Bauschinger effect

Last updated
The Bauschinger effect. Bauschinger effect.svg
The Bauschinger effect.

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. [1]

Contents

While more tensile cold working increases the tensile yield strength, the local initial compressive yield strength after tensile cold working is actually reduced. The greater the tensile cold working, the lower the compressive yield strength.

It is a general phenomenon found in most polycrystalline metals. Based on the cold work structure, two types of mechanisms are generally used to explain the Bauschinger effect:

  1. Local back stresses may be present in the material, which assist the movement of dislocations in the reverse direction. The pile-up of dislocations at grain boundaries and Orowan loops around strong precipitates are two main sources of these back stresses.
  2. When the strain direction is reversed, dislocations of the opposite sign can be produced from the same source that produced the slip-causing dislocations in the initial direction. Dislocations with opposite signs can attract and annihilate each other. Since strain hardening is related to an increased dislocation density, reducing the number of dislocations reduces strength.

The net result is that the yield strength for strain in the opposite direction is less than it would be if the strain had continued in the initial direction.

Mechanism of action

The Bauschinger effect is primarily attributed to the interaction between dislocations and the internal stress fields within the material. Initially, as external stress is applied, dislocations are generated and traverse the crystal lattice, creating internal stress fields. These fields, in turn, interact with the applied stress, leading to a phenomenon known as work hardening or strain hardening. With the accumulation of dislocations, the material's yield strength rises, hindering further plastic deformation. When stresses are applied in the reverse direction, the dislocations are now aided by the back stresses that were present at the dislocation barriers previously and also because the back stresses at the dislocation barriers in the back are not likely to be strong compared to the previous case. Hence the dislocations glide easily, resulting in lower yield stress for plastic deformation for reversed direction of loading. [2] [3]

Bauschinger effect, varies in magnitude based on factors like material composition, crystal structure, and prior plastic deformation. Materials with a higher density of dislocations and more internal stress fields tend to exhibit a more obvious Bauschinger effect. Additionally, the Bauschinger effect often accompanies other phenomena, such as Permanent Softening and Transient effects. [3] [4]

There is also a considerable amount of contribution of residual lattice stresses/strains to the Bauschinger Effect in materials that is associated with anisotropy in deformation. During loading-unloading cycles, dislocations do not return to their original position after unloading, which leaves residual strains in the lattice. These strains interact with stresses applied in the opposite direction which affect the materials response to subsequent loading-unloading cycles. The biggest effect observed is plastic yield asymmetry wherein the material will yield at different values in different loading directions. [1]

There are three types of residual stresses - type I, type II and type III that contribute to the Bauschinger effect in polycrystalline materials. Type I residual stresses arise during manufacturing due to thermal gradients and usually self-equilibrate over the length comparable to the macroscopic dimension of the material. So, they do not contribute significantly to the Bauschinger effect [2]. However, type II stresses equilibrate at the grain size scale and thus, contribute significantly to the Bauschinger effect. They result from strain incompatibility between neighboring grains due to plastic and elastic anisotropy. Thus, they are responsible to change the material's yield behavior along different directions by affecting dislocation motion along these differently oriented grains [3]. Type III stresses on the other hand arise due to mismatch between the soft matrix material and hard precipitates or dislocation cell walls (microstructural elements). They last over extremely short distances but significantly affect areas having microstructural heterogeneity. Dislocations pile-ups or stress concentration at the grain boundaries are examples of these type of residual stresses [4], [5].

As a whole, these three types of residual stresses impact properties like strength, flexibility, fatigue and durability. Thus, understanding the mechanism of residual stresses is important to mitigate the influence of the Bauschinger effect.

References

[1]  A. A. Mamun, R. J. Moat, J. Kelleher, and P. J. Bouchard, “Origin of the Bauschinger effect in a polycrystalline material,” Mater. Sci. Eng. A, vol. 707, pp. 576–584, Nov. 2017, doi: 10.1016/j.msea.2017.09.091.

[2]  J. Hu, B. Chen, D. J. Smith, P. E. J. Flewitt, and A. C. F. Cocks, “On the evaluation of the Bauschinger effect in an austenitic stainless steel—The role of multi-scale residual stresses,” Int. J. Plast., vol. 84, pp. 203–223, Sep. 2016, doi: 10.1016/j.ijplas.2016.05.009.

[3]  B. Chen et al., “Role of the misfit stress between grains in the Bauschinger effect for a polycrystalline material,” Acta Mater., vol. 85, pp. 229–242, Feb. 2015, doi: 10.1016/j.actamat.2014.11.021.

