Fact-checked by Grok 2 weeks ago

Stacking fault

A stacking fault is a planar defect in the atomic structure of a crystalline , characterized by an interruption in the regular sequence of closely packed atomic planes. These defects commonly occur in materials with close-packed lattices, such as face-centered cubic (FCC) crystals where the ideal stacking sequence is ABCABC... along the {111} planes, or hexagonal close-packed (HCP) structures with an ABAB... arrangement. Stacking faults are classified into two primary types: intrinsic and extrinsic. An intrinsic stacking fault arises from the removal or collapse of a single atomic plane, effectively shifting the sequence (e.g., from ABCABC... to ABCABABC...), and is often associated with vacancy clustering or processes. In contrast, an extrinsic stacking fault involves the insertion of an additional plane, creating a sequence like ABCABACABC..., typically due to atoms or effects. Both types are bounded by partial dislocations, such as Frank partials with a Burgers vector of a/3<111> or Shockley partials with a/6<112>, which can dissociate from perfect dislocations in FCC metals. The presence of stacking faults influences key material properties, including mechanical behavior and deformation mechanisms. The stacking fault energy (SFE), typically on the order of a few hundred mJ/m² (e.g., 51.74 mJ/m² in ), quantifies the cost of the fault and determines the width of the fault ribbon between partial ; lower SFE values promote wider ribbons, enhanced twinning, and reduced dislocation mobility, affecting , hardening, and thermal conductivity in alloys. In nanoscale structures, stacking faults act as barriers to dislocation motion, playing a critical role in strengthening and development.

Fundamentals

Definition and Types

A stacking fault is a planar defect in crystalline materials in which the regular stacking sequence of atomic s is locally interrupted, resulting in a disruption of the periodic order in close-packed structures such as face-centered cubic (FCC) or hexagonal close-packed (HCP) lattices. This interruption typically spans a two-dimensional , distinguishing stacking faults from point defects (zero-dimensional, affecting single atomic sites), line defects like dislocations (one-dimensional, extending linearly), and volume defects (three-dimensional, involving larger regions of misfit). The defect arises in modular or layered crystal structures, where atoms are arranged in repeating close-packed layers, and it often bounds regions of altered symmetry without altering the overall lattice orientation beyond the fault . Stacking faults are broadly classified into intrinsic and extrinsic types, depending on whether the fault corresponds to the removal or insertion of atomic layers relative to the ideal sequence. Intrinsic stacking faults form by the removal (or effective absence) of a single atomic plane in the stacking sequence, creating a local shift that maintains order on either side of the fault but repeats a layer position. In FCC crystals, with the normal stacking sequence ABCABC..., an intrinsic fault disrupts this to ABCBCABC..., where the layer following C is B instead of the expected A, effectively introducing a hexagonal-like segment (BCB) amid the cubic order. This type of fault is commonly bounded by Shockley partial dislocations and carries relatively low energy, influencing local atomic rearrangements. Extrinsic stacking faults, in contrast, arise from the insertion of an extra atomic plane or, equivalently, a combination of an intrinsic fault and a sessile loop, resulting in a more complex double-layer disruption. In FCC crystals, this produces a such as ABCACBA..., where two consecutive deviations occur (e.g., after ABC, the next layers are ACB instead of ABC or CBA), often forming a thin of twinned or hcp-like stacking. Extrinsic faults are less common than intrinsic ones due to their higher associated energy and are typically observed in materials under specific deformation conditions. The recognition of stacking faults as distinct defects emerged in the , alongside foundational work on dislocations during investigations of plastic deformation in metals, providing key insights into mechanisms.

