Skip to main content
Polymers logoLink to Polymers
. 2024 Aug 14;16(16):2300. doi: 10.3390/polym16162300

Robust Recovery of Optimally Smoothed Polymer Relaxation Spectrum from Stress Relaxation Test Measurements

Anna Stankiewicz 1
Editor: Dagmar R D’hooge1
PMCID: PMC11359835  PMID: 39204520

Abstract

The relaxation spectrum is a fundamental viscoelastic characteristic from which other material functions used to describe the rheological properties of polymers can be determined. The spectrum is recovered from relaxation stress or oscillatory shear data. Since the problem of the relaxation spectrum identification is ill-posed, in the known methods, different mechanisms are built-in to obtain a smooth enough and noise-robust relaxation spectrum model. The regularization of the original problem and/or limit of the set of admissible solutions are the most commonly used remedies. Here, the problem of determining an optimally smoothed continuous relaxation time spectrum is directly stated and solved for the first time, assuming that discrete-time noise-corrupted measurements of a relaxation modulus obtained in the stress relaxation experiment are available for identification. The relaxation time spectrum model that reproduces the relaxation modulus measurements and is the best smoothed in the class of continuous square-integrable functions is sought. Based on the Hilbert projection theorem, the best-smoothed relaxation spectrum model is found to be described by a finite sum of specific exponential–hyperbolic basis functions. For noise-corrupted measurements, a quadratic with respect to the Lagrange multipliers term is introduced into the Lagrangian functional of the model’s best smoothing problem. As a result, a small model error of the relaxation modulus model is obtained, which increases the model’s robustness. The necessary and sufficient optimality conditions are derived whose unique solution yields a direct analytical formula of the best-smoothed relaxation spectrum model. The related model of the relaxation modulus is given. A computational identification algorithm using the singular value decomposition is presented, which can be easily implemented in any computing environment. The approximation error, model smoothness, noise robustness, and identifiability of the polymer real spectrum are studied analytically. It is demonstrated by numerical studies that the algorithm proposed can be successfully applied for the identification of one- and two-mode Gaussian-like relaxation spectra. The applicability of this approach to determining the Baumgaertel, Schausberger, and Winter spectrum is also examined, and it is shown that it is well approximated for higher frequencies and, in particular, in the neighborhood of the local maximum. However, the comparison of the asymptotic properties of the best-smoothed spectrum model and the BSW model a priori excludes a good approximation of the spectrum in the close neighborhood of zero-relaxation time.

Keywords: viscoelasticity, relaxation time spectrum, linear relaxation modulus, optimally smoothed model, identification algorithm, model error, noise robustness

1. Introduction

The relaxation spectrum is vital for constitutive models and for providing insight into the mechanical properties of polymers since, from the relaxation spectrum, other material functions used to describe rheological properties can be uniquely derived [1,2,3,4,5]. It is applied for description, analysis, and to accomplish the pre-assumed mechanical properties of different polymers [3,6,7,8].

The spectrum is not directly accessible via measurement and must be recovered from relaxation stress or oscillatory shear data. Numerous different methods have been proposed during the last seven decades for relaxation spectrum identification using data both from the stress relaxation experiment [6,9,10,11,12,13,14,15,16,17,18,19,20] and dynamic modulus tests [5,21,22,23,24,25,26,27,28,29,30,31]. The problem of relaxation spectrum identification is the ill-posed inverse problem of solving a system of Fredholm integral equations of the first kind obtained for discrete measurements of the relaxation modulus or the storage and loss modulus data. Therefore, the solutions, if any, are very sensitive to even small changes in the measurement data, which can lead to arbitrarily large changes in the determined relaxation spectrum. In consequence, robustly stable algorithms are required to solve it. The regularization of the original problem and/or constraining the set of admissible solutions is often necessary to construct such algorithms.

In the first works concerning the relaxation spectrum determination, the sets of the spectrum models were constrained to rather narrow classes of models. In 1948 Macey [9], while examining the viscoelastic properties of ceramic material, described the spectrum by the exponential–hyperbolic model, which corresponds to the modified Bessel function of the second kind and zero-order modeling the relaxation modulus. To describe the mechanical properties of polyisobutylene, Sips [10] introduced in 1950 a simple relaxation spectrum model given by the difference between two exponential functions, which implied a logarithmic model of the relaxation modulus. This model was next augmented to consider a long-term modulus by Yamamoto [11] and applied to study the rheological properties of the plant cell wall. The relaxation spectrum modeling in [9,10,11] is based on the known pairs of Laplace transforms.

The relaxation spectrum identification based on the Post–Widder formula [12] for the inverse Laplace transform was initiated by Alfrey and Doty [13], who proposed a simple differential model based on the first-order Post–Widder formula. Ter Haar [14] approximated the spectrum of relaxation frequencies using the modulus multiplied by time, the inverse of the relaxation frequency, which is, in fact, the Post–Widder inversion formula of the zero order. After many years, Bažant and Yunping [15] and Goangseup and Bažant [16] introduced a two-stage approach of approximating the stress relaxation data via multiple differentiable models of the relaxation modulus and, next, by applying the Post–Widder formula to designate the related model of the spectrum. The effectiveness of this approach depended, among other aspects, on the function applied to approximate the relaxation modulus. In [15], a logarithmic–exponential model of the relaxation modulus was proposed, for which the authors stated the third-order Post–Widder approximation to be satisfactory.

Both the algorithms based on the Post–Widder formula [13,14,15,16], using the least-squares approximation to guarantee the best fitting of the relaxation modulus measurement data, and those using the pairs of Laplace transforms [9,10,11], did not take into account the ill-posed nature of the relaxation spectrum determination problem.

Baumgaertel and Winter [21] used a nonlinear least-squares method for the recovery of a discrete relaxation time spectrum based on storage and loss modulus data, in which the number of discrete model modes was adjusted during the scheme iterations to avoid an ill-posedness of the problem and to enhance the model fit. Regularization was not applied here, as in several of the works discussed subsequently. Malkin [22] approximated a continuous relaxation spectrum using three constants: the maximum relaxation time, the slope in the logarithmic scale, and the form factor. Malkin et al. [23] derived a method of continuous relaxation spectrum calculations using the Mellin integral transform. The algorithm for the relaxation time spectrum approximation by power series was developed by Cho [24], which, using the regression of the dynamic modulus, provided a relaxation time spectrum as stable as the regularization method. The least squares identification without regularization was also applied by Babaei et al. [17] to determine the discrete Maxwell relaxation spectrum based on the stress relaxation data from the ramp test. Lv et al. [5] applied the extended least squares method (without regularization) to dynamic experiment data. Lee et al. [25] used the Chebyshev polynomials of the first kind to approximate dynamic moduli data and next derived a spectrum equation using the complex decomposition method and the Fuoss–Kirkwood relation without any regularization. Also, a derivative-based algorithm for continuous spectrum recovery, which is also appropriate for the experimental situation where oscillatory shear data are only available for a finite range of frequencies, as proposed by Anderssen et al. [26], does not use regularization.

