Skip to main content
Polymers logoLink to Polymers
. 2023 Feb 15;15(4):958. doi: 10.3390/polym15040958

A Class of Algorithms for Recovery of Continuous Relaxation Spectrum from Stress Relaxation Test Data Using Orthonormal Functions

Anna Stankiewicz 1
Editor: Wenbo Luo1
PMCID: PMC9958823  PMID: 36850241

Abstract

The viscoelastic relaxation spectrum provides deep insights into the complex behavior of polymers. The spectrum is not directly measurable and must be recovered from oscillatory shear or relaxation stress data. The paper deals with the problem of recovery of the relaxation spectrum of linear viscoelastic materials from discrete-time noise-corrupted measurements of relaxation modulus obtained in the stress relaxation test. A class of robust algorithms of approximation of the continuous spectrum of relaxation frequencies by finite series of orthonormal functions is proposed. A quadratic identification index, which refers to the measured relaxation modulus, is adopted. Since the problem of relaxation spectrum identification is an ill-posed inverse problem, Tikhonov regularization combined with generalized cross-validation is used to guarantee the stability of the scheme. It is proved that the accuracy of the spectrum approximation depends both on measurement noises and the regularization parameter and on the proper selection of the basis functions. The series expansions using the Laguerre, Legendre, Hermite and Chebyshev functions were studied in this paper as examples. The numerical realization of the scheme by the singular value decomposition technique is discussed and the resulting computer algorithm is outlined. Numerical calculations on model data and relaxation spectrum of polydisperse polymer are presented. Analytical analysis and numerical studies proved that by choosing an appropriate model through selection of orthonormal basis functions from the proposed class of models and using a developed algorithm of least-square regularized identification, it is possible to determine the relaxation spectrum model for a wide class of viscoelastic materials. The model is smoothed and robust on measurement noises; small model approximation errors are obtained. The identification scheme can be easily implemented in available computing environments.

Keywords: viscoelasticity, relaxation spectrum, linear relaxation modulus, identification algorithm, orthonormal functions, Tikhonov regularization, singular value decomposition

1. Introduction

Viscoelasticity denotes the joint property of elasticity and viscosity and, hence, describes materials with both fluid and solid properties at the same time. Viscoelastic relaxation or retardation spectra are commonly used to describe, analyze, compare and improve the mechanical properties of polymers [1,2,3,4,5]. The spectra are vital for constitutive models and for the insight into the properties of a viscoelastic material, since, from the relaxation or retardation spectrum, other material functions used to describe rheological properties of various polymers can be uniquely determined [6,7,8,9]. However, the spectra are not directly accessible by measurement. The relaxation and retardation spectra can be recovered from oscillatory shear data and from the time measurements of the relaxation modulus or creep compliance obtained in standard stress relaxation or retardation experiments [1,2,7]. Many different methods have been proposed during the last five decades for relaxation spectrum computation using data from dynamic modulus tests. Baumgaertel and Winter [10] applied a nonlinear regression for identification of discrete relaxation and retardation time spectra based on dynamic data, in which the number of relaxation times adjusts during the iterative calculations to avoid ill-posedness and to improve the model fit; regularization is not applied here. Honerkamp and Weese [11,12], for relaxation spectrum identification, combined nonlinear regression with Tikhonov regularization and proposed a specific viscoelastic model described by the two-mode log-normal function. Malkin [13] approximated a continuous relaxation spectrum using three constants: the maximum relaxation time, slope in the logarithmic scale and form factor. Malkin et al. [14] derived a method of continuous relaxation spectrum calculations using the Mellin integral transform. An approach proposed by Stadler and Bailly [15] is based on the relaxation spectrum approximation by a piecewise cubic Hermite spline. In turn, Davies and Goulding [16] approximated the relaxation spectrum by a sum of scaling kernel functions located at appropriately chosen points. The algorithm for the relaxation time spectrum approximation by power series was developed by Cho [17]. Anderssen et al. [1] proposed a derivative-based algorithm for continuous spectrum recovery, being also appropriate for the experimental situation where the oscillatory shear data are only available for a finite range of frequencies. These works, but also many others, using different models, approaches, algorithms and computational techniques, have opened new directions of research on discrete and continuous relaxation spectra identification based on dynamic moduli data, which are still being conducted [2,18,19,20,21].

However, a classical manner of studying viscoelasticity is also through a two-phase stress relaxation test, where time-dependent shear stress is studied for the step increase in strain [4,6,7]. There are only a few papers, e.g., [22,23,24,25,26,27,28], that deal with the spectrum determination from time measurements of the relaxation modulus; additionally, only some of them are addressed to polymers. Therefore, the computationally efficient algorithms to determine the relaxation spectrum applied to time measurements of the relaxation modulus are still desirable. The objective of the present paper was to develop a class of models and an identification algorithm for the continuous relaxation spectrum determination based on discrete-time measurements of the relaxation modulus, which, taking into account the ill-posedness of the original problem of the spectrum recovery, will provide: (a) good approximation of the relaxation spectrum and modulus; (b) smoothness of the spectrum fluctuations, even for noise-corrupted measurements; (c) noise robustness; (d) applicability to a wide range of viscoelastic materials due to the choice of respective model from the considered set of models; (e) ease of implementation of the models and identification algorithm in available computing packages. Thus, the goal of this work was the synthesis of the respective models and general identification scheme, and the analysis of their properties. Approximation errors, convergence, noise robustness, smoothness and the applicability ranges were studied analytically. Further, the numerical verification of the models and algorithm for exemplary theoretical relaxation spectrum, and their applicability to spectrum of real material, polydisperse polymer, was the purpose of this work.

The approach proposed is based on the approximation of the spectrum by a finite linear combination of the basis orthonormal functions. A quadratic identification index, related to the data of the relaxation modulus, is adopted as a measure of the model quality. As a result, the primary infinite dimensional dynamic inverse problem of the continuous relaxation spectrum identification is reduced to the static linear-quadratic programming task. Next, Tikhonov regularization is used to guarantee the well-posed solution. Thus, the approach proposed integrates the technique of an expansion of a function into a series in an orthogonal basis with the least-squares regularized identification [29].

It is demonstrated that due to the choice of appropriate special functions as the basis functions for the unknown relaxation spectrum model, the components in the relaxation modulus model are given by compact analytical or recursive formulas. The technique of expanding an unknown viscoelastic function into a series of orthogonal functions or polynomials has already been used to describe various rheological models of polymers, especially in the time domain. For example, Aleksandrov et al. [30] applied Laguerre polynomials to describe experimentally obtained polyethylene deformation in the creep under diffusion in a liquid environment. Cao et al. [31] used orthogonal expansion based on shifted Legendre polynomials to solve a fractional-order viscoelastic model of polymethyl methacrylate. Abbaszadeh and Dehghan [32] employed a new class of basis based upon the Legendre polynomials to solve a two-dimensional viscoelastic equation. Kim et al. [33] used Chebyshev polynomials for direct conversion of creep data to dynamic moduli.

The idea of using a series of orthogonal functions has also been used to approximate the relaxation spectrum. Lee et al. [34] used the Chebyshev polynomials of the first kind to approximate dynamic moduli data. Stankiewicz [27,28,35] applied orthogonal functions for relaxation spectrum recovery from the stress relaxation data, but these articles use a different definition of the relaxation modulus, according to which the modulus is directly given by the Laplace integral of the spectrum. In this paper, it is shown, for the dominant literature definition of the relaxation spectrum, that the application of the concept of expanding the unknown relaxation spectrum into a series of orthonormal basis functions combined with the least-squares regularized identification allows one to determine the smoothed model of the relaxation spectrum, robust on the measurement noises, with small approximation errors of the relaxation spectrum and modulus. The selection of appropriate orthonormal basis functions, the selection of their time-scale factor and the determination of the optimal regularization parameter using standard generalized cross-validation technique enables the application of the proposed approach to a wide class of viscoelastic materials.

2. Materials and Methods

2.1. Relaxation Spectrum

The uniaxial, nonaging and isothermal stress–strain equation for a linear viscoelastic material can be represented by a Boltzmann superposition integral [7]:

σt=tGtλε˙λdλ, (1)

where σt and εt denote the stress and strain at the time t and Gt is the linear relaxation modulus. Modulus Gt is given by [1,7,36,37]:

Gt=0Hττet/τdτ, (2)

or equivalently by

Gt=0Hvvetvdv, (3)

where Hτ and Hv characterize the distributions of relaxation times τ and relaxation frequencies v, respectively. The continuous relaxation spectra Hτ and Hv, related by Hv=H1v, are generalizations of the discrete Maxwell spectrum [1,7] to a continuous function of the relaxation times τ and frequencies v. Although other definitions of the relaxation spectrum are used in the literature, for example, in [5,13,28,38], the definition introduced by Equation (2) dominates. The main symbols are summarized in Nomenclature, Appendix C.

The problem of relaxation spectrum determination is the practical problem of re-constructing the solution of the Fredholm integral equation of the first kind (2) or (3) from discrete-time measured data. Time measurements of the relaxation modulus data are considered in this paper. This problem is known to be severely Hadamard ill-posed [39,40]. In particular, small changes in the measured relaxation modulus can lead to arbitrarily large changes in the determined relaxation spectrum. In remedy, some reduction of the set of admissible solutions or appropriate regularization of the original problem can be used. Here, both techniques are used simultaneously.

2.2. Models

The modified spectrum is introduced:

HMv=Hvv, (4)

where the upper index of HMv means ‘modified’. Then, (3) can be rewritten as

Gt=0HMvetvdv, (5)

i.e., the modulus is directly the Laplace integral of the spectrum HMv.

Assume that HMvL20,, where L20, is the space of real-valued square-integrable functions on the interval 0,. The respective sufficient conditions are given by Theorem 3 in [41]. Assume that the set of the linearly independent orthonormal functions h0v,h1v,h2v, form a basis of the space L20,. Thus, the modified relaxation spectrum can be expressed as

HMv=k=0gkhkv, (6)

where the Fourier coefficients are [42]

gk=0HMvhkvdv.

For practical reasons and in order to reduce the set of admissible solutions, it is convenient to replace the infinite summation in Equation (6) with a finite one of K terms, i.e., to approximate the relaxation spectrum HMv by a model of the form

HKMv=k=0K1gkhkv, (7)

where the lower index of HKMv is the number K of model summands. Then, using (5), the respective model of the relaxation modulus is described by:

GKt=0HKMvetvdv=k=0K1gkϕkt, (8)

where the functions

ϕkt=0hkvetvdv. (9)

Note, that function ϕkt is Laplace transform of hkv for real argument t, i.e.,

Lhkv=ϕkt,

with the notation Lfv used for Laplace transform. For basis functions, hkv applied in the developed algorithms, the components ϕkt of the relaxation modulus model GKt are given by analytical or recursive formulas. This avoids quadrature errors occurring in the numerical calculation of the integrals (9). The following special functions are considered as basis functions: Laguerre, Legendre, Chebyshev and Hermite. The respective basis functions hkv and ϕkt are described in Section 3.3, Section 3.4, Section 3.5 and Section 3.6. All basis functions depend on one parameter—time-scaling factor α.

