Slip bands in metals

Last updated
A slip band formed on a ferrite grain in an age hardened duplex stainless steel. The slip band at the centre of the image was observed at a certain load, then the load was increased with a burst of dislocations coming out of the slip band tip as a response to the load increment. This burst of dislocations and topographic change ahead of the slip band was observed across different slip bands. image length is 10 um. S4 loaded slip band.gif
A slip band formed on a ferrite grain in an age hardened duplex stainless steel. The slip band at the centre of the image was observed at a certain load, then the load was increased with a burst of dislocations coming out of the slip band tip as a response to the load increment. This burst of dislocations and topographic change ahead of the slip band was observed across different slip bands. image length is 10 um.

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 (e.g., roughness due persistent slip bands during fatigue) 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. [3] [4]

Contents

Formation of slip bands under cyclic conditions is addressed as persistent slip bands (PSBs) where formation under monotonic condition is addressed as dislocation planar arrays (or simply slip-bands, see Slip bands in the absence of cyclic loading section). [5] Slip-bands can be simply viewed as boundary sliding due to dislocation glide that lacks (the complexity of ) PSBs high plastic deformation localisation manifested by tongue- and ribbon-like extrusion. And, where PSBs normally studied with (effective) Burgers vector aligned with the extrusion plane because a PSB extends across the grain and exacerbates during fatigue; [6] a monotonic slip-band has a Burger’s vector for propagation and another for plane extrusions both controlled by the conditions at the tip.

Persistent slip bands (PSBs)

PSB structure (adopted from ) PSBStructure.tif
PSB structure (adopted from )

Persistent slip-bands (PSBs) are associated with strain localisation due to fatigue in metals and cracking on the same plane. Transmission electron microscopy (TEM) and three-dimensional discrete dislocation dynamics (DDD [8] ) simulation were used to reveal and understand dislocations type and arrangement/patterns to relate it to the sub-surface structure. PSB – ladder structure – is formed mainly from low-density channels of mobile gliding screw dislocation segments and high-density walls of dipolar edge dislocation segments piled up with tangled bowing-out edge segment and different sizes of dipolar loops scattered between the walls and channels. [9] [10]

One type of dislocation loop forms the boundary of a completely enclosed patch of slipped material on the slip plane which terminates at the free surface. Widening of the slip band: Screw dislocation can have high enough resolved shear stress for a glide on more than one slip plane. Cross-slip can occur. But this leaves some segments of dislocation on the original slip plane. Dislocation can cross-slip back on to a parallel primary slip plane. where it forms a new dislocation source, and the process can repeat. These walls in PSBs are a ‘dipole dispersion’ form of stable arrangement of edge dislocations with minimal long-range stress field which has a minimal long-range stress field.[ clarification needed ] This is different to slip-bands that is a planar stack of a stable array that has a strong long-range stress field.[ clarification needed ] Thus, – in the free surface – cut and open (elimination) of dislocation loops at the surface cause the irreversible/persistent surface step associated with slip-bands. [10] [11] [12]

Surface relief through extrusion occurs on the Burger's vector direction and extrusion height and PSB depth increase with PSB thickness. [13] PSB and planar walls are parallel and perpendicularly aligned with the normal direction of the Critical resolved shear stress, respectively. [14] And once dislocation saturate and reach its sessile configuration, cracks were observed to nucleate and propagate along PSB extrusions. [15] [16] [17] To summarise, contrary to 2D line defects, the field at the slip-band tip is due to three-dimensional interactions where the slip band extrusion simulates a sink-like dislocation blooming along the slip band axis. The magnitude of the gradient deformation field ahead of the slip band depends on the slip height and the mechanical conditions for propagation is influenced by the emitted dislocations long range field.A surface marking, or slip band, appears at the intersection of an active slip plane and the free surface of a crystal. Slip occurs in avalanches separated in time. Avalanches from other slip systems crossing a slip plane containing an active source led to the observed stepped surface markings, with successive avalanches from the given source displaced relative to each other. [18]

Dislocations are generated on a single slip plane They point out that a dislocation segment (Frank–Read source), lying in a slip plane and pinned at both ends, is a source of an unlimited number of dislocation loops. In this way the grouping of dislocations into an avalanche of a thousand or so loops on a single slip plane can be understood. [19] Each dislocation loop has a stress field that opposes the applied stress in the neighbourhood of the source. When enough loops have been generated, the stress at the source will fall to a value so low that additional loops cannot form. Only after the original avalanche of loops has moved some distance away can another avalanche occur.

Generation of the first avalanche at a source is easily understood. When the stress at the source reaches r*, loops are generated, and continue to be generated until the back-stress stops the avalanche. A second avalanche will not occur immediately in polycrystals, for the loops in the first avalanche are stopped or partially stopped at grain boundaries. Only if the external stress is increased substantially will a second avalanche be formed. In this way the formation of additional avalanches with rising stress can be understood.

It remains to explain the displacement of successive avalanches by a small amount normal to the slip plane, thereby accounting for the observed fine structure of slip bands. A displacement of this type requires that a Frank–Read source move relative to the surface where slip bands are observed.

