To see the other types of publications on this topic, follow the link: Stacking fault energy.

Journal articles on the topic 'Stacking fault energy'

Create a spot-on reference in APA, MLA, Chicago, Harvard, and other styles

Select a source type:

Consult the top 50 journal articles for your research on the topic 'Stacking fault energy.'

Next to every source in the list of references, there is an 'Add to bibliography' button. Press on it, and we will generate automatically the bibliographic reference to the chosen work in the citation style you need: APA, MLA, Harvard, Chicago, Vancouver, etc.

You can also download the full text of the academic publication as pdf and read online its abstract whenever available in the metadata.

Browse journal articles on a wide variety of disciplines and organise your bibliography correctly.

1

Fan, Dawei, Qingzhou Zhang, Touwen Fan, Mengdong He, and Linghong Liu. "A New Anti-Alias Model of Ab Initio Calculations of the Generalized Stacking Fault Energy in Face-Centered Cubic Crystals." Crystals 13, no. 3 (March 8, 2023): 461. http://dx.doi.org/10.3390/cryst13030461.

Full text
Abstract:
The anti-alias model is an effective method to calculate the generalized stacking fault energy of the hexagonal close-packed crystals, but it has not been applied to the face-centered cubic crystals due to two different stacking faults occurring in the supercell during the sliding process. Based on the symmetry of these two stacking faults and the existing single analytic formula of the generalized stacking fault energy, we successfully extend the anti-alias model to compute the generalized stacking fault energy of face-centered cubic crystals, and the common fcc metals Al, Ni, Ag and Cu are taken as specific examples to illustrate the computational details. Finally, the validity of the proposed model is verified by data comparison and analysis. It is suggested that the anti-alias model is a good choice for the researchers to obtain more accurate generalized stacking fault energy of face-centered cubic metals.
APA, Harvard, Vancouver, ISO, and other styles
2

Wu, Yu-Chuan, Sea-Fue Wang, and Hong-Yang Lu. "Stacking Faults and Stacking Fault Energy of Hexagonal Barium Titanate." Journal of the American Ceramic Society 89, no. 12 (December 2006): 3778–87. http://dx.doi.org/10.1111/j.1551-2916.2006.01305.x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
3

Martin, Stefan, Christiane Ullrich, Daniel Šimek, Ulrich Martin, and David Rafaja. "Stacking fault model of ∊-martensite and itsDIFFaXimplementation." Journal of Applied Crystallography 44, no. 4 (June 28, 2011): 779–87. http://dx.doi.org/10.1107/s0021889811019558.

Full text
Abstract:
Plastic deformation of highly alloyed austenitic transformation-induced plasticity (TRIP) steels with low stacking fault energy leads typically to the formation of ∊-martensite within the original austenite. The ∊-martensite is often described as a phase having a hexagonal close-packed crystal structure. In this contribution, an alternative structure model is presented that describes ∊-martensite embedded in the austenitic matrixviaclustering of stacking faults in austenite. The applicability of the model was tested on experimental X-ray diffraction data measured on a CrMnNi TRIP steel after 15% compression. The model of clustered stacking faults was implemented in theDIFFaXroutine; the faulted austenite and ∊-martensite were represented by different stacking fault arrangements. The probabilities of the respective stacking fault arrangements were obtained from fitting the simulated X-ray diffraction patterns to the experimental data. The reliability of the model was proven by scanning and transmission electron microscopy. For visualization of the clusters of stacking faults, the scanning electron microscopy employed electron channelling contrast imaging and electron backscatter diffraction.
APA, Harvard, Vancouver, ISO, and other styles
4

Seo, Won-Seon, and Kunihito Koumoto. "Kinetics and mechanism of stacking fault annihilation and grain growth in porous ceramics of β–SiC." Journal of Materials Research 8, no. 7 (July 1993): 1644–50. http://dx.doi.org/10.1557/jmr.1993.1644.

Full text
Abstract:
Beta–SiC specimens possessing 15% stacking fault density were annealed at various temperatures for various time periods under an Ar or a N2 atmosphere, and the mechanisms of stacking fault annihilation and grain growth were investigated. The values of the geometric factor in the Avrami–Erofeev equation indicated that the rate of stacking fault annihilation is controlled by the atomic diffusion process. On the other hand, the rate of grain growth was found to be limited by surface diffusivity. Coincidence in the values of activation energy for stacking fault annihilation and grain growth within experimental errors firmly suggested that the annihilation of stacking faults is an apparent phenomenon resulting from the microstructural development in which the grain growth is controlled by surface diffusivity. Incorporation of nitrogen during heating suppressed the surface diffusivity and, hence, the rate of stacking fault annihilation.
APA, Harvard, Vancouver, ISO, and other styles
5

Wang, Y. Q., W. S. Liang, and G. G. Ross. "Stacking Fault Energy of Si Nanocrystals Embedded in SiO2." ISRN Nanotechnology 2011 (May 25, 2011): 1–3. http://dx.doi.org/10.5402/2011/639714.