Stacking Sequences in Common Crystal Structures

In close-packed crystal structures, atoms are arranged in layers of hexagonal packing, where each layer occupies one of three possible positions labeled A, B, or C, depending on the lateral shift relative to the layer below. These positions ensure maximal atomic density, with atoms touching their nearest neighbors. The overall three-dimensional structure arises from the repeating sequence of these layers, which determines the crystal's symmetry and properties. The face-centered cubic (FCC) structure features a stacking sequence of ABCABC..., where each successive close-packed plane is shifted to occupy the depressions not used by the previous two layers. This repeating pattern every three layers results in a cubic with high , equivalent to the cubic close-packed (CCP) arrangement. Atoms in the A position align directly above those in the first layer, B in the second, and C in the third, before restarting the cycle. In contrast, the hexagonal close-packed (HCP) structure follows an ABABAB... stacking sequence, where the third layer returns to the A position, aligning directly above the first layer's atoms. This alternation produces a hexagonal with lower than FCC, though both structures achieve the same packing efficiency of 74%. The HCP configuration corresponds to a local minimum for many metals, differing from FCC primarily in the periodic repetition that breaks the cubic . Body-centered cubic (BCC) structures lack distinct close-packed planes, as atoms are coordinated in a less dense arrangement with eight nearest neighbors at varying distances. Stacking faults in BCC are thus less prevalent than in close-packed lattices and typically manifest on {112} planes rather than basal planes. Other non-close-packed structures, such as simple cubic or diamond cubic, exhibit even more irregular layer arrangements, making traditional stacking fault concepts less applicable. These ideal stacking sequences provide the foundational patterns against which disruptions—such as those leading to intrinsic or extrinsic faults—can be understood in subsequent discussions of defects. To visualize them, standard notation uses A, B, and C to denote the threefold positions in the hexagonal layer net:
  • FCC: Planes stacked as A (layer 1), B (layer 2), C (layer 3), A (layer 4), etc.
  • HCP: Planes stacked as A (layer 1), B (layer 2), A (layer 3), B (layer 4), etc.
This notation highlights how minor changes in repetition create distinct lattices.

Formation and Energetics

Mechanisms of Formation

Stacking faults in face-centered cubic (FCC) crystals primarily form through the dissociation of perfect dislocations into partial dislocations during plastic deformation. A perfect dislocation with Burgers vector \mathbf{b} = \frac{a}{2}\langle 110 \rangle, where a is the lattice parameter, dissociates into two Shockley partial dislocations with Burgers vectors \mathbf{b}_1 = \frac{a}{6}\langle 121 \rangle and \mathbf{b}_2 = \frac{a}{6}\langle 21\overline{1} \rangle, satisfying the relation \mathbf{b} = \mathbf{b}_1 + \mathbf{b}_2. These partials glide on the same {111} plane, separated by a stacking fault ribbon where the local stacking sequence is disrupted, such as from ABCABC... to ABCACABC.... This dissociation reduces the total elastic energy of the system, as the shorter Burgers vectors of the partials lower the strain field interaction, though it is balanced by the creation of the faulted region. The separation distance between the partial dislocations, known as the equilibrium width of the stacking fault ribbon, depends on the balance of repulsive forces and the attractive force from the stacking fault, but the initial of the fault occurs dynamically under applied . Shear stresses during deformation provide the driving force for dislocation motion, enabling the splitting , particularly in metals with low stacking fault where wider ribbons are stable. Higher temperatures further promote this process through thermal activation, which lowers the barrier for and facilitates the of partial dislocations, as observed in or stages of deformation. Stacking faults can also nucleate during crystal processes, such as vapor deposition or solidification, due to irregular atomic attachment at the growth interface. In physical vapor transport of materials like (), deviations in the layer-by-layer addition of atoms—often influenced by doping levels or surface —lead to localized disruptions in the stacking sequence, forming faulted . For instance, heavy doping in 4H-SiC boules induces surface hillocks on the (000\overline{1}) facet, which act as sites for basal plane stacking faults during at elevated temperatures around 2300 °C. Similarly, in melt solidification of compounds like CdTe, kinetic instabilities at the solid-liquid interface cause incomplete atomic , resulting in fault incorporation as the crystal advances. In annealed metals, such as or aluminum, stacking faults form during processes where activates rearrangements, leading to partial and fault ribbons in the microstructure. During deformation under high , these faults are commonly observed in low stacking fault energy alloys like austenitic stainless steels, where they contribute to mechanisms such as twinning-induced .