2.3. Identification Problem

Identification consists of selecting, within the given class of models defined by (7), (8) such a model, which ensures the best fit to the measurement results. Suppose a certain identification experiment (stress relaxation test [4,6,7]) performed on the specimen of the material under investigation resulted in a set of measurements of the relaxation modulus {Gti=Gti+zti} at the sampling instants ti0, i=1,,N, where zti is additive measurement noise. It is assumed that the number of measurements NK. As a measure of model (8) accuracy, the square index is taken

QNgK=i=1NGtiGKti2, (10)

where gK=g0gK1T is a K-element vector of unknown coefficients of the models (7) and (8). Using the vector-matrix notation

ΦN,K=ϕ0t1ϕK1t1ϕ0tNϕK1tN,GN=Gt1GtN (11)

the identification index (10) can be rewritten in compact form as

QNgK=GNΦN,KgK22, (12)

where ·2 denotes the square norm in the real Euclidean space RN. Thus, the optimal identification of relaxation spectrum in the class of functions defined by (7) and (8) consists of solving, with respect to the model parameter gK, the following least-squares problem:

mingKRK GNΦN,KgK22. (13)

The matrix ΦN,K is usually ill-conditioned. Thus, the optimization problem (13) is still, like the original problem of solving Fredholm’s equation of the 1st kind (3), incorrectly posed in the sense of Hadamard. Consequently, the solution of (13) is not unique, i.e., there exist many optimal model parameters minimizing the identification index QNgK (12). However, even the normal (with the lowest Euclidean norm) solution of (13) is non-continuous and unbounded function of the measurement vector GN. This means that when the data are noisy, even small changes in GN would lead to arbitrarily large artefacts in any optimal model parameter. To deal with the ill posedness, the Tikhonov regularization method is used, as presented below.

2.4. Regularization

Regularization aims to replace the ill-posed problem with a nearby well-posed problem. Tikhonov regularization [43] strives to stabilize the computation of the least-squares solution by minimizing a modified square functional of the form:

mingKRK GNΦN,KgK22+λgK22, (14)

where λ>0 is a regularization parameter. The above problem is well-posed; that is, the solution always exists, is unique, and continuously depends on both the matrix ΦN,K and on the measurement data GN. The parameter vector minimizing (14) is given by:

gKλ=ΦN,KTΦN,K+λIK,K1ΦN,KTGN, (15)

where IK,K is K dimensional identity matrix.

The choice of regularization parameter λ is crucial to identify the best model parameters. Here, we apply the generalized cross-validation GCV [39,44], which does not depend on a priori knowledge about the noise variance. The GCV technique relies on choosing, as a regularization parameter, λ, which minimizes the GCV functional defined by [44]

VGCVλ=ϱλ22/trΞλ2, (16)

where the matrix

Ξλ=IN,NΦN,KΦN,KTΦN,K+λIK,K1ΦN,KT,

and

ϱλ=ΞλGN=GNΦN,KgKλ,

are the residual vector for the regularized solution (15); trΞλ denotes the trace of Ξλ. The problem of choosing the optimal regularization parameter

λGCV=minλ:λ=arg minλ0VGCVλ. (17)

has a unique solution and the resulting parameter gKλGCV differs the least from the normal solution of problem (14) that we would obtain for the ideal (not noise corrupted) measurements of the relaxation modulus [44].

2.5. Algebraic Background

Formula (15) is generally unsuitable for computational purposes. The singular value decomposition (SVD, [45]) technique will be used. Let SVD of the N×K dimensional matrix ΦN,K take the form [45]:

ΦN,K=UΣVT, (18)

where Σ=diagσ1,,σr,0,,0ϵRN,K is diagonal matrix containing the non-zero singular values σ1,,σr of the matrix ΦN,K [45], matrices VRK,K and URN,N are orthogonal and r=rankΦN,K<N. Taking advantage of the diagonal structure of Σ and the matrices V and U orthogonality, it may be simply proved that the regularized optimal parameter gKλ (15) is given by

gKλ=VΛλUTGN, (19)

where K×N diagonal matrix Λλ is as follows:

Λλ=diagσ1/σ12+λ,,σr/σr2+λ,0,,0. (20)

Using SVD (18) and introducing N dimensional vector Y=UTGN, the GCV function (16) can be expressed by a convenient analytical formula

VGCVλ=i=1rλ2yi2σi2+λ2+i=r+1Nyi2/Nr+i=1rλσi2+λ2, (21)

as a function of the singular values σi and elements yi of the vector Y. The function VGCVλ is differentiable for any λ; thus, an arbitrary gradient optimization method can be implemented to solve the GCV minimization task (17).

3. Results and Discussion

In this section, a general scheme of the relaxation spectrum identification is given. The most important results for the evaluation of the effectiveness of the algorithm and models are presented, concerning the smoothing of the models, their accuracy for ideal and noisy measurements of the relaxation modulus and the linear convergence to the model that we would obtain for the noise-free measurements. Next, examples of orthonormal basis functions hkv and corresponding functions ϕkt are given.

3.1. Identification Algorithm

The determination of the model of relaxation spectrum involves the following steps.

  1. Perform the experiment (stress relaxation test [4,7,46,47]) and record the measurements Gti, i=1,,N, of the relaxation modulus at times ti0.

  2. Choose the time-scaling factor α and the number K of model components comparing, for different values of α, a few first functions from the sequence ϕkt with the experiment results {Gti}.

  3. Compute the matrix ΦN,K (11) and next determine SVD (18).

  4. Determine GCV function VGCVλ (21), and next compute the optimal regularization parameter λGCV minimizing VGCVλ, i.e., solving the optimization task (17).

  5. Compute the regularized solution gKλ according to (19) for λ=λGCV.

  6. For λ=λGCV, using gKλ computed above, determine the modified spectrum of relaxation frequencies HKMv according to:
    HKMv=k=0K1gkλhkv. (22)
  7. Determine the spectrum of relaxation frequencies HKv according to
    HKv=HKMvv=k=0K1gkλhkvv. (23)

Two Remarks

  1. Only the SVD of the matrix, ΦN,K, of computational complexity ONK2 [45] is a space- and time-consuming task of the scheme. However, the SVD must be computed only once and is accessible in the form of optimized numerical procedures in most commonly used computational packets.

  2. The matrix ΦN,K depends on the choice of the basis functions as well as the measurement points ti; however, it does not depend on the relaxation modulus measurements Gti. Thus, when the identification scheme is applied for successive samples of the same material, step 3 should not be repeated while the same time instants ti are kept and the same model parameters α and K are selected in step 2.

3.2. Analysis

In the context of the ill-posed inverse problem, for which the model quality index refers to the measured relaxation modulus but not directly to the unknown relaxation spectrum Hv and the modified spectrum HMv (4), we cannot estimate the error HKMvHMv directly. As a reference point for the determined model HKMv (22), we will consider several characteristics, as follows:

  • (a)

    The model of the relaxation spectrum that we would obtain on the basis of ideal (undisturbed) measurements of the relaxation modulus:

H~KMv=k=0K1g~kλhkv, (24)

where g~Kλ is the vector of regularized solution of (14)

g~Kλ=ΦN,KTΦN,K+λIK,K1ΦN,KTGN, (25)

for noise-free measurements of relaxation modulus GN=Gt1GtNT; c.f., Equation (11)

  • (b)

    The model of the relaxation spectrum that we obtain on the basis of the normal solution gKN=ΦN,KGN of the linear-quadratic problem (13) for noise measurements of the relaxation modulus:

HKNv=k=0K1gkNhkv, (26)

where ΦN,K is the Moore–Penrose pseudoinverse [48] of matrix ΦN,K, and gkN is the elements of the vector gKN

  • (c)

    The model of the relaxation spectrum that we would obtain on the basis of the normal solution g~KN=ΦN,KGN of the linear-quadratic problem (13) for noise-free measurements of the relaxation modulus:

H~KNv=k=0K1g~kNhkv. (27)

Most of the results are formulated in terms of algebraic tools of the algorithm, i.e., SVD decomposition (18) of the matrix ΦN,K. Such an analysis enables a deeper insight into the properties of the algorithm and the resulting model. It shows not only the influence of the regularization parameter and measurement errors, but also the impact and significance of the selection of the basis functions, including their parameters and measurement points ti, on which the singular values σi of the matrix ΦN,K depend.

3.2.1. Smoothness

The purpose of Tikhonov regularization relies on stabilization of the resulting vector gKλ. Due to the orthonormality of the basis functions hkv in the Hilbert space L20,, for an arbitrary HKMv of the form (22), the following equality holds

HKMv22=k=0K1j=0K1gkλgjλ0hkvhjvdv=k=0K1gkλ2=gKλ22, (28)

where ·2 means the square norm, both in the real Euclidean space as well as in L20,. Therefore, the smoothness of the optimal solution gKλ of discrete problem (14) guarantees that the fluctuations in the respective spectrum of relaxation, in particular the resulting spectrum of relaxation HKMv (22), are also bounded. In view of the above, due to orthonormality of the elements the basis system hkv, the function HKMv is the approximation of the real modified spectrum HMv in the class of functions HKMv (7), optimal in the sense of the square identification index QNgK (10) of the bounded norm.

For any regularized gKλ (19), bearing in mind the definition of the vector Y=UTGN and orthogonality of V, we have gKλ22=YTΛλTVTVΛλY=YTΛλTΛλY. Thus, due to the diagonal structure of Λλ (20) and based on (28), the model smoothing efficiency can be evaluated by the following relation:

HKMv22=i=1rσi2yi2σi2+λ2<i=1ryi2σi2=HKNv22, (29)

which holds for an arbitrary regularization parameter λ>0, where the spectrum HKNv is given by (26). The last equality in (29) holds, since for ΦN,K (18), the Moore–Penrose pseudoinverse is ΦN,K=VΣUT, where K×N matrix Σ=diag1/σ1,,1/σr,0,,0. Keeping in mind (26) and (28), the above result can be derived directly from the following inequality, proved in [28]:

gKλ22=i=1ryi2σi+λ2<i=1ryi2σi2=gKN22,

where σ1,,σr are the non-zero singular values of the matrix ΦN,KTΦN,K and yi are elements of the K dimensional vector Y=VTΦN,KTGN=ΣTY, since σi=σi2 and yi=σiyi.

