Fact-checked by Grok 2 weeks ago

Viscoplasticity

Viscoplasticity is a theory in continuum mechanics that describes the rate-dependent inelastic behavior of solids, where materials exhibit time-dependent irreversible deformations under applied stress, combining viscous flow with plastic yielding. Unlike classical rate-independent plasticity, which features a distinct yield surface, viscoplasticity involves continuous deformation without such a boundary, with the plastic strain rate directly dependent on the applied stress level and characterized by overstress functions. The concept of viscoplasticity has historical roots in observations of inelastic material responses dating back over 150 years, with early experimental foundations documented in the and significant theoretical advances emerging in the mid-20th century through studies of dislocation dynamics in the and . Key developments include the Perzyna overstress model introduced in 1966, which formulates viscoplastic as a function of excess beyond a static limit, and the unified elastic-viscoplastic framework by Bodner and Partom in 1975, which integrates , , and viscous effects using internal state variables for hardening and . The Chaboche model, developed in the 1970s and refined subsequently, extends kinematic hardening to capture cyclic loading effects through nonlinear backstress evolution, often combined with isotropic hardening for comprehensive material response prediction. Viscoplastic models are essential for simulating behaviors such as in metals at elevated temperatures and high-strain-rate deformation in polycrystalline materials like ceramics and alloys. Applications span engineering fields including structures for predicting permanent deformations under dynamic loads, geotechnical of soils and clays exhibiting rate-sensitive , and biomedical modeling of biofluids like that display viscoplastic characteristics. These models enable accurate forecasting of , , and long-term durability in environments where and loading history significantly influence material performance. Recent advancements as of 2025 include integration with for parameter optimization in complex simulations.

Fundamentals

Definition and Key Characteristics

Viscoplasticity describes the mechanical behavior of materials that combine rate-independent plastic deformation with rate-dependent viscous flow, resulting in time-dependent phenomena such as under constant and under constant . This dual nature arises from mechanisms like motion in crystalline solids or molecular rearrangements in amorphous materials, where the deformation rate influences the effective strength and flow resistance. Unlike pure , which lacks explicit time-dependence, viscoplasticity accounts for viscous dissipation that slows or accelerates permanent deformation based on loading duration. Key characteristics of viscoplastic materials include a nonlinear stress-strain response, where the material yields gradually without a sharp transition, and pronounced strain-rate sensitivity that causes higher flow stresses at faster deformation rates. Upon unloading, the material recovers its elastic strain component elastically, but retains permanent viscoplastic strain, often leading to hysteresis in cyclic loading. A typical stress-strain curve for a viscoplastic material features an initial linear elastic region up to a yield-like point, followed by a curving viscoplastic regime where the tangent modulus decreases, with the curve shifting upward for increasing strain rates to reflect rate sensitivity; for instance, at low rates, the response approaches rate-independent plasticity, while high rates show enhanced hardening. The fundamental description of viscoplastic deformation is captured by the viscoplastic strain rate equation, given in general form as \dot{\epsilon}^{vp} = f(\sigma, T), where \dot{\epsilon}^{vp} is the viscoplastic strain rate, \sigma is the applied stress, T is temperature, and f represents a nonlinear function incorporating these dependencies, often derived from overstress concepts or flow rules. This formulation highlights the interplay between stress-driven plasticity and viscous drag, enabling predictions of time-dependent inelasticity. Representative examples of viscoplastic materials encompass metals at elevated temperatures, where thermal activation facilitates via climb; granular soils, exhibiting rate-sensitive shearing and ; polymers, showing time-dependent yielding due to chain entanglement; and biological tissues, such as or arterial walls, which display inelastic flow under physiological loads to accommodate dynamic function. Viscoplasticity differs fundamentally from in that it incorporates permanent, irrecoverable plastic deformation once a yield threshold is exceeded, whereas features time-dependent recoverable strains with no permanent deformation under typical loading conditions. In viscoelastic materials, such as polymers below their , the total consists of an elastic component that recovers instantly and a viscous component that recovers gradually over time, leading to phenomena like and without residual . Conversely, viscoplastic materials, like metals at elevated temperatures or certain soils, exhibit viscous only after yielding, resulting in irreversible accumulation that persists even after load removal. This distinction is evident in stress- responses: shows closed loops during cyclic loading with full recovery, while viscoplasticity displays open loops with progressive and permanent offset. The following table summarizes key differences in stress-strain responses and time-dependency between and viscoplasticity:
AspectViscoplasticity
Strain RecoveryFull or partial recovery upon unloading; no permanent deformation.Irrecoverable plastic strain; permanent deformation after .
Stress-Strain ResponseTime-dependent, nonlinear; closed in cycles; /recovery without .Rate-dependent yielding; open with ; leads to .
Time-DependencyViscous dissipation causes delayed response; recoverable over time (e.g., Prony series models).Viscous overstress above ; time-dependent flow rate, irrecoverable (e.g., Perzyna-type models).
Example Behaviors in rubbers; frequency-dependent in oscillatory tests. rupture in alloys; rate-sensitive yielding in soils.
Data adapted from comparative modeling studies. In contrast to classical , which assumes rate-independent deformation where flow initiates abruptly at a fixed and proceeds without time effects, viscoplasticity introduces viscous that renders the effective and post- flow rate-dependent. Classical models, such as J2 , define a sharp beyond which unlimited plastic straining occurs instantaneously, independent of , leading to path-dependent hardening but no under constant . Viscoplasticity extends this by incorporating a viscous term, often via overstress functions, allowing deformation even below the static but accelerating it above, which captures time-dependent phenomena like secondary in sustained loading. This rate sensitivity is crucial for applications involving high s or prolonged holds, where classical models fail to predict viscous dissipation. Compared to hypoplasticity, which is primarily a rate-independent incremental model for granular materials emphasizing changes in tangent without a distinct , viscoplasticity highlights the coupled viscous and plastic mechanisms that introduce explicit time-dependency in deformation s. Hypoplastic models, such as those for sands or clays, relate stress directly to through nonlinear tensorial equations incorporating state variables like , focusing on pre-failure evolution and critical state behavior in cohesionless soils. In viscoplasticity, the viscous component drives time-dependent flow, particularly in fine-grained or soft soils, enabling better simulation of and relaxation, whereas hypoplasticity requires extensions (e.g., visco-hypoplastic variants) to account for such effects. This makes viscoplasticity more suitable for viscous-dominated scenarios, while hypoplasticity excels in capturing directional variations in granular flows. A hallmark of viscoplasticity is its pronounced during loading-unloading cycles, where energy dissipation arises from both viscous flow and plastic irreversibility, unlike the purely recoverable in . This leads to path-dependency, as the material's response evolves with deformation history due to hardening or softening from the plastic component, compounded by rate effects from , resulting in phenomena like under asymmetric cycling. Such behavior distinguishes viscoplasticity in , enabling accurate modeling of cumulative damage in engineering components under variable loading.

