Crystal twinning

Last updated
Quartz - Japan twin Quartz-rhqtz-109b.jpg
Quartz – Japan twin
Diagram of twinned crystals of albite. On the more perfect cleavage, which is parallel to the basal plane (P), is a system of fine striations, parallel to the second cleavage (M). Twinned crystals of Albite.png
Diagram of twinned crystals of albite. On the more perfect cleavage, which is parallel to the basal plane (P), is a system of fine striations, parallel to the second cleavage (M).

Crystal twinning occurs when two or more adjacent crystals of the same mineral are oriented so that they share some of the same crystal lattice points in a symmetrical manner. The result is an intergrowth of two separate crystals that are tightly bonded to each other. The surface along which the lattice points are shared in twinned crystals is called a composition surface or twin plane.

Contents

Crystallographers classify twinned crystals by a number of twin laws, which are specific to the crystal structure. The type of twinning can be a diagnostic tool in mineral identification. There are three main types of twinning. The first is growth twinning which can occur both in very large and very small particles. The second is transformation twinning, where there is a change in the crystal structure. The third is deformation twinning, in which twinning develops in a crystal in response to a shear stress, and is an important mechanism for permanent shape changes in a crystal.

Definition

Twin lattice (2D). The long horizontal red line is the composition plane where the two crystal segments join. The upper crystal lattice is the reflection of the lower crystal lattice. The red points are the shared crystal lattice points. Twin lattice (2D).png
Twin lattice (2D). The long horizontal red line is the composition plane where the two crystal segments join. The upper crystal lattice is the reflection of the lower crystal lattice. The red points are the shared crystal lattice points.

Twinning is a form of symmetrical intergrowth between two or more adjacent crystals of the same mineral. It differs from the ordinary random intergrowth of mineral grains in a mineral deposit, because the relative orientations of the two crystal segments show a fixed relationship that is characteristic of the mineral structure. The relationship is defined by a symmetry operation called a twin operation. [1] [2]

The twin operation is not one of the normal symmetry operations of the untwinned crystal structure. For example, the twin operation may be reflection across a plane that is not a symmetry plane of the single crystal. [1] [2]

On the microscopic level, the twin boundary is characterized by a set of atomic positions in the crystal lattice that are shared between the two orientations. [1] [2] These shared lattice points give the junction between the crystal segments much greater strength than that between randomly oriented grains, so that the twinned crystals do not easily break apart. [3]

Twin laws

Twin laws are symmetry operations that define the orientation between twin crystal segments. These are as characteristic of the mineral as are its crystal face angles. For example, crystals of staurolite show twinning at angles of almost precisely 90 degrees or 30 degrees. [3] A twin law is not a symmetry operation of the full set of basis points. [2]

Twin laws include reflection operations, rotation operations, and the inversion operation. Reflection twinning is described by the Miller indices of the twin plane (i.e. {hkl}) while rotational twinning is described by the direction of the twin axis (i.e. <hkl>). Inversion twinning is typically equivalent to a reflection or rotation symmetry. [1]

Rotational twin laws are almost always 2-fold rotations, though any other permitted rotation symmetry (3-fold, 4-fold, 5-fold or 6-fold) is possible. The twin axis will be perpendicular to a lattice plane. [4] It is possible for a rotational twin law to share the same axis as a rotational symmetry of the individual crystal if the twin law is a 2-fold rotation and the symmetry operation is a 3-fold rotation. This is the case for spinel law twinning on <111>: The spinel structure has a 3-fold rotational symmetry on <111> and spinel is commonly twinned by 2-fold rotation on <111>. [1]

The boundary between crystal segments is called a composition surface or, if it is planar, a composition plane. The composition plane is often, though not always, parallel to the twin law plane of a reflection law. If this is the case, the twin plane is always parallel to a possible crystal face. [2]

Common twin laws

Spinel law contact twinning. A single crystal is shown at left with the composition plane in red. At right, the crystal has effectively been cut on the composition plane and the front half rotated by 180deg to produce a contact twin. This creates reentrants at the top and lower left of the composition plane. Spinel twin.png
Spinel law contact twinning. A single crystal is shown at left with the composition plane in red. At right, the crystal has effectively been cut on the composition plane and the front half rotated by 180° to produce a contact twin. This creates reentrants at the top and lower left of the composition plane.

In the isometric system, the most common types of twins are the Spinel Law (twin plane, parallel to an octahedron) <111>, where the twin axis is perpendicular to an octahedral face, and the Iron Cross <001>, which is the interpenetration of two pyritohedrons, a subtype of dodecahedron. [5]

In the hexagonal system, calcite shows the contact twin laws {0001} and {0112}. Quartz shows the Brazil Law {1120}, and Dauphiné Law <0001>, which are penetration twins caused by transformation, and Japan Law {1122}, which is often caused by accidents during growth. [5]

In the tetragonal system, cyclical contact twins are the most commonly observed type of twin, such as in rutile titanium dioxide and cassiterite tin oxide. [5]

In the orthorhombic system, crystals usually twin on planes parallel to the prism face, where the most common is a {110} twin, which produces cyclical twins, such as in aragonite, chrysoberyl, and cerussite. [5]

In the monoclinic system, twins occur most often on the planes {100} and {001} by the Manebach Law {001}, Carlsbad Law [001], Baveno Law {021} in orthoclase, and the Swallow Tail Twins (Manebach law) {001} in gypsum. [5]

In the triclinic system, the most commonly twinned crystals are the feldspar minerals plagioclase and microcline. These minerals show the Albite and Pericline Laws. [4] [5]

The most common twin operations by crystal system are tabulated below. This list is not exhaustive, particularly for the crystal systems of lowest symmetry, such as the triclinic system. [6] [1] [5]

30deg twin of staurolite Twin of Staurolite (cropped).JPG
30° twin of staurolite
90deg twins of staurolite Staurolite 002.jpg
90° twins of staurolite
Iron pyrite cross twin Pyrite-254452.jpg
Iron pyrite cross twin
SystemLawOperationExamples
TriclinicAlbite law
Pericline law
Carlsbad law
Baveno law
Manebach law
{010}
<010>
<001>
{021}
{001}
Plagioclase
MonoclinicCarlsbad law
Baveno law
Manebach law
<001>
{021}
{001}
{100}
<031>
<231>
Orthoclase

