Fact-checked by Grok 2 weeks ago

Generalized Maxwell model

The generalized Maxwell model, also known as the Maxwell–Wiechert model, is a linear viscoelastic constitutive model that describes the time-dependent deformation of materials exhibiting both elastic recovery and viscous flow under mechanical stress. It represents the most general form of linear viscoelasticity by combining a single elastic spring in parallel with multiple Maxwell elements, where each Maxwell element consists of a spring (elastic component) and a dashpot (viscous component) connected in series, enabling the model to capture a broad spectrum of relaxation times corresponding to different molecular or structural scales in the material. Developed as an extension of James Clerk Maxwell's 1867 simple viscoelastic model, the generalized version was introduced by Emil Wiechert in the early to better approximate the complex relaxation behaviors observed in real materials like polymers, where a single relaxation time is insufficient. The model's structure allows it to reduce to simpler forms, such as the standard Maxwell model (one element, fluid-like) or the Kelvin-Voigt model (parallel spring-dashpot, solid-like), by setting appropriate elastic constants or viscosities, making it versatile for both solid and fluid viscoelastic responses. Mathematically, the total stress is the sum of contributions from the equilibrium spring and each Maxwell branch, with the relaxation expressed as a Prony series: G(t) = G_\infty + \sum_{i=1}^N G_i e^{-t/\tau_i}, where G_\infty is the long-term , G_i are branch , and \tau_i = \eta_i / G_i are relaxation times. In creep tests under constant , the model predicts an initial elastic followed by time-dependent viscous that approaches a if an equilibrium is included. Conversely, in relaxation tests under constant , decays exponentially across multiple timescales, reflecting the material's ability to dissipate over time. This multi-time-scale capability makes the model superior to basic viscoelastic representations for accurately fitting experimental data from materials such as , rubber, biological tissues, and composites. The generalized Maxwell model finds wide applications in engineering simulations, including finite element analysis for , geomechanics, and , where it is implemented via integral formulations or state-variable evolution equations to handle deviatoric while keeping volumetric behavior elastic. Parameter identification often involves optimizing relaxation times and moduli to match Prony series fits from or creep-recovery tests, ensuring in simulations by constraining time steps to fractions of the shortest relaxation time.

Background Concepts

Viscoelasticity Fundamentals

is a that exhibits both viscous and characteristics when undergoing deformation, allowing materials to store and dissipate in a time-dependent manner. This behavior arises from the molecular structure, particularly in polymers and biological tissues, where competes with viscous . Unlike purely materials, which return to their original shape instantaneously, or purely viscous fluids, which deform permanently, viscoelastic materials display a combination of these responses under applied or . The foundational developments in occurred in the , with contributions from physicists such as James Clerk Maxwell, who proposed a model for viscous-elastic fluids in 1867; , who explored damping and elasticity in metals; and , who in 1892 introduced a model combining and viscous elements in . These early works laid the groundwork for understanding time-dependent material responses, initially motivated by observations in gases, metals, and glasses. Linear viscoelasticity, the focus of theoretical analysis, assumes small deformations where the response is proportional to the applied load and governed by the principle of superposition. Key phenomena in linear viscoelasticity include stress relaxation, creep, and hysteresis. Stress relaxation occurs when a material is subjected to constant strain, resulting in a gradual decrease in stress over time due to internal rearrangements, quantified by the relaxation modulus E(t) = \sigma(t)/\epsilon_0. Creep describes the progressive increase in strain under constant stress, characterized by the creep compliance D(t) = \epsilon(t)/\sigma_0, often showing an initial elastic response followed by viscous flow. Hysteresis manifests in cyclic loading as a loop in the stress-strain curve, indicating energy dissipation through viscous mechanisms, with the area of the loop representing lost energy per cycle. These behaviors are analyzed within the linear regime, where the total response is the superposition of incremental effects, as per Boltzmann's principle. The general for linear is expressed through the Boltzmann superposition , which relates \sigma(t) to the of : \sigma(t) = \int_{-\infty}^{t} G(t - \tau) \frac{d\epsilon(\tau)}{d\tau} \, d\tau Here, G(t) is the relaxation , capturing the material's of past deformations, and the assumes and time-invariance. This form encapsulates the time-dependent nature of viscoelastic response under small s, where ensures scalability of and .