Historical Development

Early Observations and Experiments

Early observations of time-dependent deformation in metals under constant load date back to the , with notable reports of sagging in lead pipes used in Victorian-era systems, where sustained gravitational led to gradual curvature over decades. These engineering failures, observed in structures from the mid-1800s onward, highlighted the viscous-like flow in soft metals at ambient temperatures, prompting initial concerns in and roofing applications. Similar phenomena were noted in rocks and , but metallic gained attention amid the Industrial Revolution's push for reliable materials in load-bearing components. By the late 1800s, the rise of power technology spurred more targeted constant tests on metals, particularly in and parts exposed to elevated temperatures. Engineers conducting prolonged loading experiments on iron and components in engines observed slow, ongoing deformation even below the yield , attributing it to effects that accelerated viscous flow. These tests, though not yet systematic, revealed that high-temperature metals like those in blades exhibited measurable accumulation over time, influencing early design limits for machinery. Pivotal advancements came in the early through Andrade's comprehensive experiments on pure metals such as lead, tin, and , published in 1910. Andrade applied constant tensile stress at room and moderate temperatures, documenting the evolution of over extended periods and identifying distinct phases: an initial transient primary stage with decelerating , a steady secondary stage, and an accelerating tertiary stage leading to . These phenomenological insights, derived from wire specimens under controlled loads, established as a fundamental material response without invoking mathematical models at the time. In parallel, early 20th-century soil mechanics experiments began uncovering analogous time-dependent behaviors in granular materials under sustained shear and compressive loads. By the 1940s, tests on clays and sands at constant stress revealed primary creep-like settling and secondary flow, particularly in foundation engineering contexts. These findings extended viscoplastic insights beyond high-temperature metals to geotechnical applications, emphasizing rate-dependent deformation in cohesive soils.

Evolution of Theoretical Models

The theoretical modeling of viscoplasticity began to take shape in the mid-20th century, building on empirical observations of time-dependent deformation in metals under sustained loads. In 1929, Frederick H. Norton proposed a power-law relationship to describe steady-state creep strain rates in steel at elevated temperatures, expressing the creep rate as proportional to the applied stress raised to a power, which provided an early mathematical framework for rate-dependent inelastic flow. This Norton creep law laid foundational groundwork for subsequent viscoplastic models by quantifying the nonlinear viscous-like response in materials exhibiting creep. During the 1950s, Nicholas J. Hoff extended these ideas through analytical studies on creep deformation in structures, incorporating power-law formulations into structural mechanics problems and emphasizing the role of time-dependent plasticity in engineering applications. Hoff's contributions, including his organization of the 1960 IUTAM Symposium on "Creep in Structures," marked a pivotal timeline in the field's publications, fostering integration of creep models into broader rheological theories. Significant theoretical advances also emerged through studies of dislocation dynamics in the 1950s and 1960s, providing microstructural explanations for rate-dependent inelastic behaviors that informed later viscoplastic formulations. The and saw significant unification efforts to merge classical with viscous effects, addressing the limitations of purely or frameworks in capturing rate sensitivity. Perzyna introduced the overstress concept in the early , formulating a viscoplastic theory where the inelastic is driven by the excess beyond a , generalized from earlier linear viscoplastic ideas by Hohenemser and Prager. This approach, detailed in Perzyna's 1966 monograph, enabled a continuous transition between , , and viscous behaviors without a distinct yield point, influencing subsequent models for metals and soils under . By the , these unification trends extended to thermodynamic formulations, incorporating internal variables to describe evolving material states during viscoplastic flow. The 1980s marked a shift influenced by , with viscoplastic models adapted for finite element analysis to simulate complex engineering scenarios like turbine blades and nuclear components. Jean-Louis Chaboche advanced kinematic hardening extensions during this period, decomposing the backstress into multiple nonlinear components to better predict cyclic and mean in viscoplastic materials. His work with Rousselier on plastic and viscoplastic constitutive equations highlighted the need for unified frameworks in finite element implementations, enabling numerical predictions of rate-dependent hardening. Early models, however, often neglected and damage mechanisms, assuming isotropic responses and undamaged states, which were later addressed through coupled formulations in the to account for microstructural evolution under prolonged loading.

Experimental Phenomenology

Strain Hardening Tests

Strain hardening tests, also known as work-hardening or monotonic loading experiments, are fundamental in characterizing the viscoplastic behavior of materials under increasing deformation. These tests typically involve uniaxial tensile loading of a cylindrical or dog-bone-shaped specimen at controlled constant s, using a equipped with load cells and extensometers to measure axial force and displacement. The specimen is clamped between grips, often with hydraulic or mechanical actuators, and subjected to incremental elongation while recording stress-strain responses; a schematic setup includes the testing frame, controller, and system to capture real-time curves. In these experiments, multiple tests are conducted at varying , typically ranging from 10^{-3} to 10 s^{-1} using standard universal testing machines, though higher rates require specialized equipment, to reveal rate-dependent effects. The resulting - curves exhibit initial response followed by viscoplastic flow, where the —the stress at which permanent deformation occurs—increases nonlinearly with both and , demonstrating rate-dependent hardening. This hardening is quantified through the strain hardening exponent (n) in power-law relations of the form σ = K ε^n, where σ is , ε is plastic , and K is the strength , with n typically decreasing (indicating reduced hardening capacity) at higher due to viscous drag mechanisms. For instance, in metals like aluminum, the curves show a pronounced upward shift in with increasing rate, reflecting the material's sensitivity to deformation speed. Interpretation of these curves involves fitting the data to viscoplastic constitutive equations to extract parameters such as the coefficient (η) in overstress models, where the viscoplastic is modeled as \dot{\epsilon}^{vp} = \frac{\sigma - \sigma_y}{\eta} for simple cases, though more advanced fits use logarithmic or exponential forms. For aluminum alloys like series, experimental data at and s of 10^{-4} to 10^2 s^{-1} (using appropriate equipment for higher rates) yield coefficients typical of viscoplastic models for metals, enabling prediction of flow behavior in applications like processes. These fits are performed by plotting log() versus log() at fixed strains, yielding straight lines whose slopes indicate (m ≈ 0.01–0.05 for aluminum), thus identifying key viscoplastic parameters. A primary limitation of strain hardening tests is the assumption of isothermal conditions, which may not hold during high-rate deformation due to adiabatic heating, potentially altering the measured hardening by 10–20% in sensitive metals. Additionally, results are highly sensitive to variations, with even small increases (e.g., 50°C) reducing the flow stress and hardening exponent in aluminum alloys by up to 30%, necessitating controlled environmental chambers for accuracy.

Tests

Creep tests are a fundamental experimental method used to characterize the time-dependent deformation of materials under sustained constant , providing critical insights into viscoplastic behavior where accumulates gradually due to mechanisms such as or diffusion-controlled processes. In these tests, a uniaxial tensile or is applied to a specimen, typically at elevated temperatures, and the resulting is measured over extended periods, often spanning hours to thousands of hours. The procedure involves loading the sample to the desired level at a constant rate, followed by maintaining that stress while continuously recording the axial or using extensometers or gauges. This setup isolates the viscoplastic response by minimizing variations, allowing observation of how the material deforms without external loading changes. The curve obtained from such tests typically exhibits three distinct s that highlight the viscoplastic mechanisms at play. During the primary (or transient) , the creep rate decreases over time as hardening mechanisms, such as interactions, initially resist further deformation. This decelerating phase transitions into the secondary (or steady-state) , where the creep rate reaches a minimum constant value, \dot{\epsilon}_{\min}, balancing hardening and recovery processes like dynamic recrystallization. Finally, in the , the creep rate accelerates due to necking, void formation, or microstructural damage, leading to eventual . These s are particularly pronounced in metals and alloys, where \dot{\epsilon}_{\min} serves as a key indicator of viscoplastic flow, reflecting the material's long-term deformation capacity under service conditions. Interpretation of creep test data focuses on extracting parameters that quantify viscoplasticity, such as the creep exponent n and a reference stress \sigma_0, derived from the empirical relation \dot{\epsilon}_{\min} = A \sigma^n \exp(-Q/RT), where A is a material constant, \sigma is the applied , Q is the , R is the , and T is the absolute . The exponent n typically ranges from 3 to 8 for dislocation in crystalline materials, indicating the stress sensitivity of the rate-controlling mechanism. For instance, in nickel-based superalloys used in blades, creep tests at 800–1000°C reveal n values around 5–7, with \dot{\epsilon}_{\min} on the order of 10^{-8} to 10^{-6} s^{-1} under stresses of 200–400 , underscoring their resistance to viscoplastic deformation in high-temperature environments. The Q in this relation is closely tied to processes, such as self-diffusion of the matrix or solute atoms, often measured through Arrhenius plots of \dot{\epsilon}_{\min} versus inverse temperature, yielding Q values of 200–400 kJ/mol for such alloys. Temperature plays a pivotal role in creep behavior, with the Arrhenius form governing the exponential increase in \dot{\epsilon}_{\min} as thermal activation facilitates atomic mobility and viscoplastic flow. At homologous temperatures above 0.4–0.5 T_m (where T_m is the ), diffusion creep dominates, leading to lower n values (around 1–2), while at intermediate temperatures, power-law creep prevails with higher n. These temperature-dependent factors are essential for predicting in applications like power plant components, where creep tests guide design to minimize tertiary stage acceleration. Brief consideration of prior strain hardening from monotonic loading can influence primary stage duration, but creep tests primarily emphasize the sustained stress response.