The first equality in (29) illustrates the mechanism of stabilization. The following rule holds: the greater the regularization parameter λ is, the more highly bounded the fluctuations of the spectrum HKMv are. Thus, due to orthogonality of the basis functions, the regularization parameter controls the smoothness, not only of the parameter gKλ but also of the model HKMv. The non-zero singular values of the matrix ΦN,K and the vector of measurement data GN also affect the smoothness of the spectrum model.

3.2.2. Convergence

Relaxation spectrum HKMv (22) is only an approximation of that spectrum, which can be obtained in the class of models (7) by direct minimization (without regularization) of the quadratic index QNgK (12) for noise-free measurements, i.e., the approximation of the function H~KNv (27). Since

gKλg~KN=VΛλΣUTGN+VΣUTzN,

where zN=zt1ztNT is vector of measurement noises, based on (28); the diagonal structure of ΛλΣ and using the Schwarz inequality [49], the following bound of the relaxation spectrum approximation error can be derived:

HKMvH~KNv2=gKλg~KN2i=1rλyiσiσi2+λ+1σr2zN2. (30)

The inequality (30) yields that the accuracy of the spectrum approximation depends both on the measurement noises and the regularization parameter and on the singular values σ1,,σr of the matrix ΦN,K, which, in turn, depend on the selection of the basis orthogonal functions hkv. Using (30), the regularized vector gKλ converges to the noise-free normal solution g~KN linearly with respect to the norm zN2, as λ0 and zN20, simultaneously. Therefore, the upper bound in (30) guarantees that the spectrum HKMv tends to H~KNv in each point v, at which they are both continuous, as λ0 and zN20, simultaneously.

3.2.3. Noise Robustness

The influence of disturbances in the measurements of the relaxation modulus on the regularized solution gKλ was analyzed in detail in a two-part paper [27,50]. From Property 2 in [50], the following inequalities result:

HKMvH~KMv2=gKλg~Kλ2max1irσiσi2+λzN2σ1σr2+λzN2.

Thus, the regularized vector gKλ converges to the noise-free regularized solution g~Kλ (25), and the relaxation spectrum HKMv tends to the noise-free spectrum H~KMv (24) in each point v, where they are both continuous, linearly with respect to the norm zN2, as zN20. From the above estimations, it is also evident that the accuracy of the spectrum approximation measured by HKMvH~KMv2 depends both on the measurement noises and the regularization parameter λ and on the singular values of the matrix ΦN,K.

3.3. Legendre Model

Let us assume a basis function

hkv=2α2k+1eαvPk12e2αv,k=0,1,2,, (31)

with the time-scaling factor α, where Pkx is Legendre polynomials [51,52,53] defined by Rodrigue’s formula

Pkx=12kk!dkdxkx21k, k=0,1,2,.

The polynomials Pkx form a complete set of orthonormal basis in the interval 1,1 with the weight 2k+1/2 [51,53]. Thus, using the substitution x=12e2αv, it is easy to observe that the functions hkv defined by (31) form a complete orthonormal basis in L20, [49]. The relaxation modulus basis functions ϕkt (9) are as follows:

ϕkt=2α2k+1i=0k12i+1αti=0k2i+1α+t,k=0,1,2,, (32)

where the product i=0pxi is equal to 1 when p<0. The proof by induction is presented in Appendix A.1. The above formula can be equivalently expressed in recurrent form as

ϕk+1t=ϕkt2k+32k+1αt2k+12k+3α+t,k=0,1,2,,

starting with

ϕ0t=2αα+t.

Five first basis functions hkv are shown in Figure 1a,b for two different values of the time-scaling factor α. Figure 1c,d show the related ϕkt functions. From the last figure, it is seen that the basis functions for the relaxation modulus model are in good agreement with the real relaxation modulus obtained in the experiment.

Figure 1.

Figure 1

Basis functions hkv (31) and ϕkt (32) of the Legendre model for two time-scaling factors α: (a) hkv, α=0.05[s]; (b) hkv, α=0.5[s]; (c) ϕkt, α=0.05[s]; (d) ϕkt, α=0.5[s]; k=0,1,2,3,4.

3.4. Laguerre Model

The Laguerre polynomials can be defined via Rodrigue’s formula [51,54]:

Lkv=eαvk!dkdvkvkeαv, k=0,1,2, (33)

where α>0 is a time-scaling factor [55]. The continuous Laguerre function is the product of the Laguerre polynomial and the square root of the exponential weight function αeαv [56], i.e.,

hkv=αeαv/2Lkv,k=0,1,2, (34)

The Laguerre functions form a complete orthonormal basis in L20, [51,56]. In Appendix A.2, the following formula is derived for the modulus basis functions:

ϕkt=αtα2kt+α2k+1, k=0,1,2,. (35)

The above formula is given by Wang and Cluett [55], but as there are several definitions of the Laguerre functions in the literature, and, as a result, several formulas of the Laplace transforms, for example, in [57], the derivation of (35) is given in Appendix A.2 to avoid doubts.

A few first basis functions hkv are shown in Figure 2a,b for two different values of the time-scaling factor α; the corresponding functions ϕkt are plotted in Figure 2c,d.

Figure 2.

Figure 2

Basis functions hkv (34) and ϕkt (35) of the Laguerre model for two time-scaling factors: (a) hkv, α=1[s]; (b) hkv, α=10[s]; (c) ϕkt, α=1[s]; (d) ϕkt, α=10[s]; k=0,1,2,3,4.

3.5. Chebyshev Model

The Chebyshev polynomials of the first kind defined by the recursion relation [58,59]:

Tkx=2xTk1xTk2x, k=2,3,, (36)

starting with

T0x=1, T1x=x, (37)

are orthogonal in the interval [1,1] with the weight function 1x21/2 [58]. Specifically,

11TkxTmx1x2dx=0kmπ2k=m=1,2,πk=m=0

Thus, using the substitution x=12e2αv, it is easy to demonstrate that the set of functions

hkv=2απe2αv11/4Tk12e2αv, k=1,2,, (38)

with the first function defined as

h0v=2απe2αv11/4, (39)

form an orthonormal basis in the space L20,. Here, as previously, α is a positive time-scaling factor. The relaxation modulus basis functions ϕkt (9) are described by a useful recursive formula

ϕkt=2ϕk1tϕk2t4ϕk1t+2α, k=3,4,, (40)

and for k=0,1,2 are given by:

ϕ0t=12παΓ34Γt2α+14Γt2α+1, (41)
ϕ1t=αt2απαΓ34Γt2α+14Γt2α+2, (42)
ϕ2t=2α2+t26αt4α2παΓ34Γt2α+14Γt2α+3, (43)

where Γn is the Euler’s gamma function [60]. The proof is given in Appendix A.3, where two alternative formulas (A11) and (A12) for ϕ1t and ϕ2t, respectively, are also derived. A few first basis functions hkv are shown in Figure 3a,b for two values of the factor α; the corresponding functions ϕkt are plotted in Figure 3c,d. An earlier version of the model was presented in the paper [61].

Figure 3.

Figure 3

Basis functions hkv (38), (39) and ϕkt (40)–(43) of the Chebyshev model for two time-scaling factors: (a) hkv, α=0.1[s]; (b) hkv, α=1[s]; (c) ϕkt, α=0.1[s]; (d) ϕkt, α=1[s]; k=0,1,2,3,4.

3.6. Hermite Model

The Hermite functions defined as [52,62,63]

hkv=α2kk!π4eαv2/2Hkαv, k=0,1,, (44)

where α>0 is time-scaling factor and Hkx are the Hermite polynomials, which satisfy recursion formula [52]:

Hkx=2xHk1x2k1Hk2x, k=2,3,, (45)

with the initial

H0x=1,H1x=2x, (46)

constitute an orthonormal system in the space L2, [52,63]. The relaxation modulus basis functions ϕkt (9) are described by the recursive formula:

ϕkt=12k2k!απ4Hk10+k1kϕk2t2kαtϕk1t,k=2,3, (47)

and for k=0,1 are given by

ϕ0t=π42αet2/2α2erfct2α, (48)

and

ϕ1t=2απ42αtϕ0t, (49)

where the complementary error function erfcx is defined by [64]:

erfcx=2πxez2dz. (50)

The derivation of the above formulas is given in Appendix A.4. The initial values Hk0 of Hermite polynomials are specified by [52]:

H2k0=1k2k!/k!andH2k+10=0. (51)

A few first basis functions hkv are shown in Figure 4a,b for two factors α. The related functions ϕkt are plotted in Figure 4c,d. The function ϕ0t, and, in consequence of the recursive Formula (47), all functions ϕkt, depend on the exponential multiplier et2/2α2 rapidly moving towards infinity. The following asymptotic properties were proved in Appendix A.5 for any α>0:

limt ϕkt=0,k=0,1,2,, (52)
limt tϕkt=α2kαk!π4Hk0k=0,1,2,. (53)

However, in numerical computations, the limited values of ϕkt can be guaranteed only for ttupp, where tupp depends on the maximal real number accessible in the computing environment. For example, in Matlab, the largest finite floating-point number in IEEE double precision realmax=(2252)·210231.7977·10308. Thus, in view of (48), the range of numerical applicability of the Hermite model in the time domain, determined by the inequality

et2/2α2realmax,

is as follows

ttupp=2α2lnrealmax=α2lnrealmax37.6771α. (54)

Figure 4.

Figure 4

Basis functions hkv (44) and ϕkt (47)–(49) of the Hermite model for two time-scaling factors: (a) hkv, α=0.1[s]; (b) hkv, α=1[s]; (c) ϕkt, α=0.1[s]; (d) ϕkt, α=1[s]; k=0,1,2,3,4.

Smoothness

Since the basis functions hkv of the Hermite model presented above form an orthonormal basis of the space L2, of square integrable functions on ,, for Hermite model estimation (28) can be replaced by

HKMv22=0[HMMv]2dv[HMMv]2dv=k=0K1[gkλ]2=gKλ22.

Therefore, for the Hermite model, the algorithm may (but does not have to) provide a stronger limitation of the fluctuation in the determined relaxation spectrum HKMv than for the other models.

3.7. Choice of the Basis Functions

In the models proposed above, the parameter α>0 is the time-scaling factor. The following rule holds: the lower the parameter α is, the shorter the relaxation times are, i.e., the greater the relaxation frequencies are. The above is illustrated by Figure 1, Figure 2, Figure 3 and Figure 4. Through the optimal choice of the scaling factor, the best fit of the model to the experimental data can be achieved. However, in practice, a simple rough rule for choosing the factor α, based on the comparison of a few first functions from the sequence ϕkt for different values of α with the experimentally obtained function Gti, is quite enough. In the same manner, the number K of the series GKt (8) elements can be initially evaluated. This rough selection strategy of the model parameters was used in the examples presented below. Thus, the choice of the number K and the parameter α must be carried out a posteriori, after the preliminary experiment data analysis.

