Fact-checked by Grok 2 weeks ago

Work hardening

Work hardening, also known as hardening or , is the process by which the strength and of metallic materials increase during deformation at temperatures below the recrystallization range, typically below about 0.5 times the absolute . This phenomenon arises from the generation and accumulation of dislocations—linear defects in the crystal —that tangle and interact, creating barriers to further dislocation motion and thereby raising the stress required for continued flow. The effect is most pronounced in face-centered cubic (FCC) and body-centered cubic (BCC) metals, such as steels, aluminum, and , where dislocation density can reach up to 10^{16} m^{-2} before saturation. The theoretical foundation of work hardening was established by G. I. Taylor in 1934, who modeled plastic deformation in crystals by assuming that slip occurs along discrete planes and directions, with the shear stress proportional to the square root of the dislocation density according to the relation \tau = \alpha G b \sqrt{\rho}, where \tau is the shear stress, G is the shear modulus, b is the Burgers vector, \rho is the dislocation density, and \alpha is a constant. Subsequent models, such as those developed by U. F. Kocks and H. Mecking in the 1970s and 1980s, describe the evolution of dislocation density with plastic strain through storage and annihilation processes, explaining the characteristic stages of work hardening: easy glide with low hardening rate (Stage I), linear hardening with high rate (Stage II), and dynamic recovery with decreasing rate (Stage III). These interactions include junction formation, jogs, and long-range stress fields, which collectively impede dislocation glide and elevate the yield strength. While work hardening significantly boosts mechanical properties—such as increasing the tensile strength of low-carbon steels from around 400 to over 700 and yield strength to 600 —it simultaneously reduces , with elongation dropping to as low as 6-10% in heavily deformed states. To counteract this, processes like annealing are employed to relieve internal stresses and recrystallize the microstructure, restoring formability. In industrial applications, work hardening is intentionally exploited in cold rolling and of steels for components like wires, fasteners, and shafts in automotive and sectors, where enhanced strength without is desirable. It is also critical in aluminum alloys, such as 3000 and 5000 series, used in structures and beverage cans, where strain hardening during forming achieves strengths up to 300-400 while maintaining lightweight properties. Additionally, the process contributes to wear resistance in equipment and parts made from high-manganese steels.

Practical Applications and Effects

Undesirable Effects in Manufacturing

Undesirable work hardening refers to the unintentional strengthening of metals through plastic deformation during manufacturing processes such as cutting, , or stamping, which can lead to increased , cracking, or processing difficulties. This occurs when induces dislocation interactions that raise the material's yield strength without corresponding improvements in , complicating subsequent operations. In , work hardening creates a hardened surface layer on the workpiece, accelerating and reducing cutting efficiency. For instance, during the turning of austenitic stainless steels, the deformed layer can significantly increase surface , causing tools to dull faster and potentially leading to poor or part rejection. Similarly, in forming, work hardening exacerbates springback—the elastic recovery after deformation—resulting in dimensional inaccuracies that require additional corrections. In automotive panel stamping, this can cause significant deviations from specifications, increasing scrap rates and assembly challenges. To mitigate these effects, manufacturers employ strategies like using lubricants to minimize and heat buildup during forming, which limits excessive localization. Controlled temperatures, such as warming tools or workpieces to reduce strength without full annealing, help curb springback in aluminum sheets. Post-process annealing is also common, softening the material by recrystallizing the microstructure and restoring , as seen in applications where heating to 600–800°C relieves in cold-formed steels. In , applying coolants and optimizing feed rates prevents the hardened layer formation. Historically, in early 20th-century automotive stamping, springback and related work hardening issues in low-carbon panels prompted adjustments like overbending dies and intermediate annealing to achieve consistent body shapes in .

Intentional Processes for Strengthening

Work hardening is intentionally induced in metals through controlled plastic deformation to enhance mechanical properties without altering the . These methods exploit the increase in density to raise yield strength, often achieving significant improvements in fatigue resistance and wear properties while maintaining within limits. Common techniques include cold rolling, , cold forging, and , each applying precise strains at temperatures below the recrystallization point to prevent and maximize hardening effects. Cold rolling involves passing metal sheets or strips through rollers at , imposing compressive forces that reduce thickness and induce uniform plastic across the material. This process can increase tensile strength by up to 20% through work hardening, while also producing smoother surfaces compared to hot rolling. Typical levels correspond to 20-50% thickness reduction, with careful to avoid annealing and loss of hardening. It is widely used in producing high-strength sheets for automotive panels and structural components. Wire drawing pulls metal rods or wires through a conical die, progressively reducing the cross-sectional area and elongating the material to apply tensile plastic strain. Reductions in area of 20-40% per pass are common to balance hardening with , often requiring intermediate annealing for larger total strains to mitigate excessive work hardening that could lead to breakage. This method enhances yield strength and fatigue resistance, making it essential for high-strength cables used in bridges and elevators. Cold forging shapes metal billets or preforms using compressive dies at ambient temperatures, introducing localized plastic strains through repeated impacts or presses. Strain levels can reach high values depending on the geometry, with processes designed to limit total deformation to avoid cracking while promoting surface and subsurface hardening. The resulting improved wear properties and strength are critical for components like hardened cutting tools and fasteners in machinery. Temperature is maintained low to preserve the dislocation tangles responsible for strengthening. Shot peening bombards the metal surface with spherical media at high velocity, creating overlapping dimples that induce compressive plastic strain in a shallow layer, typically 0.1-0.5 mm deep. This process generates beneficial residual compressive stresses up to half the yield strength, enhancing fatigue life by 1.5-2 times in cyclically loaded parts. Parameters such as shot intensity and coverage (often 100-200%) are controlled to optimize hardening without distorting the component. It is routinely applied to springs, gears, and blades in and automotive industries to improve resistance to cracking and wear.