Relaxation Tests

Relaxation tests in viscoplasticity involve rapidly deforming a specimen to a predetermined fixed level at a high initial rate, followed by holding the constant while continuously measuring the resulting decay over extended periods, often spanning hours or days. This experimental , typically conducted using uniaxial on cylindrical or flat dogbone-shaped samples in a controlled load with extensometers for precise , isolates the time-dependent viscous contributions to deformation by eliminating ongoing total changes. presentation on a logarithmic facilitates of the nonlinear decay patterns characteristic of rate-sensitive flow. Key observations from these tests reveal a characteristic reduction that often follows a logarithmic form, expressed as \sigma(t) = \sigma_0 - \Delta\sigma \log(t), where \sigma_0 is the initial , \Delta\sigma represents the magnitude of relaxation, and t is time; this allows identification of relaxation time constants that indicate the onset and extent of viscous dissipation. In metals such as low-carbon (e.g., Q235), the is logarithmic with time across initial stresses of 70–100 MPa and temperatures such as 450–600°C, exhibiting three regimes: an initial rapid drop, a slower , and eventual saturation approaching a threshold . For like , relaxation can reach up to 95% of initial at elevated temperatures (e.g., 538°C) over 24 hours, contrasting with minimal (∼15%) at . The decay in relaxation tests is interpreted through the viscoplastic , derived from additivity under fixed total : \dot{\epsilon}^{vp} = -\frac{\dot{\sigma}}{[E](/page/E!)}, where [E](/page/E!) is the , linking the measured rate \dot{\sigma} directly to the evolving inelastic flow. This relation facilitates parameter estimation in viscoplastic models, such as determining rate sensitivity exponents (e.g., n) or drag coefficients by integrating the curve with constitutive equations like the Perzyna over form; for instance, in P91 , relaxation data yield estimates of parameters that align with observed saturation stresses around 150 even at ambient conditions. The extent and rate of stress decay are significantly influenced by the initial strain level, with higher pre-strains (e.g., 1.8% versus 0.6% in ) promoting greater overall relaxation due to enhanced activation of dislocation-based mechanisms. Microstructure also plays a key role, as grain boundaries, phase distributions (e.g., alpha versus alpha/beta in ), and prior deformation history alter the decay rate by modulating barriers to viscoplastic flow, with textured microstructures from processing exhibiting orthotropic relaxation responses. These tests complement experiments by analogously highlighting viscous effects, though under fixed strain rather than .

Classical Rheological Models

Norton-Hoff Model

The Norton-Hoff model represents a foundational rheological description of a perfectly viscoplastic solid, characterized by continuous without a yield threshold or elastic components. It posits that deformation occurs solely through viscoplastic mechanisms, with the viscoplastic directly proportional to a power of the applied . This model captures the nonlinear rate-dependent observed in materials under sustained loading at elevated temperatures, emphasizing steady-state conditions. The constitutive relation for the model is expressed as \dot{\epsilon}^{vp} = \left( \frac{\sigma}{\eta} \right)^n where \dot{\epsilon}^{vp} , \sigma , \eta , and n typically greater than 1, reflecting the nonlinearity of the response. For uniaxial or , \sigma reduces to the of the axial . The exponent n quantifies the stress sensitivity, often ranging from 3 to 8 for metals, while \eta incorporates material-specific resistance to flow. This power-law form arises from empirical observations of secondary creep, where the strain rate remains constant after an initial transient phase, allowing straightforward calibration from constant-load creep experiments. The derivation stems from the secondary creep stage, where microstructural steady-state conditions lead to a balance between hardening and recovery processes, resulting in the power-law dependence on stress. Temperature effects are incorporated by making \eta thermally activated, following an Arrhenius-type relation: \eta = \eta_0 \exp\left(\frac{Q}{RT}\right) here, \eta_0 is a reference viscosity, Q is the activation energy for creep (often 100–300 kJ/mol for metals), R is the universal gas constant, and T is the absolute temperature. This formulation aligns the model with diffusion-controlled mechanisms like dislocation climb, enabling predictions across temperature ranges without altering the power-law structure. Calibration involves fitting to logarithmic plots of strain rate versus stress from creep tests, yielding n as the slope and \eta from the intercept at a fixed temperature. Historically, the model traces to F.H. Norton's 1929 analysis of high-temperature deformation in steels, where he empirically established the power-law relation for steady-state rates based on experimental data from components. N.J. Hoff extended this framework in the and , adapting it for broader viscoplastic applications in structural metals under combined loading, including and collapse scenarios. In applications, the Norton-Hoff model excels at predicting steady-state creep deformation in components like turbine blades or pressure vessels, where elastic strains are negligible compared to accumulated viscoplastic flow. For instance, it facilitates lifetime estimates by integrating the strain rate into time-to-rupture calculations under constant stress. However, its limitations include the absence of elastic recovery upon unloading, leading to overprediction of total deformation in cyclic or transient loading, and neglect of primary or tertiary creep stages. These constraints make it suitable primarily for long-term, monotonic scenarios calibrated via secondary creep data from phenomenological tests.

Bingham-Norton Model

The Bingham-Norton model represents an elastic-viscoplastic constitutive framework for materials that exhibit behavior up to a yield stress \sigma_y and subsequent rate-dependent viscoplastic flow beyond it. In this model, the response is purely for stresses below \sigma_y, governed by \epsilon^e = \sigma / E, where E is the . Once the yield stress is exceeded, viscoplastic straining initiates, with the viscoplastic expressed as \dot{\epsilon}^{vp} = \frac{\langle \sigma - \sigma_y \rangle}{\eta} for the linear case, where \langle \cdot \rangle is the Macaulay bracket ensuring flow only in tension (or analogously in compression), and \eta is the material viscosity. An extension incorporates a Norton power-law option, modifying the flow rule to \dot{\epsilon}^{vp} = \left( \frac{\langle \sigma - \sigma_y \rangle}{K} \right)^n \operatorname{sign}(\sigma - \sigma_y), where K is a consistency parameter and n is the power-law exponent, allowing for nonlinear rate sensitivity above yield. The total strain decomposes additively as \epsilon = \epsilon^e + \epsilon^{vp}. This model derives from the rheological analogy to the Bingham fluid, which parallels a rigid-perfectly plastic slider in series with a linear viscous dashpot to capture yield and subsequent Newtonian flow, augmented by an elastic spring for solid-like initial response and the optional power-law dashpot inspired by the Norton creep law for enhanced nonlinearity. Unlike pure viscoplastic models such as Norton-Hoff, which assume continuous flow without a yield threshold, the Bingham-Norton enforces delayed deformation until \sigma > \sigma_y. Key strengths of the Bingham-Norton model include its simplicity in reproducing initial elastic loading and the transition to viscoplastic flow at a finite , effectively delaying permanent deformation and aligning with observed rate-dependent yielding in experiments. It finds application in for modeling clays and granular materials, where yield followed by viscous shearing governs behaviors like flows or tunneling deformations. A primary shortcoming is the absence of strain hardening, resulting in perfect viscoplasticity that overpredicts accumulated s during prolonged monotonic loading, as the remains fixed at \sigma_y.