In situ nano-compression work [20] in Transmission electron microscopy (TEM) reveals that the deformation of a-Fe at the nanoscale is an inhomogeneous process characterized by a series of short displacement bursts and intermittent large displacement bursts. The series of short bursts correspond to the collective movement of dislocations within the crystal. The large single bursts are from SBs nucleated from the specimen surface. These results suggest that the formation of SBs can be considered as a source-limited plasticity process. The initial plastic deformation is characterized by the multiplication/ movement of a few dislocations over short distances due to the availability of dislocation sources within the nano-blade. Once it has reached a stage at which the mobile dislocations along preferred slips planes have moved through the nano-blade or become entangled in sessile configurations and further dislocation movement is difficult within the crystal, plasticity is carried out by the formation of SBs, which nucleate from the surface [21] and then propagate through the nano-blade.

Frank-Read source FR10G-1.GIF
Frank–Read source

Fisher et al. [18] proposed that SBs are dynamically generated from a Frank–Read source at the specimen surface and are terminated by their own stress field in single crystals. The displacement burst behaviour reported by Kiener and Minor [22] on compressing Cu single crystal nanopillars. Obviously suppressed the progress of serrated yielding (a series of short strain bursts) relative to that without the spinodal nanostructure. The results revealed that during compression deformation, the spinodal nanostructure confined the movement of dislocations (leading to a significant increase in dislocation density), causing a notable strengthening effect, and also kept the slip band morphology planar. [23]

Dislocation activity assists the growth of austenite precipitates and provide quantitative data for revealing the stress field generated by interface migration. [24] The jerky nature of the tip moving rate is probably due to the accumulation and relaxation of stress field near the tip. After leaving from the tip, the dislocation loop expands rapidly ahead of the tip thus the change in tip velocity is concomitant with dislocation emission. It indicates that the emitted dislocation is strongly repelled by the stress field present at the lath tip. When the loop meets the foil surface, it breaks into two dislocation segments that leave a visible trace, due to the presence of a thin oxide layer on the surface. The emission of a dislocation loop from the tip may also affect tip moving rate via interaction between the local dislocation loop and the possible interfacial dislocations in the semi-coherent interface surrounding the tip. consequently, the tip halted temporarily. The net shear stress acting on each dislocation results from a combination of the stress field at the lath tip (τtip), the image stress tending to attract the dislocation loop to the surface (τimage), the line tension (τl) and the interaction stress between dislocations (τinter). This implies the strain field due to the transformation of austenite is large enough to cause the nucleation and emission of dislocations from an austenite lath tip. [2]

Slip bands in the absence of cyclic loading

Slip bands formation Slip bands.jpg
Slip bands formation

While repeatedly reversed loading commonly leads to localisation of dislocation glide, creating linear extrusions and intrusions on a free surface, similar features can arise even if there is no load reversal. These arise from dislocations gliding on a particular slip plane, in a particular slip direction (within a single grain), under an external load. Steps can be created on the free surface as a consequence of the tendency for dislocations to follow one another along a glide path, of which there may be several in parallel with each other in the grain concerned. Prior passage of dislocations apparently makes glide easier for subsequent ones, and the effect may also be associated with dislocation sources, such as a Frank-Read source, acting in particular planes.

Secondary electron images of age-hardened duplex stainless-steel observed in situ in three-point bending at applied crosshead displacements of (a) 1.2 mm and (b) 1.5 mm. Selected regions (2 and 4) are shown with higher magnification in (c) and (d). The apparent slip band height is marked as 'h.' Ferrite (
a
{\displaystyle \alpha }
) and austenite (
g
{\displaystyle \gamma }
) phases are labelled. Slip band DSS.tif
Secondary electron images of age-hardened duplex stainless-steel observed in situ in three-point bending at applied crosshead displacements of (a) 1.2 mm and (b) 1.5 mm. Selected regions (2 and 4) are shown with higher magnification in (c) and (d). The apparent slip band height is marked as ‘h.’ Ferrite () and austenite () phases are labelled.

The appearance of such bands, which are sometimes termed “persistent slip lines”, is similar to that of those arising from cyclic loading, but the resultant steps are usually more localised and have lower heights. They also reveal the grain structure. They can often be seen on free surfaces that were polished before the deformation took place. For example, the figure shows micrographs [25] (taken with different magnifications) of the region around an indent created in a copper sample with a spherical indenter. The parallel lines within individual grains are each the result of several hundred dislocations of the same type reaching the free surface, creating steps with a height of the order of a few microns. If a single slip system was operational within a grain, then there is just one set of lines, but it is common for more than one system to be activated within a grain (particularly when the strain is relatively high), leading to two or more sets of parallel lines. Other features indicative of the details of how the plastic deformation took place, such as a region of cooperative shear caused by deformation twinning, can also sometimes be seen on such surfaces. In the optical micrograph shown, there is also evidence of grain rotations – for example, at the “rim” of the indent and in the form of depressions at grain boundaries. Such images can thus be very informative.

Nature of the non-cyclic slip band local field

Schematic of a slip band, relative to the measurement axes (x1, x2, and x3), and axes related to the slip-band (x, y, and z), showing the angles that describe the relationship between these axes and the traces of the slip band (, ), and the inclination angle () of the slip trace (x) and Burgers vector (b) relative to the surface. h is the slip band height, and q is the slip band propagation direction assumed for J-integral calculation when using the virtual extension method. t describes the line vector drawn here as for an edge dislocation, i.e., b[?]t, and n is the slip band plane normal. Schematic of a slip band.tif
Schematic of a slip band, relative to the measurement axes (𝑥1, 𝑥2, and 𝑥3), and axes related to the slip-band (𝑥, 𝑦, and 𝑧), showing the angles that describe the relationship between these axes and the traces of the slip band (𝛼, 𝜃), and the inclination angle (𝜓) of the slip trace (𝑥) and Burgers vector (𝑏) relative to the surface. ℎ is the slip band height, and 𝑞 is the slip band propagation direction assumed for J-integral calculation when using the virtual extension method. 𝑡 describes the line vector drawn here as for an edge dislocation, i.e., 𝑏⊥𝑡, and 𝑛 is the slip band plane normal.

