Next Article in Journal
Design and Investigation of Mechanical Properties of Additively Manufactured Novel Coil-Shaped Wave Springs
Previous Article in Journal
Deep Neural Network-Driven Analysis of Free Vibrations in Tapered Beams
Previous Article in Special Issue
State of the Art on Empirical and Numerical Methods for Cave Stability Analysis: Application in Al-Badia Lava Tube, Harrat Al-Shaam, Jordan
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Machine-Learning-Enabled Comparative Modelling of the Creep Behaviour of Unreinforced PBT and Short-Fibre Reinforced PBT Using Prony and Fractional Derivative Models

1
Chair of Solid Mechanics, Helmut-Schmidt-University/University of the Federal Armed Forces Hamburg, Holstenhofweg 85, 22043 Hamburg, Germany
2
Institute for Artificial Intelligence (HSU-AI), Helmut-Schmidt-University/University of the Federal Armed Forces Hamburg, Holstenhofweg 85, 22043 Hamburg, Germany
*
Author to whom correspondence should be addressed.
Appl. Mech. 2025, 6(3), 60; https://doi.org/10.3390/applmech6030060
Submission received: 2 June 2025 / Revised: 26 July 2025 / Accepted: 30 July 2025 / Published: 11 August 2025

Abstract

This study presents an approach based on data-driven methods for determining the parameters needed to model time-dependent material behaviour. The time-dependent behaviour of the thermoplastic polymer polybutylene terephthalate is investigated. The material was examined under two conditions, one with and one without the inclusion of reinforcing short fibres. Two modelling approaches are proposed to represent the time-dependent response. The first approach is the generalised Maxwell model formulated through the classical exponential Prony series, and the second approach is a model based on fractional calculus. In order to quantify the comparative capabilities of both models, experimental data from tensile creep tests on fibre-reinforced polybutylene terephthalate and unreinforced polybutylene terephthalate specimens are analysed. A central contribution of this work is the implementation of a machine-learning-ready parameter identification framework that enables the automated extraction of model parameters directly from time-series data. This framework enables the robust fitting of the Prony-based model, which requires multiple characteristic times and stiffness parameters, as well as the fractional model, which achieves high accuracy with significantly fewer parameters. The fractional model benefits from a novel neural solver for fractional differential equations, which not only reduces computational complexity but also permits the interpretation of the fractional order and stiffness coefficient in terms of physical creep resistance. The methodological framework is validated through a comparative assessment of predictive performance, parameter cheapness, and interpretability of each model, thereby providing a comprehensive understanding of their applicability to long-term material behaviour modelling in polymer-based composite materials.

1. Introduction

Short-fibre-reinforced composites (SFRCs) are widely used in various engineering fields like the aerospace and automotive industries. Their mechanical properties, including high stiffness and strength, together with their suitability for cheap, large-scale automated production recently increased the interest in SFRC [1,2,3,4,5,6]. SFRC materials consist of two or more different materials that are combined during a mould injection process [7,8]. However, due to mould injection, the fibres are spatially distributed within the component [8]. This results in heterogeneity of mechanical properties at the component level [9], which is derived from its heterogeneous microstructure [10,11]. The probabilistic characteristics of this microstructure are characterised by probability density functions like fibre length, fibre diameter, and fibre orientation [11,12].
While thermoplastic polybutylene terephthalate (PBT) serves as a matrix material, the short glass fibres reinforce the structure, leading to increased stiffness and strength. However, polymer-based composites like PBT GF30 (with 30% fibre mass fraction) exhibit time-dependent deformation under sustained loads, known as viscoelasticity or creep. Typically, creep is caused by the matrix material [13,14,15]. To predict the long-term performance of components, it is important to accurately characterise the creep behaviour, especially for the automotive, aerospace, and electronics industries [1,2,3,4,5].
Viscoelastic behaviour of SFRC is typically described using the generalised Maxwell model [16], which combines springs and dampers to represent elastic and viscous elements [16,17]. The Maxwell model is standardised and widely used in commercial Finite Element Method (FEM) software packages [18,19]. Its implementation in FEM relies on the Prony series [16,20,21], where the Prony series reflects the solution of an Ordinary Differential Equation (ODE), which results from the combination of springs and dampers. As these elements are typically considered to be linear, the ODE solution can be modelled as a sum of exponential functions. The effectiveness of this approach requires a large number of parameters to capture the full spectrum of material behaviour on varying time scales. This increases the complexity of parameter identification, as well as computational cost [20,21,22].
While Prony-based models approximate the non-local nature of creep effects by the superposition of several spring and damper elements, fractional derivative models can inherently model this behaviour. A fractional derivate model, also referred as a fractional element, describes dynamics that lie between a spring and a damper [16,23,24]. However, the practical application of fractional models remains challenging: normally, these models are solved indirectly by transforming them into the Fourier domain to fit experimental data. When established, this approach requires more complex data collection setups as the material behaviour has to be measured across different frequencies in the frequency domain. This makes simple tensile tests insufficient for data collection. Furthermore, Fourier analysis is limited to linear models, which complicates the representation of non-linear behaviour [25,26].
Recent advances in machine learning (ML) have enabled us to learn fractional dynamics directly from data. Coelho et al. [27] introduced neural fractional differential equations (NFDEs). NFDEs combine a solver for Fractional Dynamical Equations (FDEs) with neural nets (NNs) to learn the underlying FDE from arbitrary data. Building on this foundation, Zimmering et al. [28] improved the solver in accuracy and resource efficiency, enabling the use of the solver in real-world scenarios.
In this study, the need for special experiments that capture the frequency spectrum of SFRCs is overcome by fitting a parametric fractional element model directly to experimental tensile test data. The optimised NFDE solver of Zimmering et al. is employed to perform this fitting. The solver’s general capability to fit arbitrary dynamical models (including NNs) additionally offers flexibility in capturing non-linear creep behaviour from arbitrary data in future studies.
Following this, the contributions of this paper are as follows.
  • First, the accuracy of the fractional element in describing the creep behaviour of PBT GF0 and PBT GF30 is compared with the generalised Maxwell model. This comparison aims to determine the extent to which fractional derivative models can effectively capture the viscoelastic response of such materials over time.
  • Second, practical methodologies are studied to efficiently learn the parameters of fractional element models. It is sought to identify computational techniques that enable a stable parameter identification process while maintaining accuracy.
  • Third, the advantages that fractional derivative models may offer over the Prony models are discussed for SFRCs. It is evaluated if fractional models provide improved predictive capabilities and enhance the representation and interpretability of material properties over extended time scales.
This paper is structured as follows: Section 2 outlines the theoretical framework for modelling creep behaviour as well as the data fitting methods for the Prony fractional model. Section 3 details the experimental setup for the data acquisition methods. In Section 4, the results are presented. Here, it is focused on the determination of Prony coefficients and coefficients derived from fractional models. Finally, in Section 5, the comparative performance of the models is discussed, and conclusions about their practical applications in material characterisation are drawn.
The developed software is based on Python 3.11 and open-source software libraries and is available along with the experimental data at https://github.com/zimmer-ing/PBTGF-Creep. This allows others to reproduce the results and apply the presented methods in their own work.

2. Materials and Methods

2.1. Theoretical Framework for Creep Behaviour

Creep is a time-dependent deformation process that occurs under constant load and temperature, involving both elastic and viscous contributions. This phenomenon is particularly relevant in polymer-based materials, which exhibit viscoelastic behaviour, which combines time-independent elasticity and time-dependent viscous flow.

2.1.1. Viscoelastic Effect

Creep occurs when a material undergoes time-dependent deformation under constant load and temperature. This effect occurs in polymer-based materials, which have both elastic and viscous behaviours, commonly referred to as viscoelastic behaviours [16,21,29,30]. In general, creep deformation typically occurs in three stages, primary, secondary, and tertiary creep.
The total strain ε total ( t ) over time of a sample can be expressed additively as
ε ( t ) = ε total ( t ) = ε linear   elastic + ε creep ( t ) .
Here, ε linear   elastic represents the immediate strain response upon the application of stress, following Hooke’s law for linear elasticity. This component remains constant over time and does not contribute to the time-dependent deformation behaviour of the material. However, the creep deformation ε creep ( t ) accounts for the viscoelastic deformation. In this work, the focus is on the primary and secondary stages because capturing the tertiary stage is time-consuming and experimentally challenging.
The mathematical models are derived using rheological elements that describe the stress–strain relations for the linear viscoelastic behaviour [16,31]. These elements are shown in Figure 1, where the spring element represents Hooke’s law with stiffness K e q . A dashpot represents a purely viscous material according to Newton’s law, characterised by dynamic viscosity η . The fractional element is described by the parameters C and α .
Generally, there are two equivalent approaches to characterise the mathematical relationships between stress and deformation in linear viscoelastic materials [16,21,29,30,32]. One approach employs integral formulations to establish these relationships, while the other uses first-order ODE to connect stresses and strains. The integral formulation is where Boltzmann first articulated the superposition principle, commonly referred to as the Boltzmann superposition principle [16,33]. This integral formulation for linear viscoelastic materials describes the response to strain as a function of past stresses and the stiffness of the material. It is written as [16]
ε ( t ) = 0 t K ( t τ ) d σ ( τ ) ,
where K ( t ) represents the creep stiffness. If the roles of stress and strain are reversed, Equation (2) can be expressed as
σ ( t ) = 0 t J ( t τ ) d ε ( τ ) ,
where J ( t ) denotes the creep compliance. The material function K ( t ) depends on the material and can be obtained experimentally. In the context of experimentally measured creep data, the creep stiffness K ( t ) can be determined as follows:
K ( t ) = σ ε ( t ) ,
where σ is the sustained load and ε ( t ) gives the measured time-dependent creep elongation.