Fundamental Mechanisms

Elastic and Plastic Deformation Basics

Elastic deformation refers to the temporary and reversible change in a material's or dimensions when subjected to an applied below its yield point, allowing the material to return to its original configuration upon removal as atomic bonds stretch but do not break. In contrast, plastic deformation involves permanent alterations to the material's structure, occurring when stresses exceed the yield point and cause irreversible atomic rearrangements, such as the sliding of crystal planes. The behavior of materials under tensile loading is commonly represented by the stress-strain curve, derived from uniaxial tension tests, which plots against . This curve features an initial linear elastic region where is proportional to according to , up to the proportional limit; beyond this, the yield point marks the transition to the nonlinear plastic region, characterized by significant permanent deformation, eventual necking (localized reduction in cross-sectional area), and ultimate fracture. Key mechanical properties define these deformation regimes: , the slope of the stress-strain curve in the elastic region, quantifies a material's , with typical values for metals ranging from about 70 GPa for aluminum to 200 GPa for . measures the material's response to axial strain, typically around 0.3 for many metals, indicating the ratio of transverse contraction to longitudinal extension during elastic loading. The onset of plastic deformation is conventionally defined by the 0.2% offset yield strength, determined by drawing a line parallel to the elastic portion of the stress-strain curve but offset by 0.2% strain (0.002 in/in), with the intersection point specifying the yield stress as per ASTM E8 standards. At the microscopic level, plastic deformation in crystalline metals and alloys is enabled by the activation of slip systems—specific combinations of close-packed crystallographic planes and directions that allow layers of atoms to glide past one another under , accommodating large strains without . This process is mediated by dislocations, line defects in the crystal lattice that facilitate slip at stresses far below those required for ideal atomic shearing.

Dislocation Dynamics and Strain Fields

Dislocations are linear defects in the crystal that facilitate plastic deformation through their motion, and they are characterized by the \vec{b}, which measures the magnitude and direction of the distortion associated with the defect. The is determined by constructing a closed , known as the Burgers circuit, around the dislocation line in a distorted ; the required to close this is \vec{b}, typically equal in magnitude to the parameter in the direction of slip. Dislocations are classified based on the orientation of \vec{b} relative to the dislocation line direction \vec{\xi}. In an edge dislocation, \vec{b} is perpendicular to \vec{\xi}, representing the insertion or removal of an extra half-plane of atoms, which terminates at the dislocation core. A screw dislocation has \vec{b} parallel to \vec{\xi}, resulting in a shear distortion that resembles a helical ramp in the lattice. Mixed dislocations combine elements of both, with \vec{b} at an angle to \vec{\xi}, and their character is described by the angle between these vectors. The strain fields surrounding dislocations extend over long ranges and govern their interactions. For an edge dislocation, the strain field features regions of compression above the extra half-plane and tension below it, with additional shear components parallel to the Burgers vector; these fields decay as $1/r with distance r from the core, leading to long-range elastic interactions between dislocations. In contrast, a screw dislocation produces a pure shear strain field in planes perpendicular to \vec{\xi}, with no dilatation, also decaying as $1/r and enabling screw dislocations to cross-slip between planes. These long-range fields cause dislocations to repel or attract based on their type and orientation, influencing overall material hardening. Dislocation motion occurs primarily through glide, a conservative process where the dislocation line advances on its slip plane under applied , preserving the volume. Glide requires resolved in the direction of \vec{b} and is the dominant for flow at low temperatures. Climb, a non-conservative motion to the slip plane, involves the absorption or emission of point defects like vacancies, enabling edge components to move out of their glide plane; this diffusion-controlled process is significant only at elevated temperatures where atomic mobility is high. The driving force for dislocation motion under an applied \vec{\sigma} is quantified by the Peach-Koehler equation, which gives the force per unit length \vec{F} on the as \vec{F} = (\vec{\sigma} \cdot \vec{b}) \times \vec{\xi}. This expression arises from the virtual work principle, where the force balances the mechanical work done by the on the during infinitesimal displacement. For glide, the relevant component is the resolved parallel to \vec{b} in the slip , while climb forces involve coupled with .

Dislocation Multiplication and Hardening Stages

During plastic deformation, dislocations multiply through mechanisms such as the Frank-Read source, where a segment of dislocation line pinned at two points bows out under applied , expands into a full loop, and detaches, allowing the process to repeat and generate successive dislocation loops on the same slip plane. This mechanism operates when the stress exceeds a determined by the pinning distance L and the G, roughly \tau \approx G b / L, where b is the , enabling efficient multiplication within the crystal lattice. Another multiplication pathway involves dislocation pile-ups at barriers like grain boundaries or precipitates, where accumulated dislocations create high local stresses that can activate new sources or cause cross-slip, leading to further generation of mobile dislocations beyond the barrier. These processes collectively increase the dislocation density \rho, which starts low in annealed metals at approximately $10^6 cm^{-2} and can rise to $10^{12} cm^{-2} or higher after extensive deformation, as measured via diffraction line broadening in various metals. The progressive increase in dislocation density drives distinct hardening stages observed in single crystal deformation curves. In Stage I, known as easy glide, dislocations multiply primarily on a single slip system with minimal interactions, resulting in a low hardening rate of about 0.01 times the G. Stage II follows, characterized by linear hardening where dislocations from multiple slip systems intersect, forming jogs, Lomer-Cottrell locks, and other obstacles that impede motion, yielding a higher, athermal hardening rate around 0.1G. In Stage III, dynamic dominates through cross-slip and climb, annihilating some dislocations and reducing the hardening rate, often leading to a stress at large strains. This strengthening is quantitatively captured by the Taylor hardening relation, \tau = \alpha G b \sqrt{\rho}, where \tau is the shear stress required for further deformation, \alpha is a constant typically between 0.3 and 0.5 depending on the material and structure, G is the , b is the magnitude, and \rho is the density; the square-root dependence arises from the long-range stress fields of dislocations impeding each other's motion.