Gypsum

Staurolite
Orthorhombic{110}
{101}
{011}
Aragonite, cerrusite; often cyclic
Tetragonal{110}
{101}
{011}
Cassiterite, rutile
Hexagonal


Brazil law
Dauphine law
Japan law
{0112}
{0001}
{1011}
{1120}
<0001>
{1122}
Calcite


Quartz
IsometricSpinel law


Iron cross law
<111>
{111}
{001}
<001>
Spinel


Pyrite

Types of twinning

Chrysoberyl showing cyclic twinning Chrysoberyl-282796 (cropped).jpg
Chrysoberyl showing cyclic twinning

Simple twinned crystals may be contact twins or penetration twins. Contact twins meet on a single composition plane, often appearing as mirror images across the boundary. Plagioclase, quartz, gypsum, and spinel often exhibit contact twinning. [3] Merohedral twinning occurs when the lattices of the contact twins superimpose in three dimensions, such as by relative rotation of one twin from the other. [7] An example is metazeunerite. [8] Contact twinning characteristically creates reentrant faces where faces of the crystal segments meet on the contact plane at an angle greater than 180°. [3]

A type of twinning involving 180° relationships is called hemitropism or hemitropy.

In penetration twins the individual crystals have the appearance of passing through each other in a symmetrical manner. [3] Orthoclase, staurolite, pyrite, and fluorite often show penetration twinning. The composition surface in penetration twins is usually irregular and extends to the center of the crystal. [2]

Contact twinning can arise from either reflection or rotation, whereas penetration twinning is usually produced by rotation. [2]

If several twin crystal parts are aligned by the same twin law they are referred to as multiple or repeated twins. If these multiple twins are aligned in parallel they are called polysynthetic twins. When the multiple twins are not parallel they are cyclic twins. Albite, calcite, and pyrite often show polysynthetic twinning. Closely spaced polysynthetic twinning is often observed as striations or fine parallel lines on the crystal face. Rutile, aragonite, cerussite, and chrysoberyl often exhibit cyclic twinning, typically in a radiating pattern. [3] [2]

For rotational twinning the relationship between the twin axis and twin plane falls into one of three types: [9]

  1. parallel twinning, when the twin axis and compositional plane lie parallel to each other,
  2. normal twinning, when the twin plane and compositional plane lie normally, and
  3. complex twinning, a combination of parallel twinning and normal twinning on one compositional plane.

Modes of formation

Parallel growth of pyromorphite Pyromorphite-171937.jpg
Parallel growth of pyromorphite

There are three modes of formation of twinned crystals.

Growth twinning (nanotwinning)

Fivefold twin in a gold nanoparticle (electron microscope micrograph). Twin2.jpg
Fivefold twin in a gold nanoparticle (electron microscope micrograph).

There are two types of twinning that can occur during growth, accidental and ones where the twinned structure has lower energy.

In accidental growth twinning an atom joins a crystal face in a less than ideal position, forming a seed for growth of a twin. The original crystal and its twin then grow together and closely resemble each other. This is characteristic enough of certain minerals to suggest that it is thermodynamically or kinetically favored under conditions of rapid growth. [3] [1]

Different from these are twins found in nanoparticles such as the image here, these fivefold or decahedral nanoparticles being one of the most common. [12] These cyclic twins occur as they are lower in energy at small sizes. [13] For the five-fold case shown, there is a disclination along the common axis [14] which leads to an additional strain energy. [15] Balancing this there is a reduction in the surface free energy, in large part due to more (111) surface facets. [16] In small nanoparticles the decahedral and a more complicated icosahedral structure (with twenty units) are lower energy, but at larger energies single crystals become lower energy. [17] [18] However, they do not have to transform into single crystals and can grow very large, and are known as fivelings, documented as early as 1831 by Gustav Rose; [19] further drawings are available in the Atlas der Kristallformen, and see also the article on fivelings. [20]

Transformation twinning

Illustration of transformation twinning in 2-D. The orthorhombic crystal at left transforms to a monoclinic crystal at right, with two parallel twinning planes (polysynthetic twinning) Transform twin.png
Illustration of transformation twinning in 2-D. The orthorhombic crystal at left transforms to a monoclinic crystal at right, with two parallel twinning planes (polysynthetic twinning)

Transformation and annealing twinning takes place when a cooling crystal experiences a displacive polymorphic transition. For example, leucite has an isometric crystal structure above about 665 °C (1,229 °F), but becomes tetragonal below this temperature. Any one of the three original axes of a crystal can become the long axis when this phase change takes place. Twinning results when different parts of the crystal break their isometric symmetry along a different choice of axis. This is typically polysynthetic twinning, which enables the crystal to maintain its isometric shape by averaging out the displacement in each direction. This produces a pseudomorphic crystal that appears to have isometric symmetry. Potassium feldspar likewise experiences polysynthetic twinning as it transforms from a monoclinic structure (orthoclase) to a triclinic structure (microcline) on slow cooling. [1]

Deformation twinning

Animation of deformation twinning of crystal Output JqJHfp.gif
Animation of deformation twinning of crystal

Deformation twinning is a response to shear stress. The crystal structure is displaced along successive planes of the crystal, a process also called glide. The twinning is always reflection twinning and the glide plane is also the mirror plane. Deformation twinning can be observed in a calcite cleavage fragment by applying gentle pressure with a knife blade near an edge. This particular glide twinning, {102}, is found almost universally in deformed rock beds containing calcite. [1]

Twinning and slip are competitive mechanisms for crystal deformation. Each mechanism is dominant in certain crystal systems and under certain conditions. [21] In fcc metals, slip is almost always dominant because the stress required is far less than twinning stress. [22]

Twinning can occur by cooperative displacement of atoms along the face of the twin boundary. This displacement of a large quantity of atoms simultaneously requires significant energy to perform. Therefore, the theoretical stress required to form a twin is quite high. It is believed that twinning is associated with dislocation motion on a coordinated scale, in contrast to slip, which is caused by independent glide at several locations in the crystal.

Compared to slip, twinning produces a deformation pattern that is more heterogeneous in nature. This deformation produces a local gradient across the material and near intersections between twins and grain boundaries. The deformation gradient can lead to fracture along the boundaries, particularly in bcc transition metals at low temperatures.