2.1.2. Generalised Maxwell Model and Prony-Series

The basic Maxwell model, shown in Figure 2a, consists of a spring and a dashpot connected in series. This approach can represent simple viscoelastic behaviour, but it cannot fully capture the complex time-dependent response of materials. To address this limitation, multiple Maxwell models are arranged in parallel with an additional spring K e q , as shown in Figure 2b. This extended configuration is known as the generalised Maxwell model [21,34,35]. A first-order ODE governs each Maxwell element [21,35].
The total stress σ is modelled as the sum of the contributions of the Maxwell elements N, each representing a viscoelastic branch of the material. The constitutive behaviour of the i-th Maxwell element is governed by a first-order differential equation that relates stress and strain through elastic and viscous parameters. The governing equation for the i-th branch is given by [34,36]
d σ i ( t ) d t + 1 τ i σ i ( t ) = K i d ε ( t ) d t ,
where σ i ( t ) R represents the stress contribution of the i-th Maxwell element, K i R + is the elastic modulus of the i-th spring, and τ i R + is the relaxation time of the i-th Maxwell element, defined as τ i = η i K i , with η i R + being the viscosity of the dashpot.
For the generalised Maxwell model, as detailed in [16,21,29,35,36,37], the stress–strain relationship for viscoelastic materials in the form of a differential equation is formulated as follows:
d σ ( t ) d t + K i η i σ ( t ) = ( K e q + K i ) d ε ( t ) d t + K e q K i η i ε ( t ) .
This differential Equation (6) effectively models the viscoelastic response of complex materials, encompassing both stress relaxation and creep behaviour [36].
A widely used approach to describe the time-dependent viscoelastic response is the Prony series, which can be derived by applying the Laplace transform to the differential formulation of the generalised Maxwell model, such as Equation (6) [16]. The Prony series effectively describes the time-dependent deformation behaviour of materials, which is written as [16]
K ( t ) = K e q + i = 1 N i K i e t / τ i .
The relaxation function K ( t ) characterises the time-dependent decrease in material stiffness due to viscoelastic effects. Each exponential term corresponds to a Maxwell element with a distinct relaxation time τ i , capturing different rates of stiffness relaxation [31,34]. In the following sections, Equation (7) at t = 0 is referred as K t o t a l .
In practical applications, the Prony series parameters K e q , K i τ i are determined by fitting the Prony series to experimental data. This fitting process typically involves optimisation techniques to minimise the deviation between model predictions and observations, yielding a viscoelastic model tailored to the material under investigation.

2.1.3. From the Generalised Maxwell Model to the Fractional Viscoelastic Element

The generalised Maxwell model is limited in its ability to model continuous memory effects. To overcome this limitation, fractional derivatives are introduced, leading to the definition of the fractional viscoelastic element.
The governing equation for the fractional viscoelastic element is obtained by replacing the integer-order derivative in the Maxwell model with a fractional derivative:
D t 0 C D t α ε ( t ) = σ ( t ) C , t R + , σ ( 0 ) = σ 0 , ε ( 0 ) = 0 ,
where D t 0 C D t α denotes the Caputo fractional derivative of order α ( 0 , 1 ) , C R + is the elastic modulus, σ ( t ) R is the stress, and ε ( t ) R is the strain in the material at time t. The initial condition ε ( 0 ) = 0 represents the initial strain.
The Caputo fractional derivative is defined as
D t 0 C D t α f ( t ) = 1 Γ ( 1 α ) 0 t ( t τ ) α d f ( τ ) d τ d τ , 0 < α < 1 ,
where Γ ( · ) is the gamma function, defined as Γ ( n ) = ( n 1 ) ! for n N . For α = 1 , the Caputo derivative reduces to the classical first-order derivative, recovering the classical Maxwell model. For α = 0 , the fractional derivative acts as the identity operator, describing purely elastic behaviour (i.e., a spring).
Applying a unit stress step to Equation (8) yields the creep–compliance [16] (Equations (6.2)–(7), p. 340)
J ( t ) = t α C Γ ( 1 + α ) .
Here, C is considered as fractional (pseudo-) modulus in the unit Pa s α . Hence C controls the amplitude of the creep curve, where a larger C implies a smaller strain under a given load. In the limits α 0 and α 1 , C reduces to the instantaneous elastic modulus and to the Newtonian viscosity, respectively [16].
The fractional differential equation can be expressed as
D t 0 C D t α ε ( t ) = f θ ( ε ( t ) ) , α ( 0 , 1 ) ,
where f θ represents the model of the system dynamics, parametrised by θ (e.g., C). f θ is not restricted to linear functions as in the case of the fractional element Equation (8).
To solve the fractional differential equation, one formulates the corresponding Initial Value Problem (IVP)
ε ( t ) = ε ( 0 ) + 1 Γ ( α ) 0 t ( t τ ) α 1 f θ ( ε ( τ ) ) d τ .
This IVP can typically only be solved numerically by fractional differential equation solvers [38]. Implementations in modern computational frameworks [28] are employed to solve this equation, allowing the estimation of model parameters like α , C. This approach enables the fractional viscoelastic model to capture the complex, continuous memory nature of creep processes in viscoelastic materials.

2.2. Parameter Fitting Approach

As described in the previous sections, the experimental tensile data is used in tests to learn parameters for two models capturing the long-term deformation of SFRCs:
  • Model 1: A Maxwell model where the model’s solution is represented by a Prony series. Here, a classical approach with numerically approximated gradients to fit the Prony series to the data is used.
  • Model 2: A fractional derivative-based model where the NFDE solver of Zimmering et al. [28] is used. This solver supports exact gradient computation and can thus be used within an optimisation procedure to fit the model’s parameters to the data.
First, the general procedure for parameter fitting and explaining the underlying optimisation concepts used in this study is reviewed. This is followed by the specific approaches for the Prony series and fractional model. Lastly, the differences between the approaches used for the Prony series and fractional derivative models are highlighted.

2.2.1. General Procedure and Concept of Optimisation

Parameter fitting in viscoelastic modelling is an optimisation problem where the objective is to identify a set of model parameters that minimises the differences between experimental measurements and predictions generated by the model. Let
y exp = { y exp ( t 1 ) , y exp ( t 2 ) , , y exp ( t n ) }
represent the experimentally measured strain values at discrete time points { t 1 , t 2 , , t n } . Furthermore, let M be a model that is parametrised by a parameter vector θ and outputs predictions
y model ( θ ) = { y model ( t 1 ; θ ) , y model ( t 2 ; θ ) , , y model ( t n ; θ ) } .
During the fitting process of M , an objective (i.e., the loss) function L is minimised. Often, L is defined as the sum of squared errors
L ( θ ) = i = 1 n y exp ( t i ) y model ( t i ; θ ) 2 .
Gradients of the parameters θ , which provide information about the direction and magnitude of the steepest decrease in L ( θ ) , guide the optimisation. The gradient vector is defined as
θ L ( θ ) = L θ 1 , L θ 2 , , L θ k ,
where θ = { θ 1 , θ 2 , , θ k } represents the model parameters.
In the simplest case, the parameters of the models are updated according to
θ ( k + 1 ) = θ ( k ) η θ L ( θ ) ,
where η > 0 is the learning rate, controlling the step size. The choice of optimisation strategy depends on the model complexity, as well as the number of parameters to be optimised.
Gradient Computation Approaches. In this work, two complementary strategies for obtaining gradient information are employed:
  • Numerical Gradient Approximation: Gradients are estimated, using finite-difference methods, as described by Nocedal and Wright [39] (Chapter 8.1): a partial derivative with respect to the i-th parameter can be approximated by perturbing only that parameter by a small step size ϵ , i.e.,
    f x i ( x ) f x 1 , , x i + ϵ , , x n f ( x 1 , , x i , , x n ) ϵ .
    This procedure is used in many standard libraries (e.g., scipy.optimize in Python). It is efficient when the number of parameters is moderate or when closed-form expressions for the equations are available.
  • Exact Gradient Computation: To compute gradients, automatic differentiation (AD) is used. AD is implemented in ML frameworks such as PyTorch [40]. It generalises backpropagation by systematically applying the chain rule [41]. Compared to other methods, it enables highly accurate gradients for complex or high-dimensional models (e.g., fractional differential solvers and/or neural networks). A good introduction into AD, also discussing its distinction from alternative approaches, can be found in Baydin et al. [42].
While both methods are able to be used in gradient-based optimisation, numerical approximations are better suited for simpler models, where the model equations are evaluated in closed form (e.g., Model 1: Prony Series). For Model 2, the fractional derivative model, AD, is better suited as a numerical solver and is therefore integrated into the gradient computation. For such a scenario, AD enables more robust and accurate gradient estimation.