Isotropic Hardening Extensions

Isotropic hardening extensions to classical viscoplastic models, such as the Norton-Hoff formulation, incorporate the evolution of and flow parameters with accumulated viscoplastic \epsilon^{vp}, enabling the capture of strain-dependent strengthening beyond static criteria. In these extensions, the is often expressed as \sigma_y(\epsilon^{vp}) = \sigma_{y0} + K (\epsilon^{vp})^m, where \sigma_{y0} is the initial , K is a hardening , and m is an exponent typically between 0 and 1, integrated into the Norton-Hoff flow rule to describe power-law strain hardening under sustained loading. This mechanism builds on the baseline Bingham-Norton by allowing the effective to vary with deformation history, improving predictions for materials exhibiting progressive stiffening. Derivations of these extensions adapt rate-independent hardening laws to viscoplastic frameworks through evolution equations that link the rate of yield stress change to the viscoplastic strain rate. For linear isotropic hardening, the evolution is given by \dot{\sigma_y} = h \dot{\epsilon}^{vp}, where h is the hardening rate parameter, ensuring thermodynamic consistency via the dissipation inequality and potentials. Saturating behaviors, such as those from the Voce law, modify the stress as \sigma_y(\epsilon^{vp}) = \sigma_{y0} + Q(1 - \exp(-b \epsilon^{vp})), with Q as the saturation stress increment and b controlling the approach to saturation; this is derived by embedding the exponential form into the overstress or flow potential of viscoplastic models. These extensions exhibit saturating or exponential hardening, where initial rapid strengthening transitions to a plateau, reflecting dislocation interactions and saturation in metals. In uniaxial creep simulations, such models predict reduced creep rates over time as the evolving yield surface limits further deformation, with strain accumulation slowing after an initial transient phase, as validated in boundary value problems for thick-walled components under internal pressure. Advancements include the incorporation of recovery terms in the evolution equations to account for cyclic loading, balancing hardening with dynamic (strain-induced) and static (time- and temperature-dependent) recovery processes, such as \dot{R} = \frac{2}{3} h \dot{\epsilon}^{vp} - \gamma R |\dot{\epsilon}^{vp}|, where R is the isotropic hardening variable and \gamma governs , enhancing model fidelity for repeated loading cycles in high-temperature applications.

Rate-Dependent Formulations

Perzyna Overstress Approach

The Perzyna overstress approach provides a foundational for modeling viscoplastic behavior by extending classical rate-independent through the introduction of a viscous overstress mechanism. In this formulation, the material is permitted to sustain stresses beyond the static , with the excess stress, or overstress, driving the rate of viscoplastic straining. This approach treats viscoplasticity as a regularization of perfect , where the viscoplastic flow is activated proportionally to the overstress magnitude. The core for the viscoplastic in the Perzyna model is given by \dot{\epsilon}^{vp} = \gamma \left< \frac{f(\sigma)}{Y} \right>^m \frac{\partial g}{\partial \sigma}, where \dot{\epsilon}^{vp} denotes the viscoplastic , f(\sigma) is the function defining the static (with f(\sigma) \leq 0 for elastic states), Y represents the drag (a akin to a viscous ), \gamma > 0 is a fluidity controlling the rate sensitivity, m > 0 is an exponent governing the nonlinearity of the overstress response (often taken as m=1 for linear cases), and g(\sigma) is the (commonly g = f for associated ). The Macaulay brackets \left< \cdot \right> ensure that viscoplastic occurs only when f(\sigma) > 0, i.e., in the overstress regime. This equation posits that the direction of viscoplastic straining aligns with the outward normal to the yield surface, while the magnitude is determined by the normalized overstress raised to the power m, scaled by \gamma. The derivation of this model conceptualizes overstress f(\sigma) > 0 as the driving force for viscous dissipation, analogous to a Newtonian viscous but regularized by the criterion. In the limit as \gamma \to \infty (or equivalently, viscosity \eta = 1/\gamma \to 0), the overstress vanishes, and the response recovers the rate-independent limit where straining occurs only on the f(\sigma) = 0. This asymptotic ensures consistency with classical while incorporating rate effects for finite loading rates. The model can incorporate isotropic hardening by allowing Y or the function f to evolve with accumulated plastic , as developed in extensions of the classical . A key advantage of the Perzyna approach lies in its ability to smooth the discontinuities inherent in rate-independent models, such as abrupt yielding and non-smooth stress-strain responses, which often lead to issues in finite element simulations. By introducing a continuous viscoplastic multiplier, the formulation yields differentiable tangent moduli, facilitating stable and efficient in explicit and implicit schemes, particularly for problems involving large deformations or . This regularization enhances mesh insensitivity and overall solver robustness in applications. Calibration of the Perzyna parameters \gamma and m is typically performed using stress relaxation tests, where a fixed total is imposed, and the resultant decay over time is measured. The relaxation response follows an form derived from the overstress evolution, allowing least-squares fitting to extract \gamma (which governs the relaxation rate) and m (which captures nonlinearity). Such tests provide direct insight into the material's viscous time scale, with typical values for metals yielding relaxation times on the order of seconds to minutes under laboratory conditions.

Duvaut-Lions Variational Approach

The Duvaut-Lions variational approach provides a rigorous framework for modeling viscoplastic behavior through , treating the evolution of viscoplastic strains as a gradient flow in the context of variational inequalities. This method, originally developed in the context of inequalities in , formulates viscoplasticity as a regularization of rate-independent , where the material response relaxes hyperbolically toward the elastic domain over a characteristic timescale. Unlike local overstress-based models, it ensures global minimization of a potential, leading to thermodynamically consistent evolution equations that respect the convexity of the elastic domain. The core formulation is given by the evolution equation for the viscoplastic strain rate: \dot{\epsilon}^{vp} = \frac{1}{\eta} \left( (\epsilon - \epsilon^{vp}) - P_K(\epsilon - \epsilon^{vp}) \right) where \epsilon is the total strain tensor, \epsilon^{vp} denotes the viscoplastic strain tensor, P_K is the orthogonal projection onto the convex elastic domain K in the appropriate dual space (often the elastic strain space, related to stress via the elastic compliance tensor), and \eta > 0 is the viscosity parameter representing the relaxation time. This equation describes how the elastic strain \epsilon^e = \epsilon - \epsilon^{vp} adjusts toward the boundary of the elastic domain K, effectively smoothing the sharp yield criterion of perfect plasticity. For cases with imposed total strain rate \dot{\epsilon}, the equation generalizes accordingly. The stems from the minimization of a potential, \Phi(\dot{\epsilon}^{vp}) = \frac{\eta}{2} \|\dot{\epsilon}^{vp}\|^2, to the constraint that the associated lies within the elastic domain, leading to a of the form \langle \eta \dot{\epsilon}^{vp} + \partial I_K(\sigma), \delta \epsilon^{vp} \rangle = 0 for admissible variations \delta \epsilon^{vp}. Equivalently, this is the of the I_K in the subdifferential sense, ensuring that the direction points toward the closest point in K. This variational structure guarantees that the model recovers rate-independent in the limit \eta \to 0, with the viscoplastic multiplier emerging naturally from the . A key advantage of this approach is its inherent thermodynamic consistency, as the dissipation rate \sigma : \dot{\epsilon}^{vp} remains non-negative due to the monotonicity of the projection operator, aligning with the second law of thermodynamics for associative rules. Numerically, it promotes stability in finite element implementations by providing an implicit characterization of the , which facilitates consistent and avoids oscillations near the —contrasting with explicit schemes that may require small time steps for accuracy. Furthermore, the relaxation nature allows for robust handling of non-smooth surfaces, such as those with corners, without singularities in the tangent operators. In comparison to the Perzyna overstress approach, which relies on an explicit multiplier proportional to the overstress beyond the , the Duvaut-Lions method employs an implicit time-stepping scheme via the , yielding a closer with variational principles and better conditioning in the inviscid limit. This implicit nature enhances in return-mapping integrations, particularly for multi-surface extensions.