Quantitative Descriptions

Work Hardening Exponents and Models

The Hollomon equation provides a foundational mathematical for quantifying work hardening through a power-law between true and true . It is expressed as \sigma = K \epsilon^n where \sigma is the true , \epsilon is the true plastic , K is the strength coefficient representing the at \epsilon = 1, and n is the hardening exponent that measures the material's capacity to harden with deformation (typically $0 < n < 1). This equation derives from the empirical assumption that the flow follows a power-law dependence on , which linearizes in a log-log plot: \log \sigma = \log K + n \log \epsilon, where n is the slope of the resulting straight line fitted to experimental - data in the uniform deformation regime. The model captures the progressive increase in strength due to dislocation interactions during plastic flow. To address the limitation of the Hollomon equation in predicting unbounded stress at high strains, the Voce equation introduces a saturation behavior, modeling the approach to a maximum flow stress as dynamic recovery balances hardening. It is given by \sigma = \sigma_0 + (\sigma_s - \sigma_0) \left(1 - e^{-k \epsilon}\right) where \sigma_0 is the initial yield stress, \sigma_s is the saturation stress, and k is a rate parameter controlling the approach to saturation. This form arises from integrating a differential hardening rate that decreases exponentially with strain, reflecting physical processes like dislocation annihilation at higher deformations. The Voce model better describes the full tensile curve, particularly in regimes where hardening tapers off. For illustrative purposes, consider annealed , where n \approx 0.5 and K \approx 320 MPa in the Hollomon framework. At a true strain \epsilon = 0.1, the true stress is \sigma = 320 \times (0.1)^{0.5} \approx 101 MPa, demonstrating a substantial increase from near-zero initial plastic stress; by \epsilon = 0.5, \sigma \approx 226 MPa, highlighting the nonlinear strengthening effect. This example underscores how the exponent governs the rate of stress buildup during deformation. Despite their utility, these models have limitations, including the assumption of isotropic hardening behavior, which overlooks directional variations in properties, and neglect of texture development that can induce anisotropy in polycrystalline materials. These simplifications make them less suitable for complex deformation paths or textured alloys without modifications.

Empirical Hardening Laws

Empirical hardening laws are experimentally derived relationships that approximate the stress-strain behavior of metals during plastic deformation, enabling practical predictions in engineering design and simulation. These laws are fitted to tensile test data and prioritize accuracy across a range of strains, particularly where simple power-law models deviate, such as at low strains or in prestrained conditions. Unlike mechanistic models based on dislocation theory, empirical laws focus on curve-fitting for direct application in process modeling. The Ludwigson relation addresses limitations of the basic power-law model, \sigma = K \epsilon^n, by incorporating an exponential term to better capture the transient region at low strains in face-centered cubic (FCC) metals and alloys. It is expressed as \epsilon = \alpha \exp(-\beta \sigma) + \frac{1}{n} \left( \frac{\sigma}{K} \right)^{1/n}, where \alpha and \beta describe the initial nonlinear yielding, K is the strength coefficient, and n is the hardening exponent. This formulation provides superior fits for materials exhibiting sigmoidal stress-strain curves, such as austenitic stainless steels, with reduced errors in the strain range of 0 to 0.05 compared to the power law. The Swift equation extends hardening modeling to predict necking instability in sheet forming by accounting for initial plastic strain from prior deformation. It takes the form \sigma = K (\epsilon + \epsilon_0)^n, where \epsilon_0 is the prestrain, allowing calculation of the uniform elongation at necking onset as \epsilon_u = n - \epsilon_0. This relation is particularly useful for analyzing formability limits in processes involving prestrained sheets, improving predictions of diffuse necking under plane stress conditions. These laws are widely applied in finite element (FE) simulations of metal forming processes, such as deep drawing and incremental forming, to model evolving material properties and anticipate defects like thinning or fracture. In simulations of single-point incremental forming of aluminum sheets, for instance, using Ludwigson or Swift laws enhances accuracy in predicting force requirements and final geometry by better representing strain-dependent hardening. Typical hardening exponents n from power-law fits vary by material, reflecting differences in ductility and work-hardening capacity; values are derived from tensile tests and used to select appropriate empirical models.
Material CategoryTypical Hardening Exponent n RangeExample K (MPa)
Low-carbon steels0.15–0.25500–600
Aluminum alloys0.20–0.30150–250
Copper0.40–0.50300–400
These ranges are based on annealed conditions and can shift with alloying or processing.

