Skip to main content
Materials logoLink to Materials
. 2024 Oct 3;17(19):4870. doi: 10.3390/ma17194870

Direct Identification of the Continuous Relaxation Time and Frequency Spectra of Viscoelastic Materials

Anna Stankiewicz 1
Editor: Laurent Chazeau1
PMCID: PMC11478369  PMID: 39410441

Abstract

Relaxation time and frequency spectra are not directly available by measurement. To determine them, an ill-posed inverse problem must be solved based on relaxation stress or oscillatory shear relaxation data. Therefore, the quality of spectra models has only been assessed indirectly by examining the fit of the experiment data to the relaxation modulus or dynamic moduli models. As the measures of data fitting, the mean sum of the moduli square errors were usually used, the minimization of which was an essential step of the identification algorithms. The aim of this paper was to determine a relaxation spectrum model that best approximates the real unknown spectrum in a direct manner. It was assumed that discrete-time noise-corrupted measurements of a relaxation modulus obtained in the stress relaxation experiment are available for identification. A modified relaxation frequency spectrum was defined as a quotient of the real relaxation spectrum and relaxation frequency and expanded into a series of linearly independent exponential functions that are known to constitute a basis of the space of square-integrable functions. The spectrum model, given by a finite series of these basis functions, was assumed. An integral-square error between the real unknown modified spectrum and the spectrum model was taken as a measure of the model quality. This index was proved to be expressed in terms of the measurable relaxation modulus at uniquely defined sampling instants. Next, an empirical identification index was introduced in which the values of the real relaxation modulus are replaced by their noisy measurements. The identification consists of determining the spectrum model that minimizes this empirical index. Tikhonov regularization was applied to guarantee model smoothness and noise robustness. A simple analytical formula was derived to calculate the optimal model parameters and expressed in terms of the singular value decomposition. A complete identification algorithm was developed. The analysis of the model smoothness and model accuracy for noisy measurements was carried out. The equivalence of the direct identification of the relaxation frequency and time spectra has been demonstrated when the time spectrum is modeled by a series of functions given by the product of the relaxation frequency and its exponential function. The direct identification concept can be applied to both viscoelastic fluids and solids; however, some limitations to its applicability have been pointed out. Numerical studies have shown that the proposed identification algorithm can be successfully used to identify Gaussian-like and Kohlrausch–Williams–Watt relaxation spectra. The applicability of this approach to determining other commonly used classes of relaxation spectra was also examined.

Keywords: viscoelasticity, relaxation spectra, linear relaxation modulus, direct spectrum approximation, identification algorithm, model integral square error, noise robustness

1. Introduction

Although the first papers concerning relaxation time and frequency spectra determination come from the late 1940s of the 20th century [1,2], the recovery of the relaxation spectrum from the measurement data is still an active area of research in rheology and the identification of time-variable viscoelastic mechanical characteristics [3,4,5,6,7,8,9,10]. Relaxation time and frequency spectra, with no direct accessible measurements, are recovered from the stress relaxation or oscillatory shear data by applying appropriate identification methods intended for determination of the spectra. The relaxation spectrum identification task is the problem of numerically solving a system of Fredholm integral equations of the first kind obtained for discrete measurements of the relaxation modulus or storage and loss modulus data. These problems are well-known to be the ill-posed inverse problems, the solutions to which, if any, are very sensitive to even small changes in the experiment data leading to arbitrarily large changes in the determined relaxation spectrum. Therefore, special stable algorithms are requisite to determine noise-robust relaxation spectrum models.

Over the last 80 years, different analytical and numerical tools have been applied to identify the relaxation spectrum. Numerous classes of algorithms were developed to determine continuous and discrete relaxation spectra models. Many theoretical papers have been devoted to the methods and algorithms for relaxation spectra determination, e.g., see [3,6,9,11,12,13,14]. In addition, experimental studies conducted for various viscoelastic materials motivated relaxation spectra models and appropriate identification algorithms, for example as seen in [4,15,16,17,18]. Reviews of these methods and algorithms can be found in many papers, for example [5,19,20] and, most recently, in [10,18].

After a few models and algorithms were derived from an application of the Post–Widder differential formula [21,22,23], many more intricate methods and models have been obtained based on the usage of the least-squares identification applied both to the relaxation modulus measurements obtained in the stress relaxation test [7,10,24,25,26,27,28] and to the measurements of the storage and loss moduli resulted from the oscillatory shear experiment [3,4,5,6,8,11,12,14,15,16,17,18]. For example, in [26,27,28], different identification algorithms were derived for the optimal regularized least-squares identification of relaxation time and frequency spectra in the classes of models defined by a finite series of different basis functions. In consequence, the quality of the spectra models was estimated by the mean sum of relaxation modulus or dynamic moduli square errors used as a measure, the minimization of which, with or without regularization, was an essential step of the identification algorithm. In some papers, e.g., [4,7,29], the pure least-squares identification was applied, while for example in [3,5,11,26,27,28] the regularized least-squares were used with various rules applied for the choice of regularization parameters to ensure the stability of the scheme and model smoothness. Recently, in [10], the best smoothed spectrum model—which reproduces the relaxation modulus measurements with a small error of the relaxation modulus model by minimizing the integral square norm of the spectrum—was found; however, here, the identification criterion is related only to the spectrum model and not to the unknown real spectrum, and the model error is assessed in terms of the measurement-available relaxation modulus.

In this paper, a new approach is proposed based on direct approximation of the real unknown relaxation time spectrum by a series of appropriately selected basis functions. It was assumed that discrete-time noise-corrupted measurements of a relaxation modulus obtained in the stress relaxation experiment are available for identification. First, a modified relaxation frequency spectrum was defined as a quotient of the real relaxation spectrum and relaxation frequency. This spectrum was expanded into a series of exponential functions forming a basis of the space of square-integrable functions [30]. Such expansion is equivalent to the expansion of the relaxation time spectrum into a series of basis functions, these being the products of the relaxation frequency and the exponential function of it. The spectra models, given by the finite series of these basis functions, were assumed. An integral square error between the real unknown spectrum and the spectrum model was taken as a measure of the model quality index. The equivalence of such defined indices for the relaxation time spectrum and the modified frequency spectrum was proved, which means an equivalence between the respective spectra approximation tasks. Next, an empirical identification index was introduced by replacing the real relaxation modulus by their noise measurements. The resulting identification problem is a linear-quadratic optimization task in which Tikhonov regularization is applied to ensure its well-posedness. Simple analytical formula for determining the optimal model parameters was derived; the singular value decomposition can be used for algebraic computations. A complete identification algorithm for determining the optimal models of the relaxation spectra has been developed. Model smoothness and noise robustness were analyzed. The results of simulation studies conducted for uni- and double-mode Gaussian-like and Kohlrausch–Williams–Watts relaxation spectra are presented. Finally, based on the congruence of the boundary conditions of the real spectra and the model basis functions, a short analysis of the applicability of the proposed approach is outlined for different classes of the real spectra, and its limitations are pointed out. It is demonstrated that the concept of direct relaxation spectrum identification can be applied both for viscoelastic fluids and viscoelastic solids. In Appendix A, the proofs and derivations of some mathematical formulas and results are given.

The idea of using a series expansion of the spectrum model has been previously applied both in the time [26,27,28] and frequency [8] domains; however, in these papers, the identification indices, being minimized, were related to the models of the relaxation or dynamic moduli and not to the unknown spectrum model. Here, the use of appropriately selected basis functions of the relaxation spectrum model allowed for linking the model quality index, related directly to the unknown spectrum, with the relaxation modulus measurements. This means that the identification index being minimized, although expressed in terms of the relaxation modulus measurements, refers directly to the unknown relaxation spectrum, not to the measured relaxation modulus. This new approach is proposed and used in this paper for the first time.

2. Materials and Methods

2.1. Relaxation Spectra

It is widely assumed in rheology [31,32,33] that the linear relaxation modulus Gt (i.e., the stress per unit strain) has a relaxation spectrum representation of the form

Gt=0Hττet/τdτ, (1)

or equivalently by

Gt=0Hvvetvdv, (2)

where the relaxation time Hτ and frequency Hv spectra, related by

Hv=H1v, Hτ=H1τ, (3)

characterize the distributions of relaxation times τ and frequencies v. They are generalizations of discrete Maxwell spectra [31,32] to continuous functions of τ and v. Although other definitions of the relaxation spectrum are used in the literature, for example, in [34,35,36], the definition introduced by Equations (1) and (2) dominates.

2.2. Models

Following [26], the modified spectrum is introduced

HMv=Hvv, (4)

where the upper index of HMv means “modified”. Model transformation defined by (4) is a bijection. Equation (2) can be rewritten as follows:

Gt=0HMvetvdv, (5)

i.e., the modulus Gt 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 [37]. The set of the linearly independent exponential functions eαkv, k=0,1,, where α>0, i.e., the kernel of the Laplace transformation, form a basis of the space L20, [30]. Thus, the modified relaxation spectrum can be expressed as:

HMv=k=0gkhkv, (6)

with basis functions defined as follows

hkv=eαkv, (7)

where parameter α>0 is a time-scaling factor expressed in seconds, while gk are constant model parameters.

By (4) and (6), for the real relaxation spectrum of the material we have:

Hv=k=0gkhkvv. (8)

The modified spectrum HMvL20,, then, HMv0 as v and the first basis function can be neglected. For practical reasons, it is convenient to replace the infinite summation in the above equation with a finite one of K first terms, i.e., to approximate the relaxation spectrum HMv (4) by a model of the form

HKMv=k=1Kgkhkv, (9)

where the lower index of HKMv is the number of model summands. Spectrum HMv (4) is expressed in Pa·s, so also Pa·s is a unit of the model’s parameters gk. The model of the original spectrum Hv related to (9) takes the form

HKv=vHKMv=k=1Kgkhkvv. (10)

The related relaxation modulus model, by (5) and (9) is described by the following:

GKt=0HKMvetvdv=k=1Kgk1t+αk=k=1Kgkϕkt, (11)

where the basis functions, expressed in s1, are as follows

ϕkt=1t+αk. (12)

By the second equality in (3), (8), and (7), we obtain the following series representation of the relaxation time spectrum

Hτ=k=0gkeαkτ1τ, (13)

Omitting, as above, the first component and considering the K next terms of the series (13), the relaxation spectrum Hτ can be approximated by a model of the form

HKτ=k=1Kgkhkτ, (14)

where the basis functions

hkτ=eαkτ1τ (15)

and the model parameters gk, expressed in Pa·s, are identical to that of model (9). By (1) and (14), the relaxation modulus model is as follows:

GKt=0HKττet/τdτ=k=1Kgk01τ2et+αkτdτ=k=1Kgkϕkt,

where the basis functions are given by (12), i.e., it is identical to the model described by (11).

2.3. Properties of the Basis Functions

A few basis functions hkv (7) of the model HKMv (9) are shown in Figure 1 for two different values of the time-scale factor α, while in Figure 2, the basis functions hkvv of the model HKv (10) are given. In Figure 3, the basis functions hkτ (15) of the relaxation time spectrum model HKτ (14) are demonstrated.

Figure 1.

Figure 1

Basis functions hkv=eαkv (7) of the relaxation spectrum model HKMv (9) for two time-scaling factors, α: (a) α=0.001 [s]; (b) α=0.1 [s]; k=1, 5, 10, 20, 100, 500.

Figure 2.