[4]  J. H. Kim, D. Kim, F. Barlat, and M.-G. Lee, “Crystal plasticity approach for predicting the Bauschinger effect in dual-phase steels,” Mater. Sci. Eng. A, vol. 539, pp. 259–270, Mar. 2012, doi: 10.1016/j.msea.2012.01.092.

[5]  C.-S. Han, R. H. Wagoner, and F. Barlat, “On precipitate induced hardening in crystal plasticity: theory,” Int. J. Plast., vol. 20, no. 3, pp. 477–494, Mar. 2004, doi: 10.1016/S0749-6419(03)00098-6.

Consequence of the Bauschinger effect

Metal forming operations result in situations exposing the metal workpiece to stresses of reversed sign. The Bauschinger effect contributes to work softening of the workpiece, for example in straightening of drawn bars or rolled sheets, where rollers subject the workpiece to alternate bending stresses, thereby reducing the yield strength and enabling greater cold drawability of the workpiece. [1] [2]

Implications

The Bauschinger effect have the applications in various fields due to its implications for the mechanical behavior of metallic materials subjected to cyclic loading. It is particularly relevant in applications involving cyclic loading or loading with changes in stress direction, facilitating the design and optimization of engineering structures.

Seismic Analysis: Earthquake engineering and seismic design are crucial aspects of geology engineering. During earthquakes, structural components endure alternating stress directions, with the Bauschinger effect influencing material response, energy dissipation, and potential damage accumulation. The Giuffré-Menegotto-Pinto model is widely utilized to accurately predict the seismic performance of structures by incorporating the Bauschinger effect. This model introduces a transition curve in the stress-strain relationship to capture both the Bauschinger effect and the pinching behavior observed in reinforced concrete structures under cyclic loading. [5] [6]

Fatigue Life Prediction: Researchers have developed methods and models to incorporate the Bauschinger effect into fatigue life prediction techniques, such as the strain-life and energy-based approaches. This plays a pivotal role in predicting and designing the fatigue life of machinery, vehicles, and engineering structures. A clear understanding of the Bauschinger effect ensures accurate predictions, enhancing the reliability and safety of components subjected to cyclic loading conditions. [7] [8]

The strain-life approach correlates the plastic strain amplitude with the number of cycles to failure, while the energy-based approach considers plastic strain energy as a driving force for fatigue damage accumulation. These models integrate the Bauschinger effect by adjusting the calculation of plastic strain energy or introducing additional energy terms to address the asymmetry in hysteresis loops caused by the effect. [7] [8] [9] [10]

Aerospace and Automotive Engineering: In aerospace engineering, materials undergo repeated loading cycles during flight, leading to fatigue and deformation. Similarly, in the automotive industry, vehicles endure cyclic loading due to road conditions and operations. Understanding the Bauschinger effect is crucial for predicting material behavior under such conditions and designing components with improved fatigue resistance. Research in this domain focuses on characterizing the Bauschinger effect in alloys and developing predictive models to assess fatigue life [11] , ensuring structural integrity and reliability.

Metal Forming: The Bauschinger effect significantly influences the material's flow behavior, strain distribution, and required forming loads during these processes. Hence, understanding the Bauschinger effect is significant for optimizing forming processes, predicting material behavior. [12] [9]

Mitigation of the Bauschinger effect

To mitigate the influence of the Bauschinger effect and enhance the performance of metallic materials, several strategies and techniques have been developed, including heat and surface treatments, the use of composite materials, and composition optimization.

Surface Treatment: This method aims to alleviate the Bauschinger effect by changing the surface properties of metallic materials. Common treatments include creating a protective layer or modifying the surface microstructure through processes such as physical vapor deposition (PVD) coatings. This treatment reduces the Bauschinger effect in the near-surface regions. Another effective approach is shot peening, where high-velocity particles impact the material's surface, inducing compressive residual stresses. These stresses counteract the internal tensile stresses associated with the Bauschinger effect to reduce its impact. [13] [14]

Heat Treatment: Heat treatment and thermomechanical processing are widely used to mitigate the Bauschinger effect by relieving residual stresses and dislocation structures within the material. Stress relief annealing is a common approach, where the material is heated to a specific temperature and held for a certain duration, allowing dislocations to rearrange and internal stresses to dissipate. This process reduces the Bauschinger effect by minimizing internal stress fields and achieving a more uniform distribution of dislocations. [15]