Stacking Fault Energy

Stacking fault energy (SFE), denoted as \gamma, is defined as the excess per unit area required to introduce a stacking fault into an otherwise perfect crystal . This thermodynamic quantity measures the energetic penalty associated with disrupting the regular stacking sequence of planes, typically ranging from 10 to 200 mJ/m² in face-centered cubic (FCC) metals. Lower SFE values indicate greater stability of the faulted configuration relative to the perfect , influencing the propensity for fault formation and propagation. Several factors govern the magnitude of SFE, including , , and the specific type of stacking sequence (e.g., intrinsic versus extrinsic faults). Alloying elements can significantly alter SFE; for instance, additions that stabilize alternative stacking motifs, such as hexagonal close-packed phases, tend to reduce \gamma. generally decreases SFE due to enhanced anharmonic vibrations and expansion, which lower the energy barrier for faulting. A lower SFE promotes wider stacking fault ribbons between dissociated partial dislocations, as the repulsive elastic forces between partials are balanced against the fault energy at larger separations. The equilibrium spacing d between partial dislocations bounding a stacking fault is determined by minimizing the total , which balances the repulsion and the fault cost. This spacing is given by the formula d = \frac{G (\mathbf{b}_1 \cdot \mathbf{b}_2)}{2 \pi \gamma}, where G is the and \mathbf{b}_1 \cdot \mathbf{b}_2 is the scalar product of the Burgers vectors of the partial dislocations (approximately b^2/6 for Shockley partials in FCC, with b the magnitude of the perfect dislocation Burgers vector). Stacking faults commonly arise from the of perfect dislocations into partials, creating a faulted region whose width d directly reflects \gamma. The value of SFE profoundly impacts dislocation dynamics, particularly mobility and cross-slip. Low SFE values lead to extended dissociation widths, restricting cross-slip from one slip plane to another and favoring planar slip, which can enhance but reduce . Conversely, high SFE suppresses extensive dissociation, enabling easier cross-slip and more three-dimensional arrangements that promote and dynamic recovery during deformation. Representative examples illustrate these effects in FCC materials. In austenitic steels, low SFE (typically 20–40 mJ/m²) stabilizes stacking faults and partial dislocations, promoting deformation twinning as an alternative plasticity mechanism over dislocation slip. In contrast, aluminum exhibits high SFE (around 120–150 mJ/m²), which minimizes partial dislocation separation and suppresses stacking fault formation, favoring undissociated dislocations and cross-slip-dominated behavior.