Figure 2

Basis functions hkvv of the relaxation spectrum model HKv (10) for two time-scaling factors, α: (a) α=0.001 [s]; (b) α=0.1 [s]; k=1, 2, 5, 10, 100, 500.

Figure 3.

Figure 3

Basis functions hτ=eαkτ/τ (15) of the relaxation time spectrum model HKτ (14) for two time-scaling factors, α: (a) α=0.001 [s]; (b) α=0.1 [s]; k=1, 5, 10, 20, 100, 500.

Figure 4 shows the hyperbolic basis functions ϕkt (12) of the relaxation modulus model GKt (11). Functions ϕkt are almost constant in time and near zero (in the considered time intervals) for k=100 and k=1000. However, for smaller indices k, they are in good agreement with the real relaxation modulus obtained in the stress relaxation test.

Figure 4.

Figure 4

Basis functions ϕkt (12) of the relaxation modulus model GKt (11) for two time-scaling factors, α: (a) α=0.001 [s]; (b) α=0.1 [s]; k=1, 2, 5, 10, 100, 500.

The basis functions of the models HKMv, HKv, and GKt are positive definite. Functions hkv (7) are monotonically decreasing, while basis functions hkvv of the model HKv (10) have global maxima equal to 1/αke for the relaxation frequencies v=1/αk. In addition, basis functions hkτ (15) have unique maxima equal to 1/αke at the relaxation times τ=αk. Relaxation modulus basis functions ϕkt (12) are monotonically decreasing. For large v, and in particular for v, the basis functions hkv and hkvv decrease exponentially to zero. Similarly, for τ functions hkτ0, faster than the exponential function. Functions ϕkt tends to zero hyperbolically as t.

3. Results

In this section, the problem of optimal spectrum approximation in the class of models defined by a finite series of the introduced basis functions is formulated and solved. First, it is demonstrated that the problem of the optimal—in the sense of an integral square error—approximation of the modified relaxation frequency spectrum is equivalent to the problem of the optimal approximation of the relaxation time spectrum. It is also proved that, due to the choice of exponential basis functions, the integral square model error can be expressed in terms of the relaxation modulus in the sampling points uniquely determined by the basis functions of the spectrum model. Next, an empirical identification index is introduced in which the values of the real relaxation modulus are replaced by their noisy, in general, measurements. The optimal models of the relaxation time and frequency spectra are determined by solving the linear-quadratic identification task. However, this problem turned out to be ill-conditioned. Therefore, Tikhonov regularization is applied resulting in the stable, noise-robust and simple identification rule. Next, the equations and functions essential for the proposed identification scheme are described in terms of the singular value decomposition of the basic matrix of this identification problem. Model smoothness is estimated, error of the relaxation modulus is evaluated, noise robustness and convergence analysis is conducted. The complete identification algorithm is presented. The results of simulation studies for Gaussian-like and Kohlrausch–Williams–Watts relaxation spectra describing many real materials are presented. Finally, based on the congruence of the boundary conditions of the real spectra and the basis functions, a rough applicability analysis of the proposed approach is outlined for different classes of the real spectra.

3.1. Spectrum Approximation

Model HKMv (9) approximates the modified spectrum HMv (4). As a measure of the model (9) accuracy the integral square index is taken

JgK=0HMvHKMv2dv, (16)

where gK=g1gKT is an K—element vector of the model (9) parameters; superscript “T” indicates transpose. In view of the additive form of the model (9), index JgK can be expressed as follows:

JgK=0HMv2dv2k=1Kgk0HMvhkvdv+k=1Km=1Kgkgmφkm, (17)

where, by (7), the coefficients

φkm=0hkvhmvdv=0eαk+mvdv=1αk+m. (18)

Since, in view of (5) and (7), we have:

0HMvhkvdv=0HMeαkvdv=Gαk,

the above index takes the form

JgK=0HMv2dv2k=1KgkGαk+k=1Km=1Kgkgmφkm. (19)

Model HKτ (14) approximates the real spectrum Hτ. As a measure of the model (14) accuracy the square index, analogous to (16), is taken

JgK=0HτHKτ2dτ, (20)

where gK is a vector of the model (14) parameters. In view of (14) and (15), the above index can be expressed as follows:

JgK=0H2τdτ2k=1Kgk0Hτhkτdτ+k=1Km=1Kgkgm0hkτhmτdτ, (21)

where, by (15) and (1)

0Hτhkτdτ=0Hτeαkτ1τdτ=Gαk,

while the integrals

0hkτhmτdτ=01τ2eαk+mτdτ=1αk+m=φkm, (22)

are identical to that given by (18). Combining the above with (21) yields

JgK=0H2τdτ2k=1KgkGαk+k=1Km=1Kgkgmφkm.

For the material relaxation spectra Hτ and HMv, by (4) and simple substitution based on (3), we have the following:

0HMv2dv=0Hv2v2dv=0H1v21v2dv=0Hτ2dτ, (23)

that is, the integral square index defined by (20) is identical to that defined by (16); therefore, the same notation was used.

3.2. Identification Problem

Suppose that a certain identification experiment (stress relaxation test [1,33,38]) performed on the specimen of the material under investigation resulted in a set of measurements of the relaxation modulus G¯tk=Gtk+ztk at the sampling instants tk=αk, k=1,,K; where ztk is additive measurement noise. Generally, identification consists of selecting within the given class of models, which ensures the best fit to the measurement results. Classically, the mean square identification index related to the measurements of the relaxation modulus G¯tk is used; compare [7,24,26,27,28]. This means that the model quality index is not related directly to the unknown relaxation spectrum, which is inaccessible by measurement, but to the measurement-available relaxation modulus. Such an approach is typical in the context of the inverse problem.

Here, as a measure of the model (9), equivalently (14), the accuracy of the index JgK of the form (16) is applied. Note, that the first component of JgK given by the right-hand side of (19) depends on the unknown relaxation spectrum; the second term is determined by model parameters gk and the values of the measurable relaxation modulus at the time instant tk=αk; and the last component is affected only by the model parameters and the times tk. Replacing in (19) the relaxation modulus Gtk=Gαk by their measurements G¯tk, we obtain the following integral-empirical index

J¯KgK=0HMv2dv2k=1KgkG¯αk+k=1Km=1Kgkgmφkm. (24)

Let us introduce the vector-matrix notation

ΦK=121K+11K+112K, G¯K=G¯t1G¯tK. (25)

Note that element (k,m) of the matrix ΦK, i.e., the entry in the k-th row and m-th column of ΦK, is equal to αφkm and is dimensionless. Therefore, the algebraic properties of the matrix ΦK do not depend on the time-scale factor α. Using the above notation and bearing in mind (18), the identification index (24) can be expressed in compact form as follows:

J¯KgK=0HMv2dv2G¯KTgK+1αgKTΦKgK. (26)

Thus, the optimal identification of the relaxation spectrum in the class of models (9), equivalently (14), consists of solving—with respect to the model parameter gK—the linear-quadratic problem

mingKRKJ¯KgK, (27)

the system of normal equations of which is as follows:

ΦKgK=αG¯K. (28)

The existence and properties of the solution to (28) depend on the properties of square symmetric matrix ΦK (25) specified by the following result, which is the simple consequence of the independence of the basis functions hkv (7), as proved in Appendix A.1.

Lemma 1. 

The matrix ΦK, Equation (25), is positive definite for an arbitrary K1.

By the above lemma, the unique solution of the minimization task (27) is as follows:

gK=αΦK1G¯K. (29)

The matrix ΦK, although of full-rank, is extremely ill-conditioned and must be used with care. Namely, the problem of the matrix ΦK inversion is ill-conditioned, therefore small perturbations in ΦK may produce large changes in ΦK1. The spectral condition number ([39] Equation (2.6.3))

κΦK=ΦK12·ΦK2, (30)

where ·2 is the spectral norm of matrices, and equal, in fact, to the ratio of the largest singular value of ΦK to the smallest, measures the sensitivity of the answer to small perturbation of the data. From the first row values in Table 1, where κΦK is given for a few values of K, we see that index κΦK exceeds the value of 105 already for K=5, the value of 1010 as early as K=8 and tends to infinity with growing K. Thus, positive definite ΦK is suspected to be very ill-conditioned, even for not very large K5, and numerical solution of (28) results in fluctuations of the parameters vector gK, which are the greater, the greater is the value of the condition number κΦK.

Table 1.

Spectral condition numbers κΦK and κΦK+αλIK,K defined according to (30).

αλ K = 2 K = 3 K = 4 K = 5 K = 8 K = 10 K = 15 K = 20 K = 100 K = 1000 K = 10,000
0 38.474 1.35 × 103 4.59 × 104 1.54 × 106 5.64 × 1010 6.23 × 1013 2.61 × 1017 6.45 × 1018 1.20 × 1019 9.59 × 1020 1.0 × 1022
10 1.071 1.087 1.098 1.106 1.122 1.129 1.14129 1.1496 1.18809 1.225 1.248
1 1.699 1.874 1.977 2.056 2.215 2.288 2.412 2.495 2.880 3.247 3.479
0.1 6.983 9.688 10.773 11.559 13.154 13.877 15.123 15.953 19.800 23.469 25.798
0.01 25.552 83.135 98.55 106.59 122.542 129.766 142.232 150.535 189.001 225.687 248.977
1 × 10−3 36.600 532.056 958.14 1.06 × 103 1.22 × 103 1.29 × 103 1.41 × 103 1.49 × 103 1.88 × 103 2.25 × 103 2.48 × 103
1 × 10−4 38.278 1.17 × 103 8.06 × 103 1.05 × 104 1.22 × 104 1.29 × 104 1.41 × 104 1.49 × 104 1.88 × 104 2.25 × 104 2.48 × 104
1 × 10−5 38.454 1.33× 103 3.12 × 104 9.88 × 104 1.22 × 105 1.29 × 105 1.41 × 105 1.49 × 105 1.88 × 105 2.25 × 105 2.48 × 105

Summarizing, the linear-quadratic identification task (27) is ill-conditioned [40] and when the data are noisy even small changes in G¯K would lead to an arbitrarily large artefact in gK given by (29). Therefore, the numerical solution of the finite-dimensional problem (27) is fraught with the same difficulties that the original continuous ill-posed problems of numerical solution of the Fredholm Equations (2) or (5).

3.3. Regularization

To deal with the ill-conditioning, we use Tikhonov regularization [41], which is classical and because of its simplicity, probably the most common method for solving ill-posed linear-quadratic problems. For the linear-quadratic task (27), Tikhonov regularization strives in minimizing a modified square functional of the form

J¯KgK+λgKTgK, (31)

where λ>0 is a regularization parameter. The unit of λ must be s1 to ensure dimensional consistency of the above index. For J¯KgK given by (26), the regularized task results in the linear-quadratic optimization task

mingKRK1αgKTΦKgK2G¯KTgK+λgKTgK; (32)

the first summand of (26), being independent on gK, does not have to be taken into account here, just as it did not affect the minimization result in the original problem (27). The set of normal equations is now as follows

ΦK+αλIKgK=αG¯K, (33)

where IK is K×K identity matrix.

The existence and properties of the solution of (33) depend on the properties of the symmetric matrix ΦK+αλIK. Based on Lemma 1, ΦK+αλIK is non-singular and positive definite for any λ0. In successive rows of Table 1, the spectral conditional numbers κΦK+αλIK are given for a few values of the dimensionless product αλ, which determines the value of κΦK+αλIK for given K. The numerical studies indicate that κΦK+αλIK is not greater than the numerically acceptable value 105 [42] for K50, whenever the parameters product αλ1.7×105, for K100 if αλ1.89×105, and for αλ2.16×105, assuming measurement points K500.

