength. Figure
The bulk modulus of MGs is usually several percents smaller than the corresponding crystalline alloys.[144] This can be understood by the fact that MGs usually have a lower density, about 0.5%–2%, than their crystalline counterparts.[145] In other words, the average interatomic distance in amorphous structure is slightly larger than in the crystalline structure, resulting in a slighter change of atomic potential when atoms are moving towards or apart from each other under external hydrostatic stress. The shear modulus of MGs is, however, about 30% smaller than the corresponding crystalline alloys.[138] The significant decreasing of shear modulus in MGs indicates a different mechanism of atomic movement under shear stress. Ma et al.[146] showed that a variety of BMGs inherit their Young’s modulus and shear modulus from the solvent components. They attributed this to preferential straining of locally solvent-rich configurations among tightly bonded atomic clusters by x-ray diffraction. The x-ray diffraction results of MGs during elastic deformation are presented in Fig.
Depending on sample condition, temperature and strain rate, the plastic deformation of metallic glasses can be roughly divided into inhomogeneous flow and homogeneous flow.
Inhomogeneous flow usually occurs when the temperature is far below glass transition temperature and the strain rate is relatively fast. The basic unit of plastic deformation is believed to be a single atom hopping according to the free volume model[147,148] or a group of atoms shearing collectively according to the shear transformation zone (STZ) model,[149] driven by external stresses. In both models, shearing will induce dilatation which is characterized by increasing free volume. When the increasing of free volume is faster than annihilation, the free volume in local regions will accumulate and result in a lower viscosity. These locally softened regions then flow more easily compared to the surrounding matrix. Upon yielding, a thin sheared region called shear band is then formed from the locally softened regions with a higher content of free volume to carry the plastic strain. This is unlike the plastic deformation in crystalline alloys, in which plastic strain is compensated by dislocation sliding, grain boundary motion, Martensite–Austenite transformation, etc. Because the shear band is soft and localized, MGs are usually brittle at room temperature. However, there are some systems showing large plasticity under uni-axial compression and high fracture toughness.[150,151] Figure
It was found that BMGs with a large poisson ratio will exhibit large compression plasticity.[182] The plasticity was also found to be influenced by fabrication processes. A higher cooling rate may result in a greater plasticity.[154] Phase separation is another factor which was proposed to enhance the plasticity because of the existence of the finely distributed amorphous network structure. However, recent study on phase separated Pd-based BMGs shows that the amorphous network with two brittle MGs cannot stop the propagation of shear bands. As indicated in Fig.
As the plastic strain is carried by shear bands, understanding their structure is the key to understanding the plasticity of MGs. The formation of shear bands can be well described by a two-stage model.[184] At the first stage, a viable band is created for shearing by structural rejuvenation. The structure in a shear band at the first stage becomes disordered and ready to flow compared with the surrounding matrix. The strain in the first stage is still very small. Right after the formation of first stage, the second stage happens by synchronized sliding and shear-off along the rejuvenated plane. The softening of shear bands may come from two reasons: first, shearing will induce dilatation which lowers the viscosity as discussed above; second, during the operation of shear band, the local temperature will rise above the glass transition temperature to reduce the viscosity. Lewandowski et al.[185] demonstrated that the temperature rises after the formation of some shear bands. As shown in Fig.
The problem that prevents the wider application of MGs as structural materials is that they always lack strain hardening. Another problem is that to date no MGs have tensile plasticity except for samples in nanometer sizes.[187] Tensile plasticity for bulk samples is only achieved in some partially crystallized MGs.[188,189]
When temperature is close to glass transition temperature and the strain rate is relatively slow, the stress induced free volume will be annihilated quickly by thermal diffusion. In such cases, the plastic deformation is dominated by homogeneous flow. Kawamura reported that when the La–Al–Ni MG was deformed at a temperature above the glass transition temperature, it elongated to 1800% of its original length, showing a superplastic deformation behavior.[190] Lu et al.[191] studied the deformation behavior of Zr-based MGs as a function of temperature. Figure
Considering the influence of stress, strain rate, and temperature, a deformation map was first constructed by Spaepen.[148] Later, Schuh et al.[138] created a redrawn deformation map which is shown in Figs.