The ranges of applicability of the four classes of models described above in the relaxation times t domain and the relaxation frequencies v domain for different values of α are summarized in Table A1 in Appendix A.6. It was assumed that the range of applicability for times is determined by the value of time t, for which the first K=11 basis functions ϕkt no longer permanently exceed, i.e., for any θ>t, ε=0.5% of its maximum value. Specifically,

tapp=max0kK1mint>0t:ϕkθ0.005·ϕkmax for any θt, (55)

where

ϕkmax=maxt0ϕkt.

Similarly, the range of applicability for the relaxation frequencies was defined on the basis of the variability in the basis functions hkv. Here,

vapp=max0kK1minv>0v:hkϑ0.005·hkmax for any ϑv, (56)

with hkmax defined by

hkmax=maxv0hkv.

In view of the problems described above concerning the numerical determination of the basis functions only for ttupp, with tupp defined in (54), for the Hermite model, ε=0.0212 was assumed.

3.8. Example 1

Consider viscoelastic material of relaxation spectrum described by Gauss-like distribution [11]

Hv=vev22/3, (57)

The corresponding modified spectrum HMv (4) is:

HMv=Hvv=ev22/3. (58)

and, therefore, using (5), the ‘real’ relaxation modulus is

Gt=0ev22/3etvdv=3π2e34t22terfc3t423 (59)

where erfc is defined by (50). In the experiment, N=1000 sampling instants ti were generated with the constant period in the time interval T=0,32 seconds selected in view of the course of the modulus Gt (59). Additive measurement noises zti were selected independently by random choice with uniform distribution on the interval 0.005,0.005Pa, i.e., maximally 6.1% of the mean value of Gt in the interval T defined as the average value of the integral of Gt over T, which is equal to 0.0820Pa. The time-scaling factors α are selected by comparison for different values of α a few first functions ϕkt with the experiment results Gti. Only for the Chebyshev model, the rough selection of α required several attempts; for the remaining classes of models, it was enough to review the data from Table A1. The basis functions hkv and ϕkt were simulated in Matlab R2022a using special functions erfc, legendreP, chebyshevT, and hermiteH. For the singular value decomposition procedure, svd was applied. For K=6,8,9,10,11, and sometimes also for K=12, the regularization parameters λGCV were determined and are given in Table 1. Next, the vectors of optimal model parameters gKλ=gKλGCV (19) were computed and are given in Table A2 in Appendix B. Only for the Laguerre models, some elements of the vectors of optimal parameters are negative; for the remaining classes of models, all the parameter vectors gKλGCV are positive. The optimal modified spectra of relaxation frequencies HKMv (22) and the ‘real’ spectrum HMv (58) are plotted in Figure 5 for selected values of K, while in Figure 6, the spectra HKv (23) and Hv (57) are presented. The optimal models of the relaxation modulus GKt computed for gK=gKλGCV according to (8) are plotted in Figure A1 in Appendix B, where the measurements Gti are also marked. Since the optimal models GKt (8) fitted the data extremely well, as indicated especially by the mean-square model errors QNgKλGCV/N, which vary between 0.80·105Pa2 and 0.117·104Pa2, models GKt for different K practically coincide with the measurement points and with each other (see Figure A1). The norms gKλ2 and HKMv2, as the measures of the solution smoothness, and the identification index QNgKλGCV (10), being a measure of the errors of the relaxation modulus models, are also given in Table 1. For the ‘real’ modified spectrum HMv (58), the norm HMv2=1.4656Pa·s1/2. The distance between the ‘real’ spectrum HMv (58) and their models HKMv (22) was estimated by integral square error, defined as:

ERR2=HKMvHMv22=0HKMvHMv2dv, (60)

and is given in the last column of Table 1.

Table 1.

The parameters of the optimal models in Example 1: time-scale factors α, numbers of model summands K, regularization parameter λGCV and the approximation error indices: identification index QNgKλGCV, Equation (10), the errors ERR (60) of the relaxation spectrum models.

Model K α [s] λGCV [-] QNg-KλGCV [Pa2] g-Kλ2 [Pa·s1/2] H-KMv2 [Pa·s1/2] ERR
[Pa·s1/2]
Legendre model 6 1 0.03378 0.0088 1.5518 1.5518 2.0782
8 1 0.03239 0.0083 1.4833 1.4833 2.0643
9 1 0.03186 0.0084 1.4657 1.4657 2.0597
10 1 0.03139 0.0087 1.4537 1.4537 2.0559
11 1 0.03098 0.0089 1.4452 1.4452 2.0529
12 1 0.03062 0.0091 1.4391 1.4391 2.0505
Laguerre model 6 6 0.03259 0.0084 1.4929 1.4929 1.9914
8 6 0.02744 0.0093 1.4222 1.4222 2.0414
9 6 0.02560 0.0092 1.4240 1.4240 2.0435
10 6 0.0239 0.0089 1.4262 1.4262 2.0448
11 6 0.02256 0.0088 1.4253 1.4253 2.0444
Chebyshev model 6 1.5 0.03103 0.00992 1.5215 1.5215 2.1119
8 1.5 0.03077 0.00981 1.4677 1.4677 2.0811
9 1.5 0.03062 0.0101 1.4543 1.4543 2.0668
10 1.6 0.03165 0.0089 1.4875 1.4875 2.0864
11 1.6 0.03148 0.0091 1.47492 1.4749 2.0775
12 1.6 0.03126 0.009246 1.4651 1.4651 2.0713
Hermite model 6 1 0.00625 0.0117 1.6405 1.6029 2.0529
8 1 0.00588 0.0081 1.5077 1.4891 2.0674
9 1 0.00599 0.0080 1.4731 1.4639 2.0592
10 1 0.00599 0.0080 1.4611 1.4569 2.0607
11 1 0.00566 0.0080 1.4600 1.4561 2.0654

Figure 5.

Figure 5

Modified relaxation spectrum HMv(58) (solid red line) from Example 1 and the approximated models HKMv: (a) Legendre, (b) Laguerre, (c) Chebyshev, (d) Hermite.

Figure 6.

Figure 6

Figure 6

Relaxation spectrum Hv (57) (solid red line) from Example 1 and the approximated models HKv (23): (a) Legendre, (b) Laguerre, (c) Chebyshev, (d) Hermite.

3.9. Example 2

Consider the spectrum of relaxation times introduced by Baumgaertel, Schausberger and Winter [36,37],

Hτ=β1ττcρ1+β2ττcρ2eττmax,

which is known to be effective in describing polydisperse polymer melts [17,34], with the parameters [34]: β1=6.276·104Pa, β2=1.27·105Pa, τc=2.481s, τmax=2.564·104s, ρ1=0.25 and ρ2=0.5. The related spectra of relaxation frequencies Hv=H1v and HMv (4) are well posed for v>0. The modified spectrum is described by

HMv=Hvv=1vβ11vτcρ1+β21vτcρ2e1vτmax (61)

and depicted in Figure 7; the corresponding ‘real’ relaxation modulus Gt is defined by (5). In the experiment, N=1000 time instants ti were sampled according to the square rule ti=t(i1)2+60 s, with parameter t=T/(N1)2, where T=107s, in the time interval T=0,T is selected in view of the course of the modulus Gt. Due to the numerical problems described above related to determining the basis functions ϕkt, Equations (47)–(49), for the Hermite model the experiment was simulated in a shorter time interval T=106s using the same sampling formula. Additive measurement noises zti were selected independently by random choice with uniform distribution on the interval 0.0005,0.0005MPa. Here, for the selection of parameter α, which guarantees a satisfactory accuracy of the modulus approximation, several or even a dozen or so attempts were necessary. This means that it will be advisable to further extend the algorithm by the level of optimal selection of the time-scaling factors. For K=6,8,10,12,14, and also for K=16, the regularization parameters λGCV were determined and are given in Table 2. Only for the Laguerre and Legendre models, the accuracy of the modulus approximation measured by QNgKλGCV, Equation (10), is comparable to that obtained in Example 1. Thus, the vectors of optimal model parameters gKλ=gKλGCV (19) for Legendre and Laguerre models are given in Table A3 in Appendix B; for the remaining models, they were omitted. The optimal spectra HKMv (22) and the ‘real’ spectrum HMv (58) are plotted in Figure 7 for Laguerre and Legendre models; a logarithmic scale is used for the frequencies v and a linear scale for the spectrum. For Legendre and Laguerre models, the optimal models GKt computed for gK=gKλGCV are plotted in Figure 8, where the measurements Gti are also marked and logarithmic scale is applied for time axis. The norms gKλ2, HKMv2 and the identification index QNgKλGCV are given in Table 2. For the ‘real’ modified spectrum HMv (58), the norm HMv2=56.9809MPa·s1/2. The distance between the ‘real’ spectrum HMv (58) and its models HKMv (22) was estimated by the integral square errors ERR, defined in (60) and given in Table 2, only for Legendre and Laguerre models. For the Chebyshev and Hermite models, a satisfactory quality of approximation was not obtained, despite the research carried out for a wide range of α values. The best found α and related identification indices QNgKλGCV are given in Table 2, but other indices are omitted. The optimal Legendre and Laguerre models GKt (8) were well fitted to the experimental data, see Figure 8a,b, as also indicated by the mean-square model errors QNgKλGCV/N, which vary in a range from 5.27·107MPa2 to 3.26·106MPa2. For the Chebyshev and Hermite models, even increasing the number of model summands K does not improve the poor approximation of the relaxation modulus data (see Figure 8c,d and the values of QNgKλGCV from Table 2). In particular, the course of the GKt model from Figure 8c shows that it is not too few components of the model but the properties of the Chebyshev model that make it ineffective for approximating the relaxation spectrum HMv (61).

Figure 7.

Figure 7

Modified relaxation spectrum HMv (61) (solid black line) from Example 2 and the approximated models HKMv: (a,b) Legendre models, (c,d) Laguerre models.

Table 2.

The parameters of the optimal models in Example 2: time-scale factors α, numbers of model summands K, regularization parameter λGCV, optimal model parameters gKλGCV, approximation error indices: identification index QNgKλGCV, Equation (10), the errors ERR (60) of the relaxation spectrum models.