Full text
Abstract:
Si nanocrystals (Si nc) were produced by the implantation of Si+ into a SiO2 film on (100) Si, followed by high-temperature annealing. High-resolution transmission electron microscopy (HRTEM) observation has shown that a perfect dislocation (Burgers vector b=(1/2)〈110〉) can dissociate into two Shockley partials (Burgers vector b=(1/6)〈112〉) bounding a strip of stacking faults (SFs). The width of the SFs has been determined from the HRTEM image, and the stacking fault energy for Si nc has been calculated. The stacking fault energy for Si nc is compared with that for bulk Si, and the formation probability of defects in Si nc is also discussed. The results will shed a light on the dissociation of dislocations in nanoparticles.
APA, Harvard, Vancouver, ISO, and other styles
6

Shen, Rui, Zengyu Ni, Siyuan Peng, Haile Yan, and Yanzhong Tian. "Effects of V Addition on the Deformation Mechanism and Mechanical Properties of Non-Equiatomic CoCrNi Medium-Entropy Alloys." Materials 16, no. 14 (July 22, 2023): 5167. http://dx.doi.org/10.3390/ma16145167.

Full text
Abstract:
Equiatomic CoCrNi medium-entropy alloys exhibit superior strength and ductility. In this work, a non-equiatomic CoCrNi alloy with low stacking fault energy was designed, and different fractions of V were added to control the stacking fault energy and lattice distortion. Mechanical properties were evaluated by tensile tests, and deformation microstructures were characterized by transmission electron microscope (TEM). The main deformation mechanisms of CoCrNiV alloy with low V content are dislocation slip, stacking faults, and deformation-induced HCP phase transformation, while the dominant deformation patterns of CoCrNiV alloy with high V contents are dislocation slip and stacking faults. The yield strength increases dramatically when the V content is high, and the strain-hardening behavior changes non-monotonically with increasing the V content. V addition increases the stacking fault energy (SFE) and lattice distortion. The lower strain-hardening rate of 6V alloy than that of 2V alloy is dominated by the SFE. The higher strain-hardening rate of 10V alloy than that of 6V alloy is dominated by the lattice distortion. The effects of V addition on the SFE, lattice distortion, and strain-hardening behavior are discussed.
APA, Harvard, Vancouver, ISO, and other styles
7

Rafaja, D., C. Krbetschek, C. Ullrich, and S. Martin. "Stacking fault energy in austenitic steels determined by usingin situX-ray diffraction during bending." Journal of Applied Crystallography 47, no. 3 (May 10, 2014): 936–47. http://dx.doi.org/10.1107/s1600576714007109.

Full text
Abstract:
A method is presented which determines the stacking fault energy in face-centred cubic materials from the critical stress that is inducedviasample bending in the early stages of plastic deformation. The critical stress is gauged byin situX-ray diffraction. This method utilizes the results of Byun's consideration of the stress dependence of the partial dislocation separation [Byun (2003).Acta Mater.51, 3063–3071]. Byun showed that the separation distance of the partial dislocations increases rapidly when the critical stress is reached and that the critical stress needed for the rapid separation of the partial dislocations is directly proportional to the stacking fault energy. In the approach presented here, the partial dislocation separation and the corresponding triggering stress are monitored by usingin situX-ray diffraction during sample bending. Furthermore, thein situX-ray diffraction measurement checks the possible interactions between stacking faults present on equivalent lattice planes and the interactions of the stacking faults with other microstructure defects. The capability of the proposed method was tested on highly alloyed austenitic steels containing chromium (∼16 wt%), manganese (∼7 wt%) and nickel as the main alloying elements. For the steels containing 5.9 and 3.7 wt% Ni, stacking fault energies of 17.5 ± 1.4 and 8.1 ± 0.9 mJ m−2were obtained, respectively.
APA, Harvard, Vancouver, ISO, and other styles
8

Seki, Kazuaki, Kai Morimoto, Toru Ujihara, Tomoharu Tokunaga, Katsuhiro Sasaki, Kotaro Kuroda, and Yoshikazu Takeda. "Stacking Faults around the Hetero-Interface Induced by 6H-SiC Polytype Transformation on 3C-SiC with Solution Growth." Materials Science Forum 645-648 (April 2010): 363–66. http://dx.doi.org/10.4028/www.scientific.net/msf.645-648.363.

Full text
Abstract:
6H-SiC hetero-epitaxially grown on a (111) 3C-SiC was observed with TEM. High-density stacking faults were formed around the hetero-interface, and the density of stacking faults decreased with increasing distance from interface. On the other hand, when 3C-SiC was homo-epitaxially grown on a 3C-SiC, any stacking faults did not exist at the interface between the grown crystal and the seed crystal. Thus, the stacking faults formation started from the 6H/3C hetero-interface. Considering the lattice-mismatch strain between 3C-SiC and 6H-SiC, the strain energy is equivalent to the stacking fault energy of 6H-SiC. This similarity suggests that the stacking faults formation could be caused by the relaxation of the lattice-mismatch strain.
APA, Harvard, Vancouver, ISO, and other styles
9

Nembach, E., T. Pretorius, and D. Rönnpagel. "Stacking-fault energy mismatch strengthening revisited." Philosophical Magazine A 78, no. 4 (October 1998): 949–63. http://dx.doi.org/10.1080/01418619808239967.

Full text
APA, Harvard, Vancouver, ISO, and other styles
10

Junes, H. J., H. Alles, M. S. Manninen, A. Y. Parshin, and I. A. Todoshchenko. "Stacking Fault Energy in 4He Crystals." Journal of Low Temperature Physics 153, no. 5-6 (October 9, 2008): 244–49. http://dx.doi.org/10.1007/s10909-008-9828-0.