Behavior in Selected Materials

Ferrous Alloys like Steel

In low-carbon steels, work hardening manifests through significant plastic deformation capacity owing to the predominantly ferritic microstructure, which provides high ductility with uniform elongations often exceeding 20%. The strain hardening behavior follows the \sigma = K \epsilon^n, where the exponent n is typically around 0.2, indicating moderate work hardening that balances strength gains with formability. Following deformation, strain aging effects arise as solute atoms such as carbon and nitrogen migrate to dislocations, pinning them and resulting in yield point return, increased lower yield strength by up to 50 MPa, and a corresponding reduction in ductility by 5-10%. This phenomenon, prominent in interstitial-free or low-alloyed variants, underscores the time- and temperature-dependent recovery of strength post-forming. High-strength ferrous alloys, particularly dual-phase (DP) steels comprising ferrite and martensite, exhibit enhanced work hardening due to the transformation-induced plasticity (TRIP) mechanism in retained austenite phases. During straining, the austenite transforms progressively to martensite, generating a high work hardening rate in the later deformation stages—often exceeding 1000 MPa in instantaneous flow stress increase—thereby delaying necking and achieving total elongations over 25% alongside ultimate tensile strengths above 800 MPa. This TRIP-assisted hardening is particularly effective in automotive-grade DP steels with 10-20% martensite volume fraction, where the phase transformation provides dispersion strengthening and geometric incompatibility that impedes dislocation motion. Microstructural evolution in ferrous alloys with mixed ferrite-pearlite structures drives pronounced Stage II work hardening, characterized by a linear increase in flow stress with strain due to intense dislocation storage and interactions. Pearlite lamellae, consisting of alternating ferrite and cementite plates, constrain deformation in adjacent ferrite grains, promoting heterogeneous dislocation accumulation at phase boundaries and elevating the hardening rate to levels where Stage II dominates up to 10-15% strain. These interactions, as observed in hypoeutectoid steels with 0.2-0.5% carbon, result in cementite refinement and ferrite substructure development, sustaining high flow stresses without early saturation. In general hardening stages, this aligns with the multi-stage dislocation multiplication framework, where Stage II reflects balanced multiplication and annihilation. A key application of work hardening in ferrous alloys is the production of cold-drawn rebar for construction, where controlled deformation enhances mechanical performance in reinforced concrete elements. Applying approximately 30% true strain through cold drawing doubles the yield strength—from around 400 MPa in hot-rolled stock to over 800 MPa—via dislocation density increases and microstructural refinement, while maintaining sufficient ductility for bonding with concrete. This process, common in grades like , exploits the ferrite-pearlite matrix to achieve cost-effective strengthening without heat treatment, supporting seismic-resistant structures.

Non-Ferrous Metals like Copper and Aluminum

Work hardening in copper, a face-centered cubic (FCC) metal, is characterized by its high stacking fault energy (SFE), typically ranging from 45 to 78 mJ/m², which facilitates easy cross-slip of dislocations on multiple slip planes, contributing to pronounced strain hardening during deformation. This high SFE promotes dynamic recovery through cross-slip, yet allows for significant accumulation of dislocations in stages I and II, leading to a work hardening exponent (n) of approximately 0.5 in the power-law description of the stress-strain relationship. Stress-strain curves for pure copper exhibit saturation in stage IV at flow stresses around 500 MPa after extensive deformation, reflecting the balance between dislocation storage and annihilation. In practical applications, such as electrical wiring, pure copper (99.9% purity) is often cold-drawn to achieve up to 99% reduction in cross-sectional area, resulting in hard-drawn wire with enhanced tensile strength exceeding 400 MPa while maintaining high electrical conductivity. This severe plastic deformation induces a high density of dislocations and a fibrous microstructure, which increases the material's resistance to further deformation without requiring alloying elements. Aluminum, another FCC metal valued for its low density (approximately 2.7 g/cm³), undergoes work hardening with a lower exponent of about 0.2, indicating less pronounced strain hardening compared to copper due to its propensity for rapid recovery even at ambient temperatures. In alloys like 6061 (Al-Mg-Si), work hardening interacts with precipitation strengthening, where dislocations pin at fine precipitates formed during aging, further elevating yield strength to over 250 MPa in the T6 temper while allowing controlled ductility. Key differences arise from aluminum's lower melting point (660°C versus 1085°C for copper), which accelerates thermally activated recovery processes and initiates stage III (cross-slip dominated recovery) at lower strains and temperatures than in copper. During rolling, both metals develop characteristic deformation textures, such as the copper-type ({112}<111>) in copper and brass-type ({110}<112>) in aluminum, but aluminum exhibits more rapid texture evolution toward stable orientations due to its higher energy relative to solute effects in alloys.

Precious Metals like Gold

Precious metals such as exhibit exceptional work hardening behavior due to their high-purity face-centered cubic (FCC) crystal structures, which enable extensive plastic deformation before significant strengthening occurs. , in particular, demonstrates extremely high with a (n) of approximately 0.4, allowing it to undergo substantial without fracturing. 's high ductility arises from its FCC structure with 12 slip systems and low Peierls stress, enabling easy initial dislocation glide and uniform distribution. The ultra-high ductility of gold enables it to be hammered into extremely thin sheets known as , achieving thicknesses less than 0.1 μm through iterative cycles of work hardening and annealing. During hammering, dislocations multiply and tangle, increasing strength and allowing further deformation without rupture; subsequent annealing at low temperatures around 200°C recrystallizes the microstructure, restoring for additional working. This process exploits work hardening to progressively refine the material while maintaining its malleability, essential for applications in and . Ancient gold beating techniques, practiced for over 5,000 years, relied on this controlled hardening through repeated manual hammering to produce durable yet thin sheets for prestige items. Other precious metals share similar FCC structures but present distinct challenges. Silver, akin to gold in ductility and work hardening response, is also highly malleable but susceptible to tarnishing from sulfide formation, which can compromise aesthetic applications unless protected. Platinum, while exhibiting higher inherent strength and a more pronounced work hardening rate than gold, is favored in jewelry for its durability after deformation processes like rolling or . In wire , work hardening imposes limits on reduction per pass for these metals, typically requiring intermediate annealing to prevent cracking, with gold and silver needing more frequent recrystallizations due to their initial softness compared to platinum.