Observation and Characterization

Electron Microscopy Methods

Transmission electron microscopy (TEM) serves as the primary technique for direct visualization of stacking faults in crystalline materials, enabling high-resolution imaging of these planar defects at the nanoscale. In TEM, stacking faults appear as regions of disrupted lattice periodicity, often bounded by partial dislocations, and their contrast arises from diffraction effects due to the local change in structure factor. Early applications in the 1950s and 1960s, pioneered by researchers like P. B. Hirsch, R. W. Horne, and M. J. Whelan, demonstrated the feasibility of observing dislocations and associated faults in thin foils of metals such as aluminum, marking the beginning of defect characterization via electron imaging. Subsequent work by A. Howie and P. R. Swann in the early 1960s extended this to direct measurements of stacking faults in copper and aluminum alloys, establishing TEM as indispensable for studying fault energetics. In conventional TEM, bright-field (BF) utilizes the direct beam to form images where stacking faults exhibit contrast variations due to the deviation from perfect Bragg conditions across the fault plane. Dark-field (DF) , formed from a diffracted beam, enhances visibility by selectively highlighting regions where the fault satisfies the diffraction condition, often revealing the fault as a bright line or fringe pattern. These fringes, particularly prominent in inclined faults, result from the overlapping partial dislocations and the varying extinction distance along the fault, providing qualitative insight into fault orientation and extent. The theoretical basis for this diffraction contrast is described by the Howie-Whelan equations, which model in deformed crystals. High-resolution TEM (HRTEM) advances this capability to atomic-scale , directly imaging the disrupted atomic stacking sequences within the fault plane. In HRTEM, phase-contrast mechanisms allow visualization of individual atomic columns, where a stacking fault appears as a shift in fringes, such as an ABCABC... sequence becoming ABCABABC... for an intrinsic stacking fault in face-centered cubic crystals. This direct observation, achievable with aberration-corrected instruments since the , has been crucial for confirming fault types and their atomic configurations in metals. Sample preparation for TEM observation of stacking faults typically involves creating electron-transparent thin foils, traditionally via to achieve thicknesses below 100 nm, which minimizes overlapping defects and surface relaxation effects. For site-specific analysis, (FIB) milling has become standard since the late 1990s, enabling precise cross-sectional lamellae from bulk samples while preserving fault integrity, though care is needed to avoid ion-induced artifacts. Quantitative analysis in TEM focuses on measuring the width of stacking faults, defined as the separation between bounding partial dislocations, to infer stacking fault energy (SFE). By imaging extended dislocations under two-beam conditions and applying elasticity theory, the equilibrium width w relates to SFE \gamma via w = \frac{\mu b^2 (2 + \nu)}{8 \pi \gamma (1 - \nu)}, where \mu is the shear modulus, b the Burgers vector magnitude, and \nu Poisson's ratio; this approach, refined from early TEM studies, yields SFE values with uncertainties around 10-20% for metals like copper.

Diffraction and Other Techniques

Stacking faults in crystalline materials can be detected and quantified through diffraction techniques that reveal perturbations in the , such as broadening, streaking, or additional peaks arising from the disruption of ideal stacking sequences. X-ray diffraction () is a primary method for identifying these defects via fault-induced broadening of Bragg peaks, particularly in face-centered cubic (fcc) structures where intrinsic and extrinsic faults cause asymmetric broadening and shifts in peak positions. For instance, in fcc metals, streaking or satellite peaks near high-index reflections indicate the presence of stacking disorder, allowing estimation of fault probabilities through line-profile analysis. Warren's method, originally developed for analyzing diffraction patterns affected by planar defects, quantifies the probability of stacking faults by modeling the coefficients of peak profiles to separate contributions from faulting, size, and microstrain. This approach has been refined to account for texture and non-uniform fault distributions, enabling accurate determination of fault densities from powder XRD data. Electron diffraction techniques, performed in scanning electron microscopy (SEM) or (TEM), provide complementary insights by capturing (SAD) patterns that exhibit forbidden reflections or streaking due to the local disruption of by stacking faults. In TEM-based , faults in close-packed structures produce extra spots or arcs in the pattern, particularly along directions perpendicular to the fault plane, which are absent in perfect . These features arise from the phase shifts introduced by the faulted layers, enabling statistical assessment of fault over larger areas than direct . While visual fringe patterns from bright-field offer spatial mapping of individual faults, diffraction patterns emphasize the collective effect on reciprocal space. Neutron diffraction is particularly useful for bulk samples, where it detects diffuse from ensembles of stacking faults that are distributed throughout the material volume. This measures the intensity of away from Bragg peaks, which is enhanced by the random or correlated positioning of faults, providing information on fault correlations and overall disorder without the surface sensitivity of X-rays. In materials with significant fault populations, such as polytypic compounds, neutron diffuse reveals short-range stacking variations that contribute to the observed linewidths and asymmetries in the pattern. Spectroscopic methods, including and (CL), indirectly probe stacking faults through shifts in vibrational or luminescent signals induced by the local strain fields surrounding the defects. In , faults alter the modes, leading to broadening or splitting of peaks, such as the transverse optical mode in , which correlates with fault density and associated strain. Similarly, CL spectra show characteristic emission peaks or shifts due to quantum confinement or strain in fault regions, as observed in wide-bandgap semiconductors where faults act as radiative recombination centers. These techniques are sensitive to the electronic and vibrational perturbations from faults but require calibration against known defect concentrations. Despite their utility, and spectroscopic methods for stacking fault characterization primarily yield ensemble-averaged information, capturing statistical distributions rather than individual defect locations, and thus offer lower compared to direct techniques. This averaging can mask spatial variations in fault density, necessitating complementary high-resolution methods for complete analysis.