The deformation field at the slip-band is due to three-dimensional elastic and plastic strains where the concentrated shear of the slip band tip deforms the grain in its vicinity. The elastic strains describe the stress concentration ahead of the slip band, which is important as it can affect the transfer of plastic deformation across grain boundaries. [27] [28] [29] An understanding of this is needed to support the study of yield and inter/intra-granular fracture. [30] [31] [32] The concentrated shear of slip bands can also nucleate cracks in the plane of the slip band, [16] [17] and persistent slip bands that lead to intragranular fatigue crack initiation and growth may also form under cyclic loading conditions. [5] [33] To properly characterise slip bands and validate mechanistic models for their interactions with microstructure, it is crucial to quantify the local deformation fields associated with their propagation. However, little attention has been given to slip bands within grains (i.e., in the absence of grain boundary interaction).

The long-range stress field (i.e., the elastic strain field) around the tip of a stress concentrator, such as a slip band, can be considered a singularity equivalent to that of a crack. [34] [35] This singularity can be quantified using a path independent integral since it satisfies the conservation laws of elasticity. The conservation laws of elasticity related to translational, rotational, and scaling symmetries were derived initially by Knowles and Sternberg [36] from the Noether's theorem. [37] Budiansky and Rice [38] introduced the J-, M-, L-integral and were the first to give them a physical interpretation as the strain energy-release rates for mechanisms such as cavity propagation, simultaneous uniform expansion, and defect rotation, respectively. When evaluated over a surface that encloses a defect, these conservation integrals represent a configurational force on the defect. [39] That work paved the way for the field of Configurational mechanics of materials, with the path-independent J-integral now widely used to analyse the configurational forces in problems as diverse as dislocation dynamics, [40] [41] misfitting inclusions, [42] propagation of cracks, [43] shear deformation of clays, [44] and co-planar dislocation nucleation from shear loaded cracks. [45] The integrals have been applied to linear elastic and elastic-plastic materials and have been coupled with processes such as thermal [46] and electrochemical [47] loading, and internal tractions. [48] Recently, experimental fracture mechanics studies have used full-field in situ measurements of displacements [49] [50] and elastic strains [51] [50] to evaluate the local deformation field surrounding the crack tip as a J-integral.

Slip bands form due to plastic deformation, and the analysis of the force on a dislocation considers the two-dimensional nature of the dislocation line defect. General definitions of the Peach–Koehler configurational force (𝑃𝑘𝑗) [52] (or the elastic energy-momentum tensor [53] ) on a dislocation in the arbitrary 𝑥1, 𝑥2, 𝑥3 coordinate system, decompose the Burgers vector (𝑏) to orthogonal components. This leads to the generalised definition of the J-integral in equations below. For a dislocation pile-up, the J-integral is the summation of the Peach–Koehler configurational force of the dislocations in the pile-up (including out-of-plane, 𝑏3 [54] ).

𝐽𝑘 = ∫ 𝑃𝑘𝑗 𝑛𝑗 𝑑𝑆 = ∫(𝑊𝑠 𝑛𝑘− 𝑇𝑖 𝑢𝑖,𝑘) 𝑑𝑆

𝐽𝑘𝑥 = 𝑅𝑘𝑗 𝐽𝑗, 𝑖,𝑗,𝑘=1,2,3

where 𝑆 is an arbitrary contour around the dislocation pile-up with unit outward normal 𝑛𝑖, 𝑊𝑠 is the strain energy density, 𝑇𝑖 = 𝜎𝑖𝑗 𝑛𝑗 is the traction on 𝑑𝑆, 𝑢𝑖 are the displacement vector components, 𝐽𝑘𝑥 is 𝐽-integral evaluated along the 𝑥𝑘 direction, and 𝑅𝑘𝑗 is a second-order mapping tensor that maps 𝐽𝑘 into 𝑥𝑘 direction. This vectorial 𝐽𝑘-integral leads to numerical difficulties in the analysis since 𝐽2 and, for a three-dimensional slip band or inclined crack, the 𝐽3 terms cannot be neglected. [1]

See also

Related Research Articles

<span class="mw-page-title-main">Ductility</span> Degree to which a material under stress irreversibly deforms before failure

Ductility refers to the ability of a material to sustain significant plastic deformation before fracture. Plastic deformation is the permanent distortion of a material under applied stress, as opposed to elastic deformation, which is reversible upon removing the stress. Ductility is a critical mechanical performance indicator, particularly in applications that require materials to bend, stretch, or deform in other ways without breaking. The extent of ductility can be quantitatively assessed using the percent elongation at break, given by the equation:

<span class="mw-page-title-main">Plasticity (physics)</span> Non-reversible deformation of a solid material in response to applied forces

In physics and materials science, plasticity is the ability of a solid material to undergo permanent deformation, a non-reversible change of shape in response to applied forces. For example, a solid piece of metal being bent or pounded into a new shape displays plasticity as permanent changes occur within the material itself. In engineering, the transition from elastic behavior to plastic behavior is known as yielding.