Therefore, the problem (32) is well-posed, that is the solution exists, is unique, and continuously depends on both the matrix ΦK and the measurements G¯K. By (33), the optimal regularized vector is given by the following formula:

g¯Kλ=αΦK+αλIK1G¯K. (34)

Elegant and compact Formula (34) is, however, unsuitable for computational purposes, for which the singular decomposition technique [39] will be used.

3.4. Algebraic Background

Let the singular value decomposition (SVD) of the matrix ΦK (25) take the form [39]

ΦK=UKΣKUKT, (35)

where the diagonal K×K matrix

ΣK=diagσ1,,σK, (36)

is composed of the non-zero singular values σ1σkσK of the matrix ΦK; matrix UKRK,K is orthogonal. The SVD (35) is uniquely determined. The singular values σk of ΦK do not depend on the time-scale factor α. Therefore, for given K, the SVD must be computed only once even if the sampling instants tk=αk dependent on the parameter α are changed in the experiment. Taking advantage of the diagonal structure of ΣK and orthogonality of the matrix UK, we have:

ΦK+αλIK1=UKΣK+αλIK1UKT=UKΩKUKTT, (37)

where diagonal K×K matrix

ΩK=ΣK+αλIK1=diag1σ1+αλ,,1σK+αλ. (38)

In view of (37), (34), and (38), the optimal regularized vector g¯Kλ is expressed as follows:

g¯Kλ=αUKΣK+αλIK1YK=αUKΩKYK, (39)

where the K dimensional vector

YK=UKTG¯K. (40)

According to (9), (10), and (14), the resulting best relaxation spectra models are as follows

H¯KMv=k=1Kg¯kλhkv, (41)
H¯Kv=k=1Kg¯kλhkvv, (42)

and

H¯Kτ=k=1Kg¯kλhkτ, (43)

where g¯kλ are elements of the vector g¯Kλ (34), or equivalent (39).

3.5. Analysis

The model smoothing and its accuracy in the case of noisy measurements of the relaxation modulus will be now analyzed. Contrary to the previous papers [26,27,28]—in which the model quality index refers to the relaxation modulus but not directly to the unknown relaxation spectrum—here we can estimate the spectra errors HMvH¯KMv2, H¯KvH~Kv2, and HτH¯Kτ2 directly, where ·2 denotes the square norm in the space L20,.

3.5.1. Model Smoothness

The purpose of the regularization applied in (31) relies on stabilization of the vector of model parameters g¯Kλ (34). The norms H¯KMv2, H¯Kv2, and H¯Kτ2 are natural measures of the spectra models’ (41)–(43) smoothness. In Appendix A.2, the following result is derived.

Proposition 1. 

Let the time-scale factor α>0 and the regularization parameter λ>0. Then, for the optimal relaxation spectra models H¯KMv (41), H¯Kv (42), and H¯Kτ (43) we have the following

1ασKg¯KTλg¯KλH¯KMv22=H¯Kτ22=1αg¯KTλΦKg¯Kλ1ασ1g¯KTλg¯Kλ, (44)

where the vector g¯Kλ is given by (39) and

2α3ςKg¯KTλg¯KλH¯Kv22=2α3g¯KTλΘKg¯Kλ2α3ς1g¯KTλg¯Kλ, (45)

where σK and σ1 are the minimal and maximal singular values of the matrix ΦK (25), while ςK and ς1 are the minimal and maximal singular values of the positive definite matrix

ΘK=1231K+131K+1312K3. (46)

The values of square roots of the smallest and largest singular values σK, σ1, ςK, and ς1 for some model summands K are summarized in Table 2. Due to the ill-conditioning of matrices ΦK and ΘK, the lower bounds in (44) and (45) are not too useful. Since σ1 and ς1 grows with K, from the analysis of Table 2 data and the right inequalities in (44) and (45), the next result follows immediately.

Table 2.

The square roots of the largest σ1, ς1 and minimal σK, ςK singular value of the matrices ΦK (25) and ΘK (46) for K model summands.

K 10 50 100 500 1000 5000 10,000
σ1 1.1348 1.3153 1.3711 1.4677 1.4989 1.5556 1.5747
σK 1.437 × 10−7 1.919 × 10−10 3.955 × 10−10 1.923 × 10−10 4.838 × 10−11 2.323 × 10−11 1.573 × 10−11
ς1 0.3737 0.3737 0.3737 0.3737 0.3737 0.3737 0.3737
ςK 6.829 × 10−8 1.151 × 10−11 1.847 × 10−12 1.017 × 10−13 2.455 × 10−14 2.769 × 10−15 6.473 × 10−15
Proposition 2. 

Let the time-scale factor α>0 and the regularization parameter λ>0. If the number of relaxation modulus measurements K104, then the optimal relaxation spectra models H¯KMv (41), H¯Kv (42), and H¯Kτ (43) are such that

H¯KMv2=H¯Kτ21.5747αg¯Kλ2, (47)
H¯Kv20.5285ααg¯Kλ2, (48)

where the vector g¯Kλ (39) and ·2 denotes here the square norm in Euclidean space RK.

Since, by the right equality in (39) and the orthogonality of UK, we obtain:

g¯KTλg¯Kλ=α2YKTΩKUKTUKΩKYKT=α2YKTΩK2YKT,

bearing in mind the diagonal structure of ΩK (38) we have the formula:

g¯Kλ22=g¯KTλg¯Kλ=α2k=1K yk2σk+αλ2, (49)

where yk are the elements of the vector YK (40), which directly illustrates the mechanism of the regularization. The following rule holds: the greater the regularization parameter λ is, the more highly bounded the fluctuations of the vector g¯Kλ are.

Summarizing, the smoothness of the optimal vector g¯Kλ of the model’s parameters guarantees that the fluctuations of the resulting relaxation spectra models H¯KMv, H¯Kτ, and H¯Kv are also bounded. Both time-scale factor α and the regularization parameter λ affect the smoothness of the spectrum models. However, it should be remembered that the inequalities (47) and (48) give only the upper bounds of the respective norms.

3.5.2. Noise Robustness and Convergence

The model of the modified spectrum that we would obtain for the same time-scale factor α and regularization parameter λ on the basis of ideal (noise-free) relaxation modulus measurements:

H~KMv=k=1Kg~kλhkv, (50)

where g~Kλ is the vector model parameters given by the following (compare (39) and (40)):

g~Kλ=αUKΩKUKTGK (51)

for the noise-free relaxation modulus GK=Gt1GtKT, which will be considered as a reference point for the model H¯KMv (41). The respective noise-free optimal regularized models of the relaxation frequency and time spectra are as follows

H~Kv=k=1Kg~kλhkvv, (52)

and

H~Kτ=k=1Kg~kλhkτ. (53)

In Appendix A.3, the following estimations are derived.

Proposition 3. 

For an arbitrary time-scale factor α and arbitrary regularization parameter λ, the errors between the relaxation spectra models H¯KMv (41), H¯Kv (42), and H¯Kτ (43) and related noise-free models H~KMv (50), H~Kv (52), and H~Kτ (53) are estimated by the following inequalities:

H¯KMvH~KMv2=H¯KτH~Kτ2αγzN2, (54)

 where parameter

γ=max1kK σkσk+αλ, (55)

 and

H¯KvH~Kv22 ς1ασK+αλzN2, (56)

where ς1 is the maximal singular value of ΘK (46); zN=zt1ztNT is the vector of measurement noises.

According to inequalities (54) and (56), the accuracy of the noise-free optimal spectra approximation depends on the measurement noises, the regularization parameter, the time-scale factor, and on the singular values of the matrices ΦK (25) and ΘK (46), this being dependent on the number of measurements. By (54) and (56), having in mind the continuity of all the spectra considered here, we conclude that the spectra H¯KMv (41), H¯Kv (42), and H¯Kτ (43) tend to their noise-free counterparts for each v>0 and τ>0 linearly with respect to the norm zN2, as zN20, and the faster the larger the regularization parameter λ.

3.5.3. Error of the Relaxation Modulus Model

The approximation of the material spectrum Hv by series of functions HKMv (9) results in the relaxation modulus Gt approximation by the series GKt (11) of basis functions ϕkt (12). Therefore, the relaxation modulus model corresponding to the relaxation spectra models (41)–(43) is described by the following equation:

G¯Kt=k=1Kg¯kλϕkt. (57)

The mean square error of the relaxation modulus model is as follows

QKg¯Kλ=1Kk=1KG¯tkG¯Ktk2. (58)

In Appendix A.4, the following result is derived.

Proposition 4. 

For an arbitrary time-scale factor α and arbitrary regularization parameter λ, the square error of the relaxation modulus model G¯Kt (57) defined by (58) for the optimal model parameter g¯Kλ (39) is given by the following formula:

QKg¯Kλ=1KG¯K1αΦKg¯KλTG¯K1αΦKg¯Kλ=k=1Kyk2σkαλ+12, (59)

where σk, k=1,,K, are the singular values of the matrix; ΦK (25) and yk are the elements of the vector YK (40).

The equality (59) yields that the accuracy of the relaxation modulus approximation depends on the following: the real relaxation modulus GK and measurement noises zK affecting the value of YK=UKTGK+UKTzK; the scaling factor α; the regularization parameter λ; and singular values of the matrix ΦK (25). These, in turn, depend on the number of measurements. Note also that only the product αλ, not α and λ independently, affects the index QKg¯Kλ. Since the first derivative

QKg¯Kλαλ=2k=1Kyk2σkαλσk+αλ3,

is positive for any αλ>0, the error of the relaxation modulus model grows with increasing regularization parameter λ, slow for very small and very large αλ.

3.6. Identification Algorithm

Allowing the above, the calculation of the relaxation spectra models involves the following steps.

  1. For the studied material, perform the preliminary experiment (stress relaxation test [1,33,38]) and record the measurements G¯ti, i=1,,N, of the relaxation modulus for pre-selected time instants (e.g., sampled with the constant period in the time interval 0,T, T<);

  2. Choose the time-scaling factor α and the number K of model components comparing, for different values of α, a few functions from the sequence ϕkt given by (12), and creating relaxation modulus model GKt (11) with the experiment results G¯ti;

  3. Perform the experiment and record the measurements G¯tk of the relaxation modulus at times tk=α·k, k=1,,K;

  4. Compute the matrix ΦK (25), and next, determine the SVD (35) with the singular values σ1,,σk,σK of ΦK;

  5. Select the regularization parameter λ such that for assumed α and K the spectral condition number is such that
    κΦK+αλIK,K=σ1+αλσK+αλ105; (60)
  6. For chosen λ, compute the regularized solution g¯Kλ according to (39);

  7. Determine the modified spectrum of relaxation frequencies H¯KMv according to (41);

  8. Determined the spectra of relaxation time H¯Kτ and frequency H¯Kv according to (43) and (42), respectively, as the linear combinations of the respective basis functions.

The matrix ΦK (25) and, in particular, the singular values σk of ΦK depend only on the number of measurements and do not depend even on the time-scale factor α and on the sampling points tk. Therefore, for the fixed K matrix ΦK and the SVD of ΦK—being the most space- and time-consuming task of the scheme of computational complexity ONK2 [39]—these must be determined only once when the identification scheme is applied for successive samples of the same material (step 4). The SVD is accessible in the form of optimized numerical procedures in most computational packets.