Of the three common crystalline structures bcc, fcc, and hcp, the hcp structure is the most likely to form deformation twins when strained, because they rarely have a sufficient number of slip systems for an arbitrary shape change. High strain rates, low stacking-fault energy and low temperatures facilitate deformation twinning. [23]

If a metal with face-centered cubic (fcc) structure, like Al, Cu, Ag, Au, etc., is subjected to stress, it will experience twinning. The formation and migration of twin boundaries is partly responsible for ductility and malleability of fcc metals. [24]

Twin boundaries are partly responsible for shock hardening and for many of the changes that occur in cold work of metals with limited slip systems or at very low temperatures. They also occur due to martensitic transformations: the motion of twin boundaries is responsible for the pseudoelastic and shape-memory behavior of nitinol, and their presence is partly responsible for the hardness due to quenching of steel. In certain types of high strength steels, very fine deformation twins act as primary obstacles against dislocation motion. These steels are referred to as 'TWIP' steels, where TWIP stands for twinning-induced plasticity. [25]

Deformation twinning crystallography

Deformation twinning crystallographic planes Deformation twinning.png
Deformation twinning crystallographic planes

Twinning is crystallographically defined by its twin plane 𝑲𝟏, the mirror plane in the twin and parent material, and 𝜼𝟏, which is the twinning shear direction. Deformation twins in Zr are generally lenticular in shape, lengthening in the 𝜼𝟏 direction and thickening along the 𝑲𝟏 plane normal. [26]

The twin plane, shear direction, and shear plane form the basis vectors of an orthogonal set. The axis-angle misorientation relationship between the parent and twin is a rotation of angle 𝜉 about the shear plane's normal direction 𝑷.

More generally, twinning can be described as a 180° rotation about an axis (𝑲𝟏 for type I twins or 𝜼𝟏 for type II twins normal direction) , or a mirror reflection in a plane (𝑲𝟏 or 𝜼𝟏 normal plane). [27]

In addition to a homogeneous shear, atomic shuffles are sometimes required to reform the correct crystal structure in the twinned lattice. For each twin variant, a reciprocal twin with swapped 𝑲𝟏 and 𝑲2, 𝜼𝟏 and 𝜼2 is possible, but one variant may appear more frequently in reality due to complexities with the required shuffles. [28]

there are only two crystallographic planes in a shearing action that do not change their shape and size as a consequence of the shear. The first 𝑲𝟏 is the plane defining the upper and lower surfaces of the sheared volume. This plane contains the shear direction. The other plane, designated C. The shear direction is shown with an arrow and labelled with its customary designation 𝜼𝟏. It follows from the above that there are three ways that a crystal lattice can be sheared while still retaining its crystal structure and symmetry:

  1. When 𝑲𝟏 is a rational plane and 𝜼2 a rational direction, a twin of the first kind
  2. When 𝑲2 is a rational plane and 𝜼𝟏 a rational direction, a twin of the second kind, rare
  3. When all four elements 𝑲𝟏, 𝑲2, 𝜼𝟏, and 𝜼2 are rational, a compound twin

Deformation twinning configuration

Regular twinning planes in calcite crystal. Crossed nicols image, magnification 10x (Field of view = 2 mm) Crystal Lines.JPG
Regular twinning planes in calcite crystal. Crossed nicols image, magnification 10× (Field of view = 2 mm)

A deformation twin embryo forms in BCC metal by accumulating stacking faults, with a variant selection governed by the local stress state. [29] [30] [31] Variation of the stress field close to twins inferred from HR-EBSD experimental [32] [33] and crystal plasticity finite element (CPFE) simulation data indicated that twins nucleate on sites with maximum strain energy density and twin resolved shear stress; thus, reducing the total elastic energy after formation. This relaxation depends on the twin thickness and is a deciding factor in the spacing between twins. [34] Experimental [35] and three-dimensional [36] analysis has focussed on the (stored) strain energy density measured along a path. This highly localised stress field can provide a sufficient driving force for concurrent twin nucleation [37] and inter/intra-granular crack nucleation.

Deformation twin growth can be perceived as a two-step process of i) thickening that is mediated by the interaction between the residual and mobile twin partials at the coherent twin-parent interface, [38] and ii) dislocation mobility along the twin shear direction. [39] The twin propagates when the homogeneous shear stress reaches a critical value, and a twin-parent interface advances inside the parent grain [240]. The propagating deformation twin generates a stress field due to its confinement by the surrounding parent crystal, and deformation twins develop a 3D oblate spheroid shape (which appears in 2D sections as a bi-convex lens) with a mixed coherent and non-coherent interface (Figure b). [39]

Kannan et al. [40] found, using in-situ ultra-high-speed optical imaging, that twin nucleation in single-crystal magnesium is stress-driven accompanied by instantaneous propagation at a speed of 1 km/s (initially) that prioritises volume lateral thickening over forward propagation, past a critical width where growth is then become faster along the shear direction. Barnett [41] also indicated that growth is due to twin tip extension. Furthermore, elastic simulations of the local stress field surrounding the ellipsoidal twin tip find that the field can be described using its lens angle () and that the stress field magnitude increases with twin thickness. [42]

(a) forescatter electron diode (FSD) image for deformation twins at grain boundary in age-hardened ferrite at I) 18 mm working distance and II) 38 mm working distance. (b) Schematic of a lenticular twin with interface dislocations and (c) Twin band. Twinforescatter diode.tif
(a) forescatter electron diode (FSD) image for deformation twins at grain boundary in age-hardened ferrite at I) 18 mm working distance and II) 38 mm working distance. (b) Schematic of a lenticular twin with interface dislocations and (c) Twin band.

In practice, plastic accommodation occurs in the parent crystal; thus, it also depends on the material’s yield stress, the anisotropic elastic stiffness of the parent crystal lattice, and the deformation twinning shear magnitude. [38] This can also be accompanied by long-range diffusion of elements and elemental segregation (e.g., Cr and Co in single crystal Ni-based superalloy MD2), which occurs at the twin boundary to facilitate twin growth by lowering the critical stacking fault energy. [44] A linear variation has been observed between twin thickness, stacking fault energy and grain size, [45] and to a lesser degree, the stress state of the twinning grain (Schmid Factor). [46] The twin thickness saturated once a critical residual dislocations’ density reached the coherent twin-parent crystal boundary. [31] [47]

Significant attention has been paid to the crystallography, [48] morphology [49] and macro mechanical effects [50] of deformation twinning. Although the criterion for deformation twin growth is not entirely understood, it is a tip-controlled phenomenon linked to the interaction between the residual and mobile twin partials at the twin interface; thermodynamically, this involves the elastic energy of the strained lattice, the interface and volume free-energy of the twin, and the dissipated energy of the growth mechanism. [51] To fully understand the interactions between microstructure (i.e., grain size, texture), temperature and strain rate on deformation twinning, it is crucial to characterise the (high) local stress and strain field associated with twin thickening and propagation. This is especially important for materials where cleavage fracture can be initiated by twinning (e.g., iron-silicon, the ferrite phase of age-hardened duplex stainless-steel, and single-crystal magnesium) as a stress-relieving mechanism.