Honerkamp and Weese [27,28] combined nonlinear regression with Tikhonov regularization and proposed a specific viscoelastic model described by the two-mode log-normal function. In turn, Davies and Goulding [29] approximated the relaxation spectrum by a sum of scaling kernel functions located at appropriately chosen points. In the Mustapha and Phillips algorithm [30], the sequence of nonlinear regularized least-squares problems, solved with respect to both the discrete relaxation times and the elastic moduli, was performed with an increasing number of discrete model modes. The approach proposed by Stadler and Bailly [31] is based on the relaxation spectrum approximation using a piecewise cubic Hermite spline with respective regularization. The regularized algorithms presented in [27,28,29,30,31] were developed for dynamic rheological tests. A methodology to calculate the relaxation spectrum of biopolymeric materials from stress relaxation data has been proposed by Kontogiorgos [32], combining Hansen’s least-squares numerical algorithm and Tikhonov regularization with the L-curve criterion chosen to select regularization parameter. Stankiewicz [18,19] and Stankiewicz et al. [20] derived different identification algorithms 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.

All the known methods for the relaxation spectrum optimal identification are based on the minimization of the quadratic model error defined directly for the measurements of the relaxation modulus or storage and loss modules. For example, in [5,17,21], the least-squares criterion was used directly, while in [18,19,20,21,27,28,30], regularized least-squares were used with various rules applied for the choice of regularization parameters to ensure the stability of the scheme and smoothness of the determined relaxation spectrum. In these papers, the mathematical formula describing the relaxation spectrum model was also, in advance, limited to the assumed class of admissible models.

In this paper, as a remedy for the ill-posed nature of the spectrum identification problem, the relaxation time spectrum model that reproduces the relaxation modulus measurements and which is the best smoothed in the class of continuous square-integrable functions was sought. This problem was formulated and solved in this paper for the first time for the spectrum of relaxation times. First, by applying the well-known Hilbert projection theorem, a new model was derived in which the best smoothing was achieved together with the simultaneous interpolation of relaxation modulus measurements. Next, to achieve noise robustness, the problem of the optimal smoothness of the spectrum model was augmented by introducing a quadratic term in the Lagrange functional of the original optimal spectrum smoothing problem. The necessary and sufficient optimality condition of the modified problem implied the best relaxation spectrum model as a finite sum of the basis functions given by the quotient of the exponential function and relaxation time. The components of the corresponding relaxation modulus model were given by simple hyperbolic functions. The permitted in advance small error of the relaxation modulus model combined with the specific modification of the Lagrange functional resulted in the model’s noise robustness. The complete computational procedure for determining the best-smoothed model was given. The singular value decomposition method was used for algebraic computations. Analytical formulas describing the relaxation modulus model error, the relaxation spectrum smoothness, and noise robustness indices were derived as quadratic positive definite forms dependent on the sampling instants applied in the experiment and the relaxation modulus measurements. The monotonicity of these indices was analyzed. The applicability of the proposed model and algorithm to determining the optimally smoothed models of polymers characterized by the short and middle relaxation times of the Gauss-like relaxation spectra and the long relaxation times of the Baumgaertel, Schausberger, and Winter spectrum was verified. The rough applicability analysis of the proposed approach to modeling the relaxation time spectra of different types, such as the Kohlrausch–Williams–Watts, fractional Maxwell, Scott–Blair, inverse power, and multiplied power–exponential laws, was also carried based on the compatibility of the boundary conditions of the real spectra and the best-smoothed model. These studies have shown the applicability of the new model and identification algorithm for the optimal recovery of the smoothed relaxation spectrum of polymers with a very wide range of relaxation times.

In summary, this paper addresses the ill-posed problem of identifying the relaxation time spectrum in a new, original way, previously unknown in the literature. In the known methods, different mechanisms are built in to obtain a smooth enough and noise-robust relaxation spectrum model. Here, a new approach is proposed where the optimally smoothed continuous square-integrable model of the relaxation spectrum, which reproduces relaxation modulus measurements with assumed acceptable small errors, is directly sought. This problem is mathematically formulated and solved, resulting in a unique, best-smoothed relaxation spectrum model and a complete identification algorithm. In the construction of known relaxation spectrum identification methods, the primary idea was the best model approximation, and the next one, implied by the ill-posed nature of the task, was the concept of smoothing the model by regularizing the original problem of the model optimal approximation. Here, the main measure of the model’s quality was the integral of the square of the relaxation spectrum, being simultaneously the measure of the model’s smoothness. The idea of the optimal model approximation is replaced here by the classical interpolation of measurement points. In the basic problem, precise interpolation is applied and is next modified to interpolation with a small error being allowed to ensure the noise robustness of the model and algorithm. The idea of the Tikhonov regularization technique results from the essence of the ill-posed problem—the lack of uniqueness for its solution or its discontinuity with respect to the measurement data. The problem of smoothing the relaxation spectrum posed here finds inspiration in the consequences of the ill-posed problem and sometimes catastrophic fluctuations of the obtained solution, and it eliminates these model-devastating effects.

In Appendix A, the proofs of the main results and derivations of some mathematical formulas are given. Some tables of numerical results were moved to Appendix B to increase the clarity of the article.

2. Materials and Methods

2.1. Relaxation Time Spectrum

The continuous relaxation time spectrum Hτ of a linear viscoelastic material is defined by the following integral [1,26]:

Gt=0Hττet/τdτ, (1)

where Gt is the linear relaxation modulus at time t. The spectrum Hτ is interpreted as a generalization of the discrete Maxwell spectrum to a continuous function [1,26] and characterizes the distributions of relaxation times τ.

2.2. Model of the Relaxation Spectrum

Assume that a model HMτ of the relaxation spectrum Hτ belongs to the space L20, of real-valued square-integrable functions on the interval 0,. Note that L20, is the Hilbert space with the norm x=x,x induced by the inner product defined by the integral

x,y=0xτyτdτ,

where the functions xτ,yτL20, [33].

2.3. Identification

A classical way of conducting the identification experiment studying viscoelasticity is the stress-relaxation test, where time-dependent stress is studied for the step increase in the strain [1,34,35,36]. Suppose a certain stress relaxation experiment resulted in a set of measurements of the relaxation modulus G¯ti=Gti+zti at the sampling instants ti>0, i=1,,N, where zti is the measurement noise. Identification, classically, consists of the selection, within the chosen class of models, of such a model that ensures the best fit to the measurement results. As a measure of the model’s accuracy, the quadratic index, used in the least squares approach, is usually taken. However, in this paper, we look, in the class of continuous square-integrable functions, for the best-smoothed model HMτ that reproduces relaxation modulus measurements G¯ti. This problem is formulated and solved in the next section.

3. Results and Discussion

In this section, the problem of optimal smoothing of the relaxation spectrum model is mathematically formulated and solved using the Hilbert projection theorem. As a result, the best-smoothed relaxation spectrum model is derived in the form of a series of specific basis functions given by the quotient of the exponential function and the relaxation time. The respective model of the relaxation modulus was found to be described by a series of hyperbolic functions. The properties of these basis functions were examined, and the identifiable property of the best-smoothed model was demonstrated. For noise measurement data, the modification of the problem for the spectrum model’s smoothness was proposed by augmenting the Lagrange functional. A dual approach was applied to solve the modified problem, resulting in the necessary and sufficient optimality condition for the optimal relaxation spectrum model. Direct analytical formulas for the relaxation spectrum and modulus models are given; their numerical realization by the singular value decomposition of the basic matrix was proposed. The model smoothness, noise robustness for noisy measurements of the relaxation modulus, and the error of the relaxation modulus model were analyzed. A simple identification scheme was proposed. Finally, the results of simulation studies for polymers described by Gaussian-like and BSW relaxation spectra distributions are presented.