The condition (60) from step 5 means that the condition number does not exceed the numerically acceptable value 105 [42]; data from Table 1 may be useful here.

3.7. Simulational Studies

This section presents the results of the approach with proposed numerical studies for three simulated materials whose viscoelastic properties are described by the Gauss-like and Kohlrausch–Williams–Watts (KWW) models. Gauss distributions of the relaxation spectra are examined while developing new identification methods; the best examples are as follows: ([5] (Figure 2)), ([14] (Figures 9, 11 and 17)) and ([43] (Figures 2, 3, 6–11 and 14)). The Gaussian-like distributions of the relaxation spectra were used to describe the viscoelastic properties of a lot of real materials, mainly polymers, for example, poly(methyl methacrylate) [44], polyacrylamide gels ([45] (Figure A4)), polyethylene [46] and carboxymethylcellulose (CMC) [47]. Gaussian nature has also the spectra of many biopolymers, e.g., fresh egg white-hydrocolloids [47], cold gel-like emulsions stabilized with bovine gelatin [48], xanthan gum water solution [47], some (potato, corn, wheat, and banana) native starch gels [49], and wood [24,50]. Recently, Gaussian-type relaxation spectra have also been determined for the modified asphalt binder blends ([51] (Figures 3a,c and 5a,c,e)).

The KWW model of the stretched exponential relaxation has been found by many researchers to be more appropriate than standard exponentials to describe viscoelastic processes of many materials, for example, polymer melts [52], the segmental dynamics and the glass transition behavior of poly(2-vinylpyridine) [53], relaxation of bone and bone collagen [54], and alginate films while considering glycerol concentration [55]. The KWW model, initially introduced to describe the viscoelastic relaxation processes [56,57], has also been used to model other relaxation processes occurring in materials, for example, enthalpy relaxation in Cu46Zr45Al7Y2 and Zr55Cu30Ni5Al10 bulk metallic glasses [58], isothermal enthalpy relaxation and density relaxation investigated for bulk Pd42.5Cu30Ni7.5P20 and Pd40Ni40P20 metallic glasses [59], and structural relaxation of a Hf-microalloyed Co-based glassy alloy [60].

Applying the proposed identification algorithm, the best spectra models H¯KMv (41), H¯Kv (42), and H¯Kτ (43) were determined for a few numbers of measurements. The smoothness of the models is estimated by their integral square norms described in Proposition 1. For the spectra H¯KMv and H¯Kτ, these norms are equal and uniquely characterized by the middle equality in (44), which yields the following:

H¯KMv2=H¯Kτ2=1αg¯KTλΦKg¯Kλ. (61)

By (23), for arbitrary relaxation spectra Hτ and HMv of the real material, the analogous equality of the norms holds, i.e., HMv2=Hτ2. The formula describing the square norm of H¯Kv (42) directly results from the equality in (45), from which the following can be obtained:

H¯Kv2=2ααg¯KTλΘKg¯Kλ, (62)

with the matrix ΘK defined by (46).

The errors of the relaxation modulus models are estimated using index QKg¯Kλ (58) and expressed for the optimal models H¯KMv (41), H¯Kv (42), and H¯Kτ (43) by Formula (59) from Proposition 4.

The errors of the relaxation spectra models are measured directly by the integral JgK, introduced by Equation (16) for model HKMv and by Equation (20) for the relaxation time spectrum model. By (19), having in mind the notation (25), this index for the optimal model H¯KMv (41) can be expressed as follows (compare to (26)):

Jg¯Kλ=0HMv2dv2GKTg¯Kλ+1αg¯KTλΦKg¯Kλ, (63)

where GK is the vector of noise-free values of relaxation modulus defined below Equation (51). The error Jg¯Kλ for the model H¯Kτ (43) is obviously identical due to the identity of indices (16) and (20) proved above.

The error Jg¯Kλ related to square of the norm of real spectra HMv22=Hτ22 is measured by the relative index

Jrelg¯Kλ=Jg¯KλHτ22. (64)

The “real” materials and the optimal models were simulated in Matlab R2023b, The Mathworks, Inc., Natick, MA, USA. For the singular value decomposition procedure, svd was applied.

3.8. Identification of Uni-Mode Gauss-like Spectrum

Consider material whose rheological properties are characterized by the uni-modal Gauss-like distribution [10,61]:

Hτ=ϑe1τm2/q/τ, (65)

where the parameters are as follows [10,27,61]: ϑ=31.52 kPa·s, m=0.0912 s1, and q=3.25×103 s2. The relaxation modulus is displayed below [10]:

Gt=πq2ϑ e14t2qmterfc12tqmq, (66)

where the complementary error function erfcx is defined as follows ([62] Equation (8.250.4)):

erfcx=2π  xez2dz. (67)

By the first equality in (3), the following spectrum of relaxation frequencies corresponds to (65)

Hv=ϑvevm2/q, (68)

whence, in view of (4), the modified spectrum is described as outlined below:

HMv=ϑevm2/q. (69)

In Appendix A.5, the analytical Formulas (A10) and (A13) are derived describing the norms of the spectra Hτ (65), Hv (68), and HMv (69). These norms are as follows: HMv2=Hτ2=8.422432kPa·s1/2 and Hv2=0.805043kPa·s1/2.

The preliminary relaxation test experiment was performed (step 1) and the measurements of the relaxation modulus Gt (66) were recorded for 200 s, selected following [10,27,61]. Then, the time scale factors α have been selected by comparison of the courses of experiment results G¯ti and basis functions ϕkt (12) for a few k. Next, to simulate the experiment, K sampling instants tk=αk were generated with the constant period α for K=20,50,100,150,200 measurements. Additive measurement noises ztk were selected independently by random choice with uniform distribution on the interval 10, 10 Pa, i.e., double stronger than noises assumed in the previous papers [10,61]. The measurements G¯tk were recorded. For successive K, the matrix ΦK (25) and SVD (35) were determined. Next, the regularization parameters λ were selected according to the spectral condition number rule (60); their values are given in Table 3. The optimal model parameters g¯Kλ (39) and the models H¯KMv (41), H¯Kv (42), and H¯Kτ (43) were determined. The best models are depicted in Figure 5, Figure 6 and Figure 7 together with the real spectra (65), (68), and (69) marked by red lines. Small subfigures show fitting near the maximum of the real spectrum. The respective relaxation modulus models G¯Kt (57) are plotted in Figure 8 for K=20 and 200, where the measurements G¯tk of the real modulus Gt (66) are also marked; the small subfigures confirm the excellent model fit. For K=200, logarithmic time scale is used. In Table 3, the norms H¯KMv2=H¯Kτ2 (61), H¯Kv2 (62), and the norms g¯Kλ2—expressed in (49)—of the optimal model parameters are given. In addition, the integral square approximation index Jg¯Kλ (63) together with the relative index Jrelg¯Kλ (64) and the mean square approximation index QKg¯Kλ (58) are provided in Table 3.

Table 3.

For the uni-mode Gauss-like spectrum, Hτ (65), and the models H¯KMv (41), H¯Kv (42), and H¯Kτ (43): time-scale factors α; numbers of model summands K; regularization parameters λ; the model’s smoothness indices H¯KMv2=H¯Kτ2 (61) and H¯Kv2 (62); the mean square relaxation modulus approximation index QKg¯Kλ (58); norms g¯Kλ2 (49) of the model parameter vectors; the integral square approximation indices Jg¯Kλ (63); and relative index Jrelg¯Kλ (64).

K α [s] λ [s1] H¯Kτ2 [kPa·s1/2] H¯Kv2 kPa·s1/2 QKg¯Kλ kPa2 g¯Kλ2 kPa·s Jg¯Kλ kPa2·s Jrelg¯Kλ
20 4.90 3.1 × 10−6 8.194688 1.000985 3.681218 × 10−5 8.752841 × 103 3.548595 0.0500
50 3.75 9.0 × 10−6 8.167268 0.847115 3.486716 × 10−5 4.639282 × 103 2.461727 0.0347
100 3.98 7.8 × 10−6 8.193556 0.861458 3.272773 × 10−5 7.334379 × 103 2.439776 0.0344
150 3.75 5.5 × 10−6 8.248417 0.865259 3.310334 × 10−5 1.281206 × 104 2.625005 0.0370
200 4.08 9.5 × 10−6 8.173553 0.869989 3.299329 × 10−5 8.550751 × 103 2.875342 0.0405

Figure 5.

Figure 5

Uni-mode Gauss-like time relaxation spectrum Hτ (65) (solid red line) and the corresponding models H¯Kτ (43) for K measurements of the relaxation modulus corrupted by additive independent noises uniformly distributed over the interval 10, 10 Pa: (a) K=20, 50, 100; (b) K=100, 150, 200.

Figure 6.

Figure 6

Modified uni-mode Gauss-like time relaxation spectrum HMv (69) (solid red line) and the corresponding models H¯KMv (41) for K measurements of the relaxation modulus corrupted by additive independent noises uniformly distributed over the interval 10, 10 Pa: (a) K=20, 50, 100; (b) K=100, 150, 200.

Figure 7.

Figure 7

Uni-mode Gauss-like time relaxation frequency spectrum Hv (68) (solid red line) and the corresponding models H¯Kv (42) for K measurements of the relaxation modulus corrupted by additive independent noises uniformly distributed over the interval 10, 10 Pa: (a) K=20, 50, 100; (b) K=100, 150, 200.

Figure 8.

Figure 8

The measurements G¯tk of uni-mode Gauss-like time relaxation modulus Gt (66) corrupted by additive independent noises uniformly distributed over the interval 10, 10 Pa (red points) and the corresponding relaxation modulus models G¯Kt (57) for K measurements of the relaxation modulus: (a) K=20; (b) K=200.

The relative spectrum approximation index Jrelg¯Kλ (64) does not exceed 5%; additionally, the values of the relaxation modulus approximation index QKg¯Kλ (58) indicate the excellent model fit; models G¯Kt (57) practically coincide with the measurement points G¯tk, see Figure 8. An inspection of Figure 5, Figure 6 and Figure 7 shows that for the number of K50 measurements, satisfactory approximation of the relaxation spectra was obtained while maintaining the consistency of the maxima of real spectra and their models.

3.9. Identification of Double-Mode Gauss-like Spectrum

Consider now the viscoelastic material of the relaxation spectrum described by the double-mode Gauss-like distribution considered in [10,27,28,46]:

Hτ=ϑ1e1τm12/q1+ϑ2e1τm22/q2/τ, (70)

where the parameters are as follows [27,28]: ϑ1=467 Pa·s, m1=0.0037 s1, q1=1.124261×106 s2, ϑ2=39 Pa·s, m2=0.045 s1, and q2=1.173×103 s2. Therefore, the corresponding spectrum of relaxation frequencies is as follows

Hv=ϑ1vevm12/q1+ϑ2vevm22/q2, (71)

and, in view of (4), the modified spectrum is described by HMv=Hv/v. By (66), the related real relaxation modulus is outlined below:

Gt=π2ϑ1q1 e14t2q1m1terfc12tq1m1q1+ϑ2q2 e14t2q2m2terfc12tq2m2q2. (72)

In Appendix A.6, the analytical Formulas (A17) and (A19) are derived to describe the square norms of the “real” spectra Hv (71), Hτ (70), and HMv, which are as follows: HMv2=Hτ2=19.257051 Pa·s1/2 and Hv2=0.394490Pa·s1/2.