2.2.2. Parameter Fitting for Model 1 (Prony Series)

As introduced in Section 2.1.1, the Prony series expresses creep stiffness as a sum of exponential terms. Here, each term is characterised by a stiffness contribution K i , a relaxation time τ i , and an equilibrium stiffness K eq . To obtain the experimentally measured stiffness K exp ( t j ) ,
K exp ( t j ) = σ ε ( t j ) , j = 1 , , M
is applied. Next, an initial guess for the parameter vector K eq , { K i } , { τ i } is defined. Let
Δ K exp = max ( K exp ) min ( K exp ) ,
and set
K eq = min ( K exp ) .
δ K exp is distributed among n exponential terms using descending weights, ensuring that larger K i values correspond to earlier terms
K i = Δ K exp w i , with w i ( n i + 1 ) and i = 1 n w i = 1 .
Each relaxation time τ i is initialised by a simple heuristic such as
τ i = 10 n 1 / ( i + 1 ) ,
which spreads relaxation timescales over several orders of magnitude and helps avoid unphysical parameter values.
The model is then fitted by minimising the weighted least-squares loss function
L = j = 1 M w ( t j ) K exp ( t j ) K model ( t j ; K eq , { K i , τ i } ) 2 , where w ( t ) = α e t / τ w + β .
Here, α = 10.0 is used to assign a high initial weight, τ w = 15 to control the decay rate, and β = 1.0 as the baseline weight for later times. This weighting shifts attention to the early time data, as the initial creep behaviour is of high importance.
To ensure physically meaningful parameter values, simple box constraints on all parameters
0 K eq , K i 10 · max ( K exp ) , 0 τ i 1000 s
are employed. The bounded variant of the Limited-memory Broyden–Fletcher–Goldfarb–Shanno method (L-BFGS-B) [39,43] is used to minimise the loss function shown above. For problems of moderate dimensions, it takes advantage of its low memory requirements and efficient convergence properties. By combining a well-chosen initial guess with a direct, closed-form model and physically reasonable parameter constraints, this approach yields a plausible and fast approximation of the underlying parameters.

2.2.3. Parameter Fitting for Model 2 (Fractional Derivative Model)

As introduced in Section 2.1.3, fractional derivatives generalise integer-order derivatives and thus capture long-term memory effects. The IVP is solved corresponding to the FDE given in Equation (8) numerically using the open-source library FDEint [28]. This PyTorch-based [40] solver supports AD and can thus be used to optimise α and C.
Here, we employ the same weighting function as used for the first model
w ( t ) = α e t / τ w + β ,
with α = 10.0 , τ w = 15 , and β = 1.0 . In distinction to the first model, a two-stage optimisation process is used.
First, a global optimisation over all samples is performed simultaneously. Next, the weighted least-squares loss
L global = i = 1 M w ( t i ) ε exp ( t i ) ε model ( t i ; α , C ) 2 ,
is minimised. Here, M is the total number of data points, ε exp ( t i ) is the experimentally measured strain, and ε model ( t i ; α , C ) is the model prediction. L-BFGS-B is used to find well-scaled initial estimates of α and C.
In a second step, the model is adopted for the individual sample characteristics. α and C are fine-tuned with sample-specific optimisation by minimising
L sample = k = 1 N w ( t k ) ε exp ( j ) ( t k ) ε model ( j ) ( t k ; α , C ) 2
for each sample j, where N is the number of data points in that sample. In this stage, the Adam optimiser [44] is used. It needs fewer model evaluations as L-BFGS-B and is more robust to local parameter variations. Each sample-specific fit is initialised with the globally optimised parameters, which ensures stable convergence. This combination of global and sample-specific optimisation balances general behaviour across all datasets with the local adaptability needed for each individual sample.
To enforce physical plausibility during optimisation, C and α are reparametrised using transformations that constrain their values to be positive. Instead of optimising over C, it is optimised over log ( C ) as well as over the inverse sigmoid of α instead of α directly. It is defined as
α raw = log α 1 α ,
such that during training,
C = exp ( log ( C ) ) , α = σ ( α raw ) ,
where σ ( x ) = 1 1 + e x denotes the sigmoid function. This ensures C > 0 and 0 < α < 1 at all times.

2.2.4. Comparing the Two Approaches

The Prony series approach uses a closed-form expression for creep stiffness. It allows gradients to be computed via numerical approximations and optimised by the L-BFGS-B optimiser. While this work is implemented in Python, this approach is also available in Matlab. The optimiser is computationally efficient for a parametric model and a standard implementation in many tools. These properties make it well suited for engineering applications.
As the fractional derivative model requires solving a fractional differential equation, AD has to be used for gradient computation. To ensure this, the FDEint solver is used in a PyTorch-based (version number 2.5.1) optimisation setup. AD computes exact gradients through the solver, enabling stable gradients and efficient optimisation but at a higher computational cost. Compared to Model 1, which is already the solution of an ODE, this approach is not limited in model complexity. Hence, the fractional model provides a powerful mechanism for capturing long-term memory effects but demands more extensive computational resources and a more involved implementation strategy.

2.3. Trend Analysis Using the Master Curve and Its Standard Deviation

To analyse the trend of the creep behaviour of the samples, a master curve is calculated from the creep curves. For each material, all M = 20 samples are averaged along the time axis. The master curve allows us to check general trends and analyse the shape of the creep curves as noise is reduced. As all time values t i j of the measured curves are consistent within the data set, the average at each discrete time step t i j = t i is computed. Formally, the master curve is written as
y ¯ i = 1 M j = 1 M y i j ,
where y i j exp = y i j gives the experimental values at each discrete time t i for the j-th measurement. Averaging these values across all measurements j = { 1 , 2 , , M } for each t i , the mean value y ¯ i is obtained.
Additionally, the standard deviation σ i for each t i is analysed. The standard deviation is defined as
σ i = 1 M j = 1 M ( y i j y ¯ i ) 2 .
Here σ i measures the deviation of individual data points y i j from the mean value y ¯ i . This metric quantifies the experimental variability.
These two additional metrics help to characterise the general material behaviour as well as the experimental quality.

3. Experimental Setup

This section details the experimental investigation conducted to characterise the creep behaviour of unreinforced composites and SFRCs under controlled sustained loading. The specimen preparation and the experimental setup are also presented here.

3.1. Materials and Specimen Preparation

In this work, standardised tensile test specimens according to DIN EN ISO 527-2 type 1A [45] were used. Figure 3 illustrates the schematic configuration of these specimens, highlighting the measurement of creep behaviour within a gauge length l 0 of 75 m m . The matrix material is a thermoplastic PBT, representing the primary component, marketed under the trade name ULTRADUR B 4520, supplied by BASF. The reinforcing component of the composite is E-Glass, designated as FGCS 3540 and provided by STW.
All specimens were produced using an injection moulding process at Kunststoff–Zentrum–Leipzig gGmbH on an Arburg Allrounder 370E machine (Arburg GmbH & Co. KG, Lossburg, Germany). Key parameters from the injection moulding process for both PBT and PBT GF30 are summarised in Table 1. These values are taken directly from the production protocols and ensure consistent processing history across specimens. Two types of material were tested, each with 20 specimens, as detailed below.
  • Specimen without fibres (PBT);
  • Specimen with fibre mass fraction of 30% (PBT GF30);
In the case of the PBT GF30 samples, the fibres were introduced during a compounding process performed before the injection moulding process [46,47]. During compounding, matrix material and fibres were combined and processed into uniform pellets containing a 30% fibre mass fraction. Subsequently, these compounded pellets were used in the injection moulding process to produce the final tensile test specimens with fibres.

3.2. Creep Testing and Measurement Setup

Tensile creep experiments were conducted according to the DIN EN ISO 899-1 [48] standard, which ensures consistency and reliability in creep testing. As specified in the standard, the full force must be applied within 1 s to 5 s to ensure consistency in initial loading conditions. For PBT samples, a force of 300 N was applied, while PBT GF30 specimens, with higher stiffness, required a force of 1025 N . This adjustment ensures comparable elongation between materials, facilitating meaningful comparisons. For both materials, the force was fully applied within 3 s . The creep behaviour test was measured over 300 s at room temperature under controlled conditions (see Table 2).
The creep behaviour of the specimens was measured using the ARAMIS 3D digital image correlation (DIC) system (Carl Zeiss GOM Metrology GmbH, Braunschweig, Germany). The experimental setup is depicted in Figure 4a. This optical measuring system is specifically designed for a non-contact, full-field deformation analysis. The setup used two high-resolution 12 MP cameras positioned with 700 m m in the specimen. The force was applied and controlled using the ZWICK/ROELL Z050 tensile test machine (ZwickRoell GmbH & Co. KG, Ulm, Germany).
The ARAMIS system uses a non-contact optical technique to analyse the displacement of a stochastic speckle pattern applied to the surface of the sample [49,50]. As illustrated in Figure 4b, the distribution of the strain ( ε y ) on the surface of the PBT specimen under a sustained load of 300   N is visualised. The strain distribution is calculated using correlation algorithms that track changes in the position of the speckle features between the reference state (unloaded) and the deformed state (loaded) [50,51].

4. Results

This section presents the results of the experimental investigation of creep behaviour for unreinforced (PBT) and fibre-reinforced (PBT GF30) specimens under sustained loading. The experimental setup and conditions are summarised in Table 2. Additionally, the predictive performance of two modelling approaches, the Prony series and the fractional derivative method, is evaluated to approximate the experimental data.