Soft-magnetic MGs have attracted a great deal of research and industry interest since the first discovery of Fe–P–C MG in the 1970s.[192] Because of the low coercivity, low magnetostriction, and high electrical resistance, soft-magnetic MGs show an advantage in energy saving when used as the cores in electric power distribution transformers and electrical motors. Figure
As indicated by definition, the magnetic moment correlation in soft-magnetic MGs is dominated by ferromagnetic interaction.[196] Because of the random atomic structure, the magnetic moments are not perfectly aligned parallel with each other but rather have some degrees of canting angle, as revealed by neutron diffraction experiments.[197,198] The magnetic structures in the medium range (from several angstroms to tens of nanometers), e.g., magnetic domains and domain walls, were studied by the Small Angle Neutron Scattering (SANS).[199–205] The magnetic domains of soft-magnetic MGs were found to be influenced by fabrication processes, surface defects, and chemical inhomogeneity.
The structure of MGs can be considered random at long range, but with short-to-medium range orders as described in the above sessions. Because of the absence of crystalline lattices, MGs do not have magneto-crystalline anisotropy. Instead, the magneto anisotropy mainly comes from the nearest neighboring atomic arrangement. The short-to-medium atomic arrangements in MGs vary with location, resulting in local anisotropies with random orientations.
The random anisotropy model (RAM)[206] is applied to explain the magnetic property in the amorphous alloys. Figure
(1) |
According to RAM, the average anisotropy is described by
(2) |
Recent years have seen the development of soft-magnetic MGs. Research in this field is targeting soft-magnetic MGs with good GFA, high saturation magnetization, low coercivity, low magnetostriction and good ductility. BMGs with a critical casting thickness larger than 1 mm and good soft-magnetic properties were reported in alloy systems based on 3d-based late transition metals (Fe, Co, Ni).[210,212–233] For example, the Fe–Mo–P–C–B–Si BMG[210] has a critical casting thickness of 4 mm with a good soft-magnetic property (
To achieve the best soft-magnetic property, MGs need to be annealed below
Nano-crystalline soft-magnetic materials are usually made by crystallization of the amorphous precursors.[238] The grain size is on the order of 10 nm. It was first developed in Fe–Si–B–Nb–Cu alloy system called FINEMET by Yoshizawa et al.[239] Soon after that, Suzuki et al. developed Fe–Zr–B nano-crystalline alloy which has zero magnetostriction.[240] As described by the RAM, the average anisotropy scales with D
6. Thus, the resultant nano-crystalline materials with small grain size show superior soft-magnetic property.[241] In recent years, Makino et al. successfully made a new type of nano-crystalline alloy by crystallization of the amorphous Fe–Si–B–P–Cu alloy.[242,243] This new nano-crystalline alloy (called NANOMET) shows a high saturation magnetization (∼1.9 T), which is comparable with silicon steels and much higher than the FINEMET alloy (∼1.3 T). Figure
The mechanism of nano-crystallization in the above systems was studied extensively. For example, it was found the FINEMET alloy shows a finest grain structure with an optimized amount of Cu by the SANS study.[244,245] Using polarized neutron scattering,[246,247] Heinemann et al.[248] studied the crystallization of the FINEMET alloy. Figure
A brief survey is provided on the multiscale structures of metallic glasses. Neutron and synchrotron scattering played an important role in elucidating the complex structures, by breaking down the structures to the fundamental unit level (the short-range order) and the packing or connectivity between the structure units (medium-range order). The response of the structures, the changes at short- and medium-range order levels, determine the properties. An example is shown for mechanical deformation. The magnetic structures are discussed in a similar context.