Maxwell Element

The Maxwell element, introduced by James Clerk Maxwell in his foundational work on the dynamical theory of gases, represents the simplest viscoelastic model combining elastic and viscous responses in series. It consists of a Hookean with E, representing instantaneous recoverable deformation, connected in series to a Newtonian with \eta, representing time-dependent irreversible flow. The constitutive equation for the Maxwell element relates stress \sigma and strain \epsilon through their rates, derived from the series configuration where total strain rate is the sum of elastic and viscous contributions: \frac{d\epsilon}{dt} = \frac{\sigma}{\eta} + \frac{1}{E} \frac{d\sigma}{dt}. This differential equation captures the interplay between viscous flow and elastic recovery, with the dashpot enforcing equal stress across elements and the spring enforcing equal strain rates in series. Under constant \epsilon = \epsilon_0 applied at t=0, the Maxwell element exhibits , where initial stress \sigma(0) = E \epsilon_0 decays exponentially over time. Substituting \frac{d\epsilon}{dt} = 0 into the yields \frac{d\sigma}{dt} = -\frac{E}{\eta} \sigma, a first-order differential equation solved as \sigma(t) = \sigma(0) \exp\left(-\frac{t}{\tau}\right), with relaxation \tau = \frac{\eta}{E}. The relaxation , defined as G(t) = \frac{\sigma(t)}{\epsilon_0}, thus follows: G(t) = E \exp\left(-\frac{t}{\tau}\right). This exponential characterizes the transition from initial to viscous dominance. Despite its simplicity, the Maxwell element has key limitations: it lacks an instantaneous elastic response upon sudden loading, as the prevents immediate deformation, and it predicts unbounded long-term flow under sustained , exhibiting fluid-like behavior rather than solid equilibrium. These traits make it suitable for modeling liquids or highly viscous materials but inadequate for solids without additional parallel elements.

Model Formulation

General Structure

The generalized Maxwell model consists of N parallel Maxwell elements, where each element is formed by a spring of modulus G_i in series with a dashpot of viscosity \eta_i (for i = 1 to N), optionally augmented by an additional spring of modulus G_{eq} connected in parallel to represent equilibrium elastic behavior in solids. This schematic arrangement, also referred to as the Maxwell-Wiechert model, allows for the representation of a discrete of relaxation processes through distinct relaxation times \tau_i = \eta_i / G_i for each arm. Each individual Maxwell arm corresponds to the basic Maxwell element, capturing combined elastic and viscous deformation in series. In this parallel configuration, all elements experience the same total \epsilon, while the total \sigma is the sum of the stresses contributed by each Maxwell arm \sigma_i and the equilibrium spring: \sigma = \sum_{i=1}^N \sigma_i + G_{eq} \epsilon. The physical interpretation lies in modeling materials with multiple characteristic timescales for , where shorter \tau_i correspond to rapid viscous dissipation and longer ones to slower recovery, enabling accurate approximation of experimental relaxation spectra in polymers and other viscoelastic media. For viscoelastic fluids, the model sets G_{eq} = 0, resulting in no long-term elastic modulus and permitting unbounded deformation under sustained stress, akin to flow behavior. In contrast, including G_{eq} > 0 yields a solid-like response with finite equilibrium stiffness. The overall constitutive relation manifests as a high-order linear ordinary differential equation (ODE) of order N, coupling the stress \sigma and strain \epsilon along with their time derivatives up to the N-th order, typically expressed in the form \left(1 + \sum_{k=1}^{N} a_k \frac{d^k}{dt^k}\right) \sigma(t) = \left( E_0 + \sum_{k=1}^{N} b_k \frac{d^k}{dt^k}\right) \epsilon(t), where the coefficients a_k, b_k, and E_0 are determined by the model parameters and ensure physical realizability.