4.1. Experimental Results of Creep Behaviour

The creep behaviour of PBT and PBT GF30 under sustained load conditions was analysed using data from 20 samples per material type. The results shown in Figure 5 provide a comparative analysis of the time-dependent deformation characteristics of these two materials. In all samples tested, the initial deformation at the beginning of the creep lies in a comparable range of around 0.26 %. This indicates a consistent elastic response prior to viscoelastic deformation. The creep responses of primary and secondary creep stages, as discussed in Section 2.1.1, are observed in both materials and exhibit characteristics. However, the second stage cannot be clearly separated from the tertiary stage of creep. Clear indicators such as an accelerating strain rate or failure are missing. However, a detailed classification of creep stages or associated mechanisms lies beyond the scope of this study and is therefore not pursued further.
The master curves in Figure 6 of the total strain ε total and Figure 7, where the creep strain ε creep is shown, indicate that PBT exhibits a slightly higher strain rate compared to PBT GF30. These master curves, which are introduced in detail in Section 2.3, provide more detail on the general trend of deformation compared to the results in Figure 5. These findings align with previous studies [52,53,54], which reported similar trends in the viscoelastic behaviour of unreinforced and fibre-reinforced materials. The viscoelastic behaviour is governed primarily by the matrix material rather than the reinforcing fibres [53,55,56].
In particular, the master curves in Figure 7 show a slightly steeper creep rate for PBT compared to PBT GF30. This suggests a reduced resistance to time-dependent deformation in the absence of fibre reinforcement, specific to the PBT system studied here. In contrast, the inclusion of short glass fibres in PBT GF30 results in lower creep rates. These findings emphasise the role of fibre reinforcement in altering the viscoelastic response of PBT under constant load conditions.

4.2. Prony Series Coefficients for PBT and PBT GF30

The viscoelastic properties of the PBT and PBT GF30 materials are characterised by fitting their experimental creep data to the Prony series model using a different number of branches N i = { 1 , 2 , 3 } with respect to the Equation (7). The coefficients K e q , K i , and τ i were determined using the fitting procedure detailed in Section 2.2. The resulting coefficients for each specimen are summarised in Table A4, Table A5, and Table A6, respectively.
The mean values and standard deviations of the Prony series coefficients for both materials are presented in Table 3. The total stiffness K total derived from the Prony series representations in these tables corresponds to the effective stiffness experimentally determined in accordance with the DIN EN ISO 527-1 standard [57]. This total stiffness characterises the initial elastic response of the materials and serves as a reference for calibrating Prony series models with varying numbers of branches, specifically N = 1 , N = 2 , and N = 3 . The corresponding Young’s moduli, determined in accordance with the DIN EN ISO 527-1 standard, were obtained from tensile tests on 10 samples per material, yielding mean values of 2674.50   M Pa for PBT and 9523.12   M Pa for PBT GF30.
Significant differences in the mean Prony series coefficients between PBT and PBT GF30 underscore the role of fibre reinforcement in the modification of viscoelastic properties. Although the matrix material primarily governs time-dependent deformation, the presence of fibres considerably enhances the overall stiffness and resistance to creep. As noted above, the mean relaxation time τ 1 and the stiffness distribution K 1 (see Table 3) for PBT GF30 are higher compared to PBT, further supporting the non-parallel nature of the master curves shown in Figure 7. These observations are consistent with findings reported in the literature [52,53,54], which similarly demonstrated the influence of fibre reinforcement on the viscoelastic behaviour of the polymer.
A comparative evaluation of experimental creep data with two predictive modelling approaches, the classical Prony series model and the fractional viscoelastic model, is presented in Figure 8 for PBT and in Figure 9 for PBT GF30. These diagrams highlight the ability of each model to reproduce the experimentally observed viscoelastic behaviour over the measured time domain.
For both materials, it is evident that the Prony model with a single branch ( N = 1 ) does not capture experimental trends with sufficient accuracy. This is particularly apparent in the early and intermediate time regimes, where deviations from the measured data are most pronounced. In contrast, the fractional viscoelastic model provides a significantly improved fit across the full time range, especially for low-order Prony approximations. This can be attributed to the inherent capacity of fractional models to represent broad relaxation spectra with fewer parameters.
However, as the number of branches in the Prony series increases, the model’s ability to replicate the experimental data improves notably. From N = 2 onward, Prony-based predictions show a better agreement with the measured curves compared to the accuracy of the fractional model. This trend is consistently observed for both material systems, demonstrating the flexibility of the Prony approach when extended with sufficient model complexity. Nevertheless, the trade-off lies in the increased number of fitting parameters and the potential risk of over-fitting or loss of physical interpretability.
In summary, while the fractional model yields an excellent initial approximation of the experimental data with a compact formulation, the Prony series model achieves superior accuracy when configured with multiple branches. This highlights the importance of selecting a modelling strategy based on the required fidelity, parameter efficiency, and application context.

4.3. Fractional Model Coefficients for PBT and PBT GF30

The results for the fractional element in Equation (8) and fitted according to the methodology in Section 2.2.3 can be found in Table 4.
Table 4. Mean values and standard deviations (Mean ± StdDev) of Fractional Element parameters.
Table 4. Mean values and standard deviations (Mean ± StdDev) of Fractional Element parameters.
ParameterPBTGF0PBTGF30
C5.34 × 105 ± 1.63 × 1053.55 × 106 ± 9.60 × 105
α 0.35 ± 0.050.40 ± 0.04
Here, C describes the immediate response of the corresponding material, similar to K t o t a l in the previous section. While K t o t a l rises by 3.5 times for PBT GF30, C rises by 6.65 times, which is significantly higher.
The fractional derivate α reflects the mix of elastic ( α = 0 ) and viscous ( α = 1 ) behaviour. Its mean values are slightly higher for PBT GF30, but in a similar range. As the corresponding values overlap, due to high standard deviation, it cannot be clearly stated that PBT GF30 shows more viscous behaviour than PBT. The slight increase could be explained by additional micro-slip or friction of the fibres inside the matrix material. Since clear evidence is missing, either more samples or an improved experimental setting is necessary to study the effect in detail.
In summary, the results show that the immediate response C significantly increases for the fibre0enhanced material while the ratio between elastic and viscous behaviour remains similar.

5. Discussion

The measured creep behaviour of fibre-reinforced polymers (PBT GF30) and unreinforced polymers (PBT) is in line with expectations. It confirms that fibre reinforcement plays a role in modulating viscoelastic responses. The PBT-GF30 samples deform less under sustained load than the unreinforced PBT samples, an expected effect, as shown in Figure 7, as the glass fibres stiffen the matrix and inhibit creep strain [52,53,54]. The master curves shown in Figure 6 indicate that while both the fractional derivative model and the Prony series provide reasonable approximations of the creep response, there are notable differences in terms of accuracy and parameter stability.
For single-term Prony models ( N = 1 ), the approximation of the creep behaviour is less accurate than the fractional model. The fractional model provides a smoother representation of the viscoelastic response. When additional terms are added to the Prony series ( N 2 ), the accuracy improves. This increased number of parameters increases potential over-fitting. The higher standard deviations of the Prony coefficients suggest more variability, which also indicates that the Prony coefficients lack robustness. This variability challenges the physical interpretability of the Prony model, as over-fitting becomes a concern with higher N. In contrast, the fractional derivative model demonstrates better approximation accuracy with a lower number of required parameters. The fractional order parameter ( α ) shows stability in different material compositions. This underscores the suitability of fractional calculus-based models to provide a compact and accurate representation of creep behaviour. However, similarly to the Prony series model, greater standard deviations of the parameter C are observed.
Determining the parameters of fractional derivative models presents a challenge due to the complexity of fractional differential equations. This study explores optimisation techniques such as numerical and automatic differentiation-based frameworks to efficiently fit experimental data. While the Prony series benefits from a closed-form solution that simplifies parameter fitting, the fractional model requires solving fractional differential equations numerically, increasing computational demands. To mitigate these challenges, gradient-based optimisation methods are employed, ensuring stable parameter convergence across multiple datasets. Moreover, the application of data-driven techniques, such as NFDE solvers, offers a promising alternative for parameter identification.
Recent developments in the application of artificial intelligence to constitutive modelling have demonstrated significant potential, particularly in areas requiring robust parameter identification from noisy or incomplete data. The novelty of this study lies not in the proposal of new constitutive equations but in the implementation of a machine learning-inspired approach, specifically a neural fractional differential equation (NFDE) solver, to identify the governing parameters of both the Prony and fractional derivative models. This technique combines automatic differentiation with gradient-based optimisation, treating the governing differential equation as an operator constraint. Although the application of AI tools to curve fitting or surrogate modelling in viscoelasticity is a well-established area of research, as evidenced by the work of [58,59], the use of a solver to extract interpretable and physically meaningful parameters from time domain data for fractional models represents a novel contribution. The approach demonstrated stable convergence and interpretability across the two material systems studied, highlighting the emerging role of data-driven methods in enhancing classical mechanical modelling.
One of the key challenges in viscoelastic modelling is to determine the model parameters efficiently. The Prony series requires multiple relaxation times and corresponding stiffness coefficients, increasing the complexity of parameter identification. As the number of branches in the Prony model increases, over-fitting becomes a concern, leading to large standard deviations in the model coefficients. In contrast, the fractional derivative model requires fewer parameters, making it computationally more efficient.
The results suggest that existing numerical optimisation techniques, such as gradient-based or evolutionary algorithms, could be refined to improve the robustness of parameter identification for both models. The need for an improved fitting procedure is evident, as the high variability in Prony coefficients indicates the sensitivity of experimental noise data.
Although the test duration was limited to 300 s, the fractional viscoelastic model used allows reliable extrapolation because of its inherent power-law behaviour, as demonstrated in recent studies [60,61]. Compared to the Prony series, it offers better long-term predictions from short-term data.
The findings of this study confirm that both the fractional derivative model and the generalised Maxwell model based on the Prony series are viable approaches to modelling the creep behaviour of fibre-reinforced and unreinforced materials. However, the results highlight a critical issue in parameter estimation, particularly for the Prony series model, which exhibits a high standard deviation in parameter values due to over-fitting. The fractional derivative model emerges as a promising alternative, providing a more compact reparametrisation while maintaining accuracy in long-term creep predictions.
It is important to emphasise that the conclusions drawn in this study are specific to the thermoplastic polyester PBT and its short-fibre-reinforced variant PBT GF30. While the viscoelastic modelling strategies developed here may generally be applicable, thermoplastic systems vary greatly in molecular structure, crystallinity, and processing sensitivity. Therefore, caution should be exercised when extrapolating these findings beyond the tested materials.
This emphasises the need for a better or more robust fitting procedure that ensures accuracy while preserving physical interpretability, mitigates over-fitting, and enhances the predictive capability of viscoelastic models. Future research should focus on refining optimisation algorithms, improving the consistency of experimental data, and exploring hybrid modelling approaches that integrate the advantages of both methodologies.