Full text
APA, Harvard, Vancouver, ISO, and other styles
11

Nembach T. Pretorius D. Ronnpagel, E. "Stacking-fault energy mismatch strengthening revisited." Philosophical Magazine A 78, no. 4 (October 1, 1998): 949–63. http://dx.doi.org/10.1080/014186198253291.

Full text
APA, Harvard, Vancouver, ISO, and other styles
12

Das, Arpan. "Revisiting Stacking Fault Energy of Steels." Metallurgical and Materials Transactions A 47, no. 2 (December 14, 2015): 748–68. http://dx.doi.org/10.1007/s11661-015-3266-9.

Full text
APA, Harvard, Vancouver, ISO, and other styles
13

Gencalp Irizalp, Simge, and Nursen Saklakoglu. "Effect of multiple laser shock processing on nano-scale microstructure of an aluminum alloy." Characterization and Application of Nanomaterials 3, no. 1 (April 21, 2020): 9. http://dx.doi.org/10.24294/can.v3i1.716.

Full text
Abstract:
In this study, nano-scale microstructural evolution in 6061-T6 alloy after laser shock processing (LSP) was studied. 6061-T6 alloy plate was subjected to multiple LSP. The LSP treated area was characterized by X-ray diffraction and the microstructure of the samples was analyzed by transmission electron microscopy. Focused Ion Beam (FIB) tools were used to prepare TEM samples in precise areas. It was found that even though aluminum had high stacking fault energy, LSP yielded to formation of ultrafine grains and deformation faults such as dislocation cells, stacking faults. The stacking fault probability (PSF) was obtained in LSP-treated alloy using X-Ray diffraction. Deformation induced stacking faults lead to the peak position shifts, broadening and asymmetry of diffraction. XRD analysis and TEM observations revealed significant densities of stacking faults in LSP-treated 6061-T6 alloy. And mechanical properties of LSP-treated alloy were also determined to understand the hardening behavior with high concentration of structural defects.
APA, Harvard, Vancouver, ISO, and other styles
14

Sun, Xun, Hualei Zhang, Wei Li, Xiangdong Ding, Yunzhi Wang, and Levente Vitos. "Generalized Stacking Fault Energy of Al-Doped CrMnFeCoNi High-Entropy Alloy." Nanomaterials 10, no. 1 (December 26, 2019): 59. http://dx.doi.org/10.3390/nano10010059.

Full text
Abstract:
Using first-principles methods, we investigate the effect of Al on the generalized stacking fault energy of face-centered cubic (fcc) CrMnFeCoNi high-entropy alloy as a function of temperature. Upon Al addition or temperature increase, the intrinsic and extrinsic stacking fault energies increase, whereas the unstable stacking fault and unstable twinning fault energies decrease monotonously. The thermodynamic expression for the intrinsic stacking fault energy in combination with the theoretical Gibbs energy difference between the hexagonal close packed (hcp) and fcc lattices allows one to determine the so-called hcp-fcc interfacial energy. The results show that the interfacial energy is small and only weakly dependent on temperature and Al content. Two parameters are adopted to measure the nano-twinning ability of the present high-entropy alloys (HEAs). Both measures indicate that the twinability decreases with increasing temperature or Al content. The present study provides systematic theoretical plasticity parameters for modeling and designing high entropy alloys with specific mechanical properties.
APA, Harvard, Vancouver, ISO, and other styles
15

Fujita, S., Tokuteru Uesugi, Yorinobu Takigawa, and Kenji Higashi. "Stacking Fault Energy of Cu-Ga Alloys from First Principles." Materials Science Forum 561-565 (October 2007): 1915–18. http://dx.doi.org/10.4028/www.scientific.net/msf.561-565.1915.

Full text
Abstract:
The segregation energy of solute Ga in the staking fault in Cu-Ga alloy was calculated from the first principles. Then, we presented numerical results of the stacking fault energy for Cu-Ga alloy using the value of the segregation energy as a input parameter to a expression in the equilibrium state. The numerical results of the stacking fault energy were in good agreement with the experimental values.
APA, Harvard, Vancouver, ISO, and other styles
16

Qi-Xun, Dai, Wang An-Dong, Cheng Xiao-Nong, and Luo Xin-Min. "Stacking fault energy of cryogenic austenitic steels." Chinese Physics 11, no. 6 (May 14, 2002): 596–600. http://dx.doi.org/10.1088/1009-1963/11/6/315.

Full text
APA, Harvard, Vancouver, ISO, and other styles
17

Li, Wei, Song Lu, Qing-Miao Hu, Börje Johansson, Se Kyun Kwon, Mikael Grehk, Jan Y. Johnsson, and Levente Vitos. "Generalized stacking fault energy of γ-Fe." Philosophical Magazine 96, no. 6 (February 12, 2016): 524–41. http://dx.doi.org/10.1080/14786435.2016.1140912.

Full text
APA, Harvard, Vancouver, ISO, and other styles
18

Takeuchi, S., K. Suzuki, K. Maeda, and H. Iwanaga. "Stacking-fault energy of II–VI compounds." Philosophical Magazine A 50, no. 2 (February 1985): 171–78. http://dx.doi.org/10.1080/01418618408244220.

Full text
APA, Harvard, Vancouver, ISO, and other styles
19

Vora, Aditya M. "Stacking fault energy in some single crystals." Journal of Semiconductors 33, no. 6 (June 2012): 062001. http://dx.doi.org/10.1088/1674-4926/33/6/062001.