Model K α [s] λGCV [-] QNgKλGCV [MPa2] gKλ2 [MPa·s1/2] HKMv2 [MPa·s1/2] ERR
MPa·s1/2
Legendre model 6 950 2.72 × 10−6 0.003261 62.717 62.717 80.604
8 620 4.968 × 10−6 0.00135 58.9351 58.9351 79.7444
10 500 6.372 × 10−6 0.0007138 57.4422 57.4422 79.7170
12 400 7.754 × 10−6 0.000557 56.8251 56.8251 79.7917
14 300 9.226 × 10−6 0.000575 56.6390 56.6390 79.8353
Laguerre model 8 8500 2.524 × 10−6 0.002722 63.0250 63.0250 82.4674
10 8000 3.668 × 10−6 0.001494 59.7092 59.7092 81.0828
12 7500 4.840 × 10−6 0.0009379 58.3032 58.3032 80.6344
14 8000 5.294 × 10−6 0.000711 57.3277 57.3277 80.3069
16 6000 7.498 × 10−6 0.000699 57.39886 57.39886 80.4441
18 6500 7.686 × 10−6 0.000527 56.7967 56.7967 80.1877
Chebyshev model 6 31,000 6.0 × 10−7 0.8259 289.32 289.32
8 31,500 6.0 × 10−7 0.7115 346.30 346.30
10 32,000 5.0 × 10−7 0.6504 376.6564 376.66
12 33,000 5.0 × 10−7 0.6344 379.0402 379.04
14 33,000 5.0 × 10−7 0.6029 375.3072 375.31
16 34,500 5.0 × 10−7 0.6136 382.0129 382.01
Hermite model 6 27,000 1.8 × 10−6 3.0971 390.24 387.13
8 26,800 1.6 × 10−6 2.1911 452.31 445.67
10 26,750 1.6 × 10−6 1.9010 440.42 436.32
12 26,700 1.8 × 10−6 1.8148 389.93 378.76
14 26,600 0.999 × 10−6 2.9235 472.33 467.32
16 26,600 0.994 × 10−6 2.7767 467.45 461.56

Figure 8.

Figure 8

Figure 8

The measurements of the relaxation modulus Gt corresponding to the relaxation spectrum HMv (61) (red points) from Example 2 and the optimal approximated models GKt (8): (a) Legendre, (b) Laguerre, (c) Chebyshev, (d) Hermite.

3.10. Remarks

The example orthogonal basis functions for relaxation spectrum models have been assumed as the products of exponentials and Legendre, Laguerre, Chebyshev and Hermite polynomials. For Legendre and Laguerre models, the basis functions ϕkt of the relaxation modulus are rational functions; for the Chebyshev model, they are determined by the quotient of the Euler’s gamma functions, which is a generalization of the factorial of a non-negative integer for the no-integer argument, while, for the Hermite model, these functions are based on the complementary error function. From Figure 1c,d and Figure 2c,d, it is evident that the basis functions ϕkt for the relaxation modulus of Legendre and Laguerre models are in good agreement with the real relaxation modulus obtained in the experiment. However, Figure 3c,d and Figure 4c,d show that, for the Chebyshev and Hermite models, so good agreement is not achieved for individual functions ϕkt. Hence, a significantly worse fit of the Chebyshev and Hermite models to the measurement data for the spectrum from Example 2, which has a much wider range of relaxation frequencies, results.

Both examples show that, generally, increasing the number of model summands improves the model quality, provided that the assumed series can provide a good approximation of the relaxation modulus. If a given series of special functions is not suitable for approximation of the relaxation modulus for a given material, then, as shown by the research conducted for the Chebyshev and Hermite models in the second example, increasing the number of series components does not improve the quality of the model.

4. Conclusions

In this paper, a class of algorithms for the relaxation spectrum identification, which combines the technique of an expansion of a function into a series in an orthonormal basis with the least-squares regularized identification, was derived. It was demonstrated that due to the choice of an appropriate special functions (Legendre, Laguerre, Chebyshev, Hermite) as the basis functions for the relaxation spectrum model, the basis functions of the relaxation modulus model are given by compact analytical or recursive formulas. Due to the choice of orthonormal basis, the smoothness of the vector of the optimal model parameters implies equivalent smoothness of the fluctuations in the model of the relaxation spectrum.

The proposed approach based on the expansion of the relaxation spectrum into a function series can be applied for arbitrary basis functions. The proven convergence and noise robustness properties of the optimal models will be retained, but the smoothing of the spectrum model will require separate analysis. As part of further research, the algorithm can be extended with a superior level of optimal selection of the time-scaling factor, so as to obtain a better fit to the measurement data. The presented scheme of the relaxation spectrum identification can be easily modified for the retardation spectrum recovery from the creep compliance measurements obtained in the standard creep test.

Nomenclature

Gt linear relaxation modulus, Equation (1), Pa
Gti relaxation modulus obtained in stress relaxation test, i=1,,N , Pa
Gti noise corrupted measurement of the modulus Gti, i=1,,N, Pa
GKt relaxation modulus model corresponding to HKMv, Equation (8), Pa
GN N dimensional vector of the measurements Gti, Equation (11)
GN N dimensional vector of noise-free measurements Gti
gk parameters of the model HKMv, k=0,1,K1, Equation (7), Pa·s1/2
gK vector of parameters gk of the models HKMv (7) and GKt (8)
gKλ solution of regularized task (14) given for regularization parameter λ by (15) or (19)
g~Kλ solution of regularized task (14) for ideal measurements, Equation (25)
gKλGCV vector of optimal model parameters determined by GCV method
gKN , g~KN normal solutions of least-squares problem (13) for noise and noise-free measurements of the relaxation modulus
Hv continuous spectrum of relaxation frequencies v, Equation (3), Pa
HMv modified relaxation spectrum, Equation (4), Pa·s
HKMv model of the modified relaxation spectrum HMv , Equation (7), Pa·s
HKMv model of modified spectrum HMv for regularized parameter gKλ, Equation (22), Pa·s
HKv model of spectrum Hv for regularized parameter gKλ, Equation (23), Pa
H~KMv model of HMv obtained for ideal measurements, Equation (24), Pa·s
HKNv , H~KNv models of HMv obtained for the normal solutions gKN, g~KN described by Equations (26) and (27), Pa
hkv basis functions of the model HKMv, k=0,1,K1, Equation (7), Pa·s
IK,K K dimensional identity matrix
K number of model HKMv summands, Equation (7), –
N number of measurements in the stress relaxation test, –
QNgK square identification index defined by (10), Pa2
r number of non-zero singular values of ΦN,K, –
ti sampling instants used in the stress relaxation test, i=1,,N , s
T time interval from which the sampling instants ti were generated
V , U orthogonal matrices of singular value decomposition of ΦN,K, Equation (18)
VGCVλ the GCV functional defined by (16) and expressed by (21)
Y N dimensional vector Y=UTGN
zti additive noise of relaxation modulus measurement, i=1,,N , Pa
zN N dimensional vector of measurement noises zti, i=1,,N, Pa
α time-scaling factor of the basis functions hkv, s
λ regularization parameter introduced in the optimization task (14), –
λGCV optimal regularization parameter determined by GCV method, Equation (17)
Λλ K×N diagonal matrix defined by (20)
σi the non-zero singular values of ΦN,K, i=1,2,,r, Equation (18)
Σ N×K diagonal matrix of singular values σ1,,σr of ΦN,K, Equation (18)
ΦN,K N×K matrix defined by (11) basis for least-squares task (13)
ΦN,K the MoorePenrose pseudoinverse of matrix ΦN,K
ϕkt basis functions defined by (9) of the model GKt , Equation (8), s1/2
ERR integral square error of HMv approximation, Equation (60), Pa·s1/2

Appendix A

Appendix A.1. Derivation of the Formula (32)

Mathematical induction will be used. In the proof, the following recurrence representation of the Legendre polynomials [52,53]:

Pkx=2k1kxPk1xk1kPk2x,k=2,3,, (A1)

is used, where the two first polynomials are as follows

P0x=1, (A2)

and

P1x=x. (A3)

In the base case of the proof by induction, for k=0 and k=1, Formula (32) follows immediately from definition (9) and properties (A2), (A3). For k=0, by (9) (31) and (A2) the basis function ϕkt:

ϕ0t=02αeαvetvdv=2α1t+α=2αi=002i+1α+t,

while for k=1, applying (A3), we obtain:

ϕ1t=6αt+α26αt+3α=2α2+1i=002i+1αti=012i+1α+t.

For k2, in the induction step, combining definitions (9) and (31), we have:

ϕkt=02α2k+1eαvPk12e2αvetvdv, (A4)

where, substituting Pkx given by recurrence Equation (A1) and having in mind (A4), we obtain:

ϕkt=2k1k2α2k+12α2k1ϕk1t22k1k2α2k+12α2k1ϕk1t+2αk1k2α2k+12α2k3ϕk2t.

Via the induction hypothesis, ϕk2t and ϕk1t are given by (32); thus, the basis function ϕkt can be rewritten as:

ϕkt=2α2k+1k2k1i=0k22i+1αti=0k12i+1α+t22k1i=0k22i+1αt2αi=0k12i+1α+t+2αk1i=0k32i+1αti=0k22i+1α+t. (A5)

Since

i=0k22i+1αt2αi=0k12i+1α+t+2α=1t+αi=0k32i+1αti=1k2i+1α+t,

function ϕkt (A5) is given by

ϕkt=2α2k+1k2k1i=0k22i+1αti=0k12i+1α+t+22k1t+αi=0k32i+1αti=1k2i+1α+tk1i=0k32i+1αti=0k22i+1α+t,

and, after algebraic manipulations, can be expressed as:

ϕkt=2α2k+1i=0k22i+1αti=0k12i+1α+t2k1αt2k+1α+t.

Hence, the direct Formula (32) follows immediately.

Appendix A.2. Derivation of Formula (35)

To derive the basis function ϕkt (35), we start with the alternative formula of the Laguerre polynomials Lkv (33) given by [65]:

Lkv=j=0kkkjαvjj!, k=0,1,2, (A6)

An easy consequence of (A6) is

hkv=αj=0kkkjαjj!vjeαv/2,k=0,1,2,

where, using the well-known formula for the Laplace transform

Lvjeαv/2=j!t+α2j+1, j=0,1,2,,

the basis functions ϕkt defined by (9) are given by

ϕkt=αj=0kkkjαjt+α2j+1

for k=0,1,2,. Hence, by binomial theorem, we have

ϕkt=αt+α2k+1j=0kkkjαjt+α2kj=αt+α2k+1tα2k,

where Formula (35) follows. This formula is known in [55] and coincides with that given in [66].

Appendix A.3. Derivation of Formulas (40)–(43)

Suppose α>0. Let us first derive Formulas (41), (42). For k=0 by (39) and (9), we have

ϕ0t=0h0vetvdv=2απ0e2αv11/4etvdv, (A7)

which, after substitution e2αv=u, is equivalent to

ϕ0t=12απ011u114ut2α1du=12απ011u14ut2α34du,

whereby, due to the well-known Euler–Poisson integral [60]:

011un1um1du=ΓnΓmΓn+m,

we immediately obtain (41).

For k=1, by (38) and (9) for the polynomial T1x=x, the integral ϕ1t is given by

ϕ1t=2απ0e2αv11/412e2αvetvdv, (A8)

and, keeping in mind the last expression in (A7), can be rewritten as

ϕ1t=2ϕ0t22ϕ0t+2α.