6. Conclusions

The present study investigated the tensile creep behaviour of unreinforced PBT and short-fibre-reinforced PBT GF30 (with a fibre mass fraction of 30%) using two viscoelastic modelling approaches, the classical Prony series and a fractional derivative formulation. A comparative analysis was conducted, which revealed that the fractional derivative model outperforms the Prony series in terms of parameter stability and compactness, especially at low model orders. While the Prony model becomes more accurate with additional terms, it also becomes more susceptible to overfitting and parameter variability. In contrast, the fractional model provides accurate results with fewer parameters and demonstrates increased robustness in predicting larger time ranges.
A central contribution of this work lies in the application of a neural fractional differential equation (NFDE) solver for parameter identification. This approach enables the data-driven extraction of physically interpretable parameters from creep data in the time domain and demonstrates stable convergence, offering a viable alternative to conventional curve-fitting techniques.
Despite the limited experimental duration (300 s), the fractional model enables credible extrapolation due to its inherent power-law behaviour. The methods and conclusions presented are specific to PBT and PBT GF30 and should be cautiously extended to other polymer systems due to variations in morphology and processing sensitivity.
Future work should focus on refining optimisation algorithms for parameter identification, improving data robustness, and exploring hybrid modelling strategies that combine the strengths of fractional and classical viscoelastic representations.

Author Contributions

Conceptualisation, E.K., B.Z., O.N. and N.R.; methodology, E.K. and B.Z.; software, B.Z. and E.K.; validation, E.K. and B.Z.; formal analysis, B.Z. and E.K.; investigation, E.K.; resources, N.R. and O.N.; data curation, B.Z. and E.K.; writing—original draft preparation, E.K. and B.Z.; writing—review and editing, E.K., B.Z., O.N. and N.R.; visualisation, E.K. and B.Z.; supervision, N.R. and O.N.; project administration, N.R. and O.N.; funding acquisition, N.R. and O.N.; All authors have read and agreed to the published version of the manuscript.

Funding

This publication has been funded by the Open Access Publication Fund of the Helmut Schmidt University/University of the Federal Armed Forces Hamburg.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All code and data necessary to reproduce the results of this study are available at the following GitHub repository: https://github.com/zimmer-ing/PBTGF-Creep. License under which the data set is made available (CC BY-NC-SA 4.0).

Acknowledgments

During the preparation of this manuscript, the authors used ChatGPT-4o for the purpose of improving English language, grammar, and style. The authors take full responsibility for the content of this publication.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
ADAutomatic Differentiation
L-BFGS-BLimited Memory Broyden–Fletcher–Goldfarb–Shanno Method
DICDigital image correlation
FEMFinite Element Method
IVPInitial Value Problem
NFDENeural Fractional Differential Equation
ODEOrdinary Differential Equation
PBTPolybutylene terephthalate (thermoplastic material)
PBTPolybutylene terephthalate without fibre content
PBT GF30Polybutylene terephthalate with 30% fibre mass fraction
SFRCShort-fibre-reinforced composite

Appendix A. Time-Dependent Loss Weighting

The weighting functions introduced in Section 2.2 applied during the parameter identification of the Prony series and the fractional viscoelastic models are illustrated in Figure A1 and Figure A2. Both diagrams show that the weighting factor is initially high and decreases rapidly within the first 50–70 s of the experimental time range. This reflects the higher sensitivity of the fitting process to the early creep stage. For reference, the experimental master curves are plotted on the secondary axis, providing visual context for the time-dependent deformation behaviour.
Figure A1. Loss weighting function together with the master curves representing the average experimental creep behaviour of PBT specimens.
Figure A1. Loss weighting function together with the master curves representing the average experimental creep behaviour of PBT specimens.
Applmech 06 00060 g0a1
Figure A2. Loss weighting function together with the master curves representing the average experimental creep behaviour of PBT GF30 specimens.
Figure A2. Loss weighting function together with the master curves representing the average experimental creep behaviour of PBT GF30 specimens.
Applmech 06 00060 g0a2

Appendix B. Detailed Results

This section provides the detailed results of the parameter identification procedure for all the investigated specimens. The parameters for each individual sample were determined using the fitting approach described in Section 2.2. The following subsections contain tabulated values for both materials, PBT and PBT GF30, based on their respective model fits. In addition, a comparison of the mean L1 loss values for both materials is presented to assess the relative fit performance between the Prony series and the fractional viscoelastic models.

Appendix B.1. Detailed Results for PBT

Table A1. Mean L1 loss comparison for PBTGF0.
Table A1. Mean L1 loss comparison for PBTGF0.
Smp.Fractional DamperProny 1%DiffProny 2%DiffProny 3%Diff
04.51 × 10−46.50 × 10−444.2%2.35 × 10−4−47.9%2.14 × 10−4−52.5%
14.06 × 10−46.68 × 10−464.4%2.08 × 10−4−48.7%1.82 × 10−4−55.2%
23.57 × 10−46.00 × 10−468.1%1.49 × 10−4−58.3%1.43 × 10−4−59.9%
33.36 × 10−47.76 × 10−4130.8%3.06 × 10−4−9.0%2.53 × 10−4−24.8%
43.97 × 10−48.16 × 10−4105.5%1.97 × 10−4−50.4%1.97 × 10−4−50.4%
53.37 × 10−46.57 × 10−495.1%2.22 × 10−4−34.0%2.02 × 10−4−39.9%
62.98 × 10−48.18 × 10−4174.5%1.87 × 10−4−37.2%1.87 × 10−4−37.2%
72.47 × 10−46.44 × 10−4161.4%2.28 × 10−4−7.7%1.91 × 10−4−22.5%
82.32 × 10−45.60 × 10−4141.0%1.64 × 10−4−29.5%1.57 × 10−4−32.5%
93.45 × 10−47.97 × 10−4130.7%1.52 × 10−4−56.1%1.52 × 10−4−56.1%
102.73 × 10−45.90 × 10−4116.2%1.47 × 10−4−46.2%1.47 × 10−4−46.2%
114.39 × 10−45.48 × 10−424.8%2.34 × 10−4−46.8%2.06 × 10−4−53.1%
124.78 × 10−46.12 × 10−428.0%1.63 × 10−4−65.9%1.63 × 10−4−65.9%
133.02 × 10−46.31 × 10−4109.0%1.99 × 10−4−34.0%1.73 × 10−4−42.8%
143.44 × 10−45.67 × 10−464.8%2.54 × 10−4−26.1%2.41 × 10−4−29.9%
153.39 × 10−47.14 × 10−4110.3%3.44 × 10−41.3%1.79 × 10−4−47.2%
165.21 × 10−46.25 × 10−419.9%2.72 × 10−4−47.7%1.79 × 10−4−65.7%
173.09 × 10−49.01 × 10−4191.6%1.93 × 10−4−37.5%1.81 × 10−4−41.4%
184.16 × 10−45.25 × 10−426.0%1.65 × 10−4−60.3%1.65 × 10−4−60.3%
194.73 × 10−45.85 × 10−423.7%1.90 × 10−4−59.9%1.80 × 10−4−61.9%
Average 3.65 × 10 4 6.64 × 10 4 91.5% 2.10 × 10 4 −40.1% 1.85 × 10 4 −47.3%
Table A2. Parameters for FractionalDamper—PBTGF0.
Table A2. Parameters for FractionalDamper—PBTGF0.
SampleC α
07.93 × 1050.44
14.19 × 1050.32
25.70 × 1050.35
34.59 × 1050.35
44.16 × 1050.32
54.60 × 1050.34
64.35 × 1050.32
74.19 × 1050.31
88.57 × 1050.41
96.31 × 1050.39
106.99 × 1050.39
115.19 × 1050.35
124.85 × 1050.35
136.11 × 1050.37
144.60 × 1050.31
152.55 × 1050.24
163.34 × 1050.29
173.55 × 1050.31
187.73 × 1050.42
197.31 × 1050.42
Table A3. Parameters for Prony 1—PBTGF0.
Table A3. Parameters for Prony 1—PBTGF0.
Sample K eq K 1 τ 1 K total
0264497462741
1262584272709
2263875322713
3253488382623
4261483252697
5262287352710
6261082292692
7265377382730
8260277512679
9260487352691
10254474452618
11260083372683
12258387342671
13264283422725
14262570272695
15263779212716
16257986252665
17259390282683
18258783442670
19261993432712
Table A4. Parameters for Prony 2—PBTGF0.
Table A4. Parameters for Prony 2—PBTGF0.
Sample K eq K 1 τ 1 K 2 τ 2 K total
02627891203482749
12609641164472716
22622601273782719
32516821183652634
425727727953102703
5260975993652720
62589711374362703
72638711103342742
825797617430102685
925688421646102698
1025226718033132622
11259169883072690
1225646215648142675
132627761193162734
1425936325243102700
15263066504222738
16257167643742675
172569791384762695
182574721133092675
1925877922350172716
Table A5. Parameters for Prony 3—PBTGF0.
Table A5. Parameters for Prony 3—PBTGF0.
Sample K eq K 1 τ 1 K 2 τ 2 K 3 τ 3 K total
026068624537292142750
125718141641232542718
22618591643812812722
324789933738202222637
425727727901053102703
525967218039151712725
62589711370104362703
7257510662335392832743
82500122100036852792685
925688421602446102698
1025226718006733132622
1125716524246201512696
1225646215671442142675
1326127521530221832735
142507141100044171512708
1526155316939143412741
1625605315548152012681
172562801771314292698
182574721130693092675
1924801729645826832718