Early studies of deformation twins arrested within grains of niobium [52] and iron [53] visualised the highly local strain concentration at the twin tip using an etch-pit procedure. More recently, high-resolution electron backscatter diffraction (HR-EBSD) has been used to investigate the strain 'singularity' ahead of a twin tip in hexagonal close-packed (HCP) zirconium alloy. A deformation twin in commercial purity titanium was characterised similarly and then quantified using a local Schmid factor (LSF) at the twin tip, [54] as described in equation below.

where σ is the stress tensor, Si is the Schmid tensor, Pi is its symmetric part, di is the shear direction and ni is the shear plane normal for ith slip system. The authors concluded that conditions at the twin tip control thickening and propagation in a manner analogous to the operation of dislocation sources ahead of a crack-tip. [55] In the analysis, a broad region of high LSF ahead of the twin tip favoured propagation, whereas a narrow region of high LSF promoted thickening. Since then, it has been argued [56] that the LSF firmly controls the twin variant selection, as twinning has strong polarity.

The LSF novelty – compared to other criteria to describe conditions at the twin [45] – lies in combining a geometrical criterion with the deformation field in the parent grain to provide an approximate indication of the local twin mode (i.e., thickening or propagation). However, the LSF analysis does not take advantage of the available full-field data, relies on global information on the applied stress, and does not consider the energy balance that drives twin growth. There have been few in-situ experiments to quantify the strain field ahead of a propagating deformation twin. [43] [33] Such observations might validate geometrical or hybrid geometrical-energy-based criteria [56] for growth. Nanoscale testing (i.e., transmission electron microscopy) may not represent the behaviour in bulk samples due to plasticity starvation, i.e., large surface area to volume ratio, [57] so a suitable analysis method is needed.

Lloyd [47] described the stress concentration field ahead of the twin tip using a two-dimensional dislocation-based model within a single magnesium grain. Wang and Li, [58] who considered microscopic phase-field (MPF) models of cracks, noted that the stress fields were similar for dislocations, deformation twinning and martensitic transformations, with differences only in the traction of the created surface, i.e., there is 100% traction recovery for dislocations and a traction-free surface for a crack. They highlighted that the stress field singularity regulates the advancement of the crack-tip and dislocations. This stress concentration can be characterised using a path-independent line integral, as shown by Eshelby for dislocations considering the contribution from the surface traction and ellipsoidal inclusions, [59] and Rice [60] for cracks and stress concentrations with traction-free surfaces. Furthermore, Venables [61] noted that the oblate spheroid shape of the twin tip is the ideal example of an ellipsoid inclusion or a notch.

See also

Related Research Articles

Precipitation hardening, also called age hardening or particle hardening, is a heat treatment technique used to increase the yield strength of malleable materials, including most structural alloys of aluminium, magnesium, nickel, titanium, and some steels, stainless steels, and duplex stainless steel. In superalloys, it is known to cause yield strength anomaly providing excellent high-temperature strength.

<span class="mw-page-title-main">Electron backscatter diffraction</span> Scanning electron microscopy technique

Electron backscatter diffraction (EBSD) is a scanning electron microscopy (SEM) technique used to study the crystallographic structure of materials. EBSD is carried out in a scanning electron microscope equipped with an EBSD detector comprising at least a phosphorescent screen, a compact lens and a low-light camera. In the microscope an incident beam of electrons hits a tilted sample. As backscattered electrons leave the sample, they interact with the atoms and are both elastically diffracted and lose energy, leaving the sample at various scattering angles before reaching the phosphor screen forming Kikuchi patterns (EBSPs). The EBSD spatial resolution depends on many factors, including the nature of the material under study and the sample preparation. They can be indexed to provide information about the material's grain structure, grain orientation, and phase at the micro-scale. EBSD is used for impurities and defect studies, plastic deformation, and statistical analysis for average misorientation, grain size, and crystallographic texture. EBSD can also be combined with energy-dispersive X-ray spectroscopy (EDS), cathodoluminescence (CL), and wavelength-dispersive X-ray spectroscopy (WDS) for advanced phase identification and materials discovery.

<span class="mw-page-title-main">Grain boundary</span> Interface between crystallites in a polycrystalline material

In materials science, a grain boundary is the interface between two grains, or crystallites, in a polycrystalline material. Grain boundaries are two-dimensional defects in the crystal structure, and tend to decrease the electrical and thermal conductivity of the material. Most grain boundaries are preferred sites for the onset of corrosion and for the precipitation of new phases from the solid. They are also important to many of the mechanisms of creep. On the other hand, grain boundaries disrupt the motion of dislocations through a material, so reducing crystallite size is a common way to improve mechanical strength, as described by the Hall–Petch relationship.

<span class="mw-page-title-main">Cottrell atmosphere</span> Concept in materials science

In materials science, the concept of the Cottrell atmosphere was introduced by A. H. Cottrell and B. A. Bilby in 1949 to explain how dislocations are pinned in some metals by boron, carbon, or nitrogen interstitials.

Dynamic recrystallization (DRX) is a type of recrystallization process, found within the fields of metallurgy and geology. In dynamic recrystallization, as opposed to static recrystallization, the nucleation and growth of new grains occurs during deformation rather than afterwards as part of a separate heat treatment. The reduction of grain size increases the risk of grain boundary sliding at elevated temperatures, while also decreasing dislocation mobility within the material. The new grains are less strained, causing a decrease in the hardening of a material. Dynamic recrystallization allows for new grain sizes and orientation, which can prevent crack propagation. Rather than strain causing the material to fracture, strain can initiate the growth of a new grain, consuming atoms from neighboring pre-existing grains. After dynamic recrystallization, the ductility of the material increases.