<span class="mw-page-title-main">Fatigue (material)</span> Initiation and propagation of cracks in a material due to cyclic loading

In materials science, fatigue is the initiation and propagation of cracks in a material due to cyclic loading. Once a fatigue crack has initiated, it grows a small amount with each loading cycle, typically producing striations on some parts of the fracture surface. The crack will continue to grow until it reaches a critical size, which occurs when the stress intensity factor of the crack exceeds the fracture toughness of the material, producing rapid propagation and typically complete fracture of the structure.

<span class="mw-page-title-main">Creep (deformation)</span> Tendency of a solid material to move slowly or deform permanently under mechanical stress

In materials science, creep is the tendency of a solid material to undergo slow deformation while subject to persistent mechanical stresses. It can occur as a result of long-term exposure to high levels of stress that are still below the yield strength of the material. Creep is more severe in materials that are subjected to heat for long periods and generally increases as they near their melting point.

<span class="mw-page-title-main">Dislocation</span> Linear crystallographic defect or irregularity

In materials science, a dislocation or Taylor's dislocation is a linear crystallographic defect or irregularity within a crystal structure that contains an abrupt change in the arrangement of atoms. The movement of dislocations allow atoms to slide over each other at low stress levels and is known as glide or slip. The crystalline order is restored on either side of a glide dislocation but the atoms on one side have moved by one position. The crystalline order is not fully restored with a partial dislocation. A dislocation defines the boundary between slipped and unslipped regions of material and as a result, must either form a complete loop, intersect other dislocations or defects, or extend to the edges of the crystal. A dislocation can be characterised by the distance and direction of movement it causes to atoms which is defined by the Burgers vector. Plastic deformation of a material occurs by the creation and movement of many dislocations. The number and arrangement of dislocations influences many of the properties of materials.

<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">Crystal twinning</span> Two separate crystals sharing some of the same crystal lattice points in a symmetrical manner

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.

<span class="mw-page-title-main">Lüders band</span> Bands of plastic deformation in metals

Lüders bands are a type of slip band or stretcher-strain mark which are formed due to localized bands of plastic deformation in metals experiencing tensile stresses, common to low-carbon steels and certain Al-Mg alloys. First reported by Guillaume Piobert, and later by W. Lüders, the mechanism that stimulates their appearance is known as dynamic strain aging, or the inhibition of dislocation motion by interstitial atoms, around which "atmospheres" or "zones" naturally congregate.

<span class="mw-page-title-main">Bauschinger effect</span>

The Bauschinger effect refers to a property of materials where the material's stress/strain characteristics change as a result of the microscopic stress distribution of the material. For example, an increase in tensile yield strength occurs at the expense of compressive yield strength. The effect is named after German engineer Johann Bauschinger.

<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.

In materials science, hardness is a measure of the resistance to localized plastic deformation, such as an indentation or a scratch (linear), induced mechanically either by pressing or abrasion. In general, different materials differ in their hardness; for example hard metals such as titanium and beryllium are harder than soft metals such as sodium and metallic tin, or wood and common plastics. Macroscopic hardness is generally characterized by strong intermolecular bonds, but the behavior of solid materials under force is complex; therefore, hardness can be measured in different ways, such as scratch hardness, indentation hardness, and rebound hardness. Hardness is dependent on ductility, elastic stiffness, plasticity, strain, strength, toughness, viscoelasticity, and viscosity. Common examples of hard matter are ceramics, concrete, certain metals, and superhard materials, which can be contrasted with soft matter.

<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.

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.

Polymer fracture is the study of the fracture surface of an already failed material to determine the method of crack formation and extension in polymers both fiber reinforced and otherwise. Failure in polymer components can occur at relatively low stress levels, far below the tensile strength because of four major reasons: long term stress or creep rupture, cyclic stresses or fatigue, the presence of structural flaws and stress-cracking agents. Formations of submicroscopic cracks in polymers under load have been studied by x ray scattering techniques and the main regularities of crack formation under different loading conditions have been analyzed. The low strength of polymers compared to theoretically predicted values are mainly due to the many microscopic imperfections found in the material. These defects namely dislocations, crystalline boundaries, amorphous interlayers and block structure can all lead to the non-uniform distribution of mechanical stress.

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.

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.

<span class="mw-page-title-main">Striation (fatigue)</span>

Striations are marks produced on the fracture surface that show the incremental growth of a fatigue crack. A striation marks the position of the crack tip at the time it was made. The term striation generally refers to ductile striations which are rounded bands on the fracture surface separated by depressions or fissures and can have the same appearance on both sides of the mating surfaces of the fatigue crack. Although some research has suggested that many loading cycles are required to form a single striation, it is now generally thought that each striation is the result of a single loading cycle.