Material-Specific Contexts

In Metals and Alloys

In face-centered cubic (FCC) metals and alloys, stacking faults play a pivotal role in by facilitating the glide of partial s, which dissociate from perfect s in materials with low stacking fault energy (SFE). This mechanism impedes cross-slip and promotes planar slip, leading to enhanced through increased interactions and density, while also contributing to improved by distributing strain more uniformly. For instance, in Cu-15Al alloys with SFE around 7 mJ/m², stacking faults form early during deformation (at strains as low as 0.02) and significantly boost the strain-hardening rate in coarse- and medium-grained structures by acting as barriers to further motion. Low SFE in FCC metals often promotes deformation twinning, where twin boundaries are composed of overlapping stacking faults, enabling alternative deformation modes that enhance strength without sacrificing . In Cu-Zn alloys, such as Cu-20Zn with SFE of 19 mJ/m², twinning dominates under high-strain conditions like surface attrition treatment, as opposed to dislocation slip in higher-SFE variants like Cu-2Zn (38 mJ/m²), resulting in refined microstructures that improve overall performance. Alloying strategies are employed to engineer SFE in metals, thereby tailoring phase stability and mechanical properties; for example, increasing nickel content in austenitic Fe-Cr-Ni steels raises SFE, stabilizing the FCC austenite phase and shifting deformation from stacking fault-mediated martensitic transformation (in 10 wt% alloys) to twinning-dominated mechanisms (in 14 wt% alloys), which elevates strength through higher accumulation. In industrial applications, stacking faults in austenitic stainless steels influence fatigue behavior by nucleating regions of high defect density that facilitate martensitic transformation, thereby promoting crack initiation at boundaries or twin interfaces under cyclic loading with amplitudes above 0.3%. During processes, rapid cycles in these low-SFE alloys can exacerbate residual stresses and crack susceptibility in heat-affected zones. Recent 2020s research highlights advances in nanotwinned metals, where dense stacking faults and twin boundaries synergistically enhance strength; for example, electrodeposited nanotwinned pure achieves 4.0 GPa yield strength at 2.9 nm twin thickness due to stable twin barriers that activate secondary nanotwins for sustained hardening, while in CrCoNi medium-entropy alloys, nanotwins and stacking faults impede dislocation motion more effectively than isolated defects, boosting tensile strength up to 1 GPa with minimal softening at fine interspacings.