<span class="mw-page-title-main">Zirconium alloys</span> Zircaloy family

Zirconium alloys are solid solutions of zirconium or other metals, a common subgroup having the trade mark Zircaloy. Zirconium has very low absorption cross-section of thermal neutrons, high hardness, ductility and corrosion resistance. One of the main uses of zirconium alloys is in nuclear technology, as cladding of fuel rods in nuclear reactors, especially water reactors. A typical composition of nuclear-grade zirconium alloys is more than 95 weight percent zirconium and less than 2% of tin, niobium, iron, chromium, nickel and other metals, which are added to improve mechanical properties and corrosion resistance.

<span class="mw-page-title-main">Slip (materials science)</span> Displacement between parts of a crystal along a crystallographic plane

In materials science, slip is the large displacement of one part of a crystal relative to another part along crystallographic planes and directions. Slip occurs by the passage of dislocations on close/packed planes, which are planes containing the greatest number of atoms per area and in close-packed directions. Close-packed planes are known as slip or glide planes. A slip system describes the set of symmetrically identical slip planes and associated family of slip directions for which dislocation motion can easily occur and lead to plastic deformation. The magnitude and direction of slip are represented by the Burgers vector, b.

<span class="mw-page-title-main">Shear band</span>

A shear band is a narrow zone of intense shearing strain, usually of plastic nature, developing during severe deformation of ductile materials. As an example, a soil specimen is shown in Fig. 1, after an axialsymmetric compression test. Initially the sample was cylindrical in shape and, since symmetry was tried to be preserved during the test, the cylindrical shape was maintained for a while during the test and the deformation was homogeneous, but at extreme loading two X-shaped shear bands had formed and the subsequent deformation was strongly localized.

Pseudoelasticity, sometimes called superelasticity, is an elastic (reversible) response to an applied stress, caused by a phase transformation between the austenitic and martensitic phases of a crystal. It is exhibited in shape-memory alloys.

In geology, a deformation mechanism is a process occurring at a microscopic scale that is responsible for changes in a material's internal structure, shape and volume. The process involves planar discontinuity and/or displacement of atoms from their original position within a crystal lattice structure. These small changes are preserved in various microstructures of materials such as rocks, metals and plastics, and can be studied in depth using optical or digital microscopy.

<span class="mw-page-title-main">Grain boundary strengthening</span> Method of strengthening materials by changing grain size

In materials science, grain-boundary strengthening is a method of strengthening materials by changing their average crystallite (grain) size. It is based on the observation that grain boundaries are insurmountable borders for dislocations and that the number of dislocations within a grain has an effect on how stress builds up in the adjacent grain, which will eventually activate dislocation sources and thus enabling deformation in the neighbouring grain as well. By changing grain size, one can influence the number of dislocations piled up at the grain boundary and yield strength. For example, heat treatment after plastic deformation and changing the rate of solidification are ways to alter grain size.

Oxide dispersion strengthened alloys (ODS) are alloys that consist of a metal matrix with small oxide particles dispersed within it. They have high heat resistance, strength, and ductility. Alloys of nickel are the most common but includes iron aluminum alloys.

Severe plastic deformation (SPD) is a generic term describing a group of metalworking techniques involving very large strains typically involving a complex stress state or high shear, resulting in a high defect density and equiaxed "ultrafine" grain (UFG) size or nanocrystalline (NC) structure.

Crack closure is a phenomenon in fatigue loading, where the opposing faces of a crack remain in contact even with an external load acting on the material. As the load is increased, a critical value will be reached at which time the crack becomes open. Crack closure occurs from the presence of material propping open the crack faces and can arise from many sources including plastic deformation or phase transformation during crack propagation, corrosion of crack surfaces, presence of fluids in the crack, or roughness at cracked surfaces.

Geometrically necessary dislocations are like-signed dislocations needed to accommodate for plastic bending in a crystalline material. They are present when a material's plastic deformation is accompanied by internal plastic strain gradients. They are in contrast to statistically stored dislocations, with statistics of equal positive and negative signs, which arise during plastic flow from multiplication processes like the Frank-Read source.

Dislocation avalanches are rapid discrete events during plastic deformation, in which defects are reorganized collectively. This intermittent flow behavior has been observed in microcrystals, whereas macroscopic plasticity appears as a smooth process. Intermittent plastic flow has been observed in several different systems. In AlMg Alloys, interaction between solute and dislocations can cause sudden jump during dynamic strain aging. In metallic glass, it can be observed via shear banding with stress localization; and single crystal plasticity, it shows up as slip burst. However, analysis of the events with orders-magnitude difference in sizes with different crystallographic structure reveals power-law scaling between the number of events and their magnitude, or scale-free flow.

Crystal plasticity is a mesoscale computational technique that takes into account crystallographic anisotropy in modelling the mechanical behaviour of polycrystalline materials. The technique has typically been used to study deformation through the process of slip, however, there are some flavors of crystal plasticity that can incorporate other deformation mechanisms like twinning and phase transformations. Crystal plasticity is used to obtain the relationship between stress and strain that also captures the underlying physics at the crystal level. Hence, it can be used to predict not just the stress-strain response of a material, but also the texture evolution, micromechanical field distributions, and regions of strain localisation. The two widely used formulations of crystal plasticity are the one based on the finite element method known as Crystal Plasticity Finite Element Method (CPFEM), which is developed based on the finite strain formulation for the mechanics, and a spectral formulation which is more computationally efficient due to the fast Fourier transform, but is based on the small strain formulation for the mechanics.

<span class="mw-page-title-main">Slip bands in metals</span> Deformation mechanism in crystallines

Slip bands or stretcher-strain marks are localized bands of plastic deformation in metals experiencing stresses. Formation of slip bands indicates a concentrated unidirectional slip on certain planes causing a stress concentration. Typically, slip bands induce surface steps and a stress concentration which can be a crack nucleation site. Slip bands extend until impinged by a boundary, and the generated stress from dislocations pile-up against that boundary will either stop or transmit the operating slip depending on its (mis)orientation.

Angus J Wilkinson is a professor of materials science based at University of Oxford. He is a specialist in micromechanics, electron microscopy and crystal plasticity. He assists in overseeing the MicroMechanics group while focusing on the fundamentals of material deformation. He developed the HR-EBSD method for mapping stress and dislocation density at high spatial resolution used at the micron scale in mechanical testing and micro-cantilevers to extract data on mechanical properties that are relevant to materials engineering.