3.1. The Problem of Optimal Smoothing of the Relaxation Spectrum Model

Consider the following problem. Find function HMτL20, that minimizes the integral square index:

HM2=0HM2τdτminHM· L20,, (2)

under the constraints

G¯ti=0HMττeti/τdτ, i=1,,N. (3)

Note, that the set of functions HMτL20, satisfying linear constraints (3) is closed and convex. Since the Hilbert projection theorem [33] implies the existence of a unique element with a minimal norm in the nonempty closed and convex subset of the Hilbert space, the existence and uniqueness of the solution to the smoothing problem (2) and (3) follows.

The Lagrange functional of the optimization problem defined by (2) and (3) is as follows:

LHM,λN=0HM2τdτ+i=1NλiG¯ti0HMττeti/τdτ, (4)

where λN=λ1,,λNT is a vector of Lagrange multipliers λi. The necessary and sufficient optimality conditions for the linear-quadratic optimization task (2) and (3) are given by the equation

HMτ=i=1Nλieti/ττ, (5)

together with the constraints of (3). Substituting HMτ, given by (5), into (3) yields the following system of equations:

G¯ti=0j=1Nλjetj/ττ2eti/τdτ=j=1Nλj01τ2etj+ti/τdτ, i=1,,N. (6)

By applying the substitution v=1/τ in the integrals of the right-hand side of (6) we obtain:

ϕij=01τ2etj+ti/τdτ=0etj+tivdv=1ti+tj. (7)

Introducing the N×N dimensional symmetric matrix composed of the elements ϕij in the i row and j column according to

ΦN=ϕiji=1,,Nj=1,, N=1ti+tji=1,,Nj=1,, N, (8)

the system of Equation (6) can be rewritten in compact form as

G¯N=ΦNλN, (9)

with the vector of the relaxation modulus measurements

G¯N=G¯t1,,G¯tNT, (10)

where superscript “T” indicates transpose. The main properties of the matrix ΦN are summarized in the following proposition shown in Appendix A.1.

Proposition 1. 

For arbitrary sampling instants ti>0, i=1,,N, such that ti+1>ti, a symmetric N×N  matrix   ΦN  defined by (7) and (8) is a positive definite Gram matrix, which can be expressed as  ΦN=ΦN1/2ΦN1/2, where  ΦN1/2  is the unique symmetric non-singular positive definite square root of  ΦN. Then, the inverse matrix  ΦN1=ΦN1/2ΦN1/2  is a positive definite too.

Since matrix ΦN is non-singular, the unique solution of (9) is given by

λN=ΦN1G¯N. (11)

Therefore, by virtue of (5), the best-smoothed model is described by the finite series

H¯Nτ=i=1Nλihiτ (12)

of the basis functions (compare (5))

hiτ=eti/ττ, i=1, 2,. (13)

The lower index ‘N’ is used in H¯Nτ to express the dependence on the number of measurements.

Introducing the notation

hNτ=h1τ,,hNτT=et1/ττ,,etN/ττT (14)

and bearing in mind (11), model (12) can be expressed in compact vector–matrix form as

H¯Nτ=λNThNτ=G¯NTΦN1hNτ. (15)

In view of (1) and (12), the related model of the relaxation modulus is as follows:

G¯Nt=0H¯Nττet/τdτ=0i=1Nλihiττet/τdτ

and can be described by the finite series

G¯Nt=i=1Nλi0eti+t/ττ2dτ=i=1Nλiφit (16)

of hyperbolic basis functions (compare (7))

φit=1t+ti, i=1,,N. (17)

Similarly, using the right equality in (16) and Equation (11), we obtain

G¯Nt=λNTφNt=G¯NTΦN1φNt,

where, in view of (17), the vector function φNt is as follows:

φNt=φ1t,,φNtT=1t+t1,,1t+tNT. (18)

The index

IN=HMτ2=0HM2τdτ (19)

minimized in (2), is a direct measure of smoothing the relaxation spectrum model. For the optimal model H¯Nτ (15), the smoothness index is as follows:

IN=0H¯N2τdτ=G¯NTΦN10hNτhNTτdτ ΦN1G¯N. (20)

Using (14), (7), and (8) we find

0hNτhNTτdτ=ΦN. (21)

Thus, Equation (20) yields

IN=G¯NTΦN1G¯N, (22)

which means that the smoothness of the optimal model depends both on the time instants ti selected for the stress relaxation experiment, affecting the matrix ΦN and on the experiment results G¯N.

As a result, the following result can be stated.

Theorem 1. 

For arbitrary sampling instants ti>0, i=1,,N, such that   ti+1>ti,  the unique optimally smoothed model of the relaxation time spectrum defined by the optimization task (2) and (3) is given by H¯Nτ=G¯NTΦN1hNτ, while the respective relaxation modulus model G¯Nt=G¯NTΦN1φNt and the optimal smoothness index IN=G¯NTΦN1G¯N, where the vector functions hNτ and φNt are defined by (14) and (18), respectively, and matrix ΦN is defined by (8) and (7).

The relaxation spectrum H¯Nτ that solves the optimization task (2) and (3), is the most smoothed model in the class of square-integrable functions that, simultaneously, guarantee the reconstruction of the measurements of the relaxation modulus. Some useful algebraic identities concerning the matrix ΦN and vector φNt are given in Appendix A.2.

3.2. Properties of the Basis Functions

The basis functions hiτ (13) and φit (17) of the relaxation spectrum and modulus models are positive definite and depend on the times ti applied in the stress relaxation experiment. The greater the sampling instants ti, the faster the basic functions φit decrease. By (13), the first derivative is as follows:

dhiτdτ=tiττ3eti/τ.

Thus, the basis functions hiτ for τ=ti have a global maximum equal to hiτ=1/eti, which decreases with increasing index i due to the assumed monotonicity of the sequence ti. This means that increasing the number of measurements N, i.e., increasing the model components, can allow for the good modeling of multimodal spectra, which is confirmed by the second example presented in the final part of this paper.

Since for τ0+, using the L’Hospital’s rule, we have

limτ0+ hiτ=limτ0+ 1τeti/τ=limτ0+ 1τ2tiτ2eti/τ=limτ0+ 1tieti/τ=0+,

and for τ the functions are hiτ0, the best model H¯Nτ tends to zero both for τ0+ and τ (zero boundary conditions). The basis functions φit given by (17) monotonically decrease to zero as t.

The five first basis functions hiτ (13) are shown in Figure 1 for the sampling instants ti=10, 30, 50, 70, 90 and ti=0.1, 0.5, 1, 1.5, 2 s. Figure 2 shows the related functions φit (17). The logarithmic scale is applied for the time axes in these figures. The basis functions hiτ and φit are expressed in s1. From Figure 2, it can be seen that the monotonicity of basis functions φit is in good agreement with the courses of the relaxation modulus obtained in an experiment for real polymers; for example, these include elastic polyacrylamide hydrogels [35] (Figures 2a,b, 4a, A5, A7 and A8a), concrete [37] (Figure 13) and rubber [38] (Figure 2).

Figure 1.

Figure 1

Basis functions hiτ (13), i=1,,5, of the relaxation spectrum model H¯Nτ (12) for time sampling instants: (a) ti=10, 30, 50, 70, 90 s and (b) ti=0.1, 0.5, 1, 1.5, 2 s.

Figure 2.