Based on the course of the modulus Gt (72), in the preliminary experiment, N=5000 sampling instants were generated with the constant period in the time interval T=0,1550 s, c.f., [27,28]. Following [27,28], additive measurement noises zti were selected independently by random choice with uniform distribution on the interval 0.005, 0.005 Pa. The same as before, several K relaxation spectra models were determined using the proposed identification algorithm. The values of selected regularization parameters λ, the norms H¯KMv2=H¯Kτ2, H¯Kv2, and g¯Kλ2 and the indices Jg¯Kλ, Jrelg¯Kλ, and QKg¯Kλ (58) are presented in Table 4. The optimal models H¯KMv (41), H¯Kv (42), and H¯Kτ (43) are depicted in Figure 9, Figure 10 and Figure 11 together with the real spectra plotted by red lines. The respective relaxation modulus models G¯Kt (57) are plotted in Figure 12 with the modulus Gt (72) measurements.

Table 4.

For the double-mode Gauss-like spectrum Hτ (70) and the models H¯KMv (41), H¯Kv (42) and H¯Kτ (43): time-scale factors α; numbers of model summands K; regularization parameters λ; the model’s smoothness indices H¯KMv2=H¯Kτ2 and H¯Kv2; the mean square relaxation modulus approximation index QKg¯Kλ (58); norms g¯Kλ2 (49) of the model parameter vectors; the integral square approximation indices Jg¯Kλ (63); and relative index Jrelg¯Kλ (64).

K α [s] λ [s1] H¯Kτ2 [Pa·s1/2] H¯Kv2 Pa·s1/2 QKg¯Kλ Pa2 g¯Kλ2 Pa·s Jg¯Kλ Pa2·s Jrelg¯Kλ
50 22.5 8 × 10−7 17.008811 0.377417 8.91256 × 10−6 2.63874 × 104 83.212916 0.224394
100 16.3 1.2 × 10−6 16.749729 0.415414 8.20297 × 10−6 2.38674 × 104 89.54064 0.241457
150 9.35 2.1 × 10−6 16.145506 0.440330 8.33036 × 10−6 1.68329 × 104 90.59845 0.244309
200 6.5 3.1 × 10−6 16.447062 0.404095 8.38748 × 10−6 1.321209 × 104 85.71129 0.2311311
300 5.2 4 × 10−6 16.650568 0.401779 8.04576 × 10−6 1.228249 × 104 87.92225 0.237093

Figure 9.

Figure 9

Double-mode Gauss-like time relaxation spectrum Hτ (70) (solid red line) and the corresponding models H¯Kτ (43) for K measurements of the relaxation modulus corrupted by additive independent noises uniformly distributed over the interval 0.005, 0.005 Pa: (a) K=50, 100, 150; (b) K=150, 200, 300.

Figure 10.

Figure 10

Modified double-mode Gauss-like time relaxation spectrum HMv related to Hv (71) (solid red line) and the corresponding models H¯KMv (41) for K measurements of the relaxation modulus corrupted by additive independent noises uniformly distributed over the interval 0.005, 0.005 Pa: (a) K=50, 100, 150; (b) K=150, 200, 300.

Figure 11.

Figure 11

Double-mode Gauss-like time relaxation frequency spectrum Hv (71) (solid red line) and the corresponding models H¯Kv (42) for K measurements of the relaxation modulus corrupted by additive independent noises uniformly distributed over the interval 0.005, 0.005 Pa: (a) K=50, 100, 150; (b) K=150, 200, 300.

Figure 12.

Figure 12

The measurements G¯tk of double-mode Gauss-like time relaxation modulus Gt (72) corrupted by additive independent noises uniformly distributed over the interval 0.005, 0.005 Pa (red points) and the corresponding relaxation modulus models G¯Kt (57) for K measurements of the relaxation modulus: (a) K=50; (b) K=300.

However, for double-mode spectrum, the relative spectrum approximation index Jrelg¯Kλ is as much as 25%, an inspection of Figure 9, Figure 10 and Figure 11 indicates a satisfactory approximation of the real spectra while maintaining the locations of both their maxima. Excellent models G¯Kt fit is confirmed by the values QKg¯Kλ and Figure 12.

3.10. Identification of KWW Relaxation Spectrum

The relaxation spectrum of the KWW model of the stretched exponential relaxation is described by the following [56]:

Gt=G0etτrβ, (73)

where the stretching exponent 0<β<1, τr is the relaxation time, and G0 is the initial shear modulus, which has a unimodal [57] relaxation spectrum described by the infinite series [56,57]:

Hτ=G0π k=11k+1k!sinπβk Γβk+1 ττrβk, (74)

where Γn is Euler’s gamma function ([62] Equation (8.310.1)). However, for the stretching exponent β=0.5, spectrum Hτ has simple analytical form [57]:

Hτ=G02πττr eτ4τr . (75)

The stretching exponent 0.5 is assumed, for which the relaxation spectrum is given by the analytical formula because the effectiveness of the identification method can only be verified when the assumed spectrum is exactly known. The exponent β=0.5 has been reported by Plazek and Ngai [63] for poly(methylphenylsiloxane) at the glass temperature Tg=200 K; however, the related relaxation time τr is not reported in [63]. In 1993, Böhmer et al. [64], based on the literature and also private communications, have presented data concerning stretched exponential relaxation from about 70 amorphous polymeric glass formers (supercooled liquids and disordered crystals). The exponent β=0.5 has been experimentally obtained for sorbitol, dehydroabietic acid, BBKDE, 1,4-cis-polyisoprene, and silicate flint glass ([64] (Table I)). The coefficients β near 0.5 have been found for toluene (25%) (β=0.52), BCDE (β=0.51), polyisobutylene (β=0.55), and several other forms of glass. However, this paper also does not contain data on the related relaxation times. Recently, Chen et al. [65] applied the KWW model to describe the viscoelastic properties of the cross-linked polystyrene estimating the following parameters of the model (73) at temperature 110 °C: G0=0.78 MPa, β=0.59, and τr=1.08 s ([65] (Table 4)). For the purpose of numerical tests of the algorithm, β=0.59 was replaced here by β=0.5. The discrepancy between the relaxation modulus Gt (73) for β=0.5 and β=0.59 is illustrated in Figure 13; the mean least-squares error for 500 equidistant sampling points between these modulus is equal to 1.2853 × 10−4 MPa2.

Figure 13.

Figure 13

The KWW relaxation modulus Gt (73) for stretching exponents β=0.5 and β=0.59; the initial shear modulus G0=0.78 MPa; and the relaxation time τr=1.08 s.

The relaxation frequency spectrum corresponding to (75) is as follows

Hv=G02π1τr·v e14 τr·v, (76)

while the modified spectrum is described by the following:

HMv=G02π1τr·v3 e14 τr·v. (77)

In the preliminary experiment, N=500 sampling instants were generated with the constant period in the time interval T=0,50 s, selected based on the course of the modulus Gt (73) in Figure 13. Additive measurement noises zti were selected independently by random choice with uniform distribution on the interval 0.5, 0.5 kPa. The values of the time scale factors α selected for a few values of K and the regularization parameters λ selected according to the rule (60) are given in Table 5. The best models H¯KMv (41), H¯Kv (42), and H¯Kτ (43) are depicted in Figure 14, Figure 15 and Figure 16 together with the real spectra (red lines). The respective models G¯Kt (57) are plotted in Figure 17, with the modulus Gt (73) measurements. In Table 5, the norms H¯KMv2=H¯Kτ2, H¯Kv2, and the norms g¯Kλ2 of the optimal parameters g¯Kλ (39) are given. Since from (75) and (A3) we have the following:

Hτ22=G024π τr0τeτ2τr dτ=G02τrπ ,

the norm Hτ2=HMv2=G0τr/π =0.457331 MPa·s1/2. In turn, the norm of the relaxation frequency spectrum (76) is infinite; however, in Table 5, the model’s norms H¯Kv2 are given. The integral indices Jg¯Kλ (63), the relative index Jrelg¯Kλ, (64) and the mean square relaxation modulus approximation index QKg¯Kλ (58) are also presented in Table 5.

Table 5.

For the KWW spectrum Hτ (75) and the models H¯KMv (41), H¯Kv (42), and H¯Kτ (43): time-scale factors α; numbers of model summands K; regularization parameters λ; the model’s smoothness indices H¯KMv2=H¯Kτ2 and H¯Kv2; the mean square relaxation modulus approximation index QKg¯Kλ (58); norms g¯Kλ2 (49) of the model parameter vectors; the integral square approximation indices Jg¯Kλ (63); and relative index Jrelg¯Kλ (64).

K α[s] λ[s1] H¯Kτ2[MPa·s1/2] H¯Kv2MPa·s1/2 QKg¯KλMPa2 g¯Kλ2MPa·s Jg¯KλMPa2·s Jrelg¯Kλ
25 0.8 2 × 10−5 0.454723 0.267896 9.46218 × 10−8 76.901664 1.83447 × 10−3 8.77098 × 10−3
50 0.65 7 × 10−5 0.456956 0.284479 8.70335 × 10−8 29.800961 9.82909 × 10−4 4.69948 × 10−3
75 0.6 7.5 × 10−5 0.4571396 0.289756 8.52022 × 10−8 33.705039 8.50259 × 10−4 4.06526 × 10−3
100 0.65 8.5 × 10−5 0.456765 0.284243 8.18454 × 10−8 33.6572199 8.26024 × 10−4 3.94938 × 10−3
150 0.6 1 × 10−4 0.457356 0.289670 8.28962 × 10−8 35.262493 8.41199 × 10−4 4.02194 × 10−3
200 0.6 1.5 × 10−4 0.456657 0.289207 8.25479 × 10−8 27.087999 7.34565 × 10−4 3.51210 × 10−3
300 0.55 1.6 × 10−4 0.456875 0.294931 7.97708 × 10−8 30.574734 7.12159 × 10−4 3.40497 × 10−3
400 0.55 1.6 × 10−4 0.456791 0.294901 8.10268 × 10−8 35.581507 7.12571 × 10−4 3.40694 × 10−3

Figure 14.

Figure 14

The KWW spectrum Hτ (75) (solid red line) and the corresponding models H¯Kτ (43) for K measurements of the relaxation modulus corrupted by additive independent noises uniformly distributed over the interval 0.5, 0.5 kPa: (a) K=25, 50, 75; (b) K=100, 150, 200.

Figure 15.

Figure 15

Modified KWW spectrum HMv (77) (solid red line) and the corresponding models H¯KMv (41) for K measurements of the relaxation modulus corrupted by additive independent noises uniformly distributed over the interval 0.5, 0.5 kPa: (a) K=25, 50, 75; (b) K=100, 150, 200.

Figure 16.

Figure 16

The KWW spectrum relaxation frequency spectrum Hv (76) (solid red line) and the corresponding models H¯Kv (42) for K measurements of the relaxation modulus corrupted by additive independent noises uniformly distributed over the interval 0.5, 0.5 kPa: (a) K=25, 50, 75; (b) K=100, 150, 200.

Figure 17.

Figure 17

The measurements G¯tk of the KWW relaxation modulus Gt (73) corrupted by additive independent noises uniformly distributed over the interval 0.5, 0.5 kPa (red points) and the corresponding relaxation modulus models G¯Kt (57) for K measurements of the relaxation modulus: (a) K=25; (b) K=400.