Flow Stress Models

Johnson-Cook Model

The Johnson-Cook model is a phenomenological constitutive relation widely used to describe the of metals under high strain rates, large deformations, and elevated temperatures, particularly in scenarios. It employs a multiplicative formulation that decouples and combines strain hardening, strain-rate hardening, and thermal softening effects, making it suitable for viscoplastic behavior in metals where these factors interact independently. The core equation for the von Mises equivalent flow stress \sigma is given by: \sigma = \left( A + B \epsilon^n \right) \left( 1 + C \ln \dot{\epsilon}^* \right) \left( 1 - T^{*m} \right) where \epsilon is the equivalent plastic strain, \dot{\epsilon}^* = \dot{\epsilon} / \dot{\epsilon}_0 is the dimensionless plastic strain rate normalized by a reference strain rate \dot{\epsilon}_0 (typically $1 \, \mathrm{s}^{-1}), and T^* = (T - T_r) / (T_m - T_r) is the homologous temperature with T as the current absolute temperature, T_r as the room temperature, and T_m as the melting temperature. The model parameters A, B, n, C, and m represent the yield stress at reference conditions, hardening modulus, hardening exponent, strain-rate sensitivity, and thermal softening exponent, respectively. This formulation arises from the assumption of multiplicative coupling, where strain hardening is captured by a power-law term derived from quasi-static tests, strain-rate effects by a logarithmic term based on observed rate sensitivity in dynamic experiments, and thermal softening by a sigmoidal function to account for adiabatic heating and recovery processes. Calibration typically involves (SHPB) tests for high strain rates (up to $10^3 \, \mathrm{s}^{-1}) combined with quasi-static tensile and tests, ensuring the model fits data across a wide range of conditions without requiring detailed microstructural mechanisms. For typical low-carbon steels, such as AISI 4340, parameter values include A = 792 \, \mathrm{[MPa](/page/MPA)} (quasi-static strength), B = 510 \, \mathrm{MPa} (hardening ), n = 0.26 (strain-hardening exponent), C = 0.014 (rate sensitivity), and m = 1.03 (thermal exponent), calibrated to match experimental stress-strain curves under . The model is often extended with a initiation criterion for prediction, where cumulative plastic strain at failure depends on , , and temperature, enabling simulation of ductile failure in elements via element erosion in finite element codes. In engineering applications, the Johnson-Cook model is extensively employed in simulating ballistic impacts, where it predicts penetration and deformation in armor materials under projectiles, and in high-speed processes, capturing chip formation and due to localized high strains and temperatures.

Zerilli-Armstrong Model

The Zerilli-Armstrong model provides a physically motivated constitutive framework for predicting the viscoplastic of metals, emphasizing -based mechanisms that differentiate behavior across crystal structures such as face-centered cubic (FCC), body-centered cubic (BCC), and hexagonal close-packed (HCP). Developed through analysis of thermal activation and athermal interactions, the model decomposes into components arising from lattice friction, strain hardening via density evolution, and rate/temperature-sensitive overcoming of short-range obstacles. This approach contrasts with purely empirical formulations by linking parameters to microstructural physics, enabling extrapolation beyond calibration data. For FCC metals, the is \sigma = C_0 + C_1 \exp\left(-C_3 T + C_4 T \ln \dot{\epsilon}^*\right) \epsilon^n, where \sigma denotes the von Mises equivalent , \epsilon is the equivalent plastic , \dot{\epsilon}^* is the dimensionless (\dot{\epsilon}/1 \, \mathrm{s}^{-1}), T is the absolute temperature, and C_0 to C_4 along with n are material-specific constants determined via to experimental data. The term C_0 captures the athermal from long-range barriers like boundaries, while C_1 \exp\left(-C_3 T + C_4 T \ln \dot{\epsilon}^*\right) \epsilon^n represents temperature- and rate-dependent hardening from forest dislocation intersections, with the exponential form reflecting thermal over obstacles. The rate dependence in the exponent stems from the Orowan adapted for high-rate , where higher rates increase at lower temperatures. The derivation integrates dislocation velocity relations from thermal activation theory, where the activation energy for obstacle surmounting scales inversely with and temperature, leading to the exponential forms; athermal contributions are derived from Taylor's relation linking to square-root dislocation density. Calibration for FCC aluminum utilized quasi-static and dynamic compression data, yielding constants like C_0 \approx 22 MPa, C_1 \approx 652 MPa, C_3 \approx 0.0013 \, \mathrm{K}^{-1}, n \approx 0.5, and C_4 \approx -0.0009 \, \mathrm{K}^{-1}, which accurately reproduced stress-strain curves across temperatures from 77 K to 600 K. For BCC metals, the model employs a distinct form to account for pronounced thermal over high Peierls-Nabarro stresses inherent to the lattice: \sigma = C_0 + C_1 \exp\left(-C_3 T + C_4 T \ln \dot{\epsilon}^*\right) + C_2 \epsilon^n. Here, the thermal term C_1 \exp\left(-C_3 T + C_4 T \ln \dot{\epsilon}^*\right) dominates at low temperatures and high rates, reflecting screw kinking as the rate-controlling mechanism, while hardening C_2 \epsilon^n is less temperature-sensitive than in FCC due to stronger athermal forest interactions. Derivation parallels the FCC case but adjusts activation volume and energy parameters for BCC's non-planar core structure, with calibration to iron data from tests showing fits to flow stresses up to 1 GPa at rates exceeding 10^3 s^{-1}. A distinguishing feature is the model's lattice-specific parameterization, which for HCP metals incorporates additional twinning contributions alongside slip, modeled as stress-driven and growth of twin lamellae that enhance at high rates. Validation across structures involved Taylor cylinder impact simulations in hydrocodes, demonstrating superior prediction of deformed shapes and final lengths compared to empirical models, with agreement to experimental data at strain rates up to 10^4 s^{-1} for both aluminum and iron.