Full text
APA, Harvard, Vancouver, ISO, and other styles
20

Chu, F., A. H. Ormeci, T. E. Mitchell, J. M. Wills, D. J. Thoma, R. C. Albers, and S. P. Chen. "Stacking fault energy of the NbCr2laves phase." Philosophical Magazine Letters 72, no. 3 (September 1995): 147–53. http://dx.doi.org/10.1080/09500839508242445.

Full text
APA, Harvard, Vancouver, ISO, and other styles
21

Medlin, D. L., G. H. Campbell, and C. Barry Carter. "Grain Boundary Dislocations and Stacking Defects in the 9R Phase at an Incoherent Twin Boundary in Copper." Microscopy and Microanalysis 4, S2 (July 1998): 774–75. http://dx.doi.org/10.1017/s1431927600023990.

Full text
Abstract:
Experiment and modeling show that there is a general mode of grain boundary dissociation, common in low stacking fault energy FCC metals, that can be well understood in terms of the emission of arrays of stacking faults from the grain boundary plane. Most extensively studied of such dissociated interfaces are the Σ=3 incoherent twin boundaries. Numerous observations now exist of grain boundary dissociation at such interfaces showing a layer that is well described as a narrow, several nanometer wide slab of 9R stacked material. The 9R stacking sequence is equivalent to a close-packed stacking of FCC ﹛111﹜ planes with an intrinsic stacking fault inserted every three planes (i.e., a stacking sequence of ABC/BCA/CBA …). The width of the 9R layer is sensitive to the local stress state. Figure 1 shows results of an atomistic calculation simulating the effect of an applied shear strain parallel to the boundary.
APA, Harvard, Vancouver, ISO, and other styles
22

Suzuki, Mayumi, and Kouichi Maruyama. "Effects of Stacking Faults on High Temperature Creep Behavior in Mg-Y-Zn Based Alloys." Materials Science Forum 638-642 (January 2010): 1602–7. http://dx.doi.org/10.4028/www.scientific.net/msf.638-642.1602.

Full text
Abstract:
Compressive creep behavior of hot-rolled (40%) Mg-Y and Mg-Y-Zn alloys are investigated at 480 ~ 650 K. Creep strength is substantially improved by the simultaneous addition of yttrium and zinc. The minimum creep rate of Mg-0.9mol%Y-0.04mol%Zn (WZ301) alloy decreases to 1/10 lower than that of Mg-1.1mol%Y (W4) alloy at 650 K. Activation energy for creep in W4 and WZ301 alloys are more than 200 kJ/mol at the temperature range of 480 ~ 550 K. These values are higher than the activation energy for self-diffusion coefficient in magnesium (135 kJ/mol). Many stacking faults (planar defects, PDs) are only observed on the basal planes of the matrix in Mg-Y-Zn ternary alloys. Stacking fault energy is considered to decrease by the multiple-addition of yttrium and zinc. The size and density of these planar defects depend on solute content, aging condition. TEM observation has been revealed that the decreasing of the stacking fault energy affects the distribution of dislocations during creep. Many a-dislocations on basal planes are extended significantly. Dislocation motion is restricted significantly by both of these two types of stacking faults (planar type and extended dislocations).
APA, Harvard, Vancouver, ISO, and other styles
23

Kitashima, Tomonori. "Thermodynamic Analysis of the Effects of Alloying Elements on the Stacking Fault Energy in Ruthenium-Bearing Nickel Alloys." Advanced Materials Research 1119 (July 2015): 580–84. http://dx.doi.org/10.4028/www.scientific.net/amr.1119.580.

Full text
Abstract:
The effects of Al, Co, Re, and Ru on the stacking fault energy in Ni alloys were analyzed using computational thermodynamics. The effects of adding up to 5 at% Re or Ru to a Ni-15at%Co system were found to be weak at 300 °C, 700 °C, and 900 °C. However, Al addition decreased the stacking fault energy in a Ni-15at%Co-Xat%Ru system, where X = 0, 3, 5. In addition, this decrease in the stacking fault energy due to Al addition became more significant as the amount of Ru increased. Furthermore, in Ni–Co–Al–Ru alloys containing 9at%Al, the addition of 5at%Ru decreased the stacking fault energy as much as the addition of 12.5at%Co at 900 °C. The effects of Co and Ru addition on the γ/γ’ microstructure of Ni-based superalloys were also discussed.
APA, Harvard, Vancouver, ISO, and other styles
24

Agarwal, Anant K., Sumi Krishnaswami, Jim Richmond, Craig Capell, Sei Hyung Ryu, John W. Palmour, Bruce Geil, Dimos Katsis, Charles Scozzie, and Robert E. Stahlbush. "Influence of Basal Plane Dislocation Induced Stacking Faults on the Current Gain in SiC BJTs." Materials Science Forum 527-529 (October 2006): 1409–12. http://dx.doi.org/10.4028/www.scientific.net/msf.527-529.1409.