Emerging Materials and Polymers

In thermoplastics such as , strain hardening arises primarily from the alignment and orientation of polymer chains within the amorphous regions during deformation. As stress is applied, entangled molecular chains in the non-crystalline phase straighten and align along the direction of strain, forming a more ordered network that resists further deformation and increases the material's tensile strength and . This mechanism is particularly pronounced in , where the strain hardening is governed by the properties of the entangled chain network, leading to a characteristic upturn in the stress-strain curve after initial yielding. In glassy polymers, work hardening often manifests through , a localized deformation process that creates a network of microvoids interconnected by highly stretched . These , formed by the pulling apart of chains under tensile , bridge the voids and provide additional load-bearing capacity, resulting in an overall increase in the material's resistance to deformation. The work-hardening rate in such systems quantifies this non-elastic response, reflecting the energetic contributions from chain entanglements and the transition from initial softening to hardening as mature. Unlike chain alignment in semi-crystalline thermoplastics, crazing in glassy polymers like involves significant volumetric dilation but contributes to enhanced before . Nanocrystalline metals exhibit enhanced work hardening attributed to effects, where the high density of boundaries impedes motion more effectively than in coarse-grained counterparts. This strengthening follows the Hall-Petch relation, \sigma_y = \sigma_0 + k d^{-1/2}, with yield strength \sigma_y increasing as d decreases due to the accumulation of at boundaries and the requirement for higher stresses to activate slip across grains. For grain sizes above approximately 100 nm, this leads to significant hardening during deformation, though the relation may soften at ultra-small scales below 30 nm due to boundary-mediated mechanisms like sliding. Surveys of pure metals confirm the robustness of this enhancement, with grain refinement via severe plastic deformation yielding hardness values up to several GPa in materials like nanocrystalline . Post-2010 developments have extended work hardening to metallic glasses, traditionally prone to softening, through composite architectures and processing innovations. Bulk metallic glass composites incorporating transformation-mediated , such as those with β-Ti dendrites in a Zr-based amorphous matrix, demonstrate pronounced work hardening with tensile elongations exceeding 10%, as phase transformations generate compatible bands and back stresses that stabilize deformation. techniques, involving high-energy treatments to increase free volume, further enable hardening in macroscopic samples, achieving hardening rates up to 44 times the stress and plastic strains of 2.7–3.8% by suppressing catastrophic banding. These advances, highlighted in reviews of metallic glass mechanics, underscore their potential for structural applications despite limited overall . Additive manufacturing introduces anisotropic work hardening in alloys owing to the and microstructural gradients inherent in layer-by-layer fabrication. In 3D-printed Al-Si alloys like AlSi10Mg, the hardening response varies with build direction: horizontal orientations often exhibit higher uniform elongations and work-hardening capacities due to finer, more isotropic microstructures, while vertical builds show reduced hardening from columnar grains aligned with the build axis. This manifests in variable work-hardening exponents, with values differing by up to 20% between directions in , as layer interfaces act as barriers to propagation, inducing hetero-deformation effects. Such tailored hardening enhances strength-ductility synergies, as seen in AM-produced Ti6Al4V-316L composites reaching 1.3 GPa ultimate strength with 9% .