Figure 2

Basis functions φit (17), i=1,,5, of the relaxation modulus model G¯Nt (16) for time sampling instants: (a) ti=10, 30, 50, 70, 90 s and (b) ti=0.1, 0.5, 1, 1.5, 2 s.

3.3. Identifiability

The basic and obvious requirement for any identification method is that if the real characteristic is described by a model from the considered class of models, and the measurements are noise-free, then the method should guarantee the unique determination of the real characteristic, i.e., ensure its identifiability [39,40].

Assume that the real spectrum is of the form

Hτ=j=1Najetj/ττ, (23)

where aj represents the real parameters. Introducing the vector a=a1,,aNT and bearing in mind (13) and (14), spectrum (23) can be expressed as

Hτ=aThNτ. (24)

Assume that the measurements of the relaxation modulus are noise-free. Thus, for i=1,,N, by virtue of (1), (7) and (23), we have

G¯ti=Gti=0j=1Najetj/ττ2eti/τdτ=j=1Naj1tj+ti=j=1Najϕij.

Therefore, using (8) and (10), the vector of the relaxation modulus measurements can be expressed as

G¯N=GN=ΦNa,

whence, according to (15) and (24), the relaxation spectrum model is as follows:

H¯Nτ=G¯NTΦN1hNτ=aTΦNΦN1hN=aThN=Hτ,

i.e., the model smoothing identification results in the determination of the real relaxation spectrum (23).

3.4. Modification

The value of the Lagrange multiplier λi is the dual price [41], which in problem (2) and (3) is “paid” for satisfying the i-th constraint, i=1,,N. The higher the value of λi (precisely, the modulus of λi), the more difficult it is to meet this constraint and the stronger the chains it imposes. The impact of the fluctuations in the measurements of the relaxation modulus G¯ti, i.e., the impact of changes in the left side of each of Equation (3), on the smoothness of the spectrum is then greater. Really, by (22) and (11), we have

ING¯N=2ΦN1G¯N=2λ¯N.

Therefore, the vector of the optimal Lagrange multipliers is the measure of the model’s smoothness index sensitivity with respect to the fluctuations of the relaxation modulus measurements. To reduce it, the value of the multiplier λi should be reduced; precisely, the values of the modulus of λi should be reduced.

The Lagrange functional of the optimization problem defined by (2) and (3) is described in Equation (4). In order to decrease the values of λi and the functional (4), being maximized with respect to λi according to the dual approach, is modified by introducing the quadratic term i=1Nλi2=λNTλN, which means that the modified Lagrange functional is defined as follows:

LmHM,λN=0HM2τdτ+i=1NλiG¯ti0HMττeti/τdτγi=1Nλi2, (25)

where γ is a small positive constant and the weight that represents the relative importance of the square component λNTλN with respect to the original (non-modified) Lagrange functional LHM,λN (4). The parameter γ has no physical interpretation as the regularization parameter in the classical Tikhonov regularization.

In Appendix A.3, the unique saddle point of the modified Lagrange functional (25) is found. The saddle point defines the modified best-smoothed model H¯Nγτ, which depends on the measurements ti,G¯ti and the parameter γ introduced in (25).

Theorem 2. 

For arbitrary sampling instants  ti>0, i=1,,N, such that  ti+1>ti  and the arbitrary non-negative parameter  γ, the model of the relaxation time spectrum  H¯Nγτ  defined by the unique saddle point of the modified Lagrange functional (25) is given by

H¯Nγτ=G¯NTΦN+4γIN,N1hNτ, (26)

while the respective relaxation modulus model

G¯Nγt=G¯NTΦN+4γIN,N1φNt, (27)

where the vector functions  hNτ  and  φNt  are defined by (14) and (18), respectively, matrix  ΦN  is defined by (7) and (8) and  IN,N  is the  N×N  dimensional unit matrix.

The upper index γ in the notations H¯Nγτ and G¯Nγt indicates the dependence on the parameter γ introduced in the modified Lagrange functional (25).

To achieve the dimensional homogeneity of the components of the Lagrange functional (25), the multipliers λi are expressed in Pa·s, while the unit of the parameter γ is s1. The dimensional homogeneity of the matrix ΦN+4γIN,N is then achieved.

It is demonstrated in Appendix A.4 that for the optimal vectors of the Lagrange multipliers, λN in (11) for original optimization task (2) and (4) and the vector λNγ (A13) of the saddle point of the modified Lagrange functional (25), the following inequality holds:

λNγTλNγ<λNTλN, (28)

which means that λNγ<λN, i.e., the purpose of the modification introduced into the Lagrange functional at the beginning of this section was achieved.

3.5. Model Error

Let us introduce, by analogy to the measurements vector G¯N (10), the vector of the values of the relaxation modulus model (27) for all sampling instants ti:

G¯Nγ=G¯Nγt1,,G¯NγtNT.

For any t=ti, by virtue of (27),

G¯Nγti=G¯NTΦN+4γIN,N1φNti=φNTtiΦN+4γIN,N1G¯N,

whence, bearing in mind Equation (A4), we have

G¯Nγ=ΦNΦN+4γIN,N1G¯N.

Therefore, for the model parameterized by γ>0, the relaxation modulus equations in (3) are not satisfied and the error of these equations is as follows:

εN=G¯NG¯Nγ=G¯NΦNΦN+4γIN,N1G¯N. (29)

Through the identity (A6), the model error εN can be expressed as

εN=4γΦN+4γIN,N1G¯N, (30)

and, bearing in mind (A13), can be equivalently expressed as εN=2γλNγ. Therefore, for γ=0 the model error εN=0N, which is clear since in the original task (2) and (3) the constraints in Equation (3) are exactly satisfied; here, 0N denotes the N dimensional vector of zero elements.

By (30), the square model error is given as follows:

εNTεN=16γ2G¯NTΦN+4γIN,N2G¯N. (31)

In Appendix A.5, the following result is proved.

Proposition 2. 

For arbitrary sampling instants  ti>0,  i=1,,N, such that  ti+1>ti,  and the arbitrary non-negative parameter  γ, the square model error  εNTεN (31) of the relaxation modulus equations monotonically increases as the function of the parameter  γ , which is strictly convex for  γ  such that  ΦN8γIN,N,  and strictly concave in the case  ΦN8γIN,N.

3.6. Smoothness

For model H¯Nγτ (26), the smoothness index IN (19) is as follows:

INγ=0H¯Nγτ2dτ=G¯NTΦN+4γIN,N10hNτhNTτdτ ΦN+4γIN,N1G¯N,

and, bearing in mind (21), can be expressed as

INγ=G¯NTΦN+4γIN,N1ΦNΦN+4γIN,N1G¯N, (32)

or, by applying identity (A7), an equivalent form useful for further differential analysis can be obtained

INγ=G¯NTΦN1/2ΦN+4γIN,N2ΦN1/2G¯N. (33)

Since, according to (22), the smoothness index for the original optimization task (2) and (3) can be expresses as IN=G¯NTΦN1/2ΦN2ΦN1/2G¯N, by inequality (A5), the next estimation follows:

INγ<IN

for any γ>0 and arbitrary measurement data, i.e., the smoothness of the spectrum model H¯Nγτ (26) is stronger that the smoothness of the original model H¯Nτ (15); this was the idea of the modification introduced in the Lagrange functional.