In Semiconductors and Ceramics

In semiconductors, stacking faults act as recombination centers that trap charge carriers, thereby reducing and essential for efficient device operation. In (Si), these faults, often observed in epitaxial layers, promote non-radiative recombination and contribute to shunt formation under electrical stress, degrading photovoltaic performance. Similarly, in (GaAs) nanowires and epitaxial structures, stacking faults scatter electrons, limiting to values significantly below bulk levels, with defect-free growth techniques achieving up to twofold enhancements in mobility through fault suppression. These defects are particularly prevalent in nanostructures like nanowires, where they arise during vapor-liquid-solid growth and hinder high-speed electronics applications. Polytypic stacking faults frequently occur during (CVD) of wide-bandgap semiconductors such as () and zinc oxide (ZnO), resulting in mixtures of hexagonal and cubic phases that compromise material uniformity. In epitaxial layers, lower growth temperatures stabilize the 4H polytype while minimizing fault density, as faults nucleate due to stacking sequence disruptions at step edges on off-axis substrates. For ZnO nanowires grown via CVD on substrates like , a high density of basal plane stacking faults on (0001) planes leads to local phase variations between and zinc blende structures, altering optical and mechanical properties. In ceramics, stacking faults influence both electrical and mechanical behaviors, particularly in oxides like alumina (Al₂O₃). These faults at interfaces, such as in /Al₂O₃ heterostructures, create disordered pathways that enhance ionic conductivity by orders of magnitude compared to pure phases, enabling superionic transport. Regarding mechanical properties, faulted structures in nanocrystalline Al₂O₃ promote room-temperature through nanoscale deformation mechanisms, increasing from typical brittle values of ~3 ·m¹/² to over 5 ·m¹/² by facilitating motion and crack deflection. Stacking faults pose significant device implications in , inducing leakage currents that reduce efficiency in light-emitting diodes (LEDs) and solar cells. In GaN-based LEDs on sapphire substrates, interfacial stacking faults and inversion domains elevate reverse-bias leakage by providing conductive paths for carriers. In thin-film CdTe solar cells, partial stacking faults act as critical recombination sites, lowering ; chlorine-based treatments remove these faults, boosting efficiency from ~15% to over 22%. Mitigation often involves substrate selection, such as compliant or off-axis substrates in SiC , which reduce fault nucleation by accommodating lattice mismatch and promoting uniform stacking. In modern two-dimensional materials like , stacking faults have become increasingly relevant since post-2010 discoveries, where they induce local bandgap openings up to 0.1 eV in bilayer structures by disrupting the dispersion and enabling tunable electronic properties for .