Relaxation Modulus Expression

The relaxation modulus G(t) of the generalized Maxwell model arises from the applied to the parallel configuration of multiple Maxwell elements, each consisting of a and in series, optionally augmented by an equilibrium . For a step input \gamma_0 applied at t=0, the response \sigma(t) in each Maxwell arm decays exponentially as \sigma_i(t) = G_i \gamma_0 \exp(-t / \tau_i), where G_i is the and \tau_i is the relaxation time of the i-th arm. The total is the sum of contributions from all arms and the equilibrium , yielding the relaxation modulus via \sigma(t) = G(t) \gamma_0. The primary expression for the relaxation modulus is thus G(t) = G_{\mathrm{eq}} + \sum_{i=1}^N G_i \exp\left( -\frac{t}{\tau_i} \right), where G_{\mathrm{eq}} is the equilibrium modulus (zero for purely viscous responses), G_i are the moduli of the N Maxwell arms, and \tau_i = \eta_i / G_i are the corresponding relaxation times with viscosities \eta_i. For viscoelastic fluids, the model omits the equilibrium spring, setting G_{\mathrm{eq}} = 0, so G(t) = \sum_{i=1}^N G_i \exp(-t / \tau_i) and G(t) \to 0 as t \to \infty, reflecting complete over long times. This form is often normalized and expressed in Prony series notation as \frac{G(t)}{G_0} = g_{\infty} + \sum_{i=1}^N g_i \exp\left( -\frac{t}{\tau_i} \right), where G_0 = G_{\mathrm{eq}} + \sum_{i=1}^N G_i is the instantaneous , g_{\infty} = G_{\mathrm{eq}} / G_0, and g_i = G_i / G_0 with \sum g_i + g_{\infty} = 1. The moduli G_i and G_{\mathrm{eq}} have units of pascals (), while the relaxation times \tau_i are in seconds (s).

Mathematical Properties

Time-Domain Behavior

In the time domain, the generalized Maxwell model predicts stress relaxation through the relaxation modulus G(t), which decays as a sum of exponentials: G(t) = G_\mathrm{e} + \sum_{i=1}^N G_i \exp(-t / \tau_i), where G_\mathrm{e} is the equilibrium modulus, G_i are the moduli of individual Maxwell elements, and \tau_i = \eta_i / G_i are the relaxation times. For N > 1, this multi-exponential form results in a multi-modal decay, with distinct relaxation regimes corresponding to short and long timescales; on a logarithmic plot of G(t) versus time, the curve exhibits steps or plateaus if the \tau_i are logarithmically spaced, reflecting sequential relaxation of molecular segments or entanglements. The creep J(t) is derived from the inverse relation to G(t) via the and satisfies the t = \int_0^t G(s) J(t - s) \, ds. For a solid configuration (finite G_\mathrm{e} > 0), it can be represented using a Prony series in terms of parameters: J(t) = J_g + \sum_{k=1}^M J_k \left(1 - \exp\left(-\frac{t}{\lambda_k}\right)\right), where J_g = 1/(G_\mathrm{e} + \sum G_i) is the glassy (instantaneous) , J_k > 0 are the strengths, and \lambda_k are the times, which are related to the relaxation parameters \{G_i, \tau_i\} through interconversion methods such as numerical inversion or approximation schemes. In the general case for fluids (with G_\mathrm{e} = 0), an additional linear term t / \eta_0 is included to capture steady-state viscous flow, where \eta_0 = \sum_{i=1}^N G_i \tau_i is the zero-shear obtained by integrating G(t). Long-term behavior differs markedly between solids and fluids: in solids, G(t) \to G_\mathrm{e} > 0 and J(t) \to 1/G_\mathrm{e}, yielding a finite response with no permanent ; in fluids (G_\mathrm{e} = 0), G(t) \to 0 and J(t) \sim t / \eta_0, indicating unbounded viscous under sustained . For step responses, a sudden \epsilon_0 elicits an initial \sigma(0) = [G_\mathrm{e} + \sum G_i] \epsilon_0 that relaxes according to G(t), while a sudden \sigma_0 produces an initial \epsilon(0) = \sigma_0 / [G_\mathrm{e} + \sum G_i] that evolves via J(t), transitioning from glassy to rubbery compliance. Numerical simulations of multi-order systems (N > 1) involve time-stepping methods to solve the coupled ordinary differential equations for each element's \sigma_i(t), given by \dot{\sigma}_i + \sigma_i / \tau_i = G_i \dot{\epsilon}(t), with total \sigma(t) = G_\mathrm{e} \epsilon(t) + \sum \sigma_i(t); explicit or implicit schemes ensure stability for stiff systems with widely varying \tau_i.