In Appendix A.6, the following formulas describing the first and second derivatives of INγ with respect to γ are derived:

γINγ=8G¯NTΦN1/2ΦN+4γIN,N3ΦN1/2G¯N, (34)

and

2γ2INγ=96 G¯NTΦN1/2ΦN+4γIN,N4ΦN1/2G¯N. (35)

Thus, the smoothness index INγ is the monotonically decreasing convex function of the parameter γ>0. The following rule holds: the greater the parameter γ is, the more highly bounded the fluctuations of the spectrum model H¯Nγτ (26) are.

3.7. Noise Robustness

Following [18,19], as a reference point for the model H¯Nγτ described by Equation (26), the model of the spectrum that can obtain for the same parameter γ, the same number of measurements N and the same time instants ti on the basis of ideal measurements of the relaxation modulus is considered, which is described as follows:

H~Nγτ=GNTΦN+4γIN,N1hNτ, (36)

where GN is the vector of the noise-free relaxation modulus, i.e., GN=Gt1GtNT. In view of (26) and (36), we find

H¯NγτH~Nγτ=zNTΦN+4γIN,N1hNτ, (37)

where zN=zt1ztNT is the vector of measurement noises.

Consider the square integral index

QNγ=0H¯NγτH~Nγτ2dτ. (38)

By (37), this index can be expressed as

QNγ=zNTΦN+4γIN,N10hNτhNTτdτ ΦN+4γIN,N1zN,

whence, by virtue of (21), we obtain

QNγ=zNTΦN+4γIN,N1ΦNΦN+4γIN,N1zN, (39)

which, using the Gram property of ΦN and using the identity (A7), can by rewritten as follows:

QNγ=zNTΦN1/2ΦN+4γIN,N2ΦN1/2zN. (40)

Therefore, the noise robustness depends on the parameter γ, the measurement noises and the sampling instants ti that uniquely determine the matrix ΦN. Since both models are continuous with respect to the relaxation time τ, by virtue of (40), for any non-negative γ, the spectrum H¯Nγτ tends to the noise-free spectrum H~Nγτ for each time τ linearly with respect to the norm zN, as zN0, and the faster this is, the larger the parameter γ.

By (A5), we have

QNγ<QN0=zNTΦN1zN, (41)

which means better noise robustness than for the original model H¯Nτ (15) for any γ>0.

Similarly, as for the smoothness index INγ (compare indices (33) and (40)), the following formulas describing derivatives of QNγ with respect to γ were derived:

γQNγ=8zNTΦN1/2ΦN+4γIN,N3ΦN1/2zN,

and

2γ2QNγ=96 zNTΦN1/2ΦN+4γIN,N4ΦN1/2zN,

which means that QNγ is the monotonically decreasing convex function of the parameter γ>0 that takes the maximal value equal to QN0 (41) for γ=0.

3.8. Algebraic Background of the Computational Algorithm

The singular value decomposition (SVD, [42]) technique can be used in numerical computations in order to determine the inverse matrix ΦN+4γIN,N1 in (26). Let SVD of the N×N dimensional matrix ΦN (8) take the following form [42]:

ΦN=UNΣNUNT, (42)

where ΣN=diagσ1,,σNϵRN,N is the diagonal matrix containing the singular values σi of the matrix ΦN [42], and the matrix UNRN,N is orthogonal. Thus,

ΦN+4γIN,N1=UNΣN+4γIN,N1UNT, (43)

where the N×N diagonal matrix ΣN+4γIN,N1 is as follows:

ΣN+4γIN,N1=diag1/σ1+4γ,,1/σN+4γ.

Therefore, the optimal spectrum model (26) can be described by

H¯Nγτ=g¯NγThNτ, (44)

while the respective model (27) of the relaxation modulus is as follows:

G¯Nγt=g¯NγTφNt,

where the vector of model parameters is

g¯Nγ=UNΣN+4γIN,N1UNTG¯N. (45)

Using (42) and (43), the smoothness index INγ (32) is expressed as

INγ=G¯NTUNΩNUNTG¯N, (46)

where the N×N diagonal matrix ΩN=ΣN+4γIN,N1ΣNΣN+4γIN,N1 takes the form

ΩN=diagσ1/σ1+4γ2,,σN/σN+4γ2.

Similarly, using (42) and (43), the noise robustness index QNγ (39) can be rewritten as

QNγ=zNTUNΩNUNTzN.

Combining (31) and (43), we obtain the formula describing the square model error

εNTεN=16γ2G¯NTUNΣN+4γIN,N2UNTG¯N. (47)

3.9. Algorithm

The determination of the best-smoothed model of the relaxation time spectrum involves the following steps:

  1. Choose the parameter γ>0.

  2. Perform the experiment (stress relaxation test [1,34,35,36]) and record the measurements G¯ti, i=1,,N, of the relaxation modulus at times ti>0, such that ti+1>ti.

  3. Compute the matrix ΦN and next determine SVD (42).

  4. Compute the vector of model parameters g¯Nγ (45).

  5. Determine the spectrum of relaxation times H¯Nγτ according to (44).

  6. Determine the square model error εNTεN according to (47) and the smoothness index INγ using Formula (46).

  7. Check if the smoothness of the spectrum model H¯Nγτ measured by INγ and the error of the relaxation modulus model G¯Nγt measured by εNTεN are, simultaneously, satisfactory. If not, increase the parameter γ and repeat the computations starting from step 4. If yes, accept the current H¯Nγτ as the best-smoothed relaxation spectrum model.

Only the SVD of the matrix ΦN of computational complexity ON3 [42] is a space- and time-consuming task in the scheme. However, for given sampling points, the SVD must be computed only once in step 3. The matrix ΦN does not depend on the relaxation modulus measurements G¯ti. Therefore, when the identification scheme is applied for successive samples of the same material, step 3 should not be repeated whenever the same time instants ti are kept in the experiment. This is because, using (44) and (45), we have

H¯Nγτ=G¯NTϑNγτ,

where the vector function

ϑNγτ=UNΣN+4γIN,N1UNThNτ,

depends only on the sampling points ti and does not depend on the relaxation modulus measurements G¯ti. Therefore, the function ϑNγτ must be computed only once and used to determine the model H¯Nγτ for many samples whenever the same ti is kept.

3.10. Numerical Studies

For numerical studies, it is assumed that the viscoelastic properties are described by the Gaussian-like distribution of the relaxation spectrum, which is used to represent the rheological properties of numerous polymers, e.g., polyacrylamide gels [35], poly(methyl methacrylate) [43], polyethylene [44] and carboxymethylcellulose (CMC) [45]. Also, the spectra of many biopolymers have a Gaussian nature, for example, cold gel-like emulsions stabilized with bovine gelatin [46], fresh egg-white-hydrocolloids [45], some (wheat, potato, corn, and banana) native starch gels [47], the xanthan gum water solution [45] and wood [48,49]. The Baumgaertel, Schausberger, and Winter (BSW) spectrum [50,51] used to describe the viscoelasticity of polydisperse polymer melts [24,25], polybutadiene (PBD) [52], polymethylmethacrylate (PMMA) [52] and polymer pelts [53] is also considered.

The best-smoothed spectra models, the values of the square model error εNTεN (31), where the error εN is defined by (29), the smoothness index INγ=0H¯Nγτ2dτ given by Formula (32), and the square noise robustness index QNγ (38) expressed by Equation (39) are examined for different number of the measurements N and different values of the parameter γ.