The rapid development of the next-generation synchrotron and neutron sources and the associated cutting-edge sample environments could provide greater opportunities for study of multiscale structures of metallic glasses. New experiments will be enabled, particularly in the study of dynamic behaviors. For example, using x-ray free electron laser (XFEL), it is possible to capture motion of individual atoms during a phase transformation or a chemical reaction process in a few femtoseconds (1 fs=
[1] | |
[2] | |
[3] | |
[4] | |
[5] | |
[6] | |
[7] | |
[8] | |
[9] | |
[10] | |
[11] | |
[12] | |
[13] | |
[14] | |
[15] | |
[16] | |
[17] | |
[18] | |
[19] | |
[20] | |
[21] | |
[22] | |
[23] | |
[24] | |
[25] | |
[26] | |
[27] | |
[28] | |
[29] | |
[30] | |
[31] | |
[32] | |
[33] | |
[34] | |
[35] | |
[36] | |
[37] | |
[38] | |
[39] | |
[40] | |
[41] | |
[42] | |
[43] | |
[44] | |
[45] | |
[46] | |
[47] | |
[48] | |
[49] | |
[50] | |
[51] | |
[52] | |
[53] | |
[54] | |
[55] | |
[56] | |
[57] | |
[58] | |
[59] | |
[60] | |
[61] | |
[62] | |
[63] | |
[64] | |
[65] | |
[66] | |
[67] | |
[68] | |
[69] | |
[70] | |
[71] | |
[72] | |
[73] | |
[74] | |
[75] | |
[76] | |
[77] | |
[78] | |
[79] | |
[80] | |
[81] | |
[82] | |
[83] | |
[84] | |
[85] | |
[86] | |
[87] | |
[88] | |
[89] | |
[90] | |
[91] | |
[92] | |
[93] | |
[94] | |
[95] | |
[96] | |
[97] | |
[98] | |
[99] | |
[100] | |
[101] | |
[102] | |
[103] | |
[104] | |
[105] | |
[106] | |
[107] | |
[108] | |
[109] | |
[110] | |
[111] | |
[112] | |
[113] | |
[114] | |
[115] | |
[116] | |
[117] | |
[118] | |
[119] | |
[120] | |
[121] | |
[122] | |
[123] | |
[124] | |
[125] | |
[126] | |
[127] | |
[128] | |
[129] | |
[130] | |
[131] | |
[132] | |
[133] | |
[134] | |
[135] | |
[136] | |
[137] | |
[138] | |
[139] | |
[140] | |
[141] | |
[142] | |
[143] | |
[144] | |
[145] | |
[146] | |
[147] | |
[148] | |
[149] | |
[150] | |
[151] | |
[152] | |
[153] | |
[154] | |
[155] | |
[156] | |
[157] | |
[158] | |
[159] | |
[160] | |
[161] | |
[162] | |
[163] | |
[164] | |
[165] | |
[166] | |
[167] | |
[168] | |
[169] | |
[170] | |
[171] | |
[172] | |
[173] | |
[174] | |
[175] | |
[176] | |
[177] | |
[178] | |
[179] | |
[180] | |
[181] | |
[182] | |
[183] | |
[184] | |
[185] | |
[186] | |
[187] | |
[188] | |
[189] | |
[190] | |
[191] | |
[192] | |
[193] | |
[194] | |
[195] | |
[196] | |
[197] | |
[198] | |
[199] | |
[200] | |
[201] | |
[202] | |
[203] | |
[204] | |
[205] | |
[206] | |
[207] | |
[208] | |
[209] | |
[210] | |
[211] | |
[212] | |
[213] | |
[214] | |
[215] | |
[216] | |
[217] | |
[218] | |
[219] | |
[220] | |
[221] | |
[222] | |
[223] | |
[224] | |
[225] | |
[226] | |
[227] | |
[228] | |
[229] | |
[230] | |
[231] | |
[232] | |
[233] | |
[234] | |
[235] | |
[236] | |
[237] | |
[238] | |
[239] | |
[240] | |
[241] | |
[242] | |
[243] | |
[244] | |
[245] | |
[246] | |
[247] | |
[248] | |
[249] | |
[250] | |
[251] |