Composition Optimization and Composite Materials: Optimizing material composition is another effective approach for mitigation, as certain compositions and microstructures exhibit reduced Bauschinger effect. Materials with high stacking fault energy, such as aluminum alloys and austenitic stainless steels, tend to show less pronounced Bauschinger effect due to their enhanced ability to accommodate dislocations. Additionally, hybrid and composite materials offer mitigation potential. Metal-matrix composites (MMCs), for instance, consist of a metallic matrix reinforced with ceramic particles or fibers, which can reduce the Bauschinger effect by constraining dislocation motion in the matrix. Moreover, laminated or graded composite structures strategically combine different materials to mitigate the Bauschinger effect in critical regions while maintaining desired properties elsewhere. [16]

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">Fracture</span> Split of materials or structures under stress

Fracture is the appearance of a crack or complete separation of an object or material into two or more pieces under the action of stress. The fracture of a solid usually occurs due to the development of certain displacement discontinuity surfaces within the solid. If a displacement develops perpendicular to the surface, it is called a normal tensile crack or simply a crack; if a displacement develops tangentially, it is called a shear crack, slip band, or dislocation.

The field of strength of materials typically refers to various methods of calculating the stresses and strains in structural members, such as beams, columns, and shafts. The methods employed to predict the response of a structure under loading and its susceptibility to various failure modes takes into account the properties of the materials such as its yield strength, ultimate strength, Young's modulus, and Poisson's ratio. In addition, the mechanical element's macroscopic properties such as its length, width, thickness, boundary constraints and abrupt changes in geometry such as holes are considered.

<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">Fracture mechanics</span> Study of propagation of cracks in materials

Fracture mechanics is the field of mechanics concerned with the study of the propagation of cracks in materials. It uses methods of analytical solid mechanics to calculate the driving force on a crack and those of experimental solid mechanics to characterize the material's resistance to fracture.

<span class="mw-page-title-main">Work hardening</span> Strengthening a material through plastic deformation

In materials science, work hardening, also known as strain hardening, is the strengthening of a metal or polymer by plastic deformation. Work hardening may be desirable, undesirable, or inconsequential, depending on the context.

The Portevin–Le Chatelier (PLC) effect describes a serrated stress–strain curve or jerky flow, which some materials exhibit as they undergo plastic deformation, specifically inhomogeneous deformation. This effect has been long associated with dynamic strain aging or the competition between diffusing solutes pinning dislocations and dislocations breaking free of this stoppage.

<span class="mw-page-title-main">Fracture toughness</span> Stress intensity factor at which a cracks propagation increases drastically

In materials science, fracture toughness is the critical stress intensity factor of a sharp crack where propagation of the crack suddenly becomes rapid and unlimited. A component's thickness affects the constraint conditions at the tip of a crack with thin components having plane stress conditions and thick components having plane strain conditions. Plane strain conditions give the lowest fracture toughness value which is a material property. The critical value of stress intensity factor in mode I loading measured under plane strain conditions is known as the plane strain fracture toughness, denoted . When a test fails to meet the thickness and other test requirements that are in place to ensure plane strain conditions, the fracture toughness value produced is given the designation . Fracture toughness is a quantitative way of expressing a material's resistance to crack propagation and standard values for a given material are generally available.

<span class="mw-page-title-main">Yield (engineering)</span> Phenomenon of deformation due to structural stress

In materials science and engineering, the yield point is the point on a stress-strain curve that indicates the limit of elastic behavior and the beginning of plastic behavior. Below the yield point, a material will deform elastically and will return to its original shape when the applied stress is removed. Once the yield point is passed, some fraction of the deformation will be permanent and non-reversible and is known as plastic deformation.

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.

Methods have been devised to modify the yield strength, ductility, and toughness of both crystalline and amorphous materials. These strengthening mechanisms give engineers the ability to tailor the mechanical properties of materials to suit a variety of different applications. For example, the favorable properties of steel result from interstitial incorporation of carbon into the iron lattice. Brass, a binary alloy of copper and zinc, has superior mechanical properties compared to its constituent metals due to solution strengthening. Work hardening has also been used for centuries by blacksmiths to introduce dislocations into materials, increasing their yield strengths.

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

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

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.

In continuum mechanics, ratcheting, or ratchetting, also known as cyclic creep, is a behavior in which plastic deformation accumulates due to cyclic mechanical or thermal stress.