Frequency-Domain Representation

In the frequency domain, the generalized Maxwell model describes the viscoelastic response under oscillatory or sinusoidal loading, where the material is subjected to harmonic deformations at angular frequency \omega. The behavior is characterized by the complex modulus G^*(\omega) = G'(\omega) + i G''(\omega), with the storage modulus G'(\omega) representing the elastic energy storage and the loss modulus G''(\omega) representing the viscous energy dissipation. For the discrete form of the model, consisting of multiple Maxwell elements in parallel with an optional equilibrium spring, the storage modulus is given by G'(\omega) = G_\text{eq} + \sum_i G_i \frac{(\omega \tau_i)^2}{1 + (\omega \tau_i)^2}, and the loss modulus by G''(\omega) = \sum_i G_i \frac{\omega \tau_i}{1 + (\omega \tau_i)^2}, where G_\text{eq} is the equilibrium modulus (zero for fluids), G_i are the moduli of the springs in the i-th Maxwell element, and \tau_i are the corresponding relaxation times. These expressions arise from the continuous relaxation spectrum H(\tau), where the sums approximate the integrals G'(\omega) = G_\text{eq} + \int_{-\infty}^{\infty} H(\tau) \frac{\omega^2 \tau^2}{1 + \omega^2 \tau^2} \, d\ln\tau and G''(\tau) = \int_{-\infty}^{\infty} H(\tau) \frac{\omega \tau}{1 + \omega^2 \tau^2} \, d\ln\tau. The complex modulus G^*(\omega) is derived from the time-domain relaxation modulus G(t) via the Fourier transform, specifically G^*(\omega) = i\omega \int_0^\infty G(t) e^{-i\omega t} \, dt, or equivalently using the Laplace transform by substituting s = i\omega. This transformation connects the transient relaxation behavior to steady-state harmonic responses, enabling analysis in dynamic mechanical spectroscopy where sinusoidal strains reveal frequency-dependent properties. The loss tangent, defined as \tan \delta(\omega) = G''(\omega) / G'(\omega), quantifies the ratio of dissipated to stored energy per cycle and peaks at frequencies corresponding to relaxation transitions, providing into mechanisms in viscoelastic materials. To capture broad relaxation spectra over limited experimental frequency ranges, master curves are constructed using time-temperature superposition, shifting data from different temperatures by a factor a_T to form a single composite curve at a reference temperature; this principle, rooted in the thermorheologically simple behavior of many polymers, reveals frequency-dependent stiffening at high \omega (dominated by glassy modulus) and softening at low \omega (approaching rubbery or equilibrium response). This frequency-domain formulation excels in modeling polymers, where it accurately represents the progressive stiffening with increasing frequency due to chain segment immobilization and softening from cooperative motions, as observed in of amorphous and entangled systems.