Then, from Equation (41), we obtain

ϕ1t=1παΓ34Γt2α+14Γt2α+121παΓ34Γt+2α2α+14Γt+2α2α+1=Γ34παΓt2α+14Γt2α+12Γt2α+14+1Γt2α+2, (A9)

Using the well-known property of the Gamma function [60]:

Γx+1=xΓx, (A10)

Equation (A9), after algebraic manipulations, can be expressed as

ϕ1t=αtπαt+2αΓ34Γt2α+14Γt2α+1, (A11)

or in equivalent form as Formula (42).

For k=2, by (36), (37) and (9), we have

ϕ2t=2απ0e2αv11/4212e2αvT112e2αv1etvdv,

which, keeping in mind (A8) and (A7), can be rewritten as follows

ϕ2t=2ϕ1t4ϕ1t+2α2ϕ0t,

and where, by (41), (A11) and (A10), after algebraic manipulations, we obtain

ϕ2t=2α2+t26αtπαt+2αt+4αΓ34Γt2α+14Γt2α+1, (A12)

or, equivalently, Equation (43).

Now, it remains to show that (40) holds for any k3. Recalling the identity (36) and using (38) for k3, we have

hkv=212e2αvhk1vhk2v=2hk1vhk2v4e2αvhk1v

where, bearing in mind definition (9), we obtain (40), which concludes the proof.

Appendix A.4. Derivation of Formulas (47)–(49)

Let us first derive the formulas (48) and (49). Since H0x=1, for k=0, by applying in the integral ϕ0t the substitution u=αv+t/α/2, we have:

ϕ0t=απ40eαv2/2etvdv=2απ4et2/2α2t/2αeu2du, (A13)

where, by definition (50) of the complementary error function, Formula (48) is immediately obtained.

For k=1, the polynomial H1αv=2αv; then, the integral ϕ1t is as follows

ϕ1t=2ααπ40veαv2/2etvdv.

Comparing the first integral in Formula (A13) and the integral ϕ1t, it is easy to notice that

ϕ1t=2αϕ0t. (A14)

Since by (50), we have

derfct/2αdt=2απet2/2α2, (A15)

Equations (A14) and (48) yield

ϕ1t=π4αtαet2/2α2erfct2α2πet2/2α2et2/2α2,

where the next formula follows

ϕ1t=π4α2πtαet2/2α2erfct2α, (A16)

and, in view of (48), Formula (49) results.

The proof of the recursive Equation (47) is based on (45), which, by definitions (44) and (9) for k2, yields

ϕkt=2αα2kk!π40veαv2/2Hk1αvetvdv2k1α2kk!π40eαv2/2Hk2αvetvdv. (A17)

By the integration by parts, the first integral in (A17) is as follows

Ikt=0veαv2/2Hk1αvetvdv=1α2eαv2/2Hk1αvetv0+1α0eαv2/2Hk1αvetvdvtα20eαv2/2Hk1αvetvdv,

which, bearing in mind that

limveαv2/2Hk1αvetv=0,

and taking into account the known property of Hermite polynomials [52]:

Hkx=2kHk1x

and definitional Formulas (44), (9), can be expressed as

Ikt=1α2Hk10+2k12k2k2!π4ααϕk2t2k1k1!π4α2αtϕk1t,

which, combined with (A17), finally yields

ϕkt=2αα2kk!π41α2Hk10+2k12k2k2!π4ααϕk2t2k1k1!π4α2αtϕk1t2k1α2kk!π42k2k2!π4αϕk2t,

where, after algebraic manipulations, Formula (47) follows.

Appendix A.5. Proof of Formulas (52) and (53)

Mathematical induction will be used again. In the base step, we prove (52) and (53), successively, for k=0, k=1 and k=2. By (48), we have

ϕ0t=π42αet2/2α2erfct2α=π42αerfct/2αet2/2α2, (A18)

where both the numerator and denominator of the expression on the right hand side of (A18) become zero when t. Thus, using the L’Hospital’s rule, and taking into account (A15), we obtain

lim tϕ0t=π42αlimt2απet2/2α2tα2et2/2α2=π42αlimt2απt=0.

Analogically,

limttϕ0t=π42αlimterfct/2α1tet2/2α2=π42αlimt2απ1α2+1t2=π42α2πα=απ4, (A19)

where, since by (51) the initial condition H00=1, we have

limt tφ0t=απ4H00.

Now, by (A19) and (49), we have

lim tϕ1t=2απ42ααπ4=0.

In order to prove (53) for k=1, it is enough to prove that tϕ1t0, when t, since by (51), the initial condition H10=0. On the basis of (A16), we have

lim ttϕ1t=π4αlimt2π1tet2/2α21αerfct/2αet2/2α2t2. (A20)

Thus, using the L’Hospital’s rule and keeping in mind (A15), we obtain

limttϕ1t=2απ4limt11α2t+2t=0,

where, since based on (47) and (51), we have

ϕ2t=12ϕ0t1αtϕ1t, (A21)

we immediately obtain lim tϕ2t=0. By (A21) and (A19), the limit

lim ttϕ2t=limt12tϕ0t1αt2ϕ1t=α2π41αlimtt2ϕ1t. (A22)

Based on (A16)

lim tt2ϕ1t=π4αlimt2π1tet2/2α21αerfct2αet2/2α2t3.

Hence, similarly as in the case of (A20), using the L’Hospital’s rule and including Formula (A15), after simple transformations, we obtain

lim tt2ϕ1t=π4αlimt2πt21α2t43t2t6=π4αlimt2π1α23t2=π4α2π1α2=π4α2π1α2=α22απ4. (A23)

Next, substituting (A23) into (A22) yields

lim ttϕ2t=α2π41αα22απ4=α2απ4=2α22α2!π4,

and bearing in mind that by (51) H20=2, Equation (53) is proved for k=2.

Now, in the induction step, let us assume that Formulas (52) and (53) hold for k1, k2 and k3, where k3. We prove that they hold also for k. Consider the recursive Formula (47). By the induction hypothesis, including (52) for k2 and (53) for k1, from (47), we immediately have

lim tϕkt=12k2k!απ4Hk102kαα2k1αk1!π4Hk10=0.

It is only necessary to prove the correctness of Formula (53) for k. The limit limttϕkt can be rewritten equivalently as follows

lim ttϕkt=limt1tet2/2α2ϕktet2/2α2t2, (A24)

where both the numerator and denominator of the expression on the right expression of (A24) become zero, when t. Using the L’Hospital’s rule, we have

lim ttϕkt=limt1t+1α2tϕktϕkt1α2+2t2. (A25)

On the basis of (47):

ϕkt=k1kϕk2t2kαϕk1t2kαtϕk1t (A26)

To determine the derivative ϕk2t, note that based on (9) and (44), we have

ϕk2t=α2k2k2!π40veαv2/2Hk2αvetvdv. (A27)

Since by (45)

Hk1αv=2αvHk2αv2k2Hk3αv (A28)

Equation (A27) can be expreseed as

ϕk2t=α2α2k2k2!π40eαv2/2Hk1αvetvdv+2k20eαv2/2Hk3αvetvdv

where, in view of (44) and (9), after algebraic manipulations, the next formula results

ϕk2t=k12αϕk1t+k22αϕk3t. (A29)

Similarly, for k1, we have

ϕk1t=k2αϕkt+k12αϕk2t. (A30)

Now, combining (A26), (A30) and (A29), after algebraic manipulations, yields

ϕkt=k12kαϕk1tk1k22kαϕk3t2kαϕk1t+1α2tϕkt+k1kα2tϕk2t. (A31)

Substituting the derivative ϕkt (A31) into (A25), and bearing in mind that ϕk3t, ϕk1t and ϕkt tend to zero as t, we obtain

limttϕkt=limtk1kα2tϕk2t1α2+2t2.

By induction hypothesis, (53) holds for k2. Thus, we have

lim ttϕkt=k1kα21α2α2k2αk2!π4Hk20=αk12k2αk!π4Hk20, (A32)

where, since by (45), the initial conditions are related by Hk0=2k1Hk20, the correctness of Formula (53) for k immediately follows from (A32).

Appendix A.6. Applicability of the Models for Various Time-Scale Parameters

Table A1.

Ranges of the applicability of the models for various time-scale parameters.

Model Time-Scale Factor α [s] Range 1 of Relaxation Frequency v [s1] Range 1 of Time t [s]
Legendre model 0.001 5301 3.615
0.01 530 36.15
0.1 53 361.5
1 5.3 3614.6
10 0.53 36,144.5
100 0.05301 361,445
1000 0.005301 3,614,450
10,000 0.0005301 36,144,500
Laguerre model 0.01 5256 0.995
0.1 525.36 9.95
1 52.5359 99.5
10 5.2536 995
100 0.528 9950
1000 0.0526 99,500
10,000 0.00526 995,000
100,000 0.000526 9,950,000
Chebyshev model 0.001 10,020 0.34125
0.01 1002 3.4125
0.1 100.2 34.125
1 10.02 341.25
10 1.002 3412.5
100 0.1002 34,125
1000 0.01002 341,250
10,000 0.001002 3,412,500
100,000 0.0001002 34,125,000
Hermite model 0.001 4797.5 0.037639
0.01 479.75 0.376394
0.1 47.975 3.7639
1 4.7975 37.63944
10 0.47975 376.3944
100 0.047975 3763.944
1000 0.0047975 37,639.443
10,000 0.00047975 376,394.436
100,000 0.000047975 3,763,944.359

1 Only the upper bounds tapp (55) and vapp (56) of the respective intervals 0,tapp and 0,vapp are given.

Appendix B

Table A2.

Optimal parameters gKλGCV of the relaxation spectrum models from Example 1; the elements of the vectors gKλGCV are expressed in Pa·s1/2.