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">475 °C embrittlement</span> Loss of plasticity in ferritic stainless steel

Duplex stainless steels are a family of alloys with a two-phase microstructure consisting of both austenitic and ferritic phases. They offer excellent mechanical properties, corrosion resistance, and toughness compared to other types of stainless steel. However, duplex stainless steel can be susceptible to a phenomenon known as 475 °C (887 °F) embrittlement or duplex stainless steel age hardening, which is a type of aging process that causes loss of plasticity in duplex stainless steel when it is heated in the range of 250 to 550 °C. At this temperature range, spontaneous phase separation of the ferrite phase into iron-rich and chromium-rich nanophases occurs, with no change in the mechanical properties of the austenite phase. This type of embrittlement is due to precipitation hardening, which makes the material become brittle and prone to cracking.

References

  1. 1 2 3 4 Koko, Abdalrhaman; Elmukashfi, Elsiddig; Becker, Thorsten H.; Karamched, Phani S.; Wilkinson, Angus J.; Marrow, T. James (2022-10-15). "In situ characterisation of the strain fields of intragranular slip bands in ferrite by high-resolution electron backscatter diffraction". Acta Materialia. 239: 118284. Bibcode:2022AcMat.23918284K. doi: 10.1016/j.actamat.2022.118284 . ISSN   1359-6454. S2CID   251783802. Creative Commons by small.svg  This article incorporates textfrom this source, which is available under the CC BY 4.0 license.
  2. 1 2 Koko, A. Mohamed (2022). In situ full-field characterisation of strain concentrations (deformation twins, slip bands and cracks) (PhD thesis). University of Oxford. Archived from the original on 2023-02-01. Retrieved 2023-03-02. Creative Commons by small.svg  This article incorporates textfrom this source, which is available under the CC BY 4.0 license.
  3. Smallman, R. E.; Ngan, A. H. W. (2014-01-01), Smallman, R. E.; Ngan, A. H. W. (eds.), "Chapter 9 - Plastic Deformation and Dislocation Behaviour", Modern Physical Metallurgy (Eighth Edition), Oxford: Butterworth-Heinemann, pp. 357–414, doi:10.1016/b978-0-08-098204-5.00009-2, ISBN   978-0-08-098204-5, archived from the original on 2022-10-04, retrieved 2022-10-04
  4. Sangid, Michael D. (2013-12-01). "The physics of fatigue crack initiation". International Journal of Fatigue. Fatigue and Microstructure: A special issue on recent advances. 57: 58–72. doi:10.1016/j.ijfatigue.2012.10.009. ISSN   0142-1123.
  5. 1 2 Lukáš, P.; Klesnil, M.; Krejčí, J. (1968). "Dislocations and Persistent Slip Bands in Copper Single Crystals Fatigued at Low Stress Amplitude". Physica Status Solidi B (in German). 27 (2): 545–558. Bibcode:1968PSSBR..27..545L. doi:10.1002/pssb.19680270212. S2CID   96586802. Archived from the original on 2022-10-03. Retrieved 2022-10-03.
  6. Schiller, C.; Walgraef, D. (1988-03-01). "Numerical simulation of persistent slip band formation". Acta Metallurgica. 36 (3): 563–574. doi:10.1016/0001-6160(88)90089-2. ISSN   0001-6160. Archived from the original on 2023-03-25. Retrieved 2023-03-25.
  7. Erel, Can; Po, Giacomo; Crosby, Tamer; Ghoniem, Nasr (December 2017). "Generation and interaction mechanisms of prismatic dislocation loops in FCC metals". Computational Materials Science. 140: 32–46. doi: 10.1016/j.commatsci.2017.07.043 .
  8. "grain boundray, dislocation simulation, discrete dislocation dynamics, dislocation dynamics, grain boundary, dislocation cell, cross slip, creep, deformation, strain hardening, superalloy, climb". www.dierk-raabe.com (in German). Archived from the original on 2022-12-03. Retrieved 2023-03-05.
  9. Differt, K.; Essmann, U. (1993), "Dynamical model of the wall structure in persistent slip bands of fatigued metals", Fundamental Aspects of Dislocation Interactions, Elsevier, pp. 295–299, doi:10.1016/b978-1-4832-2815-0.50048-6, ISBN   978-1-4832-2815-0, archived from the original on 2018-06-30, retrieved 2022-10-03
  10. 1 2 Verdier, M; Fivel, M; Groma, I (1998-11-01). "Mesoscopic scale simulation of dislocation dynamics in fcc metals: Principles and applications". Modelling and Simulation in Materials Science and Engineering. 6 (6): 755–770. Bibcode:1998MSMSE...6..755V. doi:10.1088/0965-0393/6/6/007. ISSN   0965-0393. S2CID   250889422. Archived from the original on 2022-10-03. Retrieved 2022-10-03.