Preston-Tonks-Wallace Model

The Preston-Tonks-Wallace (PTW) model is a physically motivated viscoplastic constitutive developed for simulating the plastic deformation of metals under extreme conditions, including high rates up to $10^4 s^{-1} and temperatures ranging from cryogenic to near-melting levels, as encountered in shock loading and explosive events. It emphasizes the role of microstructural evolution in , distinguishing it from models like Zerilli-Armstrong by tracking dynamic internal state variables such as defect density rather than relying on fixed-parameter representations of microstructure. The core of the model lies in its expression for flow stress, given by \sigma = \theta(\epsilon, T, \dot{\epsilon}, \rho_d) \epsilon + \sigma_i, where \sigma is the flow stress, \theta is a state-dependent hardening modulus influenced by plastic strain \epsilon, temperature T, strain rate \dot{\epsilon}, and defect density \rho_d, and \sigma_i represents the initial yield stress. The defect density \rho_d, which primarily accounts for dislocations, evolves according to the differential equation \dot{\rho_d} = \alpha \sqrt{\rho_d} \dot{\epsilon} - \beta \rho_d, where \alpha and \beta are material-specific coefficients governing dislocation multiplication (the storage term proportional to \sqrt{\rho_d} \dot{\epsilon}) and dynamic recovery (the annihilation term). This evolution equation captures the competition between defect generation during deformation and their thermal or athermal annihilation, leading to a characteristic saturation of hardening at large strains where \rho_d reaches a steady state. The model's derivation stems from fundamental mechanisms of driven by multiplication and recovery, integrated into a scale-invariant that couples the viscoplastic response to a hydrodynamic equation-of-state () for density and pressure effects. Specifically, the hardening modulus \theta incorporates thermal activation for motion at moderate rates and drag at ultra-high rates (>10^9 s^{-1}), ensuring applicability across regimes from quasistatic to overdriven shocks. This coupling enables seamless incorporation into Eulerian hydrocodes, where plastic work contributes to heating and phase changes. Key features include the asymptotic approach to a saturation stress at high strains, reflecting microstructural stabilization, and robust rate and temperature sensitivity without ad hoc multipliers. The PTW model has been implemented in Los Alamos National Laboratory (LANL) simulation codes, such as FLAG, for modeling explosives detonation and high-velocity impacts on metals. Parameters, including \alpha, \beta, and components of \theta (e.g., reference stresses and rate exponents), are calibrated to plate-impact experiments on tantalum, achieving predictions within 10% of measured Hugoniot elastic limits and release stresses at strain rates up to 4000 s^{-1} and temperatures from 77 K to 1273 K. For tantalum, typical fits yield initial defect densities around $10^{10} m^{-2} and saturation values near $10^{15} m^{-2}, validated against symmetric plate-impact data.

Applications and Implementations

Engineering Applications

Viscoplasticity plays a critical role in aerospace engineering, particularly in predicting the long-term performance of components exposed to high temperatures and sustained loads, such as turbine blades in jet engines. These blades experience creep deformation due to thermal and centrifugal stresses, where viscoplastic models like the Norton-Hoff formulation are employed to simulate secondary creep behavior and estimate remaining service life. For instance, in analyses of Inconel 718 superalloy blades, the Norton-Bailey variant of the Norton law has been used to model uniaxial creep strain accumulation under operational conditions, enabling accurate life predictions by integrating creep curves from high-temperature tests. This approach helps engineers design blades that withstand prolonged exposure to high temperatures exceeding 650°C without excessive deformation. In , viscoplastic models are essential for analyzing the time-dependent behavior of soils in geotechnical structures, especially foundations and during seismic events. Bingham models, which capture the yield and viscous of cohesive soils like clays, are widely applied to simulate in pile foundations under sustained loads, accounting for the non-linear between the pile and surrounding clay layers. During earthquakes, these models describe the and lateral spreading of saturated soils as viscoplastic Bingham media, allowing prediction of ground deformation and foundation stability by incorporating dynamic rates and pore buildup. Such simulations aid in designing resilient in seismically active regions. Manufacturing processes involving high temperatures and deformation rates, such as hot forming and , rely on viscoplasticity to model material flow and prevent defects. The Johnson-Cook model, which incorporates strain-rate hardening and thermal softening, is commonly used to predict the behavior of during hot , where rapid deformation at elevated temperatures leads to viscoplastic flow. This enables optimization of process parameters to achieve uniform microstructures and minimize cracking, as demonstrated in simulations of A356 relevant to extrusion forming. By capturing rate-dependent effects, the model supports efficient of complex shapes in industries like components. In the automotive sector, viscoplastic models enhance the accuracy of crash simulations by accounting for strain-rate hardening in high-speed . Materials like high-strength steels exhibit significant rate sensitivity during collisions, where viscoplastic formulations predict enhanced strength and compared to quasi-static conditions. For example, studies on dual-phase steels use viscoplasticity to model behavior, improving finite element predictions of deformation and . This integration allows for better design of energy-absorbing structures, reducing injury risks in real-world accidents. In , viscoplastic models are applied to simulate the and deformation of biofluids, such as , which exhibits yield-stress and time-dependent behavior. Thixo-elastoviscoplastic formulations capture the complex of under physiological conditions, including , , and plastic in vessels and clots, aiding in the of cardiovascular and risk. Despite these applications, challenges persist in viscoplastic modeling, particularly in parameter identification from experimental data and handling material anisotropy. Identifying parameters like viscosity coefficients and yield stresses requires careful calibration from diverse tests, as uniaxial data alone may not capture full behavior, leading to non-unique solutions in inverse analyses. Anisotropy, arising from manufacturing processes or microstructural features, is often inadequately addressed in isotropic models, complicating predictions for textured materials like rolled sheets or composites. Advanced strategies, such as Bayesian optimization or multi-scale approaches, are emerging to mitigate these issues and improve model reliability across engineering fields.

Numerical Simulation Methods

Numerical simulations of viscoplasticity require robust computational strategies to handle the rate-dependent and nonlinear nature of material deformation, often implemented within finite element frameworks to solve coupled mechanical problems. These methods focus on integrating constitutive equations over time while ensuring and accuracy, particularly for models exhibiting overstress or behaviors. Key challenges include managing the implicit-explicit balance in time stepping and deriving consistent linearizations for iterative solvers. Time integration schemes are central to simulating viscoplastic response, with the choice depending on the model's formulation and loading conditions. For overstress-based models like Perzyna, the implicit is widely adopted due to its unconditional stability, allowing larger time steps without oscillations in low-rate regimes; this fully implicit approach solves the nonlinear equations at each increment to update plastic strain and overstress consistently. In contrast, for high-strain-rate applications such as those using the Johnson-Cook model, explicit integration schemes are preferred to capture dynamic wave propagation and inertial effects efficiently, as they avoid solving large systems of equations per step but require small time increments for stability. In finite element implementations, viscoplasticity introduces significant challenges in the global solution process, particularly for implicit analyses using Newton-Raphson iteration. Deriving the consistent tangent modulus is essential, as it provides the for quadratic convergence; this involves linearizing the integrated stress-strain relations, often through algorithmic differentiation of the return mapping procedure, which projects trial states back to the while accounting for viscous flow. Return mapping algorithms further facilitate local integration at integration points, ensuring that the viscoplastic is satisfied incrementally, though they demand careful handling of multi-axial states and potential non-uniqueness in direction. Commercial software facilitates practical implementation of these methods via user-defined material subroutines. In , the UMAT interface enables custom viscoplastic rheological models, such as those based on Perzyna over, by providing updates and operators to the solver, supporting both implicit and explicit analyses for quasi-static to . Similarly, incorporates viscoplastic flow models like Johnson-Cook through built-in material cards or user subroutines, optimized for explicit dynamics in high-rate simulations such as impact and forming processes. Recent advancements extend these methods to multi-scale simulations, coupling macroscopic viscoplastic finite element models with microscopic crystal plasticity to capture grain-level rate effects and texture evolution in polycrystalline materials. Additionally, techniques have emerged for calibrating viscoplastic , using neural networks to map experimental data to model coefficients, thereby reducing computational cost and improving accuracy over traditional optimization. These developments address limitations in standalone phenomenological approaches by incorporating physics-informed surrogates for .