The relative integral square index of the spectra approximation Jrelg¯Kλ does not exceed 0.5% for K50 measurements, which means a better approximation of the assumed relaxation spectrum in the whole range of time/frequency relaxation variation, i.e., from zero to infinity, than in the case of Gaussian spectra. Also related to the relaxation modulus index, QKg¯Kλ, not exceeding 10−7, confirms the perfect approximation of the relaxation modulus measurements. In the case of this unimodal spectrum, increasing the number of measurements, i.e., the components of the series that create the models, does not significantly affect the quality of these models, which, in addition to the indices in Table 5, is also confirmed by a review of Figure 14, Figure 15 and Figure 16. For K100, the courses of the spectra models for increasing K remain practically almost identical, although a slight improvement in the fit to the real spectra can be seen in the values of the indices Jg¯Kλ and Jrelg¯Kλ. The relative index of the spectrum Hτ approximation for K100 falls below 0.41%.

3.11. Applicability of the Approach for Identification of Relaxation Spectra of Different Types

The rough condition of the approach’s successful applicability follows from the boundary properties of the optimal models H¯Kτ (43) and H¯Kv (42), yielded by the properties of the basis functions hkτ (15) and hkvv, where hkv is given by (7). Since for τ0+ and τ, the basis functions hkτ0, the best model H¯Kτ (43) also tends to zero as the relaxation time τ tends to zero and to infinity, which limit the scope of applicability of this model to real relaxation time spectra that satisfy zero boundary conditions. For the relaxation frequencies v=0 and v, the basis functions hkvv0, that is the basis functions hkvv of the relaxation frequency model H¯Kv (42) also have zero boundary conditions. Therefore, in terms of the relaxation frequency, the scope of applicability of the model and method to real relaxation frequency spectra is confined to the spectra of zero boundary conditions, too.

The real relaxation time and frequency spectra and the known spectra models tend to zero as the relaxation time τ and the relaxation frequency v tend to infinity. Therefore, the properties of the spectra for τ0+ and v0+ are essential here.

The examples presented above showed that the approach proposed can be applied for Gauss-like relaxation spectra, both uni- and double mode, and for the KWW spectrum of the stretching exponent β=0.5. However, it is easy to check that for the relaxation spectrum Hτ (74) both zero boundary conditions are satisfied. Therefore, the proposed identification method can be applied to determine the spectrum of materials whose relaxation processes have KWW stretched exponential nature. This is also important that the optimal model H¯Kτ (43), given by a finite series, may prove to be more useful than the original KWW infinite series spectrum (74) for many applications.

A multiplicative model that combines the power law with the stretched exponential relaxation described by Equation (8) in [66]:

Hτ=nαGcτταnα eτταβ, (78)

where τα is the longest relaxation time, Gc is the plateau modulus, the stretching parameter 0<β1, and the exponent 0<nα<1, was applied for modeling spectrum of bitumen in the vicinity of the glass transition [66]. The unimodal spectrum (78), named by the authors as the broadened power-law spectrum model [66], satisfies both zero boundary conditions—compare to ([66] (Figure 11a))—and therefore, is also within the scope of the proposed algorithm’s applicability.

However, the well-known Baumgaertel, Schausberger, and Winter (BSW) spectrum [15,67] used to describe the viscoelasticity of polybutadiene (PBD) [68], polydisperse polymer melts [8], polymethylmethacrylate (PMMA) [68], and many other materials, is described by the following model:

Hτ=β1ττcρ1+β2ττcρ2eττmax,

with positive coefficients β1, β2 and relaxation times τc, τmax, which tends to infinity for τ0 whenever at least one of the parameters ρ1 and ρ2 is negative. Therefore, this is the case for real material models, compare [8,15,67,68], when the optimal model H¯Kτ (43) cannot well-approximate this spectrum.

Likewise, the real relaxation spectra modeled by pure inverse power laws [69], for example, a combined four-interval power model with fractional exponents describing a solution-polymerized styrene butadiene rubber [70] or a power type spectrum with an exponent of −1/2 describing the cross-linking polymers at their gel point [71], cannot be successfully identified by the proposed approach. The relaxation time spectra of the fractional Maxwell model and the elementary fractional Scott–Blair model also lose the zero boundary condition at zero relaxation time, see [61] (Proposition 2, Equation (19)).

3.12. Direct Identification of the Relaxation Spectra of Viscoelastic Solid Materials

For isotropic viscoelastic solids [31]

limtGt=G>0,

where G is the material equilibrium modulus. Then, Equation (1) takes the form presented below [31]:

Gt=0Hττet/τdτ+G. (79)

Analogously, Equation (5)—basic for the direct approach and related to the modified frequency spectrum HMv—can be rewritten as follows:

Gt=0HMvetvdv+G. (80)

The relaxation spectra models HKMv (9), HKv (10), and HKτ (14) do not require modification, while the related relaxation modulus model GKt (11) should be replaced by the following:

GKt=0HKMvetvdv+G=k=1Kgkϕkt+G,

which, however, does not affect the identification procedure itself.

The square integral index JgK, given by Equation (16) for the model HKMv and by (20) for HKτ, is defined as above. However, by (80) and (7) we have:

0HMvhkvdv=0HMeαkvdv=GαkG=Gαk. (81)

Therefore, by (17) and (81), the index JgK is given by the following expression (compare to (19)):

JgK=0HMv2dv2k=1KgkGαk+k=1Km=1Kgkgmφkm.

As above, from (79) and (20), the analogous formula results for the relaxation time spectrum model. As a consequence, the integral-empirical index J¯KgK (24) is now as follows

J¯KgK=0HMv2dv2k=1KgkG¯αk+k=1Km=1Kgkgmφkm,

where, compare Gαk (81), the relaxation modulus increment is defined as follows:

G¯αk=G¯αkG. (82)

Since real materials may relax over a very long time, two cases can occur.

  • Case 1. If the duration of the relaxation test can be extended so as to experimentally record a time-constant relaxation modulus (in practice, constant stress), then G is experimentally evaluated and the proposed identification algorithm can be simply applied by replacing the measurements G¯αk with their increments G¯αk (82) in relation to known G.

  • Case 2. For identification purposes, only time-varying relaxation modulus measurements are available, i.e., the steady-state stress was not recorded during the experiment. In such a situation, non-negative G is an additional model parameter that should be extrapolated beyond the experiment time horizon limited by the upper bound tK=αK. The linear-quadratic problem (32) of optimal identification needs to be reformulated, re-regularized and solved, which creates a new research problem.

However, for many materials, the equilibrium modulus is accessible by experiment; then, the algorithm of direct relaxation spectra identification can be applied with the simple modification as described above.

4. Conclusions

Summarizing, this paper addresses the relaxation spectrum identification problem in a new original way. The novelty of the paper is that it directly takes into account the unknown spectrum in the model quality index being minimized. The main result is based only on the definition of the relaxation spectrum, which relates the spectrum to the measurable relaxation modulus, and on the fact that the set of exponential functions, i.e., a kernel of the Lagrange transform constitute a basis of the space of square-integrable functions. The analytical and numerical studies demonstrated that by applying the proposed relaxation spectra models and identification algorithm, it is possible to determine the spectra models for a wide range of relaxation times and frequencies of real materials.

The concept of direct relaxation spectrum identification can be applied both for viscoelastic fluids and viscoelastic solids; however, for solid materials, a respective modification of the algorithm may be required whenever the equilibrium relaxation modulus is not available by measurement, the development of which will be the subject of further research.

It is generally accepted that the choice of respective regularization parameters is important to identify the best model. The well-studied techniques for computing a good regularization parameter such as the discrepancy principle, generalized cross-validation, and the L-curve technique have been developed for classical least-squares task and hence they cannot be directly applied here. Therefore, the regularized minimization problem (32) should be reformulated to the classic form of the linear least-squares problem. Then, the applicability of the known techniques can be verified. An alternative approach is to develop a new method of selecting the regularization parameter, specifically addressing the problem of direct spectrum identification. Although the numerical studies have shown that the simple rule based on the condition number of the basic matrix for the linear-quadratic identification problem is sufficient in many cases, the example of a two-mode Gaussian-like spectrum motivates the search for a better rule for the regularization parameter selection, dedicated for this specific identification task. This will be the subject of further research.

The impact of the molecular weight distributions (MWD) on the viscoelastic properties is intensively studied in polymer rheology. Generic analytical formulas describing the relationship between MWD and the relaxation time spectrum are known. Future research directions may include the determination of the MWD, which can be obtained from the relaxation time spectrum model and recovered from experimental results by the proposed method.

Appendix A

Appendix A.1. Proof of Lemma 1

According to (6), (17) and (25), the quadratic form gKTΦKgK is expressed as follows:

gKTΦKgK=αk=1Km=1Kgkgm1k+mα=α0k=1Kgkeαkv2dv.

Thus, gKTΦKgK0 for an arbitrary vector gK, and gKTΦKgK=0, if and only if k=1Kgkeαkv=0 for almost all v>0. Since the basis functions hkv,α=eαkv are independent, the last equality holds, if and only if gk=0 for all k=1,,K, i.e., only if the vector gK=0, which yields the positive definiteness of ΦK. This finishes the proof. □

Appendix A.2. Proof of Proposition 1

For the model H¯KMv (41) of the modified spectrum HMv (4), by (18), we have the following:

H¯KMv22=0H¯KMv2dv=k=1Km=1Kg¯kλgmφkm=1αg¯KTλΦKg¯Kλ, (A1)

with the vector of model parameters g¯Kλ (34). Similarly, (43) and (18), yield

H¯Kτ22=0H¯Kτ2dτ=k=1Km=1Kg¯kλg¯mλφkm=1αg¯KTλΦKg¯Kλ. (A2)

By the following the integral formula ([62] Equation (3.351.3))

0τneβτdτ=n!βn+1, (A3)

for the model H¯Kv (42) of the real spectrum Hv we have the following:

H¯Kv22=0H¯Kv2dv=2α3k=1Km=1Kg¯kλg¯mλ1k+m3=2α3g¯KTλΘKg¯Kλ, (A4)

where the K×K positive definite (the proof is analogous to that of Lemma 1) matrix ΘK of the elements θkm=1k+m3 is defined by Equation (46). The equalities in (44) and (45) are proved.

According the known Rayeigh–Ritz inequalities [72], (Lemma I):

λminXxTxxTXxλmaxXxTx, (A5)

which holds for any xRm and any symmetric matrix X=XTRm,m, where λminX and λmaxX are minimal and maximal eigenvalues of the matrix X. Since for positive definite ΦK and ΘK their eigenvalues are identical to the singular values [39] (p. 77), in view of (A5) Equations (A1), (A2), and (A4), imply the lower and upper bounds in (44) and (45). Proposition is proved. □

Appendix A.3. Proof of Proposition 3

The error between the spectra H¯KMv (41) and H~KMv (50) is given by the formula below:

H¯KMvH~KMv=k=1Kg¯kλg~kλhkv,

therefore, the integral square error between these spectra is as follows:

H¯KMvH~KMv22=0k=1Kg¯kλg~kλhkv2dv,

and, in view of (18), is described by the next formula:

H¯KMvH~KMv22=k=1Km=1Kg¯kλg~kλg¯mλg~mλφkm,

whence, having in mind the notation (25), the quadratic form is obtained

H¯KMvH~KMv22=1αg¯Kλg~KλTΦKg¯Kλg~Kλ. (A6)

By (51) and (39)

g¯Kλg~Kλ=αUKΩKUKTzN (A7)

with the vector of the measurement noises zN=G¯KGK, which, substituted into (A6) and combined with the SVD (35), yields

H¯KMvH~KMv22=αzNTUKΩKΣKΩKUKTzN.