Applications in Materials

Solid Viscoelastic Models

The generalized Maxwell model applied to solid viscoelastic materials incorporates a positive G_{eq} > 0, which represents the long-term elastic response and ensures that the material does not fully relax under sustained stress, thereby capturing the finite elasticity characteristic of solids such as rubbers and glassy polymers. This component arises from a parallel spring in the model structure, preventing complete stress decay and maintaining structural integrity over extended timescales, in contrast to purely viscous behaviors. In the relaxation expression G(t), the term G_{eq} provides the asymptotic value as t \to \infty, enabling accurate prediction of residual stiffness in solid-like responses. A prominent special case of the generalized Maxwell model for solids is the standard linear solid (SLS), achieved with N=1 Maxwell arm in parallel with the equilibrium spring G_{eq}, which is mathematically equivalent to the Zener model introduced in 1948. The SLS effectively describes materials exhibiting both instantaneous elasticity and time-dependent viscous effects, with the equilibrium modulus ensuring a non-zero long-term suitable for solid deformation analysis. This three-parameter configuration balances simplicity and realism, making it ideal for initial modeling of solid before extending to more complex forms. For broader relaxation spectra in heterogeneous solids, the model employs multiple arms (N=2 or more), allowing each arm to contribute distinct relaxation mechanisms that span wide time scales and capture the distributed viscoelastic response in materials like composites or soft biological tissues. In composites, these additional arms account for interfacial effects and filler-induced broadening of the relaxation , enhancing fidelity in predicting dynamic mechanical properties. Similarly, in soft tissues, multi-arm configurations model the hierarchical structure of and networks, providing a representation of and recovery behaviors under physiological loads. Applications of the generalized Maxwell model in solid include finite element simulations in , where it facilitates stress-strain analysis of arterial walls and other load-bearing tissues by incorporating multi-modal relaxation to match experimental data. In , the model supports stress analysis in rubbers, such as styrene-butadiene rubber (SBR) composites, by parameterizing relaxation to evaluate fatigue and under cyclic deformation. These implementations often rely on for time-domain solutions, enabling predictive design in both fields. Typical relaxation times \tau_i in the model for solid materials range from seconds to hours, reflecting the diverse molecular mobilities in polymers and elastomers; for instance, short \tau_i (around 0.01–10 seconds) dominate glassy transitions, while longer ones (up to hours) govern rubbery plateau behaviors. Parameter identification from ensures these values align with experimental spectra, optimizing model accuracy for specific solids.

Fluid Viscoelastic Models

The generalized Maxwell model applied to fluid viscoelastic behaviors sets the equilibrium modulus G_{eq} = 0, resulting in a relaxation modulus G(t) that decays to zero over long times, which captures the dissipative viscous dominant in materials like melts and solutions without residual elasticity. This configuration aligns with parallel arrangements of Maxwell elements, enabling the description of time-dependent leading to complete flow, as opposed to partial in . Such models are essential for predicting the rheological response of where elastic effects are transient, influenced by molecular chain entanglements that relax under sustained deformation. A foundational example is the three-parameter Jeffreys model, proposed by in 1929, which consists of a Maxwell element in parallel with a (Newtonian solvent), extending the simple Maxwell model by including a retardation time. This structure models the interplay between viscous damping and elastic storage in fluids, predicting steady under constant , as derived from irreversible for geophysical and rheological contexts. The Jeffreys model formulation allows accurate representation of fluids with finite retardation times that exhibit delayed compliance. The generalized Oldroyd-B model emerges as a multi-mode extension, analogous to multiple parallel Maxwell arms combined with a Newtonian , providing a comprehensive framework for complex with distributed relaxation spectra. Originally formulated for non-Newtonian flows, its generalized form incorporates several relaxation times to mimic polydisperse systems, enhancing predictions of -thinning and normal stress effects in dilute to semi-dilute solutions. In steady flows at low rates, these models yield a Newtonian \eta_0 = \sum \eta_i, where \eta_i = G_i \tau_i sums the contributions from each Maxwell arm's , establishing the baseline flow resistance before nonlinear deviations. These fluid models find practical utility in processes involving viscoelastic liquids, such as , where the transient elasticity affects droplet ejection and satellite formation in polymer-based inks, requiring precise tuning of relaxation times for high-fidelity deposition. In food processing, they describe the flow of viscoelastic solutions or melts during and mixing, optimizing texture and stability in products like gels or sauces by accounting for shear-induced relaxation.