Full text
Abstract:
SiC BJTs show instability in the I-V characteristics after as little as 15 minutes of operation. The current gain reduces, the on-resistance in saturation increases, and the slope of the output characteristics in the active region increases. This degradation in the I-V characteristics continues with many hours of operation. It is speculated that this phenomenon is caused by the growth of stacking faults from certain basal plane dislocations within the base layer of the SiC BJT. Stacking fault growth within the base layer is observed by light emission imaging. The energy for this expansion of the stacking fault comes from the electron-hole recombination in the forward biased base-emitter junction. This results in reduction of the effective minority carrier lifetime, increasing the electron-hole recombination in the base in the immediate vicinity of the stacking fault, leading to a reduction in the current gain. It should be noted that this explanation is only a suggestion with no conclusive proof at this stage.
APA, Harvard, Vancouver, ISO, and other styles
25

Zhang, Jing, and Pavel A. Korzhavyi. "First Principles Investigation on Thermodynamic Properties and Stacking Fault Energy of Paramagnetic Nickel at High Temperatures." Metals 10, no. 3 (February 28, 2020): 319. http://dx.doi.org/10.3390/met10030319.

Full text
Abstract:
Reliable data on the temperature dependence of thermodynamic properties of alloy phases are very useful for modeling the behavior of high-temperature materials such as nickel-based superalloys. Moreover, for predicting the mechanical properties of such alloys, additional information on the energy of lattice defects (e.g., stacking faults) at high temperatures is highly desirable, but difficult to obtain experimentally. In this study, we use first-principles calculations, in conjunction with a quasi-harmonic Debye model, to evaluate the Helmholtz free energy of paramagnetic nickel as a function of temperature and volume, taking into account the electronic, magnetic, and vibrational contributions. The thermodynamic properties of Ni, such as the equilibrium lattice parameter and elastic moduli, are derived from the free energy in the temperature range from 800 to 1600 K and compared with available experimental data. The derived temperature dependence of the lattice parameter is then used for calculating the energies of intrinsic and extrinsic stacking faults in paramagnetic Ni. The stacking fault energies have been evaluated according to three different methodologies, the axial-next-nearest-neighbor Ising (ANNNI) model, the tilted supercell approach, and the slab supercell approach. The results show that the elastic moduli and stacking fault energies of Ni decrease with increasing temperature. This “softening” effect of temperature on the mechanical properties of nickel is mainly due to thermal expansion, and partly due to magnetic free energy contribution.
APA, Harvard, Vancouver, ISO, and other styles
26

Morishige, Taiki, Yuto Suzuki, and Toshihide Takenaka. "Extra-Hardening of SPD-Processed Al-Mg Alloy with Minimum Grain Sizes." Materials Science Forum 1016 (January 2021): 952–56. http://dx.doi.org/10.4028/www.scientific.net/msf.1016.952.

Full text
Abstract:
Severe plastic deformation (SPD) processing of Al alloys could obtain high strength by grain refinement mechanism. The minimum grain size of Al alloy, obtained at higher strain rate at low temperature, is determined the stacking fault energy of the alloy. SPD-processed pure Al metal, has high stacking fault energy, has relatively large grain size. During SPD processing, large strain is introduced, and the dislocation is rearranged in the specimen. The re-arrangement of dislocation in SPD-processed Al alloy with intermediate stacking fault energy significantly delayed, thus the strain remains in the grain interior. The extra-hardening, a kind of strain hardening, results from an incomplete of dynamic recrystallization during SPD processing. Al-Mg solid solution alloy has intermediate stacking fault energy and the minimum grain size of this alloy approaches about 200 nm after SPD. The mechanical property of this alloy is remarkably higher than the predictable strength by Hall-Petch relationship due to the extra-hardening. In addition, the increase in strength by the extra-hardening varies with the Mg content of Al-Mg alloy. In this study, the effect of Mg content, i.e. the stacking fault energy of the alloy, on the degree of the extra-hardening of SPD-processed Al-Mg alloy was investigated in terms of the dislocation density and low-angle grain boundary of the alloy.
APA, Harvard, Vancouver, ISO, and other styles
27

Kibey, S., J. B. Liu, M. J. Curtis, D. D. Johnson, and H. Sehitoglu. "Effect of nitrogen on generalized stacking fault energy and stacking fault widths in high nitrogen steels." Acta Materialia 54, no. 11 (June 2006): 2991–3001. http://dx.doi.org/10.1016/j.actamat.2006.02.048.

Full text
APA, Harvard, Vancouver, ISO, and other styles
28

Chew, E., H. H. Kim, C. Ferraris, Yong Hao Zhao, Enrique J. Lavernia, and C. C. Wong. "Effect of Ppm Level Dopant on Ductility of Ultrafine Grained Gold Wires." Materials Science Forum 633-634 (November 2009): 449–57. http://dx.doi.org/10.4028/www.scientific.net/msf.633-634.449.

Full text
Abstract:
The addition of calcium (Ca) simultaneously improves the ductility and strength of UFG Au wires. Based on the observation on stacking faults, microstructures, simulation results and significant effect of Ca on grain boundary related properties, it is inferred that segregation of Ca to stacking faults and grain boundaries has occurred to induce effective stacking fault energy (SFE) reduction and properties improvement. Considering the known greater impact of SFE in UFG/ NC metals, segregating dopants are proposed to be an effective strategy for achieving dual improvement in this class of materials. Also, dopant selection criteria for this purpose is also suggested and verified.
APA, Harvard, Vancouver, ISO, and other styles
29

Al-Ani, Oras A., J. P. Goss, N. E. B. Cowern, Patrick R. Briddon, Meaad Al-Hadidi, Raied Al-Hamadany, and M. J. Rayson. "A Density Functional Study of Iron Segregation at ISFs and Σ5-(001) GBs in mc-Si." Solid State Phenomena 242 (October 2015): 224–29. http://dx.doi.org/10.4028/www.scientific.net/ssp.242.224.