References

  1. [1]
    5.4.1 Partial Dislocations and Stacking Faults - Dallas R. Trinkle
    Stacking faults by themselves are simple two-dimensional defects. They carry a certain stacking fault energy γ; very roughly around a few 100 mJ/m2. The disc of ...
  2. [2]
    LAMMPS Intrinsic Stacking-Fault Energy
    Dec 12, 2023 · A stacking fault is an irregularity in the planar stacking sequence of atoms in a crystal. The natural stacking arrangement for FCC metals ...
  3. [3]
    [PDF] CHAPTER 4: IMPERFECTIONS IN SOLIDS
    Shockley partial dislocations with b=a/6·[211] separated by ribbon of staking faults of width d. The width of d is large for the alloys with low stacking fault ...
  4. [4]
    Stacking Fault - an overview | ScienceDirect Topics
    A stacking fault is a planar defect and, as its name implies, it is a local region in the crystal where the regular stacking sequence is interrupted. Stacking ...
  5. [5]
    Stacking fault - Online Dictionary of Crystallography
    Jul 14, 2021 · A stacking fault is a type of planar defect that occurs in modular crystal structures, and especially (but not exclusively) in layer structures.
  6. [6]
    [PDF] Lecture Outline Real Structure
    □ Point defects being “OD” entities, dislocations are a line of defects defect, 1D; stacking faults/grain boundaries are a plane of defects, 2D; alien ...
  7. [7]
    nglos324 - stackingfault
    Stacking fault are local regions of incorrect stacking of crystal planes associated with the presence of partial dislocations.
  8. [8]
    Intrinsic Stacking Fault - an overview | ScienceDirect Topics
    Intrinsic stacking fault is defined as a disruption in the normal stacking sequence of atomic layers within a crystal, where the stacking order is ...
  9. [9]
    Effect of strain on the stacking fault energy of copper: A first ...
    Aug 26, 2013 · An intrinsic stacking fault occurs when the stacking sequence is altered by removal of one of the planes in this stacking sequence, e.g., ...
  10. [10]
    Diffraction by fcc Crystals with Intrinsic and Extrinsic Faults
    A model is given for diffraction by an fcc lattice which contains a randomly arranged mixture of intrinsic and extrinsic stacking faults.
  11. [11]
    The nature of deformation stacking faults in F.C.C. alloys
    stacking sequences indicate that an extrinsic fault can be regarded as consisting of two intrinsic faults on neighbouring planes and hence (neglect- ing ...<|separator|>
  12. [12]
    A brief history of dislocation theory | Metallurgical and Materials ...
    Early developments leading to the concept of a dislocation are discussed. The discoveries of 1934 are described. The subsequent evolution of various aspect.
  13. [13]
    [PDF] Structure of Crystalline Solids and Materials Synthesis
    The difference between FCC and HCP close-packed structures lies in the stacking sequence. A: Centers of all the atoms. B: Vertex up triangle (gap). C: Vertex ...
  14. [14]
    [PDF] Chapter 1 Crystal Structures - Physics
    Oct 23, 2016 · The most often found are fcc and hcp structures. If the stacking is arranged in ABABAB… then the structure is hcp. If the 3rd layer goes ...
  15. [15]
    [PDF] 3d Crystallization - FCC structures
    An hcp-lattice is obtained if the particles in the third layer are on top of the particles in the first layer. An fcc-lattice is obtained if the particles in ...
  16. [16]
    [PDF] ME531-Twinning Model BCC - High Temperature Materials Laboratory
    The thinnest stable multilayer stacking fault on the {112} plane is a three-layer fault, and can be considered as the twin nucleus. The increase in additional ...
  17. [17]
    Mechanisms of Plasticity (all content) - DoITPoMS
    Partial dislocations in an fcc system​​ Slip by this two-step process creates a stacking fault ABCAC ABC in the stacking sequence. The two partial dislocations, ...
  18. [18]
    Stacking fault energy in concentrated alloys | Nature Communications
    Jun 11, 2021 · Stacking faults are created during the dissociation of lattice dislocations into partial dislocations to reduce the elastic energy according to ...
  19. [19]
    [PDF] Dislocations in fcc Metals - IAM, KIT
    The dissociation of dislocations is therefore restricted to special cases. Shockley partial dislocations border an intrinsic stacking fault. stacking sequence ...
  20. [20]
    Analysis of the stacking faults of dissociated glide dislocations in ...
    In f.c.c. crystals the dislocations gliding on { l l l } planes are split into partial dislocations bounding a stacking fault ribbon.
  21. [21]
    Temperature Effect on Stacking Fault Energy and Deformation ...
    Feb 20, 2020 · Generally speaking, dislocation behaviours like dislocation climb and dislocation cross-slip are easier to occur at high temperature; as ...
  22. [22]
    None
    Nothing is retrieved...<|control11|><|separator|>
  23. [23]
    The behaviour of stacking fault energy upon interstitial alloying - PMC
    Sep 11, 2017 · The excess stuctural energy induced in a material by the process defines the so-called stacking fault energy (SFE, γ sf ). Because a stacking ...
  24. [24]
    [PDF] Effect of Stacking Fault Energy on Mechanism of Plastic Deformation ...
    The stacking fault energy (SFE) plays a critical role in the deformation properties of face- centered cubic (fcc) metals and alloys. The SFE influences such ...