Appendix B.2. Detailed Results for PBT GF30

Table A6. Mean L1 loss comparison for PBTGF30.
Table A6. Mean L1 loss comparison for PBTGF30.
Smp.Fractional DamperProny 1%DiffProny 2%DiffProny 3%Diff
03.65 × 10−45.44 × 10−448.8%1.40 × 10−4−61.7%1.39 × 10−4−61.8%
12.79 × 10−43.67 × 10−431.4%1.19 × 10−4−57.4%1.18 × 10−4−57.6%
22.18 × 10−45.52 × 10−4153.9%1.51 × 10−4−30.8%1.51 × 10−4−30.8%
32.06 × 10−43.87 × 10−487.8%1.26 × 10−4−38.8%1.23 × 10−4−40.6%
42.96 × 10−43.82 × 10−429.3%1.61 × 10−4−45.5%1.61 × 10−4−45.5%
52.06 × 10−43.14 × 10−452.1%2.17 × 10−45.3%2.03 × 10−4−1.7%
63.23 × 10−43.85 × 10−419.2%1.54 × 10−4−52.5%1.54 × 10−4−52.3%
72.36 × 10−44.47 × 10−489.1%1.51 × 10−4−36.3%1.51 × 10−4−36.3%
83.54 × 10−43.83 × 10−48.1%2.41 × 10−4−31.8%2.15 × 10−4−39.4%
92.10 × 10−43.74 × 10−477.7%1.49 × 10−4−29.0%1.49 × 10−4−29.0%
102.06 × 10−45.77 × 10−4179.6%1.52 × 10−4−26.2%1.46 × 10−4−29.4%
113.62 × 10−45.83 × 10−461.2%1.60 × 10−4−55.8%1.60 × 10−4−55.8%
122.41 × 10−43.73 × 10−454.6%1.31 × 10−4−45.6%1.31 × 10−4−45.7%
132.32 × 10−44.75 × 10−4104.3%1.43 × 10−4−38.5%1.29 × 10−4−44.5%
143.10 × 10−45.17 × 10−466.5%1.70 × 10−4−45.1%1.68 × 10−4−45.8%
152.62 × 10−43.17 × 10−421.3%2.05 × 10−4−21.8%1.83 × 10−4−30.3%
162.51 × 10−44.60 × 10−483.6%1.59 × 10−4−36.4%1.59 × 10−4−36.4%
172.59 × 10−45.80 × 10−4124.1%2.15 × 10−4−16.8%1.94 × 10−4−25.0%
182.50 × 10−43.73 × 10−449.6%2.28 × 10−4−8.5%2.28 × 10−4−8.5%
192.67 × 10−45.01 × 10−487.5%1.91 × 10−4−28.6%1.61 × 10−4−39.9%
Average 2.67 × 10 4 4.44 × 10 4 71.5% 1.68 × 10 4 −35.1% 1.61 × 10 4 −37.8%
Table A7. Parameters for FractionalDamper—PBTGF30.
Table A7. Parameters for FractionalDamper—PBTGF30.
SampleC α
02.33 × 1060.35
13.89 × 1060.40
22.95 × 1060.37
33.55 × 1060.39
44.84 × 1060.44
53.84 × 1060.38
64.27 × 1060.44
75.16 × 1060.47
82.94 × 1060.38
95.83 × 1060.46
102.53 × 1060.35
113.13 × 1060.42
124.38 × 1060.44
132.77 × 1060.37
142.76 × 1060.37
152.92 × 1060.36
163.25 × 1060.39
172.63 × 1060.36
184.41 × 1060.40
192.64 × 1060.35
Table A8. Parameters for Prony 1—PBTGF30.
Table A8. Parameters for Prony 1—PBTGF30.
Sample K eq K 1 τ 1 K total
09203232269435
19463197399660
29311209369520
39259198469456
49295196479491
59201171689372
69320228499548
79262222539485
89225224499449
99340198599538
109231206369436
119290263389553
129312221549533
139212213389425
149316230349546
159285203529487
169236216439452
179348219389567
189370174489544
199300201359501
Table A9. Parameters for Prony 2—PBTGF30.
Table A9. Parameters for Prony 2—PBTGF30.
Sample K eq K 1 τ 1 K 2 τ 2 K total
09123177214148119448
19373184276112169668
2923120119910699539
392261831136769477
489654081000123239496
591601861545249398
69148280436128219556
7919722217282109502
8919619610769109461
9910234560999219546
1091741951439559464
119200232200137129568
12924620818790159544
1391791911028359453
149244191185125129560
159268195835629520
1691762041589199471
179213256298115109584
1893381661226259566
1992621761138769525
Table A10. Parameters for Prony 3—PBTGF30.
Table A10. Parameters for Prony 3—PBTGF30.
Sample K eq K 1 τ 1 K 2 τ 2 K 3 τ 3 K total
09105187265131149129454
1935618135522100110169668
292311791992319810699539
3921817913263102219483
489654081000017123239496
58929341100080714839398
6914123353557226125209556
791971511727117182109502
88976330764139313219477
990473229267834299219546
10915120219680103729470
11920021720015197137129568
129193234320106231729550
13916217714758195939456
149215204262120162449563
1590822481000138545229521
169176160160441529199471
1788885601000122162719597
189338123122421216259566
19923517318876164639530