References

  1. [1]
    Viscoplasticity - an overview | ScienceDirect Topics
    Viscoplasticity refers to the mechanical response of solids involving time-dependent, irreversible (inelastic) strains.
  2. [2]
    Viscoplastic Material - an overview | ScienceDirect Topics
    A viscoplastic material is defined as one that exhibits irreversible deformation in response to applied loading, with permanent deformation linked to the ...
  3. [3]
    Review of a Unified Elastic—Viscoplastic Theory - SpringerLink
    Although inelastic response of solid materials at low stress levels has been observed and measured for over a century and a half (an account of the early ...
  4. [4]
    [PDF] 19910004448.pdf - NASA Technical Reports Server (NTRS)
    In this discussion the modelsare reviewed only in uniaxial form because in virtually all cases they are converted to multiaxial form by using J2.
  5. [5]
    A review of some plasticity and viscoplasticity theories - ResearchGate
    Aug 6, 2025 · The purpose of the present review article is twofold: recall elementary notions as well as the main ingredients and assumptions of developing macroscopic ...
  6. [6]
    Constitutive laws - 3.8 Viscoplasticity - Applied Mechanics of Solids
    Viscoplastic constitutive equations are used to model the behavior of polycrystalline materials (metals and ceramics) that are subjected to stress at high ...
  7. [7]
    Viscoplasticity – Knowledge and References - Taylor & Francis
    Viscoplasticity refers to the property of a material to exhibit permanent deformations over time, similar to plasticity, but with a rate-dependent behavior.
  8. [8]
    Viscoplasticity: Theory, Modelling and Applications
    Jun 26, 2025 · Biofluids, such as blood and saliva, and geophysical materials, e.g., sea ice, snow, and lava, are also viscoplastic. A remarkable number of ...
  9. [9]
  10. [10]
    Viscoplasticity - an overview | ScienceDirect Topics
    Viscoplasticity is defined as the irreversible but rate-dependent deformation of a solid material under load, where the magnitude of the load influences the ...
  11. [11]
    None
    ### Summary of Isotropic Hardening in Viscoplasticity from NASA TM-102388
  12. [12]
    Mixed hardening hyper-viscoplasticity model for soils incorporating ...
    In this paper, a new mixed hardening hyper-viscoplasticity model is proposed for the derivation of the time-dependent constitutive behaviour of soils, with the ...
  13. [13]
    Elastic-viscoplastic modeling of soft biological tissues using a mixed ...
    The characteristic highly nonlinear, time-dependent, and often inelastic material response of soft biological tissues can be expressed in a set of ...
  14. [14]
  15. [15]
    [PDF] Comparison of Viscoelastic/Viscoplastic Models for Describing the ...
    Apr 20, 2023 · ❖ First assume viscoelasticity only. ❖ Viscoplasticity is activated when load is beyond yield surface. ❖ Damage factors are activated under ...
  16. [16]
    Viscoplastic Behavior - an overview | ScienceDirect Topics
    Viscoplastic behavior is when materials show both viscous and plastic behavior, resulting in irrecoverable deformations, and is rate-dependent.
  17. [17]
    [PDF] Hypoplastic models for fine-grained soils A dissertation submitted for ...
    ... and visco- plasticity and may be seen as different extensions of the ... Soil behaviour and critical state soil mechanics. Cambridge University. Press ...
  18. [18]
    [PDF] Phenomenological modelling of viscoplasticity - HAL
    two different "flow stress" curves for the same rate depending on the prior history. In materials science a steady-state condition is postulated at sufficiently ...
  19. [19]
    CHAPTER 19 - Deformation-Mechanism Maps
    19.1 Lead pipes on a 75-year-old building in southern New England. The creep-induced curvature of these pipes is typical of Victorian lead water piping.
  20. [20]
    When did the problem of creep in metals first get attention?
    May 10, 2017 · To elaborate:- Creep (deformation) was well known observation for a long time in deformation of rocks, ice, sagging of lead pipes, failure of W ...
  21. [21]
    Damage Control: High-Temperature Creep Detection
    Oct 27, 2022 · Time-dependent creep deformation in metals was first observed in the nineteenth century [2]. ... This early experimental work involved testing ...Missing: 19th | Show results with:19th
  22. [22]
    Historical Survey of the Development of Creep Mechanics from its ...
    During the 19th century observations of creep in engineering structures must have been made, but its real nature was discovered only in the early 20th century ...
  23. [23]
    On the viscous flow in metals, and allied phenomena - Journals
    Kennedy A (2002) A Method of Fitting the Andrade Creep Equation to Experimental Results, Proceedings of the Physical Society, 10.1088/0959-5309/61/6/304, 61 ...
  24. [24]
    Review: developments in the creep of materials over a period of ...
    The most comprehensive early experiments on the creep of metals were undertaken by Andrade and published in detailed reports in 1910 [4] and 1914 [5]. These ...
  25. [25]
    Soil Mechanics - an overview | ScienceDirect Topics
    Soil mechanics is the study of soil's mechanical behavior, including its structure, stability, and resistance to deformation under stress.
  26. [26]
    [PDF] f?V- 35 - DTIC
    The separation of time-dependent creep deformation from time-independent plastic deformation ... tests ... Time-dependent behavior of soils has been shown to be ...
  27. [27]
    Approximate analysis of structures in the presence of moderately ...
    Approximate analysis of structures in the presence of moderately large creep deformations · N. .. Hoff · Published 1 April 1954 · Engineering, Materials Science, ...
  28. [28]
  29. [29]
    A review of some plasticity and viscoplasticity constitutive theories
    The purpose of the present review article is twofold: recall elementary notions as well as the main ingredients and assumptions of developing macroscopic ...
  30. [30]
    [PDF] Viscoplastic Characterization of Ti-6-4: Experiments
    Tensile, creep, and stress relaxation tests were performed over a wide range of temperatures and strain rates to engage various amounts of time-dependent ...
  31. [31]
    Stress relaxation behavior of low carbon steel at different temperatures
    Oct 31, 2023 · In this paper, the stress relaxation behavior of Q235 with the initial tensile stress of 70, 85 and 100 MPa were investigated at different temperature.
  32. [32]
    Logarithmic versus Andrade's transient creep: Role of elastic stress ...
    Mar 13, 2023 · A logarithmic increase of strain under constant stress (creep test), ɛ ∼ ln ( c + t ) , can be mirrored by a logarithmic decrease of the stress ...
  33. [33]
    [PDF] Determination of material parameters for a unified viscoplasticity ...
    A procedure for obtaining estimates of Z and n from this stress relaxation curve will be outlined here. During stress relaxation stage, total strain rate can be ...
  34. [34]
    Applied Mechanics of Solids (A.F. Bower) Section 3.8: Viscoplasticity
    3.8 Small Strain Viscoplasticity · 3.9 Large Strain Viscoplasticity · 3.10 Large Strain Viscoelasticity · 3.11 Soils · 3.12 Crystal Plasticity · 3.13 Surfaces ...
  35. [35]
    (PDF) Study of Viscoplastic Flows Governed by the Norton-Hoff ...
    Aug 6, 2025 · We deal with viscoplastic flows. The fluid motion is governed by the nonlinear incompressible Norton-Hoff operator with homogeneous boundary ...
  36. [36]
    Viscoplasticity with creep and plasticity bounds - ScienceDirect.com
    Abstract. A viscoplastic theory is developed that reduces analytically to creep theory under steady-state conditions and becomes plasticity theory at its rate- ...Missing: Hoff original papers
  37. [37]
    The Creep of Steel at High Temperatures - Frederick Harwood Norton
    F. H. Norton No preview available - 2015. The Creep of Steel at High ... The Creep of Steel at High Temperatures Issue 35, Part 1 of Bulletin, Cornell ...
  38. [38]
    Shape Sensitivity and Large Deformation of the Domain for Norton ...