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

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

References

  1. 1 2 Soboyejo, Wole O. (2003). "7.12 Dislocation Pile-ups and Bauschinger Effect". Mechanical properties of engineered materials. Marcel Dekker. ISBN   0-8247-8900-8. OCLC   300921090.
  2. 1 2 Dieter, George E. (1988). Mechanical Metallurgy. McGraw Hill Book Company. pp. 236, 237. ISBN   0-07-084187-X.
  3. 1 2 Yoshida, Fusahito; Uemori, Takeshi; Fujiwara, Kenji (2002-10-01). "Elastic–plastic behavior of steel sheets under in-plane cyclic tension–compression at large strain". International Journal of Plasticity. 18 (5): 633–659. doi:10.1016/S0749-6419(01)00049-3. ISSN   0749-6419.
  4. Chaboche, J. L. (2008-10-01). "A review of some plasticity and viscoplasticity constitutive theories". International Journal of Plasticity. Special Issue in Honor of Jean-Louis Chaboche. 24 (10): 1642–1693. doi:10.1016/j.ijplas.2008.03.009. ISSN   0749-6419.
  5. Ibarra, Luis F.; Medina, Ricardo A.; Krawinkler, Helmut (2005-06-13). "Hysteretic models that incorporate strength and stiffness deterioration". Earthquake Engineering & Structural Dynamics. 34 (12): 1489–1511. Bibcode:2005EESD...34.1489I. doi:10.1002/eqe.495. ISSN   0098-8847.
  6. Nguyen, Van Tu; Vu, Ngoc Quang; Nguyen, Xuan Dai (2023). "Inelastic behavior of reinforced concrete buildings subjected to earthquakes". American Institute of Physics Conference Series. Advances in Sustainable Construction Materials. 2759 (1): 020056. Bibcode:2023AIPC.2497b0056N. doi:10.1063/5.0103457 . Retrieved 2024-05-12.
  7. 1 2 Lemaitre, Jean; Chaboche, Jean-Louis (1990). Mechanics of Solid Materials. Cambridge: Cambridge University Press. doi:10.1017/cbo9781139167970. ISBN   978-0-521-32853-1.
  8. 1 2 Jiang, Y.; Sehitoglu, H. (1996). "Modeling of Cyclic Ratchetting Plasticity, Part I: Development of Constitutive Relation". Journal of Applied Mechanics. 63 (3): 720–725. Bibcode:1996JAM....63..720J. doi:10.1115/1.2823355 . Retrieved 2024-05-10.
  9. 1 2 Banabic, Dorel (2010). Sheet Metal Forming Processes. doi:10.1007/978-3-540-88113-1. ISBN   978-3-540-88112-4.
  10. Buciumeanu, M.; Palaghian, L.; Miranda, A.S.; Silva, F.S. (2010-08-01). "Fatigue life predictions including the Bauschinger effect". International Journal of Fatigue. 33 (2): 145–152. doi:10.1016/j.ijfatigue.2010.07.012. ISSN   0142-1123.
  11. Zhu, Xianghui; Yang, Xusheng; Huang, Weijiu; Liu, Mofang; Wang, Xin; Li, Mengdi (2024-05-04). "Modeling Bauschinger effect through the reversibility of dislocation slip in an Al–Cu–Li alloy". Materials Science and Engineering: A. 902: 146574. doi:10.1016/j.msea.2024.146574. ISSN   0921-5093.
  12. Jiang, Y.; Sehitoglu, H. (1996). "Modeling of Cyclic Ratchetting Plasticity, Part I: Development of Constitutive Relation". Journal of Applied Mechanics. 63 (3): 720–725. Bibcode:1996JAM....63..720J. doi:10.1115/1.2823355 . Retrieved 2024-05-10.
  13. Li, Songbai; Liang, Wei; Yan, Hongzhi; Wang, Yuhang; Gu, Chu (2022-06-01). "Prediction of fatigue crack propagation behavior of AA2524 after laser shot peening". Engineering Fracture Mechanics. 268: 108477. doi:10.1016/j.engfracmech.2022.108477. ISSN   0013-7944.
  14. Rakita, Milan; Wang, Meng; Han, Qingyou; Liu, Yanxiong; Yin, Fei (2013). "Ultrasonic shot peening". International Journal of Computational Materials Science and Surface Engineering. 5 (3): 189. doi:10.1504/IJCMSSE.2013.056948. ISSN   1753-3465.
  15. Totten, George (28 September 2006). Steel Heat Treatment: Metallurgy and Technologies. CRC Press. ISBN   978-0-8493-8455-4.{{cite book}}: CS1 maint: date and year (link)
  16. Snitkoff, Joshua; Abdelfattah, Tarik; Russell, Ronnie (2016-09-26). "Flow-Biased Inflow Control Device Enhances Injection Rates While Still Maintaining Desired Production Resistance". Day 1 Mon, September 26, 2016. SPE. doi:10.2118/181492-ms.