Model gKλGCV
K=6 K=8 K=9 K=10 K=11 K=12
Legendre model 0.8222 0.8225 0.8229 0.8232 0.8235 0.8238
0.7699 0.7695 0.7705 0.7714 0.7723 0.7731
0.6138 0.5979 0.5963 0.5959 0.5963 0.5970
0.5299 0.4503 0.4347 0.4263 0.4215 0.4188
0.5448 0.3564 0.3215 0.3021 0.2889 0.2801
0.4300 0.3190 0.2887 0.2647 0.2490 0.2378
0.2807 0.2442 0.2227 0.2074 0.1966
0.2520 0.2185 0.1945 0.1781 0.1663
0.1973 0.1742 0.1578 0.1462
0.1562 0.1398 0.1280
0.1264 0.1146
0.1030
Laguerre model 0.3277 0.3265 0.3265 0.3266 0.3267
−0.5817 −0.5814 −0.5804 −0.5795 −0.5805
0.7416 0.7908 0.7897 0.7891 0.7856
−0.6558 −0.7164 −0.7193 −0.7163 −0.7109
0.7138 0.5043 0.5076 0.5013 0.5015
−0.5418 −0.36570 −0.3658 −0.3777 −0.3717
0.2063 0.2083 0.2121 0.2282
−0.0935 −0.0918 −0.1178 −0.1254
−0.0054 0.0122 0.0446
0.0517 0.0269
−0.0712
Chebyshev model 0.8227 0.8265 0.8316 0.7988 0.8023 0.7991
0.8350 0.8226 0.8139 0.7962 0.7905 0.7944
0.5313 0.5312 0.5382 0.5234 0.5287 0.5248
0.5168 0.4545 0.4362 0.4502 0.4397 0.4416
0.4625 0.3392 0.3233 0.3549 0.3426 0.3294
0.4209 0.2912 0.2623 0.3195 0.2992 0.2828
0.2444 0.2138 0.2768 0.2542 0.2403
0.2148 0.1899 0.2509 0.2317 0.2146
0.1603 0.2270 0.2047 0.1899
0.2112 0.1917 0.1752
0.1734 0.1585
0.1487
Hermite model 0.5504 0.6391 0.5909 0.6044 0.6051
0.7337 0.7670 0.8046 0.8004 0.8009
0.7965 0.6140 0.6703 0.6826 0.6853
0.6881 0.5241 0.4749 0.5005 0.5075
0.4916 0.5137 0.4089 0.3846 0.3885
0.7075 0.4046 0.3715 0.2971 0.2875
0.2833 0.2769 0.2366 0.2181
0.3419 0.2239 0.2318 0.2207
0.2576 0.2218 0.2178
0.1315 0.1381
0.0236

Figure A1.

Figure A1

The measurements of the relaxation modulus Gt (59) (red points) from Example 1 and the approximated models GKt (8): (a) Legendre, (b) Laguerre, (c) Chebyshev, (d) Hermite.

Table A3.

Optimal parameters gKλGCV of the relaxation spectrum models from Example 2; the elements of the vectors gKλGCV are expressed in MPa·s1/2.

Model gKλGCV
K=6 K=8 K=10 K=12 K=14 K=16
Legendre model 43.80739 39.6666 37.4618 35.1566 32.2127
−29.51378 −31.6618 −32.1025 −32.1325 −31.6435
8.10418 15.9237 19.1145 21.6308 23.8515
−2.20279 −6.1369 −8.1839 −10.9840 −14.3753
−3.819709 0.0198 1.5667 3.9792 7.5618
32.529089 10.4212 5.8515 2.9019 −0.9375
−10.2725 −5.9794 −4.2456 −1.7853
19.8011 11.2957 8.1554 5.3793
−8.6168 −6.9530 −5.7798
12.6105 9.3899 7.7029
−7.0433 −6.9429
8.7841 7.9943
−6.6946
7.3213
K=8 K=10 K=12 K=14 K=16 K=18
Laguerre model 53.0965 53.0609 53.0093 53.0664 52.5153 52.7585
−1.4762 0.2502 1.8283 0.4988 6.3911 4.8727
6.9102 6.5809 6.4712 5.2459 8.0061 6.8685
−0.2900 −3.7059 −4.5349 −6.4194 −2.4884 −4.0192
−10.8544 −6.2248 −3.8522 −3.7269 −0.8334 −1.5307
−10.7846 −6.6303 −6.0659 −6.4034 −4.9840 −5.6574
−8.9629 −8.7827 −7.0219 −6.1421 −4.2199 −4.2360
−28.0790 −12.2858 −8.2334 −6.5918 −6.1999 −6.0076
−6.1049 −6.7165 −5.7225 −5.4599 −4.9298
−18.5624 −10.7187 −7.3355 −6.9589 −6.0759
−4.8687 −4.0801 −5.6896 −4.7692
−13.0786 −8.3498 −7.4589 −6.0712
−2.0153 −5.4359 −4.2447
−9.4015 −7.7715 −6.0352
−4.9662 −3.5890
−7.9368 −5.9731
−2.9159
−5.8840

Appendix C

Mathematical Terminology and Special Functions

erfcx complementary error function, Equation (50)
Hkx Hermite polynomials, k=0,1,2,, Equations (45) and (46)
Lkx Laguerre polynomials, k=0,1,2,, Equation (33)
Pkx Legendre polynomials, k=0,1,2,, Equations (A1)–(A3)
Tkx Chebyshev polynomials of the first kind, k=0,1,2,, Equations (36) and (37)
L20, space of real-valued square-integrable functions on the interval 0,
Lfv Laplace transform of the function fv
Γn Euler’s gamma function
·2 square norm in the real Euclidean space RN and in L20,
minxfx find the value of x, which minimizes the function fx

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The author declares no conflict of interest.

Funding Statement

This research received no external funding.

Footnotes

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.