Full text
Abstract:
Removal of the dilaterous effects of iron in silicon is critical for the performance of multicrystalline silicon (mc-Si) solar cells, with internal gettering at extended defects including stacking faults and grain boundaries being one possibility. We present the results of a density function study of the behavoiur of iron at the intrinsic stacking fault and (001)–Σ 5 twist grain boundary, which both represent examples of fully bonded systems. Our results show iron is bound more strongly to the grain-boundary than the stacking fault, which we ascribe to a combination of Si-Fe chemistry and strain relaxation. However, we find that the binding energy of a single Fe atom to these extended defects is modest, and less than 0.5 eV.
APA, Harvard, Vancouver, ISO, and other styles
30

Li, Juan, and Christoph Kirchlechner. "Does the stacking fault energy affect dislocation multiplication?" Materials Characterization 161 (March 2020): 110136. http://dx.doi.org/10.1016/j.matchar.2020.110136.

Full text
APA, Harvard, Vancouver, ISO, and other styles
31

Fouchier, Marc, and John J. Boland. "Energy of Si(111) dimer-stacking-fault structures." Physical Review B 57, no. 15 (April 15, 1998): 8997–9002. http://dx.doi.org/10.1103/physrevb.57.8997.

Full text
APA, Harvard, Vancouver, ISO, and other styles
32

Zhi'an, Yang, and Wang Zhongguang. "The dependence of creep on stacking-fault energy." Philosophical Magazine A 63, no. 1 (January 1991): 87–94. http://dx.doi.org/10.1080/01418619108204594.

Full text
APA, Harvard, Vancouver, ISO, and other styles
33

Ferreira, P. J., and P. Müllner. "A thermodynamic model for the stacking-fault energy." Acta Materialia 46, no. 13 (August 1998): 4479–84. http://dx.doi.org/10.1016/s1359-6454(98)00155-4.

Full text
APA, Harvard, Vancouver, ISO, and other styles
34

Momeen and M. Y. Khan. "Stacking Fault Energy of Silicon Carbide (SiC) Polytypes." Crystal Research and Technology 30, no. 8 (1995): 1127–33. http://dx.doi.org/10.1002/crat.2170300821.

Full text
APA, Harvard, Vancouver, ISO, and other styles
35

Lee, Dong Nyung. "Effect of Stacking Fault Energy on Evolution of Recrystallization and Grain Growth Textures of Metals." Materials Science Forum 558-559 (October 2007): 93–100. http://dx.doi.org/10.4028/www.scientific.net/msf.558-559.93.

Full text
Abstract:
The drawing textures of aluminum, copper, gold, silver, and Cu-7.3% Al bronze wires are approximated by major <111>+minor <100>, except silver wire, which can have the <100> texture at extremely high reductions. The <111> component in the drawing textures of aluminum, copper, gold, and silver transform to the <100> component after recrystallization. On the other hand, the <111> deformation texture of the Cu-7.3% Al bronze wire, which has very low stackingfault- energy, remains unchanged after recrystallization. The <100> + <111> recrystallization textures change to the <111> texture after abnormal grain growth. The Brass component {110}<112> in rolling textures of high stacking-fault-energy metals such as aluminum, copper, Cu- 16% Mn, and Cu-1% P changes to the Goss orientation {110}<001> after recrystallization. However, the Brass orientation in rolling textures of low stacking-fault-energy fcc metals such as brass and silver appears to change to an orientation approximated by the {236}<385> orientation after annealing. The texture changes are discussed based on the strain-energy-release-maximization model for medium to high stacking-fault-energy metals and on grain growth for low stacking-fault energy metals.
APA, Harvard, Vancouver, ISO, and other styles
36

Liu, Lili, Liwan Chen, Youchang Jiang, Chenglin He, Gang Xu, and Yufeng Wen. "Temperature Effects on the Elastic Constants, Stacking Fault Energy and Twinnability of Ni3Si and Ni3Ge: A First-Principles Study." Crystals 8, no. 9 (September 14, 2018): 364. http://dx.doi.org/10.3390/cryst8090364.

Full text
Abstract:
The volume versus temperature relations for Ni 3 Si and Ni 3 Ge are obtained by using the first principles calculations combined with the quasiharmonic approach. Based on the equilibrium volumes at temperature T, the temperature dependence of the elastic constants, generalized stacking fault energies and generalized planar fault energies of Ni 3 Si and Ni 3 Ge are investigated by first principles calculations. The elastic constants, antiphase boundary energies, complex stacking fault energies, superlattice intrinsic stacking fault energies and twinning energy decrease with increasing temperature. The twinnability of Ni 3 Si and Ni 3 Ge are examined using the twinnability criteria. It is found that their twinnability decrease with increasing temperature. Furthermore, Ni 3 Si has better twinnability than Ni 3 Ge at different temperatures.
APA, Harvard, Vancouver, ISO, and other styles
37

Tochigi, E., Naoya Shibata, Atsutomo Nakamura, Takahisa Yamamoto, and Yuichi Ikuhara. "TEM Characterization of 2º Tilt Grain Boundary in Alumina." Materials Science Forum 561-565 (October 2007): 2427–30. http://dx.doi.org/10.4028/www.scientific.net/msf.561-565.2427.