  11. Déprés, C.; Robertson, C. F.; Fivel, M. C. (January 2006). "Low-strain fatigue in 316L steel surface grains: a three dimension discrete dislocation dynamics modelling of the early cycles. Part 2: Persistent slip markings and micro-crack nucleation". Philosophical Magazine. 86 (1): 79–97. Bibcode:2006PMag...86...79D. doi:10.1080/14786430500341250. ISSN   1478-6435. S2CID   135953582. Archived from the original on 2021-05-17. Retrieved 2022-10-03.
  12. Déprés, C.; Robertson, C. F.; Fivel, M. C. (August 2004). "Low-strain fatigue in AISI 316L steel surface grains: a three-dimensional discrete dislocation dynamics modelling of the early cycles I. Dislocation microstructures and mechanical behaviour". Philosophical Magazine. 84 (22): 2257–2275. Bibcode:2004PMag...84.2257D. doi:10.1080/14786430410001690051. ISSN   1478-6435. S2CID   137329770. Archived from the original on 2022-10-07. Retrieved 2022-10-03.
  13. Man, J.; Obrtlík, K.; Polák, J. (June 2009). "Extrusions and intrusions in fatigued metals. Part 1. State of the art and history†". Philosophical Magazine. 89 (16): 1295–1336. Bibcode:2009PMag...89.1295M. doi:10.1080/14786430902917616. ISSN   1478-6435. S2CID   136919859. Archived from the original on 2022-10-06. Retrieved 2022-10-03.
  14. Amodeo, R. J.; Ghoniem, N. M. (1990-04-01). "Dislocation dynamics. II. Applications to the formation of persistent slip bands, planar arrays, and dislocation cells". Physical Review B. 41 (10): 6968–6976. Bibcode:1990PhRvB..41.6968A. doi:10.1103/PhysRevB.41.6968. ISSN   0163-1829. PMID   9992953. Archived from the original on 2023-03-25. Retrieved 2022-10-03.
  15. Wood, W. A. (July 1958). "Formation of fatigue cracks". Philosophical Magazine. 3 (31): 692–699. Bibcode:1958PMag....3..692W. doi:10.1080/14786435808237004. ISSN   0031-8086. Archived from the original on 2022-10-07. Retrieved 2022-10-03.
  16. 1 2 Koss, D.A.; Chan, K.S. (September 1980). "Fracture along planar slip bands". Acta Metallurgica. 28 (9): 1245–1252. doi:10.1016/0001-6160(80)90080-2. Archived from the original on 2018-06-28. Retrieved 2022-10-03.
  17. 1 2 Mughrabi, H. (September 1983). "Dislocation wall and cell structures and long-range internal stresses in deformed metal crystals". Acta Metallurgica. 31 (9): 1367–1379. doi:10.1016/0001-6160(83)90007-X. Archived from the original on 2022-06-16. Retrieved 2022-10-03.
  18. 1 2 Fisher, John C.; Hart, Edward W.; Pry, Robert H. (1952-09-15). "Theory of Slip-Band Formation". Physical Review. 87 (6): 958–961. Bibcode:1952PhRv...87..958F. doi:10.1103/PhysRev.87.958. ISSN   0031-899X.
  19. Frank, F. C.; Read, W. T. (1950-08-15). "Multiplication Processes for Slow Moving Dislocations". Physical Review. 79 (4): 722–723. Bibcode:1950PhRv...79..722F. doi:10.1103/PhysRev.79.722. ISSN   0031-899X. Archived from the original on 2023-03-25. Retrieved 2022-10-03.
  20. Xie, Kelvin Y.; Wang, Yanbo; Ni, Song; Liao, Xiaozhou; Cairney, Julie M.; Ringer, Simon P. (December 2011). "Insight into the deformation mechanisms of α-Fe at the nanoscale". Scripta Materialia. 65 (12): 1037–1040. doi:10.1016/j.scriptamat.2011.08.023. Archived from the original on 2018-06-02. Retrieved 2022-10-03.
  21. Zheng, He; Cao, Ajing; Weinberger, Christopher R.; Huang, Jian Yu; Du, Kui; Wang, Jianbo; Ma, Yanyun; Xia, Younan; Mao, Scott X. (December 2010). "Discrete plasticity in sub-10-nm-sized gold crystals". Nature Communications. 1 (1): 144. Bibcode:2010NatCo...1..144Z. doi:10.1038/ncomms1149. ISSN   2041-1723. PMC   3105591 . PMID   21266994.
  22. Kiener, D.; Minor, A.M. (February 2011). "Source-controlled yield and hardening of Cu(100) studied by in situ transmission electron microscopy". Acta Materialia. 59 (4): 1328–1337. Bibcode:2011AcMat..59.1328K. doi:10.1016/j.actamat.2010.10.065. Archived from the original on 2022-06-17. Retrieved 2022-10-03.
  23. Hsieh, Yi-Chieh; Zhang, Ling; Chung, Tsai-Fu; Tsai, Yu-Ting; Yang, Jer-Ren; Ohmura, Takahito; Suzuki, Takuya (December 2016). "In-situ transmission electron microscopy investigation of the deformation behavior of spinodal nanostructured δ-ferrite in a duplex stainless steel". Scripta Materialia. 125: 44–48. doi:10.1016/j.scriptamat.2016.06.047. Archived from the original on 2022-06-20. Retrieved 2022-10-03.
  24. Du, Juan; Mompiou, Frédéric; Zhang, Wen-Zheng (March 2018). "In-situ TEM study of dislocation emission associated with austenite growth". Scripta Materialia. 145: 62–66. doi:10.1016/j.scriptamat.2017.10.014. Archived from the original on 2018-06-30. Retrieved 2022-10-03.
  25. Campbell, JE; Thompson, RP; Dean, J; Clyne, TW (2019). "Comparison between stress-strain plots obtained from indentation plastometry, based on residual indent profiles, and from uniaxial testing". Acta Materialia. 168: 87–99. Bibcode:2019AcMat.168...87C. doi:10.1016/j.actamat.2019.02.006.