Model Comparisons

With Single Maxwell Model

The single Maxwell model, consisting of a spring and dashpot in series, is limited by its single relaxation time constant \tau, which restricts its ability to describe materials with broad relaxation spectra, such as polymers exhibiting multiple molecular relaxation processes. This model also lacks an equilibrium modulus, predicting complete relaxation to zero under constant , which does not align with real materials that retain some response at long times. Furthermore, it exhibits unlimited under sustained , leading to infinite over time, a behavior unsuitable for solids or semi-solids. In contrast, the generalized Maxwell model overcomes these shortcomings by incorporating multiple Maxwell elements in parallel, each with distinct relaxation times \tau_i, enabling it to capture non-exponential relaxation typical of complex materials like polymers and biological tissues. An additional equilibrium spring with modulus G_{eq} can be included in parallel, allowing for partial relaxation where stress approaches a finite value rather than zero, thus better representing materials with long-term elasticity. For instance, while the single model forecasts full dissipation of stress in a step strain experiment, the generalized model can fit experimental data showing residual modulus, as seen in rubber-like polymers. The single Maxwell model remains appropriate for simple fluids with narrow relaxation spectra, such as dilute solutions exhibiting primarily viscous flow on long timescales. However, for broader applications, the generalized form provides superior realism. Historically, the evolution from the single to the generalized Maxwell model gained prominence in modeling after the , driven by the need to represent diverse viscoelastic behaviors in synthetic materials; (1950) demonstrated that any linear viscoelastic response could be equivalently expressed using such multi-element configurations, facilitating their adoption in rheological studies.

With Kelvin-Voigt Model

The Kelvin-Voigt model consists of a and connected in parallel, combining elastic and viscous responses to exhibit delayed elasticity under applied stress. This configuration allows the model to effectively capture behavior, where gradually increases over time under constant stress, mimicking the retarded elastic deformation observed in certain materials. However, it fails to describe , as the parallel arrangement results in a constant stress response under fixed , with no decay to zero over time. In contrast, the generalized Maxwell model, comprising multiple Maxwell elements (each a spring-dashpot in series) arranged in parallel, excels at modeling while providing a more nuanced representation of viscoelastic dynamics. Unlike the Kelvin-Voigt model, which lacks an instantaneous elastic response due to the dashpot's resistance to sudden deformation, the generalized Maxwell model incorporates immediate upon loading followed by relaxation through its distributed relaxation times. This makes it superior for materials where time-dependent stress decay is prominent, as the Kelvin-Voigt predicts an infinite relaxation time, whereas the generalized Maxwell yields finite relaxation times \tau_i = \eta_i / G_i for each branch, enabling accurate fitting to experimental relaxation spectra. The generalized Maxwell model's ability to hybridize multiple relaxation modes provides deeper insights into materials dominated by relaxation processes, such as , where the Kelvin-Voigt's simplistic parallel structure oversimplifies the spectrum of time scales. For instance, in polymer melts or networks, the generalized Maxwell captures the broad distribution of molecular relaxations that lead to eventual stress equilibration, a feature absent in the Kelvin-Voigt framework. In practical applications, the Kelvin-Voigt model is often employed for porous materials like polymeric foams, where under is the primary concern, such as in cushioning or . Conversely, the generalized Maxwell model is preferred for amorphous materials like glasses, which exhibit pronounced viscoelastic relaxation during thermal processing or mechanical loading near the .