Full text
Abstract:
Dislocation structure of {1120}/<1100> 2º tilt grain boundary in alumina was observed by transmission electron microscopy (TEM). The grain boundary consisted of periodical array of basal dislocations, which were dissociated into pairs of 1/3<1010> and 1/3<0110> partials with {1120} stacking-fault in between. The relationship between the separation distance of partials and the stacking-fault was modeled by considering the force balances of periodical dislocations. The estimated stacking-fault energy for 2o tilt grain boundary was consistent with the previous reports.
APA, Harvard, Vancouver, ISO, and other styles
38

Ikeda, Yuji, Fritz Körmann, Isao Tanaka, and Jörg Neugebauer. "Impact of Chemical Fluctuations on Stacking Fault Energies of CrCoNi and CrMnFeCoNi High Entropy Alloys from First Principles." Entropy 20, no. 9 (August 30, 2018): 655. http://dx.doi.org/10.3390/e20090655.

Full text
Abstract:
Medium and high entropy alloys (MEAs and HEAs) based on 3d transition metals, such as face-centered cubic (fcc) CrCoNi and CrMnFeCoNi alloys, reveal remarkable mechanical properties. The stacking fault energy (SFE) is one of the key ingredients that controls the underlying deformation mechanism and hence the mechanical performance of materials. Previous experiments and simulations have therefore been devoted to determining the SFEs of various MEAs and HEAs. The impact of local chemical environment in the vicinity of the stacking faults is, however, still not fully understood. In this work, we investigate the impact of the compositional fluctuations in the vicinity of stacking faults for two prototype fcc MEAs and HEAs, namely CrCoNi and CrMnFeCoNi by employing first-principles calculations. Depending on the chemical composition close to the stacking fault, the intrinsic SFEs vary in the range of more than 150 mJ/m 2 for both the alloys, which indicates the presence of a strong driving force to promote particular types of chemical segregations towards the intrinsic stacking faults in MEAs and HEAs. Furthermore, the dependence of the intrinsic SFEs on local chemical fluctuations reveals a highly non-linear behavior, resulting in a non-trivial interplay of local chemical fluctuations and SFEs. This sheds new light on the importance of controlling chemical fluctuations via tuning, e.g., the annealing condition to obtain the desired mechanical properties for MEAs and HEAs.
APA, Harvard, Vancouver, ISO, and other styles
39

Kondo, Y., and N. Miura. "Dislocation Substructure of Single Crystal Ni-Based Superalloy, CMSX-4, Crept at 1073-1273K and 250-600MPa." Materials Science Forum 638-642 (January 2010): 2268–73. http://dx.doi.org/10.4028/www.scientific.net/msf.638-642.2268.

Full text
Abstract:
The rate controlling mechanism for creep deformation of the single crystal Ni-based superalloy, CMSX-4, at 1073-1273K and 250-700MPa was investigated. Constant load tensile creep tests and creep interrupted tests up to the minimum creep stage, were conducted in air. And TEM observations carried out on creep interrupted specimens. Stacking faults in the ’ were observed on creep interrupted specimens at the temperature lower than 1223K and the stress higher than 500MPa. The number of stacking faults increases monotonously with a decrease in temperature, and remarkably with an increase in stress. The stacking fault formation depends on creep temperature and stress conditions. The stress exponent of minimum creep rate, n value, and the activation energy for creep, Qc, were constant for all creep test condition range. From these results, the stacking fault formation has no influence on creep resistance and the rate controlling mechanism for creep deformation at the low temperature and high stress condition was not thought to be shearing the ’, but movement of the mobile dislocations in the  channel as well as at the high temperature and low stress condition.
APA, Harvard, Vancouver, ISO, and other styles
40

Molnár, Dávid, Göran Engberg, Wei Li, and Levente Vitos. "Deformation Properties of Austenitic Stainless Steels with Different Stacking Fault Energies." Materials Science Forum 941 (December 2018): 190–97. http://dx.doi.org/10.4028/www.scientific.net/msf.941.190.

Full text
Abstract:
In FCC metals a single parameter – stacking fault energy (SFE) – can help to predict the expectable way of deformation such as martensitic deformation, deformation twinning or pure dislocation glide. At low SFE one can expect the perfect dislocations to dissociate into partial dislocations, but at high SFE this separation is more restricted. The role of the magnitude of the stacking fault energy on the deformation microstructures and tensile behaviour of different austenitic steels have been investigated using uniaxial tensile testing and electron backscatter diffraction (EBSD). The SFE was determined by using quantum mechanical first-principles approach. By using plasticity models we make an attempt to explain and interpret the different strain hardening behaviour of stainless steels with different stacking fault energies.
APA, Harvard, Vancouver, ISO, and other styles
41

Tian, Su Gui, Ben Jiang Qian, Yong Su, Hui Chen Yu, and Xing Fu Yu. "Influence of Stacking Fault Energy on Creep Mechanism of a Single Crystal Nickel-Based Superalloy Containing Re." Materials Science Forum 706-709 (January 2012): 2474–79. http://dx.doi.org/10.4028/www.scientific.net/msf.706-709.2474.