Diagonal structure of the matrices ΣK (36) and ΩK (38) implies the structure of the next matrix

ΩKΣKΩK=diagσ1σ1+αλ2,,σKσK+αλ2,

whence, by the right inequality in (A5) and since for orthogonal UK we have zNTUKUKTzN=zNTzN, the next upper bound is obtained

H¯KMvH~KMv22αmax1kKσkσk+αλ2zNTzN,

whence the inequality (54) with parameter γ (55) for the models H¯KMv and H~KMv directly follows.

Similarly, for the spectra H¯Kτ (43) and H~Kτ (53) we have the following:

H¯KτH~Kτ=k=1Kg¯kλg~kλhkτ,

whence, having in mind (22), we obtain

H¯KτH~Kτ22=1αg¯Kλg~KλTΦKg¯Kλg~Kλ,

that is, this norm is identical to (A6); the second inequality in (54) follows.

Finally, for the spectra H¯Kv (42) and H~Kv (52) we have the following:

H¯KvH~Kv=k=1Kg¯kλg~kλhkvv,

whence, having in mind the matrix ΘK introduced in (A4), we obtain

H¯KvH~Kv22=2α3g¯Kλg~KλTΘKg¯Kλg~Kλ,

and next, by (A7),

H¯KvH~Kv22=2αzNTUKΩKUKTΘKUKΩKUKTzN.

By applying the right inequality in (A5) and including the orthogonality of UK we have:

H¯KvH~Kv222ας1zNTUKΩKΩKUKTzN.

whence in view of the structure of the matrix ΩK (38) we immediately obtain

H¯KvH~Kv222 ς1ασK+αλ2zNTzN,

which implies (56) and completes the proof. □

Appendix A.4. Proof of Proposition 4

Since for any tk=αk and any m, by (12) and (18), we have the following:

ϕmtk=1αk+αm=φkm=φmk,

the value of the relaxation modulus model G¯Kt (57) for t=tk=αk can be described by the equation below:

G¯Ktk=m=1Kg¯mλϕmtk=m=1Kg¯mλφkm=m=1Kg¯mλφmk.

Therefore, index QKg¯Kλ (58) can be expressed as follows:

QKg¯Kλ=1Kk=1KG¯tk2+1Kk=1Km=1Kg¯mλφmkφkmg¯kλ2Kk=1Km=1Kg¯mλφmkG¯tk,

whence, due to (22) and (25), i.e., having in mind that elements of the matrix ΦK are equal to αφkm, the equivalent matrix-vector form follows

QKg¯Kλ=1KG¯KTG¯K+1K1α2g¯KTλΦKΦKg¯Kλ2K1αG¯KTΦKg¯Kλ,

which in compact form is given by the following:

QKg¯Kλ=1KG¯K1αΦKg¯KλTG¯K1αΦKg¯Kλ.

The first equality in (59) is derived.

By the SVD (35), including Formula (39), the above can be rewritten as outlined below:

QKg¯Kλ=1KG¯K1αUKΣKUKTUKΩKYKTG¯K1αUKΣKUKTUKΩKYK;

whence, due to orthogonality of UK, we obtain

QKg¯Kλ=1KG¯KTG¯K2YKTΣKΩKYK+YKTΩKΣKΣKΩKYK.

The diagonal structure of the matrices ΣK (36) and ΩK (38) yields

ΣKΩK=diagσ1σ1+αλ,,σKσK+αλ,

whence, remembering that YK=UKTG¯K (40), we have the following:

QKg¯Kλ=k=1K12σkσk+αλ+σk2σk+αλ2yk2,

and, after algebraic manipulations, equivalently,

QKg¯Kλ=k=1Kαλ2yk2σk+αλ2.

whence second Equation in (59) directly follows. □

Appendix A.5. Norms of the Spectra Hτ (65), Hv (68), and HMv (69)

By (69),

HMv22=0HMv2dv=ϑ20e2vm2/qdv,

which can be written as follows:

HMv22=ϑ2e2m2/q0e2v2/q+4vm/qdv.

Therefore, using the known integral ([62] Equation (3.322.2)),

0ex24βχxdx=πβ eβχ2erfcχβ (A8)

and lying β=q/8 and χ=4m/q, we immediately obtain

HMv22=ϑ2πq/22 erfcmq/2, (A9)

whence and in view of the equality of the norms HMv2=Hτ2, the next equality follows

HMv2=Hτ2=ϑπq/242 erfcmq/2. (A10)

By (68), we have the following:

Hv22=0Hv2dv=ϑ20v2e2vm2/qdv,

which, to facilitate determination of the integral, can be rewritten as follows

Hv22=ϑ2e2m2/q0v2e2v2/q+4vm/qdv.

Whence, using the known integral ([62] Equation (3.462.7))

0x2eμx22χxdx=χ2μ2+πμ5 2χ2+μ4eχ2/μerfcχμ, (A11)

by lying μ=2/q and χ=2m/q, we immediately obtain

Hv22=14ϑ2mq e2m2/q+πq2 4m2+q2erfc2mq, (A12)

whence

Hv2=ϑ2qπ24m2+q2erfc2 mq+mq e2m2/q. (A13)

Appendix A.6. Norms of the Double-Mode Gauss Spectra Hτ (70), Hv (71), and HMv

We obtain the desired result presenting the spectrum Hv (71) as sum of two uni-mode Gauss spectra

Hv=ϑ1vevm12/q1+ϑ2vevm22/q2=H1v+H2v. (A14)

Therefore, we have the following:

Hv22=H1v22+H2v22+20H1vH2vdv, (A15)

where

0H1vH2vdv=ϑ1ϑ2ea0v2eq¯v2e2vm¯dv,

with the parameters m¯, q¯ and a defined by the following formula:

m¯=m1q1+m2q2, q¯=1q1+1q2, a=m12q1+m22q2. (A16)

Therefore, by (A11), lying μ=2/q and χ=2m/q, we immediately obtain

0H1vH2vdv=ϑ1ϑ2eam¯2q¯2+πq¯5 2m¯2+q¯4em¯2/q¯erfcm¯q¯,

which combined with Formula (A12) applied for H1v and H2v, in view of (A15) yields

Hv22=14ϑ12m1q1 e2m12/q1+πq12 4m12+q12erfc2m1q1+14ϑ22m2q2 e2m22/q2+πq22 4m22+q22erfc2m2q2+2ϑ1ϑ2eam¯2q¯2+πq¯5 2m¯2+q¯4em¯2/q¯erfcm¯q¯. (A17)

By (A14) and (4)

HMv=ϑ1evm12/q1+ϑ2evm22/q2=H1Mv+H2Mv.

Therefore, as above,

HMv22=H1Mv22+H2Mv22+20H1MvH2Mvdv, (A18)

where

0H1MvH2Mvdv=ϑ1ϑ2ea0eq¯v2e2vm¯dτ,

with the parameters m¯, q¯ and a defined by (A16), which, by Equation (A8), lying β=14q¯ and χ=2m¯, can be expresses as follows:

0H1MvH2Mvdv=12ϑ1ϑ2eaπq¯ em¯2q¯erfcm¯1q¯.

Substituting the above into (A18) and combining with (A9) applied for the partial spectra, H1Mv and H2Mv, we obtain

HMv22=ϑ12πq1/22 erfcm1q1/2+ϑ22πq2/22 erfcm2q2/2+ϑ1ϑ2eaπq¯ em¯2q¯erfcm¯1q¯,

whence, for Hτ2=HMv2, the next formula follows

Hτ2=πq1/2 ϑ122 erfcm1q1/2+πq2/2 ϑ222 erfcm2q2/2+ϑ1ϑ2πq¯ em¯2q¯aerfcm¯q¯12 (A19)

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article; further inquiries can be directed to the corresponding author.

Conflicts of Interest

The author declares no conflicts of interest.

Funding Statement

The cost was partially incurred from funds financed by the IDUB University Development Strategy for 2024–2026 in the discipline of Mechanical Engineering as part of the task “Stage: 1, payment from funds: SUBB.RNN.24.019”.