  25. [25]
    Thermodynamic modeling of the stacking fault energy of austenitic ...
    The ideal stacking fault energy γSFE is then expressed by γ SFE = 2 ρ Δ G γ → ε + 2 σ where ρ is the molar surface density along {1 1 1} planes, Δ G γ → ε is ...
  26. [26]
    Temperature dependence of the stacking-fault Gibbs energy for Al ...
    Dec 26, 2018 · The stacking fault energy (SFE) for Al, Cu, and Ni decreases with temperature. The quasiharmonic approximation captures the largest ...
  27. [27]
    07mae324 - PD59
    where γ is the stacking fault energy. · The dislocations and stacking fault reach an equilibrium width when these two forces are equal, the separation of ...Missing: formula | Show results with:formula
  28. [28]
    Stacking Fault Energy Determination in Fe-Mn-Al-C Austenitic Steels ...
    The horizontal planes represent the SFE values in which the literature reports a change in the deformation mechanism; the SFE values less than 20 mJ / m 2 ...
  29. [29]
    Tuning stacking fault energy and enhancing mechanical properties ...
    The intrinsic stacking fault energy (γsf) of pure fcc Al was experimentally calculated to be 121 mJ/m². Published calculations for the γsf range from 112 to 142 ...
  30. [30]
    Atomic scale characterization of complex stacking faults and their ...
    Oct 15, 2020 · Atomic-resolution scanning transmission electron microscopy is used to study stacking faults and their configurations formed in cold deformed samples.
  31. [31]
    Application of FIB in the preparation of TEM specimens of ...
    Therefore, taking advantage of FIB's feature of being a site-specific method is indispensable in preparing thin foils comprising WC grains but also WC-WC and WC ...
  32. [32]
    X-ray Diffraction Analysis of Stacking and Twin Faults in f.c.c. Metals
    Dec 28, 2015 · It is discussed how stacking- and twin-fault probabilities (and crystallite sizes) can be determined from diffraction line-profile measurements.
  33. [33]
    Satellite intensity and forbidden reflections from stacking faults in a ...
    These additional reflections are most commonly attributed to the presence of chemical short-range order, or to thin foil effects in transmission electron ...
  34. [34]
    [PDF] stacking faults in a deformed superalloy - University of Cambridge
    Sep 17, 2025 · reflections in electron diffraction patterns that do not correspond to allowed reflections from the A1. 306 structure. These have most ...
  35. [35]
    Nuclear and magnetic diffuse scattering of neutrons from β-ferrous ...
    This approach shows qualitatively the effects of stacking faults seen by neutron diffraction through nuclear and magnetic diffuse scattering. In spite of ...
  36. [36]
    Diffuse scattering of neutrons in thermally oxidized silicon
    The diffuse scattering can be connected with the growth of the stacking faults in the oxidized silicon. Previous article in issue; Next ...
  37. [37]
    Detection of stacking faults in 6H-SiC by Raman scattering
    Nov 27, 2000 · Raman spectra of 6H-SiC crystals including stacking faults have been examined for the c face using backscattering geometry.
  38. [38]
    Cathodoluminescence study of the properties of stacking faults in
    Mar 3, 2009 · In-grown stacking faults in n-type 4H-SiC epitaxial layers have been investigated by real-color cathodoluminescence imaging and spectroscopy ...
  39. [39]
    X-Ray Diffraction Line Broadening: Modeling and Applications ... - NIH
    For instance, stacking faults and twins will contribute to broadening similar to size effects. Warren-Averbach Method. This method was developed originally ...
  40. [40]
    Significant contribution of stacking faults to the strain hardening ...
    Nov 19, 2015 · For FCC metals and alloys with low SFE, SFs and deformation twins are favored in contrast to those materials with high SFE under plastic ...
  41. [41]
    Correlation of stacking fault energy with deformation mechanism in ...
    Jan 1, 2022 · Research results indicate that the deformation mechanism directly correlated with the SFE in Cu–Zn alloys.
  42. [42]
    The impact of nickel concentration and stacking fault energy on ...
    In this study, we use two high-purity Fe-Cr-Ni alloys to investigate how Ni concentration and SFE interact to alter deformation mechanisms and induce ...
  43. [43]
    Fatigue mechanisms in an austenitic steel under cyclic loading
    The atomistic simulations show the role of regions of a high density of stacking faults for the martensitic transformation. ... crack initiation and crack growth ...
  44. [44]
    Welding-induced microstructure in austenitic stainless steels before ...
    Irradiated austenitic stainless steels, because of their low stacking fault energy and high shear modulus, should exhibit a high ratio of stacking fault ...
  45. [45]
    Ultrastrong nanotwinned pure nickel with extremely fine twin thickness
    Jun 30, 2021 · An unprecedented strength of 4.0 GPa was achieved at extremely fine twin thickness of 2.9 nm, which is about 12 times stronger than that of conventional coarse ...
  46. [46]
    Effects of nanotwins and stacking faults on the mechanical ...
    It is found that the strength of MEAs is enhanced by the introduction of planar defects through impeding dislocation motion.