<span class="mw-page-title-main">Precipitate-free zone</span> Region around a material grain boundary free of solid impurities

In materials science, a precipitate-free zone (PFZ) refers to microscopic localized regions around grain boundaries that are free of precipitates. It is a common phenomenon that arises in polycrystalline materials where heterogeneous nucleation of precipitates is the dominant nucleation mechanism. This is because grain boundaries are high-energy surfaces that act as sinks for vacancies, causing regions adjacent to a grain boundary to be devoid of vacancies. As it is energetically favorable for heterogeneous nucleation to occur preferentially around defect-rich sites such as vacancies, nucleation of precipitates is impeded in the vacancy-free regions immediately adjacent to grain boundaries

References

  1. 1 2 3 4 5 6 7 8 9 Nesse, William D. (2000). Introduction to mineralogy. New York: Oxford University Press. pp. 87–91. ISBN   9780195106916.
  2. 1 2 3 4 5 6 7 8 Klein, Cornelis; Hurlbut, Cornelius S. Jr. (1993). Manual of mineralogy : (after James D. Dana) (21st ed.). New York: Wiley. pp. 102–106. ISBN   047157452X.
  3. 1 2 3 4 5 6 7 8 Sinkankas, John (1964). Mineralogy for amateurs. Princeton, N.J.: Van Nostrand. pp. 96–105. ISBN   0442276249.
  4. 1 2 Nelson, Stephen A. (2013). "Twinning, Polymorphism, Polytypism, Pseudomorphism". Tulane University. Retrieved 19 February 2022.
  5. 1 2 3 4 5 6 7 Klein & Hurlbut 1993, pp. 104–106.
  6. Nesse 2000, pp. 89, 214–216.
  7. Yeates, Todd O. (1997). "[22] Detecting and overcoming crystal twinning". Macromolecular Crystallography Part A. Methods in Enzymology. Vol. 276. pp. 344–358. doi:10.1016/S0076-6879(97)76068-3. ISBN   9780121821777. PMID   27799105.
  8. Locock, A. J.; Burns, P. C. (1 April 2003). "Crystal Structures and Synthesis of the Copper-Dominant Members of the Autunite and Meta-Autunite Groups: Torbernite, Zeunerite, Metatorbernite and Metazeunerite". The Canadian Mineralogist. 41 (2): 489–502. Bibcode:2003CaMin..41..489L. doi:10.2113/gscanmin.41.2.489.
  9. Tobi, Alexander C. (1961). "The recognition of plagioclase twins in sections normal to the composition plane". American Mineralogist. 46 (11–12): 1470–1488. Retrieved 19 February 2022.
  10. Klein & Hurlbut 1993, p. 101.
  11. Klein & Hurlbut 1993, p. 167.
  12. Hofmeister, H. (1998). <3::aid-crat3>3.0.co;2-3 "Forty Years Study of Fivefold Twinned Structures in Small Particles and Thin Films". Crystal Research and Technology. 33 (1): 3–25. doi:10.1002/(sici)1521-4079(1998)33:1<3::aid-crat3>3.0.co;2-3. ISSN   0232-1300.
  13. Marks, Laurence. D. (2023), "Shape, thermodynamics and kinetics of nanoparticles", Encyclopedia of Nanomaterials, Elsevier, pp. 383–417, doi:10.1016/b978-0-12-822425-0.00082-8, ISBN   978-0-12-822423-6 , retrieved 2023-07-11
  14. Wit, R de (1972). "Partial disclinations". Journal of Physics C: Solid State Physics. 5 (5): 529–534. doi:10.1088/0022-3719/5/5/004. ISSN   0022-3719.
  15. Howie, A.; Marks, L. D. (1984). "Elastic strains and the energy balance for multiply twinned particles". Philosophical Magazine A. 49 (1): 95–109. doi:10.1080/01418618408233432. ISSN   0141-8610.
  16. Marks, L. D. (1984). "Surface structure and energetics of multiply twinned particles". Philosophical Magazine A. 49 (1): 81–93. doi:10.1080/01418618408233431. ISSN   0141-8610.
  17. Baletto, Francesca; Ferrando, Riccardo (2005). "Structural properties of nanoclusters: Energetic, thermodynamic, and kinetic effects". Reviews of Modern Physics. 77 (1): 371–423. doi:10.1103/RevModPhys.77.371. ISSN   0034-6861.
  18. Rogers, Blake; Lehr, Alexander; Velázquez-Salazar, J. Jesús; Whetten, Robert; Mendoza-Cruz, Ruben; Bazan-Diaz, Lourdes; Bahena-Uribe, Daniel; José Yacaman, Miguel (2023). "Decahedra and Icosahedra Everywhere: The Anomalous Crystallization of Au and Other Metals at the Nanoscale". Crystal Research and Technology. 58 (4). doi:10.1002/crat.202200259. ISSN   0232-1300.
  19. Rose, Gustav (1831). "Ueber die Krystallformen des Goldes und des Silbers". Annalen der Physik. 99 (10): 196–204. doi:10.1002/andp.18310991003. ISSN   0003-3804.
  20. Goldschmidt, V (1918). Atlas der Kristallformen, V4, Plates. Carl Winters Universitätsbuchhandlung. pp. Tafel 48, Fig. 21. (Gold). hdl:2027/nyp.33433084029887.
  21. Mahajan, S.; Williams, D. F. (June 1973). "Deformation Twinning in Metals and Alloys". International Metallurgical Reviews. 18 (2): 43–61. doi:10.1179/imtlr.1973.18.2.43.
  22. Beyerlein, Irene J.; Mara, Nathan A.; Bhattacharyya, Dhriti; Alexander, David J.; Necker, Carl T. (January 2011). "Texture evolution via combined slip and deformation twinning in rolled silver–copper cast eutectic nanocomposite". International Journal of Plasticity. 27 (1): 121–146. doi:10.1016/j.ijplas.2010.05.007.
  23. Courtney, Thomas H. (2000) Mechanical Behavior of Materials, 2nd ed. McGraw Hill. ISBN   1-57766-425-6
  24. Nurul Akmal Che, Lah; Trigueros, Sonia (2019). "Synthesis and modelling of the mechanical properties of Ag, Au and Cu nanowires". Sci. Technol. Adv. Mater. 20 (1): 225–261. Bibcode:2019STAdM..20..225L. doi:10.1080/14686996.2019.1585145. PMC   6442207 . PMID   30956731.