References

  1. [1]
    Generalized Maxwell Viscoelastic Models - PyLith - Read the Docs
    The most general form of linear viscoelastic model is the generalized Maxwell model, which can be represented by a spring in parallel with a number of Maxwell ...
  2. [2]
    [PDF] ENGINEERING VISCOELASTICITY - MIT
    Oct 24, 2001 · This document is intended to outline an important aspect of the mechanical response of polymers and polymer-matrix composites: the field of ...
  3. [3]
  4. [4]
    [PDF] 10 Viscoelasticity
    The generalized Maxwell model consists of N different Maxwell units in parallel, each unit with different parameter values. The absence of the isolated spring ...<|control11|><|separator|>
  5. [5]
    Maxwell Model - an overview | ScienceDirect Topics
    The Maxwell model is a type of simplest and basic mathematical model to describe the mechanics characteristic of viscoelastic solid material.
  6. [6]
    5.4: Linear Viscoelasticity - Engineering LibreTexts
    Mar 28, 2025 · This page presents an overview of linear viscoelasticity, detailing the mechanical response of polymers and composites, including molecular ...
  7. [7]
    [PDF] Rheology: Tools and Methods
    A simple model that captures the essential features of linear viscoelastic behavior is the Maxwell model, originally proposed by James Clerk Maxwell in 1867.
  8. [8]
    [PDF] Wormlike Micelles: Solutions, Gels, or Both?
    been associated with 19th-century scientists like James Clerk Maxwell, Lord. Kelvin, and Woldemar Voigt who developed the first models for viscoelastic. Soft ...
  9. [9]
    None
    ### Summary of Linear Viscoelasticity Content
  10. [10]
    IV. On the dynamical theory of gases - Journals
    Zakora D (2020) Model of the Maxwell Compressible Fluid, Journal of Mathematical Sciences, 10.1007/s10958-020-05030-6, 250:4, (593-610), Online publication ...
  11. [11]
    Linear Viscoelastic Materials - COMSOL Documentation
    Figure 3-1: Generalized Maxwell model. Hence, G is the stiffness of the main elastic branch, Gm represents the stiffness of the spring in branch ...
  12. [12]
    [PDF] Time domain analysis of viscoelastic models
    In the first section, the four viscoelastic models under study are presented. ... The generalised Maxwell model, initially introduced by Weichert [8], consists of ...Missing: Wiechert | Show results with:Wiechert
  13. [13]
    [PDF] γ τ γ γ γ γ η τ γμ τ
    Predictions of the Generalized Maxwell Model (GMM) and Generalized Linear-Viscoelastic Model (GLVE). © Faith A. Morrison, Michigan Tech U. Small-amplitude.
  14. [14]
    [PDF] Determining a Prony Series for a Viscoelastic Material From Time ...
    The measured data includes ramp loading, relaxation, and unloading stress-strain data. The resulting. Prony series, which captures strain rate loading and.
  15. [15]
    [PDF] Viscoelastic Properties of Polymers - Pearl HiFi
    Ferry, John D. Viscoelastic properties of polymers. I ncludes bibliographies ... The phenomenological theory of linear viscoelasticity is essentially complete.
  16. [16]
    [PDF] Step Strain Experiment STRESS RELAXATION
    viscoelastic liquid with the Generalized Maxwell Model. Page 13. Creep and Creep Recovery. SHEAR COMPLIANCE. Apply a constant stress σ for all times t > 0.
  17. [17]
    Linear Viscoelastic Materials - COMSOL Documentation
    The function Γ(t) is called the relaxation shear modulus function (or just relaxation function) and it can be found by measuring the stress evolution in time ...Missing: citation | Show results with:citation
  18. [18]
    [PDF] The generalized Maxwell model - DiVA portal
    Mar 25, 2024 · In Section 2 we present the governing equations of a dynamic viscoelastic problem based on the generalized Maxwell linear viscoelastic model, we ...
  19. [19]
    Identification of the extended standard linear solid material model by ...
    An extended Standard Linear Solid (SLS) material model is taken into account to model the material linear visco-elastic behavior.
  20. [20]
    A Generalized Maxwell Model for Creep Behavior of Artery Opening ...
    This means that the creep or relaxation function may be used to study the viscoelastic behavior of various soft tissues under general boundary conditions.
  21. [21]