References

  • 1.Anderssen R.S., Davies A.R., de Hoog F.R., Loy R.J. Derivative based algorithms for continuous relaxation spectrum recovery. J. Non-Newton. Fluid Mech. 2015;222:132–140. doi: 10.1016/j.jnnfm.2014.10.004. [DOI] [Google Scholar]
  • 2.Chang F.-L., Bin H., Huang W.-T., Chen L., Yin X.-C., Cao X.-W., He G.-J. Improvement of rheology and mechanical properties of PLA/PBS blends by in-situ UV-induced reactive extrusion. Polymer. 2022;259:125336. doi: 10.1016/j.polymer.2022.125336. [DOI] [Google Scholar]
  • 3.Ankiewicz S., Orbey N., Watanabe H., Lentzakis H., Dealy J. On the use of continuous relaxation spectra to characterize model polymers. J. Rheol. 2016;60:1115–1120. doi: 10.1122/1.4960334. [DOI] [Google Scholar]
  • 4.Dealy J.M., Read D.J., Larson R.G. Structure and Rheology of Molten Polymers. 2nd ed. Carl Hanser Verlag GmbH & Co. KG; Munich, Germany: 2018. pp. 105–145. [DOI] [Google Scholar]
  • 5.Pogreb R., Loew R., Bormashenko E., Whyman G., Multanen V., Shulzinger E., Abramovich A., Rozban A., Shulzinger A., Zussman E., et al. Relaxation spectra of polymers and phenomena of electrical and hydrophobic recovery: Interplay between bulk and surface properties of polymers. J. Polym. Sci. Part B: Polym. Phys. 2017;55:198–205. doi: 10.1002/polb.24260. [DOI] [Google Scholar]
  • 6.Owens R.G., Phillips T.N. Computational Rheology. World Scientific; London, UK: 2002. [DOI] [Google Scholar]
  • 7.Ferry J.D. Viscoelastic Properties of Polymers. 3rd ed. John Wiley & Sons; New York, NY, USA: 1980. [Google Scholar]
  • 8.Mead D.W. Numerical interconversion of linear viscoelastic material functions. J. Rheol. 1994;38:1769–1795. doi: 10.1122/1.550526. [DOI] [Google Scholar]
  • 9.Hajikarimi P., Moghadas Nejad F. Chapter 6-Interconversion of constitutive viscoelastic functions. In: Hajikarimi P., Moghadas Nejad F., editors. Applications of Viscoelasticity. Elsevier; Amsterdam, The Netherlands: 2021. pp. 107–139. [DOI] [Google Scholar]
  • 10.Baumgaertel M., Winter H.H. Determination of discrete relaxation and retardation time spectra from dynamic mechanical data. Rheol. Acta. 1989;28:511–519. doi: 10.1007/BF01332922. [DOI] [Google Scholar]
  • 11.Honerkamp J., Weese J. Determination of the relaxation spectrum by a regularization method. Macromolecules. 1989;22:4372–4377. doi: 10.1021/ma00201a036. [DOI] [Google Scholar]
  • 12.Honerkamp J., Weese J. A nonlinear regularization method for the calculation of relaxation spectra. Rheol. Acta. 1993;32:65–73. doi: 10.1007/BF00396678. [DOI] [Google Scholar]
  • 13.Malkin A.Y. The use of a continuous relaxation spectrum for describing the viscoelastic properties of polymers. Polym. Sci. Ser. A. 2006;48:39–45. doi: 10.1134/S0965545X06010068. [DOI] [Google Scholar]
  • 14.Malkin A.Y., Vasilyev G.B., Andrianov A.V. On continuous relaxation spectrum. Method of calculation. Polym. Sci. Ser. A. 2010;52:1137–1141. doi: 10.1134/S0965545X10110076. [DOI] [Google Scholar]
  • 15.Stadler F.J., Bailly C. A new method for the calculation of continuous relaxation spectra from dynamic-mechanical data. Rheol. Acta. 2009;48:33–49. doi: 10.1007/s00397-008-0303-2. [DOI] [Google Scholar]
  • 16.Davies A.R., Goulding N.J. Wavelet regularization and the continuous relaxation spectrum. J. Non-Newton. Fluid Mech. 2012;189–190:19–30. doi: 10.1016/j.jnnfm.2012.09.002. [DOI] [Google Scholar]
  • 17.Cho K.S. Power series approximations of dynamic moduli and relaxation spectrum. J. Rheol. 2013;57:679–697. doi: 10.1122/1.4789787. [DOI] [Google Scholar]
  • 18.Takeh A., Shanbhag S. A Computer Program to Extract the Continuous and Discrete Relaxation Spectra from Dynamic Viscoelastic Measurements. Appl. Rheol. 2013;23:24628. doi: 10.3933/applrheol-23-24628. [DOI] [Google Scholar]
  • 19.Liu H., Luo R., Lv H. Establishing continuous relaxation spectrum based on complex modulus tests to construct relaxation modulus master curves in compliance with linear viscoelastic theory. Constr. Build. Mater. 2018;165:372–384. doi: 10.1016/j.conbuildmat.2017.12.204. [DOI] [Google Scholar]
  • 20.Luo L., Xi R., Ma Q., Tu C., Ibrahim Shah Y. An improved method to establish continuous relaxation spectrum of asphalt materials. Constr. Build. Mater. 2022;354:129182. doi: 10.1016/j.conbuildmat.2022.129182. [DOI] [Google Scholar]
  • 21.Poudel S., Shanbhag S. Efficient test to evaluate the consistency of elastic and viscous moduli with Kramers–Kronig relations. Korea-Aust. Rheol. J. 2022;10:2202. doi: 10.1007/s13367-022-00041-y. [DOI] [Google Scholar]
  • 22.Alfrey T., Doty P. The Methods of Specifying the Properties of Viscoelastic Materials. J. Appl. Phys. 1945;16:700–713. doi: 10.1063/1.1707524. [DOI] [Google Scholar]
  • 23.Bažant Z.P., Yunping X. Continuous Retardation Spectrum for Solidification Theory of Concrete Creep. J. Eng. Mech. 1995;121:281–288. doi: 10.1061/(ASCE)0733-9399(1995)121:2(281). [DOI] [Google Scholar]
  • 24.ter Haar D. An easy approximate method of determining the relaxation spectrum of a viscoelastic materials. J. Polym. Sci. 1951;6:247–250. doi: 10.1002/pol.1951.120060211. [DOI] [Google Scholar]
  • 25.Goangseup Z., Bažant Z.P. Continuous Relaxation Spectrum for Concrete Creep and its Incorporation into Microplane Model M4. J. Eng. Mech. 2002;128:1331–1336. doi: 10.1061/(ASCE)0733-9399(2002)128:12(1331). [DOI] [Google Scholar]
  • 26.Kurenuma Y., Nakano T. Analysis of stress relaxation on the basis of isolated relaxation spectrum for wet wood. J. Mater. Sci. 2012;47:4673–4679. doi: 10.1007/s10853-012-6335-0. [DOI] [Google Scholar]
  • 27.Stankiewicz A. Identification of the relaxation spectrum of viscoelastic materials under strong noise measurement data. Part I. Problem and example. MOTROL. Comm. Mot. Energetics Agric. 2013;15:119–126. [Google Scholar]
  • 28.Stankiewicz A. Identification of the relaxation and retardation spectra of plant viscoelastic materials using Chebyshev functions. Part II. Analysis. Teka Comm. Mot. Energetics Agric. 2010;10:372–378. [Google Scholar]
  • 29.Pillonetto G., Chen T., Chiuso A., De Nicolao G., Ljung L. Regularized System Identification: Learning Dynamic Models from Data. 1st ed. Springer International Publishing; Berlin/Heidelberg, Germany: 2022. [DOI] [Google Scholar]
  • 30.Aleksandrov A., Lekova S., Milenova M. Generalized model of viscoelastic deformation. [(accessed on 17 December 2022)];Int. J. Pure Appl. Math. 2022 81:635–646. Available online: https://ijpam.eu/contents/2012-81-4/11/index.html. [Google Scholar]
  • 31.Cao J., Chen Y., Wang Y., Cheng G., Barrière T. Shifted Legendre polynomials algorithm used for the dynamic analysis of PMMA viscoelastic beam with an improved fractional model. Chaos Solitons Fractals. 2020;141:110342. doi: 10.1016/j.chaos.2020.110342. [DOI] [Google Scholar]
  • 32.Abbaszadeh M., Dehghan M. The proper orthogonal decomposition modal spectral element method for two-dimensional viscoelastic equation. Thin-Walled Struct. 2021;161:107429. doi: 10.1016/j.tws.2020.107429. [DOI] [Google Scholar]
  • 33.Kim S., Lee J., Cho K.S. Direct conversion of creep data to dynamic moduli using point-wise method. Korea-Aust. Rheol. J. 2021;33:105–112. doi: 10.1007/s13367-021-0009-x. [DOI] [Google Scholar]
  • 34.Lee S.H., Bae J.-E., Cho K.S. Determination of continuous relaxation spectrum based on the Fuoss-Kirkwood relation and logarithmic orthogonal power-series approximation. Korea-Aust. Rheol. J. 2017;29:115–127. doi: 10.1007/s13367-017-0013-3. [DOI] [Google Scholar]
  • 35.Stankiewicz A. On determination of the relaxation spectrum of viscoelastic materials from discrete-time stress relaxation data. Teka Comm. Mot. Energetics Agric. 2012;12:217–222. [Google Scholar]
  • 36.Baumgaertel M., Schausberger A., Winter H.H. The relaxation of polymers with linear flexible chains of uniform length. Rheol. Acta. 1990;29:400–408. doi: 10.1007/BF01376790. [DOI] [Google Scholar]
  • 37.Baumgaertel M., Winter H.H. Interrelation between continuous and discrete relaxation time spectra. J. Non-Newton. Fluid Mech. 1992;44:15–36. doi: 10.1016/0377-0257(92)80043-W. [DOI] [Google Scholar]
  • 38.Povolo F., Hermida É.B. Influence of Intensity of Relaxation on Interconversion between Normalized Distribution Functions. Polym. J. 1992;24:1–13. doi: 10.1295/polymj.24.1. [DOI] [Google Scholar]
  • 39.Hansen P.C. Rank-Deficient and Discrete Ill-Posed Problems. Society for Industrial and Applied Mathematics; Philadelphia, PA, USA: 1998. [DOI] [Google Scholar]
  • 40.Friedrich C., Honerkamp J., Weese J. New ill-posed problems in rheology. Rheol. Acta. 1996;35:186–193. doi: 10.1007/BF00396045. [DOI] [Google Scholar]
  • 41.Stankiewicz A. On the existence and uniqueness of the relaxation spectrum of viscoelastic materials. Part II: Other existence conditions. Teka Comm. Mot. Energetics Agric. 2010;10:388–395. [Google Scholar]
  • 42.Datta K.B., Mohan B.M. Orthogonal Functions in Systems and Control. World Scientific; Singapore: 1995. [Google Scholar]
  • 43.Tikhonov A.N., Arsenin V.Y. Solutions of Ill-Posed Problems. John Wiley & Sons; New York, NY, USA: 1977. [Google Scholar]
  • 44.Wahba G. Practical Approximate Solutions to Linear Operator Equations When the Data are Noisy. SIAM J. Numer. Anal. 1977;14:651–667. doi: 10.1137/0714044. [DOI] [Google Scholar]
  • 45.Golub G.H., Van Loan C.F. Matrix Computations. Johns Hopkins University Press; Baltimore, MD, USA: 1996. [Google Scholar]
  • 46.Pérez-Calixto D., Amat-Shapiro S., Zamarrón-Hernández D., Vázquez-Victorio G., Puech P.-H., Hautefeuille M. Determination by Relaxation Tests of the Mechanical Properties of Soft Polyacrylamide Gels Made for Mechanobiology Studies. Polymers. 2021;13:629. doi: 10.3390/polym13040629. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Malkin A.I.A., Malkin A.Y., Isayev A.I. Rheology: Concepts, Methods and Applications. ChemTec; Deerfield Beach, FL, USA: 2006. [(accessed on 17 December 2022)]. Available online: https://books.google.pl/books?id=8rGafjhgz-UC. [Google Scholar]
  • 48.Barata J.C.A., Hussein M.S. The Moore–Penrose Pseudoinverse: A Tutorial Review of the Theory. Braz. J. Phys. 2012;42:146–165. doi: 10.1007/s13538-011-0052-z. [DOI] [Google Scholar]
  • 49.Szabatin J. Podstawy Teorii Sygnałów. Wydawnictwa Komunikacji i Łączności; Warszawa, Poland: 1982. (In Polish) [Google Scholar]
  • 50.Stankiewicz A. Identification of the relaxation spectrum of viscoelastic materials under strong noise measurement data. Part II. Identification algorithm as FIR filter. MOTROL. Comm. Mot. Energetics Agric. 2013;15:127–134. [Google Scholar]
  • 51.Szego G. Orthogonal Polynomials. American Mathematical Society; Providence, RI, USA: 1975. [(accessed on 17 December 2022)]. Available online: https://books.google.nl/books?id=ZOhmnsXlcY0C. [Google Scholar]
  • 52.Lebedev N.N. Special Functions & Their Applications. Dover Publications; New York, NY, USA: 2012. [Google Scholar]
  • 53.Tikhonov A.N., Samarskii A.A. Equations of Mathematical Physics. Dover Publications; New York, NY, USA: 2013. [Google Scholar]
  • 54.Boumenir A. Inverse Spectral Problem for the Laguerre Differential Operator. J. Math. Anal. Appl. 1998;224:218–240. doi: 10.1006/jmaa.1998.6000. [DOI] [Google Scholar]
  • 55.Wang L., Cluett W.R. Optimal choice of time-scaling factor for linear system approximations using Laguerre models. IEEE Trans. Autom. Control. 1994;39:1463–1467. doi: 10.1109/9.299635. [DOI] [Google Scholar]
  • 56.Prokhorov S.A., Kulikovskikh I.M. Unique condition for generalized Laguerre functions to solve pole position problem. Signal Process. 2015;108:25–29. doi: 10.1016/j.sigpro.2014.08.040. [DOI] [Google Scholar]
  • 57.Malti R., Maquin D., Ragot J. Some results on the convergence of transfer function expansion on Laguerre series; Proceedings of the 1999 European Control Conference (ECC); Karlsruhe, Germany. 31 August–3 September 1999; pp. 4649–4655. [DOI] [Google Scholar]
  • 58.Farikhin, Mohd I. Orthogonal Functions Based on Chebyshev Polynomials. MATEMATIKA: Malays. J. Ind. Appl. Math. 2011;27:97–107. doi: 10.11113/matematika.v27.n.299. [DOI] [Google Scholar]
  • 59.Rawitscher G., dos Santos Filho V., Peixoto T.C. An Introductory Guide to Computational Methods for the Solution of Physics Problems: With Emphasis on Spectral Methods. Springer International Publishing; Cham, Germany: 2018. Chebyshev Polynomials as Basis Functions; pp. 43–62. [DOI] [Google Scholar]
  • 60.Van Assche W. Ordinary Special Functions. In: Françoise J.-P., Naber G.L., Tsun T.S., editors. Encyclopedia of Mathematical Physics. Academic Press; Oxford, MI, USA: 2006. pp. 637–645. [DOI] [Google Scholar]
  • 61.Stankiewicz A. Identification of the relaxation and retardation spectra of plant viscoelastic materials using Chebyshev functions. Part I. Identification algorithm. Teka Comm. Mot. Energetics Agric. 2010;10:363–371. [Google Scholar]
  • 62.Celeghini E., Gadella M., del Olmo M.A. Hermite Functions and Fourier Series. Symmetry. 2021;13:853. doi: 10.3390/sym13050853. [DOI] [Google Scholar]
  • 63.Wayne C.E., Zharnitsky V. Exponential bound of the integral of Hermite functions product with Gaussian weight. J. Math. Anal. Appl. 2023;517:126544. doi: 10.1016/j.jmaa.2022.126544. [DOI] [Google Scholar]
  • 64.Deaño A., Temme N.M. Analytical and numerical aspects of a generalization of the complementary error function. Appl. Math. Comput. 2010;216:3680–3693. doi: 10.1016/j.amc.2010.05.025. [DOI] [Google Scholar]
  • 65.Dong S.-H. Factorization Method in Quantum Mechanics. 1st ed. Springer Science & Business Media; Dordrecht, The Netherlands: 2007. [Google Scholar]
  • 66.Wagner K.W. Operatorenrechnung Nebst Anwendungen in Physik und Technik. 3rd ed. J.A. Barth; Leipzig, Germany: 1962. [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Data Availability Statement

Not applicable.


Articles from Polymers are provided here courtesy of Multidisciplinary Digital Publishing Institute (MDPI)

RESOURCES