The “real” materials and the best-smoothed models were simulated in Matlab R2023b, The Mathworks, Inc., Natick, MA, USA. For the singular value decomposition procedure svd was applied. A normal distribution with zero mean value and variance σ2 as well as the uniform distribution were applied for the random independent generation of additive measurement noises.

3.11. Example I

Considering the polymer whose relaxation spectrum is described by the uni-modal Gaussian-like distribution as follows:

Hτ=ϑe1τm2/q/τ, (48)

and where the parameters are as follows [54,55]: ϑ=31,520 Pa·s, m=0.0912 s1 and q=3.25×103 s2. The related relaxation modulus is desribed by the function [55]:

Gt=πq2ϑ e14t2qmterfc12tqmq, (49)

where the complementary error function erfcx is given by [56]:

erfcx=2π  xez2dz.

In the experiment, N sampling instants ti were generated with the constant period in the time interval of T=0, 200 seconds, selected on the basis of the modulus Gt (49) course.

3.11.1. Noise-Free Measurements

For noise-free measurements of the modulus Gt the best-smoothed model solving the original task (2) and (3) was determined for N=20, 100, 150, 200, 500, 1000 measurements. Two optimal models H¯Nτ (15) and the ‘real’ spectrum Hτ (48) are plotted in Figure 3. Small subfigures confirm the excellent model fit; real spectra described by the red lines practically coincide with the blue models both for the small and large number of measurements. This shows that in the case of noise-free measurements, the practically ideal approximation of the real relaxation spectrum was obtained even for a small number of measurements (N=20). In Figure 4, the related models of the relaxation modulus G¯Nt (16) are plotted; the measurements G¯ti of the ‘real’ modulus Gt (49) are also marked. The values of the smoothness index IN (19) are given in Table 1.

Figure 3.

Figure 3

Relaxation time spectrum Hτ (48) (solid red line) from Example I and the corresponding best-smoothed models H¯Nτ (15) for N noise-free relaxation modulus measurements: (a) N=20; (b) N=1000.

Figure 4.

Figure 4

The measurements G¯ti of the ‘real’ relaxation modulus Gt (49) (red circles) from Example I and the optimal model G¯Nt (16) for N noise-free relaxation modulus measurements: (a) N=20; (b) N=1000.

Table 1.

The smoothness index IN (19) for noise-free N measurements of the relaxation modulus from Example I described by the relaxation spectrum Hτ (48).

N 20 50 100 150 200 500 1000
INkPa2·s 70.937064 70.937063 70.937025 70.9370341 70.937296 70.9378314 70.937012

3.11.2. Noise-Corrupted Measurements

Additive independent measurement noises are generated by a normal distribution with zero mean value and variance σ2. For the noise robustness analysis, the standard deviations σ=2,4,6 Pa were used. The parameters γ=5×107, 106, 5×106, 105 s1 were applied.

In Table 2, the values of the square model error εNTεN (31), the smoothness index INγ (32), and the square noise robustness index QNγ (38) are given for noises of σ=2Pa, while for the stronger noises, the same data are given in Table A1 and Table A2 in Appendix B. As previously, the exemplary courses of the spectrum models H¯Nγτ (26) for N=20 and N=500 measurements are illustrated in Figure 5 for noises of σ=2, 4, 6 Pa, while the respective relaxation modulus models G¯Nγt (27) are depicted in Figure 6.

Table 2.

For Example I, the square model error εNTεN (31), the smoothness index INγ=0H¯Nγτ2dτ, Equation (32), and the noise robustness index QNγ (38) for N measurements of the relaxation modulus corrupted by normally distributed additive independent noises with zero mean value and standard deviation σ=2 Pa; parameter γ is introduced in the modified Lagrange Functional (25).

γ [s1] Index N=20 N=50 N=100 N=500 N=1000
5 × 10−7 εNTεN kPa2 1.39645 × 10−4 2.52753 × 10−4 4.62907 × 10−4 1.92494 × 10−3 3.88455 × 10−3
INγ kPa2·s 72.99045 67.737749 68.723774 69.975844 73.348805
QNγ kPa2·s 1.596146 1.662047 1.457413 1.371245 3.277104
1 × 10−6 εNTεN kPa2 1.49648 × 10−4 2.57649 × 10−4 4.65545 × 10−4 1.92724 × 10−3 3.88936 × 10−3
INγ kPa2·s 69.48318 65.994479 67.80256 69.162430 71.604073
QNγ kPa2·s 0.891089 0.829375 0.757987 0.655363 1.298411
5 × 10−6 εNTεN kPa2 2.08947 × 10−4 2.78699 × 10−4 4.85586 × 10−4 1.94192 × 10−3 3.90908 × 10−3
INγ kPa2·s 63.29753 63.719666 65.798822 67.689507 69.559083
QNγ kPa2·s 0.212948 0.293031 0.228734 0.103584 0.188545
1 × 10−5 εNTεN kPa2 2.68258 × 10−4 3.03442 × 10−4 5.14355 × 10−4 1.96206 × 10−3 3.93448 × 10−3
INγ kPa2·s 61.213971 62.869605 64.812065 66.994669 68.681175
QNγ kPa2·s 0.102555 0.1849723 0.149022 5.95577 × 10−2 9.49178 × 10−2
Figure 5.

Figure 5

Figure 5

Relaxation time spectrum Hτ (48) (solid red line) from Example I and the corresponding models H¯Nγτ (26) for N measurements of the relaxation modulus corrupted by normally distributed additive independent noises with zero mean value and standard deviation σ: (a) σ=2 Pa and N=20; (b) σ=2 Pa and N=500; (c) σ=4 Pa and N=20; (d) σ=4 Pa and N=500; (e) σ=6 Pa and N=20; and (f) σ=6 Pa and N=500.

Figure 6.

Figure 6

The measurements G¯ti of the ‘real’ relaxation modulus Gt (49) (red circles) from Example I and the model G¯Nγt (27) for N measurements of the relaxation modulus corrupted by normally distributed additive independent noises with zero mean value and standard deviation σ: (a) σ=2 Pa and N=20; (b) σ=2 Pa and N=500; (c) σ=4 Pa and N=20; (d) σ=4 Pa and N=500; (e) σ=6 Pa and N=20; and (f) σ=6 Pa and N=500.

An inspection of Figure 5a,c,e shows that for each noise case, the number of N=20 measurements was not enough to obtain the satisfactory smoothness of the model H¯Nγτ even for the weakest noises. So, N=20, which is good for noise-free case, fails here. However, for N=500 measurements, the model H¯Nγτ is smoothed enough, and the influence of the regularization parameter γ is much weaker; see Figure 5b,d,f. Figure 6 and the values of the model errors εNTεN from Table 2, Table A1, and Table A2 confirm the excellent approximation of the relaxation modulus model, even though model imbalance is allowed.

The analytically shown monotonicity of the smoothness INγ and noise robustness QNγ indices, being the monotonically decreasing convex functions of the parameter γ, is reflected in the numerical studies. An inspection of the numerical results indicates that for N50 and any fixed parameter γ, the smoothness index INγ is a monotonically increasing function of the number of measurements; the slower this is, the larger the number N. An analysis of the asymptotic properties of the algorithm and optimal model H¯Nγτ (26) as the number of measurements grows to infinity will be the subject of future studies.