    reclaimed rubber blends using Generalized Maxwell Model (GMM)
    The relaxation times corresponding to the first, second and third ranges are denoted by τ1, τ2, and τ3 in Table 6. The Table also demonstrates that τ1< τ2 < τ3.Missing: typical | Show results with:typical
  22. [22]
    [PDF] Foundations of viscoelasticity and application to soft tissues ... - arXiv
    Aug 18, 2021 · In Figure 2 we sketch the generalized Maxwell scheme, which is used to model the viscoelastic response of solid materials. The isolated spring k ...<|control11|><|separator|>
  23. [23]
    Parameters estimation of generalized Maxwell model for SBR and ...
    It is shown that the direct measurement method is 5%−7% more accurate than the indirect measurement method, at least for the SBR and SBR-CB.Missing: biomechanics | Show results with:biomechanics
  24. [24]
    Revisiting the identification of generalized Maxwell models from ...
    This paper discussed the identification of the material parameters of a generalized Maxwell model in linear viscoelasticity. The parameters of the model are ...Missing: textbook | Show results with:textbook
  25. [25]
    Generalized vs. fractional: a comparative analysis of Maxwell ...
    Sep 21, 2024 · They predict exponential behaviour for the relaxation modulus G(t) with discrete relaxation times, a situation that is typically well obeyed by ...Missing: citation | Show results with:citation
  26. [26]
    Generalized Maxwell Viscoelastic Models - PyLith - Read the Docs
    Aug 22, 2025 · The most general form of linear viscoelastic model is the generalized Maxwell model ... 0 is always zero and we only use a single Maxwell model.
  27. [27]
    [PDF] Rheology Bulletin - MIT
    Jan 1, 2007 · The Jeffreys model contains three- parameters (one spring and two dashpots), and at steady state shows a steady rate of creep as expected in a ...
  28. [28]
    Modified Jeffreys model and its application | Rheologica Acta
    Study on the constitutive equation with fractional derivative for the viscoelastic fluids – Modified Jeffreys model and its application. ORIGINAL ...
  29. [29]
    Numerical methods and analysis for simulating the flow of a ...
    The generalized Oldroyd-B fluid model (1), the time fractional diffusion-wave equation [48] ( 2 = 4 = 0), and the generalized Maxwell fluid model [49] ( 3 ...
  30. [30]
    A tutorial review of linear rheology for polymer chemists
    Feb 19, 2024 · We observe that relaxation time τ controls the rate of stress relaxation and is formally equal to the time at which G(t) falls to 1/e of its ...
  31. [31]
    (PDF) Viscoelasticity in inkjet printing - ResearchGate
    Aug 5, 2025 · The Maxwell viscoelastic model was then employed to simulate the inks' viscoelastic behavior for higher frequencies up to 10⁶ rad·s⁻¹ and to ...
  32. [32]
    Characterization and Modeling of the Viscoelastic Behavior ... - MDPI
    The generalized Maxwell model is a typical classic viscoelastic model composed of multiple Maxwell elements connected in parallel, as shown in Figure 1 [33].
  33. [33]
    Mechanical Models for the Representation of Visco-Elastic Properties
    Aug 6, 2025 · Generalized Maxwell and Kelvin rheological models are used for modeling the microscopic behavior of the viscoelastic phase (Roscoe 1950; Mandel ...
  34. [34]
    Comparative Analysis of Nonlinear Viscoelastic Models Across ...
    Apr 14, 2021 · Overall, the fractional Kelvin-Voigt model captures the same complexity as the generalized Maxwell model (with preset assumptions about the ...<|control11|><|separator|>
  35. [35]
    Improved Maxwell Model Approach and its Applicability toward ...
    Sep 2, 2021 · The current study presents a novel approach toward describing the long-term behavior of viscoelastic polymers based on the Maxwell model.
  36. [36]
    Kelvin–Voigt model of polymeric foams. - ResearchGate
    Foams, as a type of porous materials, have found broad functional and structural application in heat and sound insulation, the mitigation of mechanical ...
  37. [37]
    Modeling of thermo‐viscoelastic material behavior of glass over a ...
    Dec 17, 2019 · Finite element simulation of the slumping process of a glass plate using 3D generalized viscoelastic Maxwell model. J Non-Cryst Solids. 2014 ...