Full text
Abstract:
By means of calculating stacking fault energy (SFE), measuring creep properties and contrast analysis of dislocation configuration, an investigation has been made into the influence of the stacking fault energy on the creep mechanism of the single crystal nickel-based superalloy. Results show that the alloy at 760¡æ has a lower stacking fault energy (SFE), and the SFE of the alloy increases with the temperatures. The deformed mechanism of the alloy during creep at 760¡æ is the cubical γ′ phase sheared by <110> super-dislocation which may be decomposed to form the configuration of (1/3)<112> super-Shockley partials dislocation plus the superlattice intrinsic stacking fault (SISF). The deformed mechanism of the alloy which possesses the higher SFE at 1070¡æ is the screw or edge super-dislocation shearing into the rafted γ′ phase. The SFE of the alloy at 980¡æ is intervenient between the ones of 760¡æ and 1070¡æ, the deformation mechanism of the alloy during creep is the rafted γ′ phase sheared by <110> screw and edge super-dislocations which may be decomposed into the configuration of (1/2)<110> partial dislocation plus APB.
APA, Harvard, Vancouver, ISO, and other styles
42

Tochigi, E., A. Nakamura, Naoya Shibata, Takahisa Yamamoto, K. P. D. Lagerlöf, and Yuichi Ikuhara. "Dislocation Structure of 10° Low-Angle Tilt Grain Boundary in α-Al2O3." Materials Science Forum 558-559 (October 2007): 979–82. http://dx.doi.org/10.4028/www.scientific.net/msf.558-559.979.

Full text
Abstract:
Dislocation structure of 10º low-angle tilt grain boundary in α-Al2O3 has been observed by high-resolution electron microscopy (HRTEM). It was found that perfect <1120> edge dislocations, which are introduced to compensate the misorientation, dissociated into two mixed partial dislocations with {1120} stacking-fault in between. The distances between the two partials were estimated by the force balances between repulsive forces of periodical dislocations and attractive forces from stacking-fault. The stacking-fault energy for 10o low-angle tilt grain boundary was estimated to be much higher than the previously reported value.
APA, Harvard, Vancouver, ISO, and other styles
43

Ding, Zhigang, Shuang Li, Wei Liu, and Yonghao Zhao. "Modeling of Stacking Fault Energy in Hexagonal-Close-Packed Metals." Advances in Materials Science and Engineering 2015 (2015): 1–8. http://dx.doi.org/10.1155/2015/639519.

Full text
Abstract:
The deformation of metals is known to be largely affected by their stacking fault energies (SFEs). In the review, we examine the theoretical background of three normally used models, supercell model, Ising model, and bond orientation model, for the calculation of SFE of hexagonal-close-packed (hcp) metals and their alloys. To predict the nature of slip in nanocrystalline metals, we further review the generalized stacking fault (GSF) energy curves in hcp metals and alloys. We conclude by discussing the outstanding challenges in the modeling of SFE and GSF energy for studying the mechanical properties of metals.
APA, Harvard, Vancouver, ISO, and other styles
44

Vitos, L., P. A. Korzhavyi, J.-O. Nilsson, and B. Johansson. "Stacking fault energy and magnetism in austenitic stainless steels." Physica Scripta 77, no. 6 (May 21, 2008): 065703. http://dx.doi.org/10.1088/0031-8949/77/06/065703.

Full text
APA, Harvard, Vancouver, ISO, and other styles
45

Madhavan, R., R. Kalsar, R. K. Ray, and S. Suwas. "Role of stacking fault energy on texture evolution revisited." IOP Conference Series: Materials Science and Engineering 82 (April 24, 2015): 012031. http://dx.doi.org/10.1088/1757-899x/82/1/012031.

Full text
APA, Harvard, Vancouver, ISO, and other styles
46

Moon, W. J., T. Umeda, and H. Saka. "Temperature dependence of the stacking-fault energy in GaAs." Philosophical Magazine Letters 83, no. 4 (January 2003): 233–47. http://dx.doi.org/10.1080/0950083031000072480.

Full text
APA, Harvard, Vancouver, ISO, and other styles
47

Heino, P., L. Perondi, K. Kaski, and E. Ristolainen. "Stacking-fault energy of copper from molecular-dynamics simulations." Physical Review B 60, no. 21 (December 1, 1999): 14625–31. http://dx.doi.org/10.1103/physrevb.60.14625.

Full text
APA, Harvard, Vancouver, ISO, and other styles
48

Zheng, J. G., Q. Li, Z. G. Liu, D. Feng, and G. Frommeyer. "Complex stacking fault energy of Cr-alloyed γ-TiAl." Physics Letters A 196, no. 1-2 (December 1994): 125–27. http://dx.doi.org/10.1016/0375-9601(94)91056-1.

Full text
APA, Harvard, Vancouver, ISO, and other styles
49

Li, Ke-Qiang, Zhen-Jun Zhang, Lin-Lin Li, Peng Zhang, Jin-Bo Yang, and Zhe-Feng Zhang. "Effective Stacking Fault Energy in Face-Centered Cubic Metals." Acta Metallurgica Sinica (English Letters) 31, no. 8 (April 19, 2018): 873–77. http://dx.doi.org/10.1007/s40195-018-0718-4.

Full text
APA, Harvard, Vancouver, ISO, and other styles
50

Zhang, Yuefei, Jin Wang, Haiquan Shan, and Kejie Zhao. "Strengthening high-stacking-fault-energy metals via parallelogram nanotwins." Scripta Materialia 108 (November 2015): 35–39. http://dx.doi.org/10.1016/j.scriptamat.2015.05.039.

Full text
APA, Harvard, Vancouver, ISO, and other styles
We offer discounts on all premium plans for authors whose works are included in thematic literature selections. Contact us to get a unique promo code!

To the bibliography