3.12. Example II

Consider the double-mode Gaussian-like distribution of the relaxation spectrum [19,20,44]

Hτ=β1e1τm12/q1+β2e1τm22/q2/τ, (50)

where the parameters are as follows: β1=467 Pa·s, m1=0.0037 s1, q1=1.124261×106 s2, β2=39 Pa·s, m2=0.045 s1 and q2=1.173×103 s2. The double-Gaussian relaxation spectra are examined while developing new identification methods in [31] (Figure 2), [29] (Figures 9, 11 and 17), and [26] (Figures 2, 3, 6, 7–11 and 14). Such spectra describe the rheological properties of various polymers [44] (Figures 4b and 8b), polyacrylamide gels [35] (Figure A4), and wood [38]. The corresponding ‘real’ relaxation modulus is composed of two summands described by formulas like that of (49). In the experiment, N=50, 100, 200, 500, 1000, 5000 sampling instants ti were generated with the constant period in the time interval T=0, 1550 seconds, selected in view of the course of the modulus. Following [19,20], additive measurement noises zti were selected independently by random choice with uniform distribution on the interval 0.005, 0.005 Pa.

In Table 3, the values of the square model error εNTεN (31), the smoothness index INγ (32), and the square noise robustness index QNγ (38) are given. The spectrum models H¯Nγτ (26) are illustrated in Figure 7 along with the real spectrum (50). Since, similar to the one-mode Gaussian relaxation spectrum, the relaxation modulus models G¯Nγt (27) for different N and γ values practically coincide, the respective figures are omitted here.

Table 3.

For the relaxation spectrum (50) from Example II described by double-mode Gaussian distribution: the square model error εNTεN (31), the smoothness index INγ (32), and the noise robustness index QNγ (38) for N measurements of the relaxation modulus corrupted by additive independent noises uniformly distributed on the interval 0.005, 0.005 Pa; parameter γ introduced in the modified Lagrange functional (25).

γ [s1] Index N=50 N=100 N=200 N=500 N=1000 N=5000
1 × 10−7 εNTεN Pa2 4.22629 × 10−4 8.08789× 10−4 1.62982 × 10−3 3.87154 × 10−3 7.89202 × 10−3 4.11897 × 10−2
INγ Pa2·s 2.97202 × 102 2.90993 × 102 2.90362 × 102 2.82977 × 102 3.04036 × 102 3.36177 × 102
QNγ Pa2·s 11.911735 10.130120 13.23408 15.429651 21.957947 19.365373
5 × 10−7 εNTεN Pa2 4.49498 × 10−4 8.23669 × 10−4 1.64435 × 10−3 3.87762 × 10−3 7.91074 × 10−2 4.12242 × 10−2
INγ Pa2·s 2.70677 × 102 2.74519 × 102 2.75150 × 102 2.75893 × 102 2.81813 × 102 2.96481 × 102
QNγ Pa2·s 2.360449 1.926742 2.624791 2.092844 2.495385 2.836702
1 × 10−6 εNTεN Pa2 4.90577 × 10−4 8.43919 × 10−4 1.66358 × 10−3 3.88467 × 10−3 7.92263 × 10−2 4.12478 × 10−2
INγ Pa2·s 2.56575 × 102 2.67642 × 102 2.68545 × 102 2.73497 × 102 2.77623 × 102 2.88112 × 102
QNγ Pa2·s 1.251268 0.969347 1.209452 0.880198 1.097522 1.369153
5 × 10−6 εNTεN Pa2 8.80287 × 10−4 1.16953 × 10−3 1.90025 × 10−3 4.02082 × 10−3 8.03088 × 10−2 4.13474 × 10−2
INγ Pa2·s 2.17932 × 102 2.37719 × 102 2.46356 × 102 2.61489 × 102 2.67451 × 102 2.77109 × 102
QNγ Pa2·s 0.244248 0.220253 0.216093 0.169113 0.215385 0.248797
1 × 10−5 εNTεN Pa2 1.37477 × 10−3 1.68512 × 10−3 2.29245 × 10−3 4.31822 × 10−3 8.23885 × 10−2 4.14497 × 10−2
INγ Pa2·s 2.00761 × 102 2.19869 × 102 2.32877 × 102 2.51385 × 102 2.60383 × 102 2.73583 × 102
QNγ Pa2·s 0.114042 0.113907 0.109599 0.090753 0.114361 0.119821

Figure 7.

Figure 7

Relaxation time spectrum Hτ (50) (solid red line) from Example II and the corresponding models H¯Nγτ (26) determined for N measurements of the relaxation modulus corrupted by additive independent noises uniformly distributed on the interval 0.005, 0.005 Pa: (a) N=50; (b) N=100; (c) N=200; (d) N=500; (e) N=1000; and (f) N=5000; the values of the parameter γ introduced in the modified Lagrange functional (25) are given in the figures.

3.13. Example III

Consider the spectrum of relaxation times introduced by Baumgaertel, Schausberger, and Winter [50,51]

Hτ=β1ττcρ1+β2ττcρ2eττmax, (51)

which is known to be effective in describing polydisperse polymer melts [24,25] with the parameters [25] β1=6.276×102 MPa, β2=0.127 MPa, τc=2.481 s, τmax=2.564×104 s, ρ1=0.25, and ρ2=0.5. As in [54], in the experiment, N time instants ti were sampled with the constant period in the time interval of T=0,105 seconds, where, following [18], the interval was selected in view of the course of the ‘real’ modulus Gt defined by (1). Additive measurement noises zti were selected independently by random choice with uniform distribution on the interval 0.005, 0.005 MPa. The results of the numerical experiment are given in Table 4 and illustrated in Figure 8.

Table 4.

For the polymer described by the BSW spectrum (51): the square model error εNTεN (31), the smoothness index INγ (32), and the noise robustness index QNγ (38) for N measurements of the relaxation modulus corrupted by additive independent noises selected according to uniform distribution from the interval 0.005, 0.005 MPa and parameters γ introduced in the modified Lagrange functional (25).

γ [s1] Index N=50 N=100 N=200 N=500 N=1000
1 × 10−7 εNTεN MPa2 4.50707 × 10−4 8.19820 × 10−4 1.65574 × 10−3 3.89542 × 10−3 7.93196 × 10−3
INγ MPa2·s 3.21654 × 103 3.24851 × 103 3.24718 × 103 3.23537 × 103 3.24043 × 103
QNγ MPa2·s 11.891956 11.219052 10.864104 8.652512 10.942303
5 × 10−7 εNTεN MPa2 5.16852 × 10−4 8.57969 × 10−4 1.67497 × 10−3 3.90799 × 10−3 7.94415 × 10−3
INγ MPa2·s 3.15685 × 103 3.21397 × 103 3.22838 × 103 3.22318 × 103 3.22606 × 103
QNγ MPa2·s 2.175434 2.067436 2.0363525 2.006206 2.710421
1 × 10−6 εNTεN MPa2 6.72858 × 10−4 9.49036 × 10−4 1.71894 × 10−3 3.93047 × 10−3 7.95659 × 10−3
INγ MPa2·s 3.10426 × 103 3.18335 × 103 3.213648 × 103 3.21555 × 103 3.22182 × 103
QNγ MPa2·s 1.076912 0.975302 0.944682 0.997097 1.417322
5 × 10−6 εNTεN MPa2 3.81118 × 10−3 2.95039 × 10−3 2.87847 × 10−3 4.44269 × 10−3 8.24008 × 10−3
INγ MPa2·s 2.82672 × 103 3.00764 × 103 3.11432 × 103 3.17172 × 103 3.19764 × 103
QNγ MPa2·s 0.169205 0.230928 0.230358 0.204836 0.308728
1 × 10−5 εNTεN MPa2 1.08157 × 10−2 7.46682 × 10−3 5.81259 × 10−3 5.84646 × 10−3 9.03681 × 10−3
INγ MPa2·s 2.58939 × 103 2.85463 × 103 3.01535 × 103 3.12462 × 103 3.17095 × 103
QNγ MPa2·s 0.078727 0.114988 0.140981 0.118423 0.169938

Figure 8.