  26. Parks, D.M. (December 1977). "The virtual crack extension method for nonlinear material behavior". Computer Methods in Applied Mechanics and Engineering. 12 (3): 353–364. Bibcode:1977CMAME..12..353P. doi:10.1016/0045-7825(77)90023-8. Archived from the original on 2022-06-29. Retrieved 2022-10-03.
  27. Benjamin Britton, T.; Wilkinson, Angus J. (September 2012). "Stress fields and geometrically necessary dislocation density distributions near the head of a blocked slip band". Acta Materialia. 60 (16): 5773–5782. Bibcode:2012AcMat..60.5773B. doi:10.1016/j.actamat.2012.07.004. hdl: 10044/1/13886 . Archived from the original on 2023-01-01. Retrieved 2022-10-03.
  28. Guo, Y.; Britton, T.B.; Wilkinson, A.J. (September 2014). "Slip band–grain boundary interactions in commercial-purity titanium". Acta Materialia. 76: 1–12. Bibcode:2014AcMat..76....1G. doi:10.1016/j.actamat.2014.05.015. hdl: 10044/1/25973 . S2CID   136904692. Archived from the original on 2022-06-20. Retrieved 2022-10-03.
  29. Andani, Mohsen Taheri; Lakshmanan, Aaditya; Sundararaghavan, Veera; Allison, John; Misra, Amit (November 2020). "Quantitative study of the effect of grain boundary parameters on the slip system level Hall-Petch slope for basal slip system in Mg-4Al". Acta Materialia. 200: 148–161. Bibcode:2020AcMat.200..148A. doi: 10.1016/j.actamat.2020.08.079 . S2CID   225279335.
  30. Livingston, J.D; Chalmers, B (June 1957). "Multiple slip in bicrystal deformation". Acta Metallurgica. 5 (6): 322–327. doi:10.1016/0001-6160(57)90044-5. Archived from the original on 2022-06-17. Retrieved 2022-10-03.
  31. Lee, T.C.; Robertson, I.M.; Birnbaum, H.K. (May 1989). "Prediction of slip transfer mechanisms across grain boundaries". Scripta Metallurgica. 23 (5): 799–803. doi:10.1016/0036-9748(89)90534-6. Archived from the original on 2022-07-10. Retrieved 2022-10-03.
  32. Luster, J.; Morris, M. A. (July 1995). "Compatibility of deformation in two-phase Ti-Al alloys: Dependence on microstructure and orientation relationships". Metallurgical and Materials Transactions A. 26 (7): 1745–1756. Bibcode:1995MMTA...26.1745L. doi:10.1007/BF02670762. ISSN   1073-5623. S2CID   137425735. Archived from the original on 2023-03-25. Retrieved 2022-10-03.
  33. Tu, S.-T.; Zhang, X.-C. (2016), "Fatigue Crack Initiation Mechanisms", Reference Module in Materials Science and Materials Engineering, Elsevier, pp. B9780128035818028526, doi:10.1016/b978-0-12-803581-8.02852-6, ISBN   978-0-12-803581-8, archived from the original on 2022-07-11, retrieved 2022-10-03
  34. Makin, M. J. (April 1970). "The mechanism of slip band growth in irradiated crystals". Philosophical Magazine. 21 (172): 815–817. Bibcode:1970PMag...21..815M. doi:10.1080/14786437008238467. ISSN   0031-8086. Archived from the original on 2022-10-07. Retrieved 2022-10-03.
  35. Rice, James R. (December 1987). "Tensile crack tip fields in elastic-ideally plastic crystals". Mechanics of Materials. 6 (4): 317–335. Bibcode:1987MechM...6..317R. doi:10.1016/0167-6636(87)90030-5. Archived from the original on 2022-12-05. Retrieved 2022-10-03.
  36. Knowles, J. K.; Sternberg, Eli (January 1972). "On a class of conservation laws in linearized and finite elastostatics". Archive for Rational Mechanics and Analysis. 44 (3): 187–211. Bibcode:1972ArRMA..44..187K. doi:10.1007/BF00250778. ISSN   0003-9527. S2CID   122386163. Archived from the original on 2023-03-25. Retrieved 2022-10-03.
  37. Noether, Emmy (January 1971). "Invariant variation problems". Transport Theory and Statistical Physics. 1 (3): 186–207. arXiv: physics/0503066 . Bibcode:1971TTSP....1..186N. doi:10.1080/00411457108231446. ISSN   0041-1450. S2CID   119019843. Archived from the original on 2022-07-15. Retrieved 2022-10-03.
  38. Budiansky, B.; Rice, J. R. (1973-03-01). "Conservation Laws and Energy-Release Rates". Journal of Applied Mechanics. 40 (1): 201–203. Bibcode:1973JAM....40..201B. doi:10.1115/1.3422926. ISSN   0021-8936. Archived from the original on 2022-10-03. Retrieved 2022-10-03.
  39. Eshelby, J. D. (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. ISSN   0080-4614. S2CID   14703976. Archived from the original on 2022-10-03. Retrieved 2022-10-03.
  40. Agiasofitou, Eleni; Lazar, Markus (May 2017). "Micromechanics of dislocations in solids: J -, M -, and L -integrals and their fundamental relations". International Journal of Engineering Science. 114: 16–40. arXiv: 1702.00363 . doi:10.1016/j.ijengsci.2017.02.001. S2CID   119531197. Archived from the original on 2022-08-15. Retrieved 2022-10-03.