References

  1. [1]
    Work Hardening - an overview | ScienceDirect Topics
    Work hardening, also known as strain hardening or cold working, is a phenomenon that occurs during the plastic deformation of metallic materials.
  2. [2]
    Hardness and work hardening - DoITPoMS
    It is observed that the hardness and strength of materials are increased in the area of plastic deformation. This process is called work hardening, strain ...
  3. [3]
    The mechanism of plastic deformation of crystals. Part I.—Theoretical
    Taylor Geoffrey Ingram. 1934The mechanism of plastic deformation of crystals. Part I.—TheoreticalProc. R. Soc. Lond. A145362–387http://doi.org/10.1098/rspa ...
  4. [4]
    [PDF] Dislocation evolution during plastic deformation - Harvard University
    Work hardening is worse”, remarked Cottrell.1. Work hardening, a mechanism that occurs in crystalline metals, manifests as a rise in the stress required for ...
  5. [5]
    nglos324 - workhardening
    Dislocations on intersecting slip planes permit both elastic interactions and dislocation reactions to contribute to work hardening.
  6. [6]
    Work Hardening of Engineering Steels - AZoM
    Apr 17, 2001 · Work hardening is the resistance to of a material to plastic flow. Within certain limits, it can improve the mechanical properties of carbon ...Missing: definition | Show results with:definition
  7. [7]
    Work Hardening Aluminum Alloys: Part One - Total Materia
    This article examines the fundamental principles and applications of strain hardening in aluminum alloys.
  8. [8]
    The Benefits of Work Hardening Wear Steel - Titus Steel
    Jul 29, 2024 · Applications of Work Hardened Wear Steel ; Mining Equipment: Crusher jaws, hammers, and grinding mills ; Construction Machinery: Excavator buckets ...
  9. [9]
    Work Hardening & Strain Hardening [Machinist's Guide]
    Feb 26, 2024 · Undesirable Strain Hardening ... If nothing else, the material will be harder to machine, cut through, or form during the manufacturing process.
  10. [10]
    How to Prevent Work Hardening in Machining - Kennametal
    Oct 30, 2025 · Work hardening can be a significant challenge in machining, leading to increased tool wear, reduced efficiency, and compromised part quality.
  11. [11]
    Tool wear and machined surface characteristics in side milling ...
    Apart from tool wear, severe work hardening and plastic deformation deteriorate the machined surface quality [2]. ... tool wear effects in machining ...
  12. [12]
    Springback - AHSS Guidelines
    Decades ago, the major concern in sheet metal forming was elimination of necks and tears. These forming problems are a function of plastic strain, and ...
  13. [13]
    Die Science: Preventing springback in thin, high-strength metals
    Mar 27, 2019 · Work Hardening – No Strain, No Gain. For a stamping to retain its shape, it must be adequately strained. This means you must form the metal in a ...<|control11|><|separator|>
  14. [14]
    Springback Mitigation and Process Optimization in Sheet Metal ...
    Oct 13, 2025 · Control of lubrication, temperature, and forming speed reduces uneven strain in the part. Multi-stage forming sequences often yield better ...Missing: annealing | Show results with:annealing
  15. [15]
    Tool temperature control to increase the deep drawability of ...
    The effect of tool temperature on the reduction of springback amount of aluminum 1050 sheet has been investigated in this study. As the springback phenomenon ...Missing: mitigation | Show results with:mitigation
  16. [16]
    Predicting springback in V-bending: Effects of load, load holding ...
    Annealing reduces springback; quenching has the greatest effect on steel. •. Synergistic effects of load, holding time, or heat treatment reduce springback.<|control11|><|separator|>
  17. [17]
    Sheet Metal Basics - Old Cars Weekly
    Jan 25, 2019 · This process is called “annealing.” Auto body sheet metal will lose the effects of work hardening if it is heated to temperatures of about 1,600 ...
  18. [18]
    Types of Cold Working Processes: Definitions, Applications, and…
    Sep 12, 2022 · Work hardening occurs when metal is put under stress and atoms in its grains dislodge and “lock” into place. When this happens over and over ...Missing: shot peening
  19. [19]
    The Different Types of Cold Work Metalworking Processes - AZoM
    Nov 23, 2022 · The four main types of cold working processes are: Squeezing, Bending, Shearing, and Drawing. Precision stamping is a complex process that ...Missing: shot peening
  20. [20]
    [PDF] PROCESS CHARACTERISTICS OF WIRE DRAWING
    Yield strength depends on material composition but typically the reduction in area through a die is in the region of 20 to 40%. If a greater reduction is needed ...
  21. [21]
    Cold Drawing - an overview | ScienceDirect Topics
    As shown in Fig. 4.7, in Ti–Ni wires, cold working of over 50% would cause work hardening, which increases yield stress and tends to reduce ductility ( ...
  22. [22]
    Cold forging | OpenLearn - Open University
    Nov 22, 2017 · Mild steel has to be annealed before cold forging and at interstage processes. Cold forging causes strain hardening in the materials, resulting ...
  23. [23]
    [PDF] SHOT PEENING OVERVIEW - shotpeener.com
    Jan 18, 2001 · Shot peening is a cold working process ... increasing the surface hardening through cold working of the metal and providing residual compressive.Missing: intentional | Show results with:intentional
  24. [24]
    [PDF] Shot Peening Applications - Curtiss-Wright Surface Technologies
    ... Peening of Al 6061-T6 of the residual stresses of a laser peened surface is less than a shot peened surface due to the reduced cold work that is involved ...
  25. [25]
    Properties of Metals | Engineering Library
    The phenomenon of elastic strain and plastic deformation in a material are called elasticity and plasticity, respectively. At room temperature, most metals ...
  26. [26]
    [PDF] ASTM E8
    of offset used should be stated in parentheses after the term yield strength. Thus: Yield strength (offset 0.2 %) = 52 000 psi. In using this method a Class ...
  27. [27]
    [PDF] Crystal structures and their slip systems; dislocation shear stress ...
    30 sep 2009 · required to begin plastic deformation or slip. □ Temperature, strain rate, and material dependent. □ The system on which slip occurs has.