  25. Steinmetz, D.R.; Jäpel, T.; Wietbrock, B.; Eisenlohr, P.; Gutierrez-Urrutia, I.; Saeed (2013), "Revealing the strain-hardening behavior of twinning-induced plasticity steels: Theory, simulations, experiments", Acta Materialia, 61 (2): 494, Bibcode:2013AcMat..61..494S, doi:10.1016/j.actamat.2012.09.064 .
  26. Christian, J.W.; Mahajan, S. (1995). "Deformation twinning". Progress in Materials Science. 39 (1–2): 1–157. doi:10.1016/0079-6425(94)00007-7.
  27. Calhoun, C.A.; Garlea, E.; Sisneros, T.A.; Agnew, S.R. (April 2018). "In-situ neutron diffraction characterization of temperature dependence deformation in α-uranium". Journal of Nuclear Materials. 502: 60–67. Bibcode:2018JNuM..502...60C. doi: 10.1016/j.jnucmat.2018.01.036 . OSTI   1478070.
  28. Bilby, B. A.; Crocker, A. G. (1965-10-26). "The theory of the crystallography of deformation twinning". Proceedings of the Royal Society of London. Series A. Mathematical and Physical Sciences. 288 (1413): 240–255. Bibcode:1965RSPSA.288..240B. doi:10.1098/rspa.1965.0216. ISSN   0080-4630. S2CID   120566509.
  29. Cahn, R.W. (Oct 1954). "Twinned crystals". Advances in Physics. 3 (12): 363–445. Bibcode:1954AdPhy...3..363C. doi:10.1080/00018735400101223. ISSN   0001-8732.
  30. Wang, S.; Schuman, C.; Bao, L.; Lecomte, J.S.; Zhang, Y.; Raulot, J.M.; Philippe, M.J.; Zhao, X.; Esling, C. (May 2012). "Variant selection criterion for twin variants in titanium alloys deformed by rolling" (PDF). Acta Materialia. 60 (9): 3912–3919. Bibcode:2012AcMat..60.3912W. doi:10.1016/j.actamat.2012.03.046.
  31. 1 2 Christian, J.W. (2002). "Deformation Twinning". The Theory of Transformations in Metals and Alloys. Elsevier. pp. 859–960. doi:10.1016/b978-008044019-4/50025-8. ISBN   978-0-08-044019-4.
  32. Abdolvand, Hamidreza; Sedaghat, Omid; Guo, Yi (Oct 2018). "Nucleation and growth of { 1122 } twins in titanium: Elastic energy and stress fields at the vicinity of twins". Materialia. 2: 58–62. doi:10.1016/j.mtla.2018.06.012. S2CID   139852253.
  33. 1 2 Guo, Y.; Schwiedrzik, J.; Michler, J.; Maeder, X. (2016-11-01). "On the nucleation and growth of {112¯2} twin in commercial purity titanium: In situ investigation of the local stress field and dislocation density distribution". Acta Materialia. 120: 292–301. doi:10.1016/j.actamat.2016.08.073. ISSN   1359-6454.
  34. Paudel, YubRaj; Barrett, Christopher D.; Tschopp, Mark A.; Inal, Kaan; El Kadiri, Haitham (July 2017). "Beyond initial twin nucleation in hcp metals: Micromechanical formulation for determining twin spacing during deformation". Acta Materialia. 133: 134–146. Bibcode:2017AcMat.133..134P. doi: 10.1016/j.actamat.2017.05.013 . ISSN   1359-6454.
  35. Paramatmuni, Chaitanya; Zheng, Zebang; Rainforth, W. Mark; Dunne, Fionn P. E. (2020-12-01). "Twin nucleation and variant selection in Mg alloys: An integrated crystal plasticity modelling and experimental approach" (PDF). International Journal of Plasticity. 135: 102778. doi:10.1016/j.ijplas.2020.102778. ISSN   0749-6419. S2CID   224887964.
  36. Paramatmuni, Chaitanya; Guo, Yi; Withers, Philip J.; Dunne, Fionn P. E. (2021-08-01). "A three-dimensional mechanistic study of the drivers of classical twin nucleation and variant selection in Mg alloys: A mesoscale modelling and experimental study". International Journal of Plasticity. 143: 103027. doi:10.1016/j.ijplas.2021.103027. hdl: 10044/1/90290 . ISSN   0749-6419. S2CID   236612852.
  37. Arul Kumar, M.; Beyerlein, I. J.; Tomé, C. N. (2016-09-01). "Effect of local stress fields on twin characteristics in HCP metals". Acta Materialia. 116: 143–154. Bibcode:2016AcMat.116..143A. doi: 10.1016/j.actamat.2016.06.042 . ISSN   1359-6454.
  38. 1 2 Beyerlein, Irene J.; Zhang, Xinghang; Misra, Amit (July 2014). "Growth Twins and Deformation Twins in Metals". Annual Review of Materials Research. 44 (1): 329–363. Bibcode:2014AnRMS..44..329B. doi: 10.1146/annurev-matsci-070813-113304 . ISSN   1531-7331.
  39. 1 2 Britton, T. B.; Dunne, F. P. E.; Wilkinson, A. J. (2015-06-08). "On the mechanistic basis of deformation at the microscale in hexagonal close-packed metals". Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences. 471 (2178): 20140881. Bibcode:2015RSPSA.47140881B. doi: 10.1098/rspa.2014.0881 . S2CID   138085929.
  40. Liu, Y.; Tang, P. Z.; Gong, M. Y.; McCabe, R. J.; Wang, J.; Tomé, C. N. (2019-07-25). "Three-dimensional character of the deformation twin in magnesium". Nature Communications. 10 (1): 3308. Bibcode:2019NatCo..10.3308L. doi:10.1038/s41467-019-10573-7. ISSN   2041-1723. PMC   6658514 . PMID   31346160.
  41. Barnett, M. R. (2007-08-25). "Twinning and the ductility of magnesium alloys: Part I: "Tension" twins". Materials Science and Engineering: A. 464 (1): 1–7. doi:10.1016/j.msea.2006.12.037. ISSN   0921-5093.
  42. Arul Kumar, M.; Kanjarla, A. K.; Niezgoda, S. R.; Lebensohn, R. A.; Tomé, C. N. (2015-02-01). "Numerical study of the stress state of a deformation twin in magnesium". Acta Materialia. 84: 349–358. doi: 10.1016/j.actamat.2014.10.048 . ISSN   1359-6454.