References

  1. Osswald, T.A.; Menges, G. Materials Science of Polymers for Engineers; Carl Hanser Verlag GmbH Co KG: Munich, Germany, 2012. [Google Scholar] [CrossRef]
  2. Mark, J. Physical Properties of Polymers Handbook; Springer: New York, NY, USA, 2007. [Google Scholar]
  3. Tomar, N.; Maiti, S.N. Mechanical properties of PBT/ABAS blends. J. Appl. Polym. Sci. 2007, 104, 1807–1817. [Google Scholar] [CrossRef]
  4. Domininghaus, H. Kunststoffe: Eigenschaften und Anwendungen; Springer: Berlin/Heidelberg, Germany, 2012. [Google Scholar] [CrossRef]
  5. Schürmann, H. Konstruieren Mit Faser-Kunststoff-Verbunden; Springer: Berlin/Heidelberg, Germany, 2007. [Google Scholar] [CrossRef]
  6. Li, Q. Three Dimensional Numerical Simulation of Melt Filling Process in Mold Cavity with Insets. Procedia Eng. 2015, 126, 496–501. [Google Scholar] [CrossRef][Green Version]
  7. von Bradsky, G.J.; Bailey, R.S.; Cervenka, A.J.; Zachmann, H.G.; Allan, P.S. Characterization of finite length composites: Part IV—Structural studies on injection moulded composites (Technical Report). Pure Appl. Chem. 1997, 69, 2523–2540. [Google Scholar] [CrossRef]
  8. Huang, C.T.; Chen, X.W.; Fu, W.W. Investigation on the Fiber Orientation Distributions and Their Influence on the Mechanical Property of the Co-Injection Molding Products. Polymers 2020, 12, 24. [Google Scholar] [CrossRef]
  9. Rauter, N. A computational modeling approach based on random fields for short fiber-reinforced composites with experimental verification by nanoindentation and tensile tests. Comput. Mech. 2021, 67, 699–722. [Google Scholar] [CrossRef]
  10. Rauter, N.; Lammering, R. Experimental Characterization of Short Fiber-Reinforced Composites on the Mesoscale by Indentation Tests. Appl. Compos. Mater. 2021, 28, 1747–1765. [Google Scholar] [CrossRef]
  11. Klatt, E.; Widera, I.M.; Rauter, N. A procedure for determination of the microstructural probabilistic characteristics from micrographs and their impact on subsequent results. PAMM 2023, 22, e202200277. [Google Scholar] [CrossRef]
  12. Rauter, N.; Lammering, R. Correlation structure in the elasticity tensor for short fiber-reinforced composites. Probabilistic Eng. Mech. 2020, 62, 103100. [Google Scholar] [CrossRef]
  13. Kaliske, M. A formulation of elasticity and viscoelasticity for fibre reinforced material at small and finite strains. Comput. Methods Appl. Mech. Eng. 2000, 185, 225–243. [Google Scholar] [CrossRef]
  14. Chen, Y.; Zhao, Z.; Li, D.; Guo, Z.; Dong, L. Constitutive modeling for linear viscoelastic fiber-reinforced composites. Compos. Struct. 2021, 263, 113679. [Google Scholar] [CrossRef]
  15. Cruz-González, O.; Ramírez-Torres, A.; Rodríguez-Ramos, R.; Otero, J.; Penta, R.; Lebon, F. Effective behavior of long and short fiber-reinforced viscoelastic composites. Appl. Eng. Sci. 2021, 6, 100037. [Google Scholar] [CrossRef]
  16. Tschoegl, N.W. The Phenomenological Theory of Linear Viscoelastic Behavior: An Introduction; Springer: Berlin/Heidelberg, Germany, 1989. [Google Scholar] [CrossRef]
  17. Wineman, A.; Rajagopal, K. Mechanical Response of Polymers: An Introduction; Cambridge University Press: Cambridge, UK, 2000. [Google Scholar]
  18. Structural Mechanics Module User’s Guide; COMSOL: Stockholm, Sweden, 2023.
  19. Abaqus 6.14 Analysis User’s Guide; Dassault Systèmes: Providence, RI, USA, 2014.
  20. Effinger, V.M. Finite nichtlinear viskoelastische Modellierung offenzelliger Polymerschäume. Ph.D. Thesis, Institut für Baustatik und Baudynamik, Universität Stuttgart, Stuttgart, Germany, 2016. [Google Scholar]
  21. Gutierrez-Lemini, D. Engineering Viscoelasticity; Springer: New York, NY, USA, 2013. [Google Scholar]
  22. Endo, V.T.; de Carvalho Pereira, J.C. Linear orthotropic viscoelasticity model for fiber reinforced thermoplastic material based on Prony series. Mech. Time-Depend. Mater. 2017, 21, 199–221. [Google Scholar] [CrossRef]
  23. Bhangale, N.; Kachhia, K.B.; Gómez-Aguilar, J.F. Fractional viscoelastic models with Caputo generalized fractional derivative. Math. Methods Appl. Sci. 2023, 46, 7835–7846. [Google Scholar] [CrossRef]
  24. Chen, W.; Sun, H.; Li, X. Fractional Derivative Modeling in Mechanics and Engineering; Springer: Singapore, 2022. [Google Scholar]
  25. Xu, Y.; Dong, Y.; Huang, X.; Luo, Y.; Zhao, S. Properties Tests and Mathematical Modeling of Viscoelastic Damper at Low Temperature with Fractional Order Derivative. Front. Mater. 2019, 6, 194. [Google Scholar] [CrossRef]
  26. Kontou, E.; Katsourinis, S. Application of a fractional model for simulation of the viscoelastic functions of polymers. J. Appl. Polym. Sci. 2016, 133. [Google Scholar] [CrossRef]
  27. Coelho, C.; P. Costa, M.F.; Ferrás, L. Neural fractional differential equations. Appl. Math. Model. 2025, 144, 116060. [Google Scholar] [CrossRef]
  28. Zimmering, B.; Coelho, C.; Niggemann, O. Optimising Neural Fractional Differential Equations for Performance and Efficiency. In Proceedings of the 1st ECAI Workshop on “Machine Learning Meets Differential Equations: From Theory to Applications”, Santiago de Compostela, Spain, 20 October 2024; Coelho, C., Zimmering, B., Costa, M.F.P., Ferrás, L.L., Niggemann, O., Eds.; Volume 255, pp. 1–22. [Google Scholar]
  29. Lakes, R.S. Viscoelastic Solids; CRC Press: Boca Raton, FL, USA, 1998; Volume 9. [Google Scholar]
  30. Phan-Thien, N. Understanding Viscoelasticity—Basics of Rheology; Springer Science & Business Media: Berlin/Heidelberg, Germany, 2013. [Google Scholar]
  31. Drozdov, A.D. Finite Elasticity and Viscoelasticity; World Scientific: Singapore, 1996. [Google Scholar] [CrossRef]
  32. Haddad, Y.M. Viscoelasticity of Engineering Materials; Springer: Dordrecht, The Netherlands, 1995. [Google Scholar]
  33. Boltzmann, L. Sitzungsber. Kgl. Akad. Wiss. Wien, Math. Naturw. Classe A 1874, 2, 70–275. [Google Scholar]
  34. Ferry, J.D. Viscoelastic Properties of Polymers; John Wiley & Sons: Hoboken, NJ, USA, 1980. [Google Scholar]
  35. Christensen, R. Theory of Viscoelasticity, 2nd ed.; Dover Civil and Mechanical Engineering; Dover Publications: Mineola, NY, USA, 2013. [Google Scholar]
  36. Kreider, D.; Kuller, R.; Ostberg, D. Elementary Differential Equations; Addison-Wesley Series in Mathematics; Addison-Wesley Publishing Company: Boston, MA, USA, 1968. [Google Scholar]
  37. Findley, W.N.; Lai, J.S.; Onaran, K. Creep and Relaxation of Nonlinear Viscoelastic Materials, with an Introduction to Linear Viscoelasticity; Elsevier: Amsterdam, The Netherlands, 1976. [Google Scholar]
  38. Diethelm, K.; Ford, N.J.; Freed, A.D. A Predictor-Corrector Approach for the Numerical Solution of Fractional Differential Equations. Nonlinear Dyn. 2002, 29, 3–22. [Google Scholar] [CrossRef]
  39. Nocedal, J.; Wright, S.J. Numerical Optimization; Springer Series in Operations Research and Financial Engineering; Springer: New York, NY, USA, 2006. [Google Scholar] [CrossRef]
  40. Paszke, A.; Gross, S.; Massa, F.; Lerer, A.; Bradbury, J.; Chanan, G.; Killeen, T.; Lin, Z.; Gimelshein, N.; Antiga, L.; et al. PyTorch: An Imperative Style, High-Performance Deep Learning Library. In Advances in Neural Information Processing Systems; Curran Associates Inc.: Red Hook, NY, USA, 2019. [Google Scholar]
  41. Rumelhart, D.E.; Hinton, G.E.; Williams, R.J. Learning Representations by Back-Propagating Errors. Nature 1986, 323, 533–536. [Google Scholar] [CrossRef]
  42. Baydin, A.G.; Pearlmutter, B.A.; Radul, A.A.; Siskind, J.M. Automatic Differentiation in Machine Learning: A Survey. J. Mach. Learn. Res. 2018, 18, 5595–5637. [Google Scholar]
  43. Nocedal, J. Updating Quasi-Newton Matrices with Limited Storage. Math. Comput. 1980, 35, 773–782. [Google Scholar] [CrossRef]
  44. Kingma, D.P.; Ba, J. Adam: A Method for Stochastic Optimization. arXiv 2014, arXiv:1412.6980. [Google Scholar]
  45. DIN EN ISO 527-2; Plastics—Determination of Tensile Properties—Part 2: Test Conditions for Moulding and Extrusion Plastics. German Version EN ISO 527-2:2012; DIN Deutsches Institut für Normung e.V.: Berlin, Germany, 2012.
  46. Chung, D.D.L. Composite Materials, 2nd ed.; Engineering Materials and Processes; Springer: London, UK, 2010. [Google Scholar]
  47. Vaidya, U.K.; Chawla, K.K. Processing of fibre reinforced thermoplastic composites. Int. Mater. Rev. 2008, 53, 185–218. [Google Scholar] [CrossRef]
  48. DIN EN ISO 899-1; Plastics—Determination of Creep Behaviour—Part 1: Tensile Creep. German Version EN ISO 899-1:2017; DIN Deutsches Institut für Normung e.V.: Berlin, Germany, 2018.
  49. Szalai, S.; Dogossy, G. Speckle pattern optimization for DIC technologies. Acta Tech. Jaurinensis 2021, 14, 228–243. [Google Scholar] [CrossRef]
  50. Reu, P. Introduction to Digital Image Correlation: Best Practices and Applications. Exp. Tech. 2012, 36, 3–4. [Google Scholar] [CrossRef]