    Norton-Hoff Penalized Problem. The model we study here was first introduced by Norton to modelize the creep of steel at high-temperature, and generalized by ...Missing: Nicholas | Show results with:Nicholas
  39. [39]
    Creep in Structures 1970 - SpringerLink
    The first IUTAM Symposium on Creep in Structures was held in Stanford, Cal. 1960 (Proceedings, ed. N. J. HOFF, Springer-Verlag 1962).
  40. [40]
  41. [41]
    [PDF] Approximation of the Norton–Hoff plasticity model with isotropic ...
    The Norton–Hoff model with isotropic hardening is a Norton–Hoff model with one additional scalar function, which is called isotropic hardening. This model.
  42. [42]
    Analysis of strain-hardening viscoplastic thick-walled sphere and ...
    Abstract. The boundary value problems of thick-walled sphere and cylinder, made from an isotropic strain-hardening viscoplastic material, and subjected to ...Missing: paper | Show results with:paper
  43. [43]
    Fundamental Problems in Viscoplasticity - ScienceDirect.com
    View PDF; Download full volume. Search ScienceDirect. Elsevier · Advances in Applied Mechanics · Volume 9, 1966, Pages 243-377. Advances in Applied Mechanics ...
  44. [44]
    Numerical simulation of finite strain viscoplastic problems
    In summary from the studies conducted we can conclude that results are quite insensitive to the finite element mesh chosen for the viscoplastic response.Numerical Simulation Of... · 2. Large Strain... · 5. Numerical Examples
  45. [45]
    (PDF) Calibration of perzyna-type elasto-viscoplastic models from ...
    In this work it is performed a calibration of Perzyna-type elastic-viscoplastic constitutive models under infinitesimal strains assumption.
  46. [46]
    [PDF] Constitutive modelling of non-cohesive soils under high-strain rates
    May 17, 2022 · At high strain rates (HSR), soils show more strength and enhanced dilation. (viscoplastic behaviour) compared to the response at low rates ( ...
  47. [47]
    An internal variable variational formulation of viscoplasticity
    In the paper evolutive processes and variational principles in viscoplasticity are examined in detail. The treatment is faced by the recourse to an internal ...
  48. [48]
    On Abstract Variational Inequalities in Viscoplasticity with Frictional ...
    May 26, 2007 · In this paper, we study quasistatic abstract variational inequalities with time-dependent constraints. We prove existence results and ...
  49. [49]
    (PDF) Comparison of viscoplasticity formats and algorithms
    Aug 5, 2025 · Algorithmic issues for the two thermodynamically consistent viscoplastic formulations of Perzyna and Duvaut–Lions are discussed.
  50. [50]
    Viscoplastic augmentation of the smooth cap model
    ### Summary of Duvaut-Lions Viscoplastic Model from the Article
  51. [51]
    a constitutive model and data for metals subjected to large strains ...
    This paper presents a constitutive model and data for materials subjected td large strains, high strain rates and high temperatures.
  52. [52]
    Johnson-Cook parameter evaluation from ballistic impact data via ...
    A methodology is presented for evaluating a strain rate sensitivity parameter for plastic deformation of bulk metallic materials.
  53. [53]
    (PDF) Johnson - Cook Strength Models for Mild and DP 590 Steels
    The J-C model constants A, B, n, C, and m for 1080 steel are 0.514 GPa, 2.83 GPa, 0.612, 0.031, and 0.890, respectively. For Vascomax 300 steel A=2.07 GPa; B= ...
  54. [54]
    A Modified Johnson–Cook Constitutive Model and Its Application to ...
    This paper aims to investigate the material dynamic deformation during high speed machining of 7050-T7451 aluminum alloy with the aid of split Hopkinson ...
  55. [55]
    [PDF] Introduction to PTW
    Mar 16, 2004 · In the PTW model, the plastic stress in a material is a function of the amount of strain ψ it has undergone, the strain rate ˙ψ, the material ...
  56. [56]
    [PDF] Modelling and creep strain analysis of inconel 718 alloy by uniaxial ...
    In the present study, Norton-Bailey power law is used to predict the creep properties of a turbine blade test specimen made of nickel-based superalloy named.
  57. [57]
    Creep-Based Reliability Evaluation of Turbine Blade-Tip Clearance ...
    Oct 29, 2019 · Creep curve of a structure. In engineering, the Norton implicit model is usually adopted to express the second constitutive relation of ...
  58. [58]
    Numerical prediction of liquefied ground characteristics from back ...
    In this paper, liquefied and laterally spreading soils triggered by seismic shaking are modeled as viscoplastic Bingham media characterized by two ...
  59. [59]
    An Enhanced Johnson–Cook Model for Hot Compressed A356 ...
    Al–Si–Mg alloy parts is performed by hot. forming technology, such as hot extrusion,. spinning, and forging, which can be con-. sidered as a practicable method ...
  60. [60]
    Study of Viscoplasticity Models for the Impact Behavior of High ...
    Aug 9, 2025 · This study reports on modeling the mechanical behavior of high-strength steels subjected to impact loading. The materials studied were steel ...<|control11|><|separator|>
  61. [61]
    [PDF] Crash Analysis of Auto-body Structures Considering the Strain-Rate ...
    The crashworthiness of vehicles with finite element methods depends on the geometry modeling and the material properties. The vehicle body structures are ...
  62. [62]
    Parameter identification for viscoplastic models based on analytical ...
    In this work a unified strategy for identification of material parameters of viscoplastic constitutive equations from uniaxial test data is presented.
  63. [63]
    An anisotropic viscoplasticity model for shale based on layered ...
    Dec 3, 2020 · This provides a significant advantage over many single-scale models that require anisotropy parameters that can be challenging to calibrate and ...
  64. [64]
    [PDF] Robust Integration Schemes for Generalized Viscoplasticity With ...
    This two-part report is concerned with the development of a general framework for the implicit time-stepping integrators for the flow and evolution equations in.Missing: Perzyna | Show results with:Perzyna
  65. [65]
    [PDF] Implicit integration of the Perzyna viscoplastic material model
    May 16, 1995 · A robust numerical integration is the Euler backward, or fully implicit, algorithm. A fully implicit algorithm is derived for the Perzyna ...
  66. [66]
    Comparative Study on High Strain Rate Fracture Modelling Using ...
    The Johnson–Cook model, with and without a modification of the strain rate ... explicit time integration—partly compensate for this. The question to be ...
  67. [67]
    [PDF] Return Mapping Algorithms (RMAs) for Two-Yield Surface ... - HAL
    Jun 11, 2019 · The classical Newton–Raphson algorithm used to solve the overall equilibrium problem requires the use of a global tangent operator which is also ...
  68. [68]
    (PDF) Finite element implementation of a certain class of elasto ...
    Jun 12, 2023 · Newton–Raphson (N-R) method was used for solving the nonlinear ... The radial return mapping algorithm is utilized to discretize the general form ...
  69. [69]
    [PDF] ABAQUS user subroutines for the simulation of viscoplastic ...
    The resulting models have been implemented into subroutines of the FE-code ABAQUS as "user-defined material models" (UMAT) and can be used to perform FE ...
  70. [70]
    [PDF] Implementation of Constitutive Equations for Viscoplasticity - Dynalook
    Johnson-Cook. Equation 5 shows the Johnson-Cook relation, giving the flow stress in terms of plastic strain, strain rate and temperature. The plasticity in ...
  71. [71]
    Multiscale modeling of plasticity based on embedding the ...
    This paper is concerned with the multiscale simulation of plastic deformation of metallic specimens using physically-based models that take into account ...
  72. [72]
    Use of machine learning in determining the parameters of ...
    The aim of this paper is to introduce an alternative method, based on artificial neural networks, for determining the parameters of a viscoplastic model. Design ...