  28. [28]
    nglos324 - dislocation
    Three generic dislocation types are identified: (1) Edge dislocations with b normal to t, (2) Screw dislocation with b parallel to t, and (3) Mixed ...
  29. [29]
    5.1.2 Volterra Construction and Consequences - Dallas R. Trinkle
    Edge- and screw dislocations (with an angle of 90° or 0°, resp., between the Burgers- and the line vector) are just special cases of the general case of a mixed ...Missing: types | Show results with:types
  30. [30]
    Theory of Dislocations - Peter M. Anderson, John P. Hirth, Jens Lothe
    Professor Lothe's research interests include nucleation theory, dislocation theory, and elastic wave theory for anisotropic media. Lothe is author or coauthor ...<|separator|>
  31. [31]
    5.2.2 Stress Field of a Straight Dislocation - Dallas R. Trinkle
    The stress field of an edge dislocation is somewhat more complex than that of a screw dislocation, but can also be represented in an isotropic cylinder by the ...
  32. [32]
    Climb and cross slip - DoITPoMS
    Climb is the mechanism of moving an edge dislocation from one slip plane to another through the incorporation of vacancies or atoms. Climb can be either ...
  33. [33]
    The Forces Exerted on Dislocations and the Stress Fields Produced ...
    The Forces Exerted on Dislocations and the Stress Fields Produced by Them. M. Peach and J. S. Koehler* ... 80, 436 – Published 1 November, 1950. DOI: https ...
  34. [34]
    III. Dislocation densities in some annealed and cold-worked metals ...
    In annealed metals the values of p range from 2 × 107 cm of dislocation line per cm3 for aluminium to 3 × 108 for tungsten and molybdenum, the majority of ...
  35. [35]
    Hardening Exponent - an overview | ScienceDirect Topics
    The simplest model for work hardening is known as power law hardening (or Ludwik/Holloman) equation, i.e.. [11.3] σ = K 1 ε n. where K1 is known as the ...<|separator|>
  36. [36]
    Metal Properties: Strain-Hardening Exponent (n-Value)
    Jan 20, 2023 · For metal alloys where the Holloman relationship accurately represents the flow curve, n-value is the slope of eq. 2, and K is the true stress ...
  37. [37]
    Constitutive equations optimized for determining strengths of ...
    The most commonly used constitutive equation is the Hollomon equation, which has the form of a power-law relation between true stress σ and strain ε; σ = K ε n ...
  38. [38]
    Analysis of elevated temperature flow and work hardening ...
    The Voce equation was used to model the tensile flow and work-hardening behaviour of a service-exposed 2.25Cr-1Mo steel. The applicability of Voce ...
  39. [39]
    Limitations of the Hollomon strain-hardening equation - IOPscience
    It is concluded that for maximum strength and strain hardening in materials obeying the Hollomon equation, large values of K and n values between 0·1 and 0·3 ...
  40. [40]
    (PDF) Limitations of Hollomon and Ludwigson stress strain relations ...
    It is shown that the deviation from the ideal Hollomon relation in describing the stress-strain behaviour is characteristic of all materials at low strains.
  41. [41]
    Elastoplastic Materials - COMSOL 6.3
    For simplicity, Equation 3-91 assumes linear isotropic hardening ... The Ludwik equation (also called the Ludwik–Hollomon equation) for isotropic hardening is ...
  42. [42]
    Effect of hardening law and process parameters on finite element ...
    The purpose of this work was to study the effect of these process parameters as well as hardening law on single point incremental forming (SPIF) process.
  43. [43]
    [PDF] Types of Strain
    FIGURE 2.8 The effect of strain-hardening exponent n on the shape of true stress- true strain curves. When n = 1, the material is elastic, and when n = 0, it is.
  44. [44]
    Multistage strain aging phenomenon of low-carbon steels with ...
    Multistage strain aging increased samples' strength by 10% and reduced their plasticity by 5%. To evaluate the effects of temperature on the time, we propose ...
  45. [45]
    (PDF) Multistage strain aging of low-carbon steels - ResearchGate
    Aug 9, 2025 · It was found that in general multistage strain aging was detrimental to the ductility of the material. It was also found that strain aging was ...
  46. [46]
    Transformation-induced plasticity (TRIP) in advanced steels: A review
    Sep 23, 2020 · This transformation substantially raises the work-hardening rate and strengthens the material effectively at the neck region (that experienced a ...
  47. [47]
    Transformation Induced Plasticity (TRIP) - AHSS Guidelines
    However, in TRIP steels the retained austenite also progressively transforms to martensite with increasing strain, thereby increasing the work hardening rate at ...
  48. [48]
    [PDF] Strain-hardening characteristics of ferrite layers in pearlite ...
    Jun 8, 2018 · Strain hardening of ferrite layers in pearlite microstructures plays a crucial role in the stability of elasto-plastic deformation of ...
  49. [49]
    Production and Characteristics of High Strength Reinforcement Bars
    Jul 16, 2020 · In this method, the reinforcement bars are submitted to a strain hardening after hot rolling. For such reinforcement bars, the yield strength ...Missing: doubling | Show results with:doubling
  50. [50]
    Enhanced ductility of strong cold-rolled ribbed rebar through ...
    Nov 15, 2021 · In this study, intermediate frequency induction heat treatment (IFIHT) is applied to cold-rolled ribbed steel rebar with low-cost Q235 as raw material.Missing: doubling | Show results with:doubling
  51. [51]
    (PDF) The Dependence of Cross-slip on Stacking-Fault Energy in ...
    Aug 5, 2025 · Abstract. The results of an experimental study of the temperature and strain-rate dependence of τIII for Cu–Zn alloys are described and ...Missing: exponent | Show results with:exponent
  52. [52]
    Experimental and Numerical Study on the Effect of Strain Hardening ...
    Dec 15, 2020 · The results were showed that the strain hardening exponent of pure Copper more than Brass (70-30) and its value reached (0.54) for pure Copper ...
  53. [53]
    High strength and high conductivity pure copper prepared by ECAP ...
    To eliminate work hardening, the copper rods were annealed at 600 °C for 1 h ... 500 MPa, but also the conductivity is maintained at about 90 %IACS. To ...