  51. Schreier, H.; Orteu, J.J.; Sutton, M.A. Image Correlation for Shape, Motion and Deformation Measurements: Basic Concepts, Theory and Applications; Springer: New York, NY, USA, 2009. [Google Scholar] [CrossRef]
  52. Obaid, N.; Kortschot, M.T.; Sain, M. Understanding the Stress Relaxation Behavior of Polymers Reinforced with Short Elastic Fibers. Materials 2017, 10, 472. [Google Scholar] [CrossRef]
  53. Drozdov, A.; Al-Mulla, A.; Gupta, R. The viscoelastic and viscoplastic behavior of polymer composites: Polycarbonate reinforced with short glass fibers. Comput. Mater. Sci. 2003, 28, 16–30. [Google Scholar] [CrossRef]
  54. Kutty, S.K.N.; Nando, G.B. Short kevlar fiber–thermoplastic polyurethane composite. J. Appl. Polym. Sci. 1991, 43, 1913–1923. [Google Scholar] [CrossRef]
  55. Fliegener, S.; Hohe, J. An anisotropic creep model for continuously and discontinuously fiber reinforced thermoplastics. Compos. Sci. Technol. 2020, 194, 108168. [Google Scholar] [CrossRef]
  56. Burgarella, B.; Maurel-Pantel, A.; Lahellec, N.; Bouvard, J.L.; Billon, N.; Moulinec, H.; Lebon, F. Effective viscoelastic behavior of short fibers composites using virtual DMA experiments. Mech. Time-Depend. Mater. 2019, 23, 337–360. [Google Scholar] [CrossRef]
  57. DIN EN ISO 527-1; Plastics—Determination of Tensile Properties—Part 1: General Principles. German Version EN ISO 527-1:2019; DIN Deutsches Institut für Normung e.V.: Berlin, Germany, December 2019.
  58. Marino, E.; Flaschel, M.; Kumar, S.; De Lorenzis, L. Automated identification of linear viscoelastic constitutive laws with EUCLID. Mech. Mater. 2023, 181, 104643. [Google Scholar] [CrossRef]
  59. Qin, B.; Zhong, Z. A Physics-Guided Machine Learning Model for Predicting Viscoelasticity of Solids at Large Deformation. Polymers 2024, 16, 3222. [Google Scholar] [CrossRef] [PubMed]
  60. Otsuki, Y.; Hashimoto, K.; Kobayashi, Y.; Nishitsuji, S.; Matsuno, H.; Ito, H. Analyzing the Tensile Creep Behavior of Different Types of Polypropylenes Using a Simple Fractional Differential Viscoelastic Model. Polymers 2025, 17, 1095. [Google Scholar] [CrossRef] [PubMed]
  61. Santi, L.; Bennison, S.; Härth, M.; Royer Carfagni, G. Fractional Viscoelastic Modelling of Polymeric Interlayers in Laminated Glass. Comparisons with Prony Series Approach; Glass Performance Days: Tampere, Finland, 2023. [Google Scholar]
Figure 1. Elementary rheological models of viscoelasticity.
Figure 1. Elementary rheological models of viscoelasticity.
Applmech 06 00060 g001
Figure 2. Schematic representation of rheological models for viscoelasticity: (a) basic Maxwell model and (b) generalised Maxwell model.
Figure 2. Schematic representation of rheological models for viscoelasticity: (a) basic Maxwell model and (b) generalised Maxwell model.
Applmech 06 00060 g002
Figure 3. Schematic representation of the tensile test specimen according to DIN EN ISO 527-2 Type 1A, with maximum dimensions of 150 mm × 20 mm × 4 mm .
Figure 3. Schematic representation of the tensile test specimen according to DIN EN ISO 527-2 Type 1A, with maximum dimensions of 150 mm × 20 mm × 4 mm .
Applmech 06 00060 g003
Figure 4. Visualisation of the experimental analysis: (a) setup with the ARAMIS digital image correlation system, (b) the tensile test specimen mounted in the tensile testing machine, and the strain distribution ( ε y ) in the x-y plane.
Figure 4. Visualisation of the experimental analysis: (a) setup with the ARAMIS digital image correlation system, (b) the tensile test specimen mounted in the tensile testing machine, and the strain distribution ( ε y ) in the x-y plane.
Applmech 06 00060 g004
Figure 5. Experimental creep behaviour of the total strain ε total results for 20 specimens of PBT and PBT GF30 over the test duration.
Figure 5. Experimental creep behaviour of the total strain ε total results for 20 specimens of PBT and PBT GF30 over the test duration.
Applmech 06 00060 g005
Figure 6. Master curves representing the average experimental creep behaviour ε total of the PBT and PBT GF30 samples.
Figure 6. Master curves representing the average experimental creep behaviour ε total of the PBT and PBT GF30 samples.
Applmech 06 00060 g006
Figure 7. Master curves representing the average experimental creep behaviour ε creep according to Equation (1) of the PBT and PBT GF30 samples.
Figure 7. Master curves representing the average experimental creep behaviour ε creep according to Equation (1) of the PBT and PBT GF30 samples.
Applmech 06 00060 g007
Figure 8. Master curves representing the average experimental creep behaviour of PBT specimens. The prediction models include the Prony series model ( N { 1 , 2 , 3 } ) with mean values from Table 3 and the fractional model with mean values from Table 4.
Figure 8. Master curves representing the average experimental creep behaviour of PBT specimens. The prediction models include the Prony series model ( N { 1 , 2 , 3 } ) with mean values from Table 3 and the fractional model with mean values from Table 4.
Applmech 06 00060 g008
Figure 9. Master curves representing the average experimental creep behaviour of PBT GF30 specimens. The prediction models include the Prony series model ( N { 1 , 2 , 3 } ) with mean values from Table 3 and the fractional model with mean values from Table 4.
Figure 9. Master curves representing the average experimental creep behaviour of PBT GF30 specimens. The prediction models include the Prony series model ( N { 1 , 2 , 3 } ) with mean values from Table 3 and the fractional model with mean values from Table 4.
Applmech 06 00060 g009
Table 1. Injection moulding parameters for the PBT and PBT GF30 specimens.
Table 1. Injection moulding parameters for the PBT and PBT GF30 specimens.
ParameterPBT (Unreinforced)PBT GF30
Material Conditioning and Machine Setup
Drying temperature [°C]120120
Drying time [h]55
Mould temperature [°C]6060
Melt temperature [°C]265265
Screw diameter [mm]3030
Number of cavities22
Clamping force [kN]550550
Injection Phase
Dosing volume [cm3]35.035.0
Back pressure [bar]8080
Dosing speed [mm/s]300300
Injection speed [cm3/s]2525
Injection pressure [bar]556 (spec.)913 (spec.)
Injection time [s]1.221.23
Dosing time [s]3.933.83
Dosing delay [s]1.001.00
Packing and Holding Phase
Switchover point [cm3]7.07.0
Holding pressure [bar]550600
Holding time [s]20.020.0
Residual cushion [cm3]3.503.64
Cooling and Cycle Summary
Cooling circuits—nozzle side22
Cooling circuits—ejector side33
Cooling time [s]24.024.0
Table 2. Summary of experimental conditions for creep behaviour testing of PBT and PBT GF30 specimens.
Table 2. Summary of experimental conditions for creep behaviour testing of PBT and PBT GF30 specimens.
Type of MaterialForce [N]Time to Apply the Force [s]Duration of Creep Measurement [s]
PBT3003300
PBT GF3010253300
Table 3. Mean values and standard deviations (Mean ± StdDev) of Prony-series coefficients for Prony models with 1 to 3 terms.
Table 3. Mean values and standard deviations (Mean ± StdDev) of Prony-series coefficients for Prony models with 1 to 3 terms.
Prony ModelParameterPBTGF0PBTGF30
Prony 1 K e q 2607.86 ± 30.74 9288.97 ± 62.72
K 1 83.29 ± 6.54 211.06 ± 20.41
τ 1 35.12 ± 7.97 44.36 ± 9.83
K total 2691.15 ± 31.03 9500.03 ± 63.12
Prony 2 K e q 2587.84 ± 32.41 9201.04 ± 83.65
K 1 72.42 ± 7.65 219.80 ± 59.49
τ 1 143.91 ± 58.90 243.57 ± 212.33
K 2 39.20 ± 7.21 96.43 ± 26.91
τ 2 8.04 ± 3.63 10.57 ± 5.83
K total 2699.47 ± 32.14 9517.27 ± 59.73
Prony 3 K e q 2561.99 ± 41.87 9140.27 ± 123.57
K 1 84.94 ± 29.50 240.49 ± 102.30
τ 1 348.40 ± 291.28 447.15 ± 343.35
K 2 27.43 ± 19.02 67.33 ± 40.98
τ 2 26.59 ± 21.48 89.64 ± 93.09
K 3 27.29 ± 12.52 72.53 ± 41.13
τ 3 5.22 ± 4.16 7.95 ± 7.11
K total 2701.65 ± 32.32 9520.62 ± 59.09
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Klatt, E.; Zimmering, B.; Niggemann, O.; Rauter, N. Machine-Learning-Enabled Comparative Modelling of the Creep Behaviour of Unreinforced PBT and Short-Fibre Reinforced PBT Using Prony and Fractional Derivative Models. Appl. Mech. 2025, 6, 60. https://doi.org/10.3390/applmech6030060

AMA Style

Klatt E, Zimmering B, Niggemann O, Rauter N. Machine-Learning-Enabled Comparative Modelling of the Creep Behaviour of Unreinforced PBT and Short-Fibre Reinforced PBT Using Prony and Fractional Derivative Models. Applied Mechanics. 2025; 6(3):60. https://doi.org/10.3390/applmech6030060

Chicago/Turabian Style

Klatt, Eduard, Bernd Zimmering, Oliver Niggemann, and Natalie Rauter. 2025. "Machine-Learning-Enabled Comparative Modelling of the Creep Behaviour of Unreinforced PBT and Short-Fibre Reinforced PBT Using Prony and Fractional Derivative Models" Applied Mechanics 6, no. 3: 60. https://doi.org/10.3390/applmech6030060

APA Style

Klatt, E., Zimmering, B., Niggemann, O., & Rauter, N. (2025). Machine-Learning-Enabled Comparative Modelling of the Creep Behaviour of Unreinforced PBT and Short-Fibre Reinforced PBT Using Prony and Fractional Derivative Models. Applied Mechanics, 6(3), 60. https://doi.org/10.3390/applmech6030060

Article Metrics

Back to TopTop