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.Macey H.H. On the Application of Laplace Pairs to the Analysis of Relaxation Curves. J. Sci. Instrum. 1948;25:251–253. doi: 10.1088/0950-7671/25/7/323. [DOI] [Google Scholar]
  • 2.Sips R. Mechanical behavior of viscoelastic substances. J. Polym. Sci. 1950;5:69–89. doi: 10.1002/pol.1950.120050103. [DOI] [Google Scholar]
  • 3.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]
  • 4.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]
  • 5.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]
  • 6.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]
  • 7.Babaei B., Davarian A., Pryse K.M., Elson E.L., Genin G.M. Efficient and optimized identification of generalized Maxwell viscoelastic relaxation spectra. J. Mech. Behav. Biomed. Mater. 2016;55:32–41. doi: 10.1016/j.jmbbm.2015.10.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.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]
  • 9.Cho G., Choi J., Lee J., Cho K.S. Application of Post–Widder inversion formula to the calculation of relaxation spectrum from relaxation modulus. Korea-Aust. Rheol. J. 2024;36:79–88. doi: 10.1007/s13367-023-00086-7. [DOI] [Google Scholar]
  • 10.Stankiewicz A. Robust Recovery of Optimally Smoothed Polymer Relaxation Spectrum from Stress Relaxation Test Measurements. Polymers. 2024;16:2300. doi: 10.3390/polym16162300. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.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]
  • 12.Syed Mustapha S.M.F.D., Phillips T.N. A dynamic nonlinear regression method for the determination of the discrete relaxation spectrum. J. Phys. D Appl. Phys. 2000;33:1219. doi: 10.1088/0022-3727/33/10/313. [DOI] [Google Scholar]
  • 13.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]
  • 14.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]
  • 15.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]
  • 16.Sun Y., Huang B., Chen J. A unified procedure for rapidly determining asphalt concrete discrete relaxation and retardation spectra. Constr. Build. Mater. 2015;93:35–48. doi: 10.1016/j.conbuildmat.2015.04.055. [DOI] [Google Scholar]
  • 17.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]
  • 18.Zhang F., Wang L., Li C., Xing Y. The Discrete and Continuous Retardation and Relaxation Spectrum Method for Viscoelastic Characterization of Warm Mix Crumb Rubber-Modified Asphalt Mixtures. Materials. 2020;13:3723. doi: 10.3390/ma13173723. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Malkin A.Y. Continuous relaxation spectrum–its advantages and methods of calculation. [(accessed on 25 August 2024)];Int. J. Appl. Mech. Eng. 2006 11:235–243. Available online: http://www.ijame.uz.zgora.pl/ijame_files/archives/v11PDF/n2/235-243_Article_02.pdf. [Google Scholar]
  • 20.Friedrich C., Loy R.J., Anderssen R.S. Relaxation time spectrum molecular weight distribution relationships. Rheol. Acta. 2009;48:151–162. doi: 10.1007/s00397-008-0314-z. [DOI] [Google Scholar]
  • 21.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]
  • 22.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]
  • 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.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]
  • 25.Kontogiorgos V. Calculation of Relaxation Spectra from Stress Relaxation Measurements. In: Elnashar M., editor. Biopolymers. Sciyo; London, UK: 2010. [DOI] [Google Scholar]
  • 26.Stankiewicz A. A Class of Algorithms for Recovery of Continuous Relaxation Spectrum from Stress Relaxation Test Data Using Orthonormal Functions. Polymers. 2023;15:958. doi: 10.3390/polym15040958. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Stankiewicz A. Two-Level Scheme for Identification of the Relaxation Time Spectrum Using Stress Relaxation Test Data with the Optimal Choice of the Time-ScaleFactor. Materials. 2023;16:3565. doi: 10.3390/ma16093565. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Stankiewicz A., Bojanowska M., Drozd P. On Recovery of a Non-Negative Relaxation Spectrum Model from the Stress Relaxation Test Data. Polymers. 2023;15:3464. doi: 10.3390/polym15163464. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Lv H., Ye W., Tan Y., Zhang D. Inter-conversion of the generalized Kelvin and generalized Maxwell model parameters via a continuous spectrum method. Constr. Build. Mater. 2022;351:128963. doi: 10.1016/j.conbuildmat.2022.128963. [DOI] [Google Scholar]
  • 30.Bachman G., Narici L. Functional Analysis. Academic Press; New York, NY, USA: 1966. [Google Scholar]
  • 31.Christensen R.M. Theory of Viscoelasticity: An Introduction. 2nd ed. Academic Press; New York, NY, USA: 2012. [Google Scholar]
  • 32.Ferry J.D. Viscoelastic Properties of Polymers. 3rd ed. John Wiley & Sons; New York, NY, USA: 1980. [Google Scholar]
  • 33.Malkin A.I.A., Malkin A.Y., Isayev A.I. Rheology: Concepts, Methods and Applications. ChemTec; Deerfield Beach, FL, USA: 2006. [(accessed on 15 June 2024)]. Available online: https://books.google.pl/books?id=8rGafjhgz-UC. [Google Scholar]
  • 34.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]
  • 35.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]
  • 36.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]
  • 37.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]
  • 38.Rao M.A. Rheology of Fluid, Semisolid, and Solid Foods: Principles and Applications. Springer; New York, NY, USA: 2013. [(accessed on 25 August 2024)]. Available online: https://books.google.pl/books?id=9-23BAAAQBAJ. [Google Scholar]
  • 39.Golub G.H., Van Loan C.F. Matrix Computations. Johns Hopkins University Press; Baltimore, MD, USA: 2013. [Google Scholar]
  • 40.Hansen P.C. Rank-Deficient and Discrete Ill-Posed Problems. Society for Industrial and Applied Mathematics; Philadelphia, PA, USA: 1998. [DOI] [Google Scholar]
  • 41.Tikhonov A.N., Arsenin V.Y. Solutions of Ill-Posed Problems. John Wiley & Sons; New York, NY, USA: 1977. [Google Scholar]
  • 42.Maćkiewicz A. Linear Algebra Algorithms. Direct Methods. Poznań University of Technology Publishing House; Poznań, Poland: 2002. (In Polish) [Google Scholar]
  • 43.Davies A.R., Anderssen R.S., de Hoog F.R., Goulding N.J. Derivative spectroscopy and the continuous relaxation spectrum. J. Non-Newton. Fluid Mech. 2016;233:107–118. doi: 10.1016/j.jnnfm.2016.01.018. [DOI] [Google Scholar]
  • 44.Muzeau E., Perez J., Johari G.P. Mechanical spectrometry of the beta-relaxation in poly(methyl methacrylate) Macromolecules. 1991;24:4713–4723. doi: 10.1021/ma00016a036. [DOI] [Google Scholar]
  • 45.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]
  • 46.Kwakye-Nimo S., Inn Y., Yu Y., Wood-Adams P.M. Linear viscoelastic behavior of bimodal polyethylene. Rheol. Acta. 2022;61:373–386. doi: 10.1007/s00397-022-01340-5. [DOI] [Google Scholar]
  • 47.Cirillo G., Spizzirri U.G., Iemma F. Functional Polymers in Food Science: From Technology to Biology, Volume 1: Food Packaging. Wiley; New York, NY, USA: 2015. [(accessed on 20 August 2024)]. Available online: https://onlinelibrary.wiley.com/doi/book/10.1002/9781119108580. [Google Scholar]
  • 48.Lorenzo G., Checmarev G., Zaritzky N., Califano A. Linear viscoelastic assessment of cold gel-like emulsions stabilized with bovine gelatin. LWT—Food Sci. Technol. 2011;44:457–464. doi: 10.1016/j.lwt.2010.08.023. [DOI] [Google Scholar]
  • 49.Carrillo-Navas H., Hernández-Jaimes C., Utrilla-Coello R.G., Meraz M., Vernon-Carter E.J., Alvarez-Ramirez J. Viscoelastic relaxation spectra of some native starch gels. Food Hydrocoll. 2014;37:25–33. doi: 10.1016/j.foodhyd.2013.10.023. [DOI] [Google Scholar]
  • 50.Bardet S., Gril J. Modelling the transverse viscoelasticity of green wood using a combination of two parabolic elements. C. R. Mécanique. 2002;330:549–556. doi: 10.1016/S1631-0721(02)01503-6. [DOI] [Google Scholar]
  • 51.Ravindra V.G., Sasikumar K.V. Synthesis and characterization of xanthan gum stabilized polyvinyl acetate-based wood adhesive. Polym. Bull. 2023;81:7423–7440. [Google Scholar]
  • 52.de Gennes P.-G. Relaxation Anomalies in Linear Polymer Melts. Macromolecules. 2002;35:3785–3786. doi: 10.1021/ma012167y. [DOI] [Google Scholar]
  • 53.Roksana Winkler R., Unni A.B., Tu W., Chat K., Adrjanowicz K. On the Segmental Dynamics and the Glass Transition Behavior of Poly(2-vinylpyridine) in One- and Two-Dimensional Nanometric Confinement. J. Phys. Chem. B. 2021;125:5991–6003. doi: 10.1021/acs.jpcb.1c01245. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Sasaki N., Nakayama Y., Yoshikawa M., Enyo A. Stress relaxation function of bone and bone collagen. J. Biomech. 1993;26:1369–1376. doi: 10.1016/0021-9290(93)90088-V. [DOI] [PubMed] [Google Scholar]
  • 55.Giz A.S., Aydelik-Ayazoglu S., Catalgil-Giz H., Bayraktar H., Alaca B.E. Stress relaxation and humidity dependence in sodium alginate-glycerol films. J. Mech. Behav. Biomed. Mater. 2019;100:103374. doi: 10.1016/j.jmbbm.2019.103374. [DOI] [PubMed] [Google Scholar]
  • 56.Wu J., Jia Q. The heterogeneous energy landscape expression of KWW relaxation. Sci. Rep. 2016;6:20506. doi: 10.1038/srep20506. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Anderssen R.S., Saiful A., Husain R., Loy J. The Kohlrausch function: Properties and applications. In: Crawford J., Roberts A.J., editors. Proceedings of the 11th Computational Techniques and Applications Conference CTAC-2003. Volume 45. ANZIAM Journal; Callaghan, NSW, Australia: 2004. pp. C800–C816. [DOI] [Google Scholar]
  • 58.Qiao J.C., Pelletier J.M. Enthalpy relaxation in Cu46Zr45Al7Y2 and Zr55Cu30Ni5Al10 bulk metallic glasses by differential scanning calorimetry (DSC) Intermetallics. 2011;19:9–18. doi: 10.1016/j.intermet.2010.08.042. [DOI] [Google Scholar]
  • 59.Haruyama O., Sakagami H., Nishiyama N., Inoue A. The free volume kinetics during structural relaxation in bulk Pd-P based metallic glasses. Mater. Sci. Eng. A. 2007;449–451:497–500. doi: 10.1016/j.msea.2006.02.325. [DOI] [Google Scholar]
  • 60.Taghvaei A.H., Shirazifard N.G., Ramasamy P., Bednarčik J., Eckert J. Thermal behavior, structural relaxation and magnetic study of a new Hf-microalloyed Co-based glassy alloy with high thermal stability. J. Alloys Compd. 2018;748:553–560. doi: 10.1016/j.jallcom.2018.03.199. [DOI] [Google Scholar]
  • 61.Stankiewicz A. On Applicability of the Relaxation Spectrum of Fractional Maxwell Model to Description of Unimodal Relaxation Spectra of Polymers. Polymers. 2023;15:3552. doi: 10.3390/polym15173552. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Jeffrey A., Zwillinger D., Gradshteyn I.S., Ryzhik I.M. Table of Integrals, Series, and Products. 7th ed. Academic Press; New York, NY, USA: 2007. [Google Scholar]
  • 63.Plazek D.J., Ngai K.L. Correlation of polymer segmental chain dynamics with temperature-dependent time-scale shifts. Macromolecules. 1991;24:1222–1224. doi: 10.1021/ma00005a044. [DOI] [Google Scholar]
  • 64.Böhmer R., Ngai K.L., Angell C.A., Plazek D.J. Nonexponential relaxations in strong and fragile glass formers. J. Chem. Phys. 1993;99:4201–4209. doi: 10.1063/1.466117. [DOI] [Google Scholar]
  • 65.Chen S., Yang Y., Wan X., Zhang X., Li G., Liu G. Modeling the viscoelastic behavior in the frequency domain of crosslinked polystyrene with different degrees of crosslinking from the perspective of relaxation. Mech. Time-Depend Mater. 2023;27:743–763. doi: 10.1007/s11043-023-09610-2. [DOI] [Google Scholar]
  • 66.Laukkanen O.V., Winter H.H., Soenen H., Seppäla J. An empirical constitutive model for complex glass-forming liquids using bitumen as a model material. Rheol. Acta. 2018;57:57–70. doi: 10.1007/s00397-017-1056-6. [DOI] [Google Scholar]
  • 67.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]
  • 68.Choi J., Cho K.S., Kwon M.K. Self-Similarity and Power-Law Spectra of Polymer Melts and Solutions. Polymers. 2022;14:3924. doi: 10.3390/polym14193924. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Bonfanti A., Kaplan J.L., Charras G., Kabla A. Fractional viscoelastic models for power-law materials. Soft Matter. 2020;16:6002–6020. doi: 10.1039/D0SM00354A. [DOI] [PubMed] [Google Scholar]
  • 70.Saphiannikova M., Toshchevikov V., Gazuz I., Petry F., Westermann S., Heinrich G. Multiscale Approach to Dynamic-Mechanical Analysis of Unfilled Rubbers. Macromolecules. 2014;47:4813–4823. doi: 10.1021/ma501159u. [DOI] [Google Scholar]
  • 71.Winter H.H., Chambon F. Analysis of Linear Viscoelasticity of a Crosslinking Polymer at the Gel Point. J. Rheol. 1986;30:367–382. doi: 10.1122/1.549853. [DOI] [Google Scholar]
  • 72.Lee C.-H. Upper and lower matrix bounds of the solution for the discrete Lyapunov equation. IEEE Trans. Autom. Control. 1996;41:1338–1341. doi: 10.1109/9.536505. [DOI] [Google Scholar]

Associated Data

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

Data Availability Statement

The original contributions presented in the study are included in the article; further inquiries can be directed to the corresponding author.


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

RESOURCES