  54. [54]
    [PDF] Influence of cold reduction on the structure and hardness of ... - UNIZG
    Deformation and reduction of the wire's cross-section is achieved by the reduction in the cross-section of the die opening during cold drawing. The 99,90 % pure ...
  55. [55]
    (PDF) Work hardening in aluminium alloys - ResearchGate
    Dec 21, 2022 · This chapter deals with the process of work hardening in aluminium and a number of its alloys. The approach taken is to develop a simple one parameter model.<|control11|><|separator|>
  56. [56]
    Tensile properties of AA6061 in different designated precipitation ...
    In most application of aluminum–magnesium–silicon alloys, both work hardening and precipitation strengthening are important contributors of strength.
  57. [57]
    The microstructural origin of work hardening stages - ScienceDirect
    Apr 15, 2018 · The strain evolution of the flow stress and work hardening rate in stages III and IV is explored by utilizing a fully described deformation microstructure.
  58. [58]
    [PDF] development of texture in copper by cold. rolling
    The texture of severely cold-rolled copper is similar to the texture of other face-centered cubic metals and alloys, such as aluminum, nickel and alpha brass.
  59. [59]
    (PDF) Texture evolution during rolling of aluminium alloys
    Aug 7, 2025 · The crystallographic texture evolution during rolling of Aluminium is described and its principles are discussed.
  60. [60]
    Effect of strain-hardening rate on the grain-to-grain variability of local ...
    Aug 9, 2025 · Micro-indentation hardness was used to assess the local equivalent plastic strain distribution in four fcc metals subjected to identical ...Missing: exponent | Show results with:exponent
  61. [61]
    Why do FCC metals with low stacking fault energy have a ... - Quora
    Nov 19, 2017 · FCC metals with low stacking fault energy have a higher strain hardening exponent (n) than FCC metals with high stacking fault energy and BCC alloys.
  62. [62]
    Deformation mechanism behind the unique texture of Kanazawa ...
    Sep 26, 2025 · ... gold leaf is produced with a thickness of approximately 0.1 µm (Fig. 1c). The second hammering stage has two key features: the leather ...
  63. [63]
    Recrystallization Annealing – A Comprehensive Guide - VIRGAMET
    Jun 30, 2025 · gold: recrystallization 200°C, melting 1063°C; cadmium: recrystallization ~20°C, melting 321°C; cobalt: recrystallization 550°C, melting 1490°C ...<|control11|><|separator|>
  64. [64]
    The ancient craft of gold beating - Gold Bulletin
    ### Summary of Ancient Gold Beating Techniques and Work Hardening
  65. [65]
    Solid solution softening and enhanced ductility in concentrated FCC ...
    Jun 27, 2018 · In this paper, the mechanical properties of additional concentrated silver solid solution phases with other solute elements, Al, Ga and Sn, have ...
  66. [66]
    The Hardening of Platinum Alloys for Potential Jewellery Application
    Alloying platinum increases its hardness significantly. However, platinum alloys used in jewellery do need to be easy to work and thus the alloy should be ...
  67. [67]
    [PDF] COLD WORK AND ANNEALING OF KARAT GOLD JEWELRY ...
    Susz recommends that we anneal this alloy roughly between 600°C and 650°C, as there is a dramatic drop in the hardness only after five minutes of annealing time ...
  68. [68]
    Strain hardening of high density polyethylene - ResearchGate
    Aug 6, 2025 · ... The strain hardening is controlled primarily by the properties of the molecular network of entangled chains in the amorphous phase, ...
  69. [69]
    Strain hardening of thermoplastics | Macromolecules
    Evaluation of microscopic structural changes during strain hardening of polyethylene solids using In situ Raman, SAXS, and WAXD measurements under step ...
  70. [70]
    Suppressed Segmental Relaxation as the Origin of Strain Hardening ...
    Jan 20, 2009 · At high strains polymer chains deform resulting in a large increase of stress known as “strain hardening” [1–8] . The putative solid undergoes ...<|separator|>
  71. [71]
    On the workhardening rate of glassy polymers - ScienceDirect.com
    The aim of this paper is to introduce a new parameter to characterize the non-elastic behaviour of glassy polymers, namely the workhardening rate.
  72. [72]
    Strain hardening of polymer glasses: Entanglements, energetics ...
    Mar 3, 2008 · The most dramatic hardening at large strains reflects increases in energy as chains are pulled taut between entanglements rather than a change in entropy.Article Text · INTRODUCTION · POLYMER MODEL AND... · RESULTS
  73. [73]
    Investigations into crazing in glassy amorphous polymers through ...
    Crazing involves massively dilatational response where a small volume of material, under tensile deformation, widens into a network of interconnected voids and ...
  74. [74]
    Six decades of the Hall–Petch effect – a survey of grain-size ...
    In this article, we gather the grain-size strengthening data from the Hall–Petch studies on pure metals and use this aggregated data to calculate best estimates ...
  75. [75]
    The Hall–Petch and inverse Hall–Petch relations and the hardness ...
    Nov 14, 2019 · In this review, we focus on how the hardness of metals is affected by the interaction between dislocations and grain boundaries.
  76. [76]
    Bulk Metallic Glass Composites with Transformation‐Mediated Work ...
    Bulk Metallic Glass Composites with Transformation-Mediated Work-Hardening and Ductility. Yuan Wu,. Yuan Wu. State Key Laboratory for Advanced ...
  77. [77]
    Metallic glasses | MRS Bulletin
    Sep 28, 2023 · Rejuvenated samples show strain hardening (Figure 3c) and suppression of shear bands up to εp = 2.7‒3.8 percent. Opposite to the effect of ...
  78. [78]
    On the work-hardening behaviour of the additively manufactured Al ...
    This study aims to investigate the work-hardening capability, uniform elongation and deformability of AlSi7Mg and AlSi10Mg samples after different post- ...
  79. [79]
    Full article: Additive manufacturing of metals and alloys to achieve ...
    Jan 30, 2024 · In this review article, we focus on four types of heterogeneous microstructures induced by AM: lamellar, gradient, laminated, and harmonic ones.