  41. Kim, Hokun; Kim, Soon; Kim, Sung Youb (March 2021). "Lattice-based J integral for a steadily moving dislocation". International Journal of Plasticity . 138: 102949. doi:10.1016/j.ijplas.2021.102949. S2CID   233799154. Archived from the original on 2023-03-25. Retrieved 2022-10-03.
  42. Markenscoff, Xanthippi; Ni, Luqun (January 2010). "The energy-release rate and "self-force" of dynamically expanding spherical and plane inclusion boundaries with dilatational eigenstrain". Journal of the Mechanics and Physics of Solids. 58 (1): 1–11. Bibcode:2010JMPSo..58....1M. doi:10.1016/j.jmps.2009.10.001. Archived from the original on 2022-02-21. Retrieved 2022-10-03.
  43. Rice, Jr; Drugan, WJ; Sham, T-L (1980-01-01), Paris, Pc (ed.), "Elastic-Plastic Analysis of Growing Cracks", Fracture Mechanics, 100 Barr Harbor Drive, PO Box C700, West Conshohocken, PA 19428-2959: ASTM International, pp. 189–189–33, doi:10.1520/stp36972s, ISBN   978-0-8031-0363-4, archived from the original on 2018-06-02, retrieved 2022-10-03{{citation}}: CS1 maint: location (link)
  44. Palmer, A. C.; Rice, J. R. (1973-04-03). "The growth of slip surfaces in the progressive failure of over-consolidated clay". Proceedings of the Royal Society of London. A. Mathematical and Physical Sciences. 332 (1591): 527–548. Bibcode:1973RSPSA.332..527P. doi:10.1098/rspa.1973.0040. ISSN   0080-4630. S2CID   7935986. Archived from the original on 2022-10-03. Retrieved 2022-10-03.
  45. Rice, James R. (January 1992). "Dislocation nucleation from a crack tip: An analysis based on the Peierls concept". Journal of the Mechanics and Physics of Solids. 40 (2): 239–271. Bibcode:1992JMPSo..40..239R. doi:10.1016/S0022-5096(05)80012-2. Archived from the original on 2022-07-16. Retrieved 2022-10-03.
  46. Kc, Amit; Kim, Jeong-Ho (May 2008). "Interaction integrals for thermal fracture of functionally graded materials". Engineering Fracture Mechanics. 75 (8): 2542–2565. doi:10.1016/j.engfracmech.2007.07.011. Archived from the original on 2022-08-07. Retrieved 2022-10-03.
  47. Haftbaradaran, Hamed; Qu, Jianmin (November 2014). "A path-independent integral for fracture of solids under combined electrochemical and mechanical loadings". Journal of the Mechanics and Physics of Solids. 71: 1–14. Bibcode:2014JMPSo..71....1H. doi:10.1016/j.jmps.2014.06.007. Archived from the original on 2023-03-25. Retrieved 2022-10-03.
  48. Walters, Matthew C.; Paulino, Glaucio H.; Dodds, Robert H. (July 2005). "Interaction integral procedures for 3-D curved cracks including surface tractions". Engineering Fracture Mechanics. 72 (11): 1635–1663. doi:10.1016/j.engfracmech.2005.01.002. Archived from the original on 2022-08-08. Retrieved 2022-10-03.
  49. Becker, T. H.; Mostafavi, M.; Tait, R. B.; Marrow, T. J. (October 2012). "An approach to calculate the J-integral by digital image correlation displacement field measurement: An Approach to Calculate the J -Integral Using Digital Image Correlation". Fatigue & Fracture of Engineering Materials & Structures. 35 (10): 971–984. doi:10.1111/j.1460-2695.2012.01685.x. Archived from the original on 2022-10-03. Retrieved 2022-10-03.
  50. 1 2 Koko, A.; Earp, P.; Wigger, T.; Tong, J.; Marrow, T.J. (May 2020). "J-integral analysis: An EDXD and DIC comparative study for a fatigue crack". International Journal of Fatigue. 134: 105474. doi:10.1016/j.ijfatigue.2020.105474. S2CID   214391445. Archived from the original on 2022-06-22. Retrieved 2022-10-03.
  51. Barhli, S.M.; Saucedo-Mora, L.; Jordan, M.S.L.; Cinar, A.F.; Reinhard, C.; Mostafavi, M.; Marrow, T.J. (November 2017). "Synchrotron X-ray characterization of crack strain fields in polygranular graphite". Carbon. 124: 357–371. Bibcode:2017Carbo.124..357B. doi:10.1016/j.carbon.2017.08.075. hdl: 1983/93385e0e-155e-49e6-9a3d-e4f881de013a . Archived from the original on 2022-06-18. Retrieved 2022-10-03.
  52. Lubarda, Vlado A. (2019-01-01). "Dislocation Burgers vector and the Peach–Koehler force: a review". Journal of Materials Research and Technology. 8 (1): 1550–1565. doi: 10.1016/j.jmrt.2018.08.014 . ISSN   2238-7854. S2CID   125242265.
  53. Eshelby, J.D. (1956), The Continuum Theory of Lattice Defects, Solid State Physics, vol. 3, Elsevier, pp. 79–144, doi:10.1016/s0081-1947(08)60132-0, ISBN   978-0-12-607703-2, archived from the original on 2022-12-19, retrieved 2022-10-03
  54. Anderson, PM; Hirth, JP; Lothe, J. Theory of Straight Dislocations. Theory of dislocations.

Further reading