  43. 1 2 Koko, Abdalrhaman; Elmukashfi, Elsiddig; Dragnevski, Kalin; Wilkinson, Angus J.; Marrow, Thomas James (2021-10-01). "J-integral analysis of the elastic strain fields of ferrite deformation twins using electron backscatter diffraction". Acta Materialia. 218: 117203. Bibcode:2021AcMat.21817203K. doi:10.1016/j.actamat.2021.117203. ISSN   1359-6454.
  44. Barba, D.; Alabort, E.; Pedrazzini, S.; Collins, D. M.; Wilkinson, A. J.; Bagot, P. A. J.; Moody, M. P.; Atkinson, C.; Jérusalem, A.; Reed, R. C. (2017-08-15). "On the microtwinning mechanism in a single crystal superalloy". Acta Materialia. 135: 314–329. Bibcode:2017AcMat.135..314B. doi: 10.1016/j.actamat.2017.05.072 . hdl: 10044/1/63276 . ISSN   1359-6454. S2CID   55924981.
  45. 1 2 Mahajan, S.; Williams, D. F. (1973-06-01). "Deformation Twinning in Metals and Alloys". International Metallurgical Reviews. 18 (2): 43–61. doi:10.1179/imtlr.1973.18.2.43. ISSN   0367-9020.
  46. Beyerlein, I.J.; Capolungo, L.; Marshall, P.E.; McCabe, R.J.; Tomé, C.N. (2010-05-28). "Statistical analyses of deformation twinning in magnesium". Philosophical Magazine. 90 (16): 2161–2190. Bibcode:2010PMag...90.2161B. doi:10.1080/14786431003630835. ISSN   1478-6435. S2CID   136111937.
  47. 1 2 Lloyd, J. T. (2018-02-28). "A dislocation-based model for twin growth within and across grains". Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences. 474 (2210): 20170709. Bibcode:2018RSPSA.47470709L. doi:10.1098/rspa.2017.0709. PMC   5832837 . PMID   29507516.
  48. Venables, J. A. (1961-03-01). "Deformation twinning in face-centred cubic metals". The Philosophical Magazine. 6 (63): 379–396. Bibcode:1961PMag....6..379V. doi:10.1080/14786436108235892. ISSN   0031-8086.
  49. Johansson, Johan; Odén, Magnus (2000-06-01). "Load sharing between austenite and ferrite in a duplex stainless steel during cyclic loading". Metallurgical and Materials Transactions A. 31 (6): 1557–1570. doi:10.1007/s11661-000-0166-3. ISSN   1543-1940. S2CID   137142133.
  50. Marrow, T. J.; King, J. E. (1994-06-15). "Fatigue crack propagation mechanisms in a thermally aged duplex stainless steel". Materials Science and Engineering: A. 183 (1): 91–101. doi:10.1016/0921-5093(94)90893-1. ISSN   0921-5093.
  51. Christian, J. W.; Mahajan, S. (1995-01-01). "Deformation twinning". Progress in Materials Science. 39 (1): 1–157. doi:10.1016/0079-6425(94)00007-7. ISSN   0079-6425.
  52. Sleeswyk, A. W (1962-08-01). "Emissary dislocations: Theory and experiments on the propagation of deformation twins in α-iron". Acta Metallurgica. 10 (8): 705–725. doi:10.1016/0001-6160(62)90040-8. ISSN   0001-6160.
  53. Spreadborough, J.; Langheinrich, D.; Anderson, E.; Brandon, D. (Dec 1964). "Etch-Pit Observations Concerning Twins in Iron and Iron Alloys". Journal of Applied Physics. 35 (12): 3585–3587. Bibcode:1964JAP....35.3585S. doi:10.1063/1.1713275. ISSN   0021-8979.
  54. Guo, Y.; Abdolvand, H.; Britton, T. B.; Wilkinson, A. J. (2017-03-01). "Growth of {1122} twins in titanium: A combined experimental and modelling investigation of the local state of deformation". Acta Materialia. 126: 221–235. doi: 10.1016/j.actamat.2016.12.066 . hdl: 10044/1/67478 . ISSN   1359-6454. S2CID   136128910.
  55. Baxevanakis, K. P.; Gourgiotis, P. A.; Georgiadis, H. G. (July 2017). "Interaction of cracks with dislocations in couple-stress elasticity. Part I: Opening mode". International Journal of Solids and Structures. 118–119: 179–191. doi: 10.1016/j.ijsolstr.2017.03.019 . ISSN   0020-7683. S2CID   55155535.
  56. 1 2 Kacher, Josh; Sabisch, Julian E.; Minor, Andrew M. (2019-07-01). "Statistical analysis of twin/grain boundary interactions in pure rhenium". Acta Materialia. 173: 44–51. Bibcode:2019AcMat.173...44K. doi: 10.1016/j.actamat.2019.04.051 . ISSN   1359-6454.
  57. Gong, Mingyu; Hirth, John P.; Liu, Yue; Shen, Yao; Wang, Jian (2017-11-03). "Interface structures and twinning mechanisms of twins in hexagonal metals". Materials Research Letters. 5 (7): 449–464. doi: 10.1080/21663831.2017.1336496 . S2CID   136448819.
  58. Wang, Yunzhi; Li, Ju (2010-02-01). "Phase field modeling of defects and deformation". Acta Materialia. 58 (4): 1212–1235. Bibcode:2010AcMat..58.1212W. doi:10.1016/j.actamat.2009.10.041. ISSN   1359-6454.
  59. Eshelby, John Douglas; Mott, Nevill Francis (1951-11-06). "The force on an elastic singularity". Philosophical Transactions of the Royal Society of London. Series A, Mathematical and Physical Sciences. 244 (877): 87–112. Bibcode:1951RSPTA.244...87E. doi:10.1098/rsta.1951.0016. S2CID   14703976.
  60. Rice, J. R. (1968-06-01). "A Path Independent Integral and the Approximate Analysis of Strain Concentration by Notches and Cracks". Journal of Applied Mechanics. 35 (2): 379–386. Bibcode:1968JAM....35..379R. doi:10.1115/1.3601206. ISSN   0021-8936.
  61. Venables, J. A (1964-07-01). "The electron microscopy of deformation twinning". Journal of Physics and Chemistry of Solids. 25 (7): 685–692. Bibcode:1964JPCS...25..685V. doi:10.1016/0022-3697(64)90177-5. ISSN   0022-3697.