Figure 8

Figure 8

The relaxation time BSW spectrum (51) (solid red line) from Example III and the corresponding models H¯Nγτ (26) for N measurements of the relaxation modulus corrupted by additive independent noises selected according to uniform distribution from the interval 0.005, 0.005 MPa and parameters γ introduced in the modified Lagrange functional (25): (a) N=50; (b) N=100; (c) N=200; (d) N=500; (e) N=1000; and (f) N=5000.

Since the real spectrum Hτ (51) tends to infinity for τ0 whenever at least one of the parameters ρ1 and ρ2 is negative, the best-smoothed model H¯Nγτ (26) cannot adequately approximate this spectrum for small relaxation times τ; in the example for 0<τ<103 s. This is well illustrated in Figure 8. However, this figure also shows that for a sufficiently large γ, the spectrum Hτ for higher frequencies and its local maximum are well approximated.

3.14. Applicability for Identification of Relaxation Spectra of Different Types

The natural condition of this approach’s successful applicability follows from the properties of the best-smoothed model H¯Nγτ (26) yielded by the properties of the basis functions hiτ (13), which compose the vector hNτ according to (14). Since for τ0+ and τ, the basis functions are hiτ0 (c.f., Section 3.2), the best model H¯Nγτ also tends to zero as the relaxation time τ tends to zero and to infinity. Therefore, zero boundary conditions limit the scope of applicability of the model and method to the real relaxation time spectra that satisfy these conditions. The example of the BSW spectrum demonstrates that the properties of the spectrum for τ0+ are essential here since the real relaxation time spectra and the known spectra models tend to zero as the relaxation time τ tends to infinity.

The Kohlrausch–Williams–Watts (KWW) model of the stretched exponential relaxation described by [57]

Gt=G0etτKWWβ, (52)

where the stretching exponential 0<β<1, τKWW is the characteristic relaxation time and G0 denotes the initial shear modulus, has been found by many researchers to be more appropriate than standard exponentials [57,58,59,60,61,62]. In spite of the simple, compact form of (52), the related unimodal [58] relaxation spectrum is described by the following infinite series [57,58]:

Hτ=G0π k=11k+1k!sinπβk Γβk+1 tτKWWβk, (53)

which is based on Pollard’s representation of the stretched exponential as a Laplace integral [63], where Γn is Euler’s gamma function [64] (Equation (A.1.1)). However, for some specific stretching exponentials, namely β=12, β=13 and β=23, the KWW spectrum has a compact form described by some special functions [58]. For (53), both zero boundary conditions are satisfied; compare [57] (Figure 1a). Therefore, the proposed approach can be used to identify the spectrum of materials whose relaxation processes are described by the KWW model, e.g., polymer melts [59], the local segmental dynamics of poly(vinylacetate) [60], the segmental dynamics and the glass transition behavior of poly(2-vinylpyridine) [61], the relaxation of bone and bone collagen [62], alginate films while considering glycerol concentration [65], and even the relaxation processes of the onion structure in sine-oscillatory shear [66]. The best-smoothed model H¯Nγτ (26) given by finite series may prove to be more useful than the original KWW spectrum (53).

In recent decades, non-integer order differential equations have increased interest in the modeling of time-dependent relaxation processes; the fractional Maxwell model (FMM) and the elementary Scott–Blair model are probably the most known rheological non-integer order models. The applicability of the FMM relaxation time spectrum, which is described by the compact analytical formula [54] (Equation (12)), to modeling the unimodal relaxation spectra of polymers was recently examined in [54]. However, it was demonstrated in [54] that the FMM relaxation spectrum tends to infinity as τ0+ [54] (Proposition 2, Equation (19)); therefore, the exact fitting of the FMM-type spectrum by the proposed best-smoothed model in the whole relaxation time domain is excluded, which is similar for the BSW spectrum. The relaxation time spectrum of the Scott–Blair model described by the inverse power of the relaxation time with the non-integer exponent, see [54] (Equation (15)), also loses the zero boundary condition at zero relaxation time.

Similarly, real relaxation spectra which are well characterized by simple inverse power laws with various exponents [67]; for example, the power-type spectrum of cross-linking polymers at their gel point described by Winter and Chambon model with an exponent of −1/2 [68] and the spectrum of solution-polymerized styrene butadiene rubber described by a combined four-interval power model with fractional exponents [69] could not be successfully identified by the proposed method in the whole relaxation times domain. In turn, Winter’s power law relaxation time spectrum with a positive exponent [70,71] (Equation (2)), which was proposed to describe relaxation in many molecular and colloidal glasses, although satisfies the zero initial condition, could not be determined by the proposed algorithm due to its confined domain.

However, the proposed approach can be successfully applied to identify the relaxation spectra of materials such as bitumen, being characterized by the broadened power law model [71] (Equation (8)):

Hτ=nαGctταnα etταβ, (54)

which multiplicative form combines the power law with an exponential of stretching parameter β. In the above model, the exponent 0<nα<1, τα is the longest relaxation time and Gc is the plateau modulus. The unimodal spectrum (54) satisfies both zero boundary conditions; compare [71] (Figure 11a).

4. Conclusions

The objective of this paper was to develop a relaxation time spectrum model that could reproduce the relaxation modulus measurements and which is the best-smoothed in the class of continuous square-integrable functions. The unique optimal relaxation spectrum model was found to be described by a finite series of specific exponential–hyperbolic functions. A new identification algorithm was proposed in which the best smoothing of the model was achieved together with the simultaneous reconstruction of relaxation modulus measurements with small model errors. The analytical and numerical studies proved that using a developed model and algorithm, it is possible to determine the relaxation spectrum model for a wide class of polymers with zero boundary conditions, in particular, Gaussian-like distributed relaxation spectra. The model is smoothed and noise robust; small relaxation modulus model errors are obtained. The applicability of this approach to determining the Baumgaertel, Schausberger, and Winter spectrum was also examined, and it was proved that, due to the asymptotic properties of this spectrum, it can be well approximated for higher frequencies and, in particular, in the neighborhood of the local maximum. The rough applicability analysis, based on the consistency of the zero boundary conditions of the real spectra and the best-smoothed model, shows the possibility of using the proposed method and model to describe the relaxation spectra of different types that are characteristic of many polymers. However, the search for such a modification of the proposed approach so that it can also be applicable to the identification of spectra with non-zero boundary conditions for relaxation times approaching zero, like the BSW spectrum, will be the subject of future research. Generally, the properties of the method, including the smoothness of the relaxation spectrum, depend both on the experiment plan, i.e., on the sampling instants used in the relaxation test, and on the relaxation modulus measurements. The results of numerical studies confirm the analytically proved monotonic dependence on the gamma parameter: monotonically increasing for the square model error and monotonically decreasing for the noise robustness and spectrum model smoothness indices. However, the dependence of these indices on the number of measurements is not so clear and must be the subject of further studies.

Summarizing the numerical studies implies the following directions for future research:

  • The asymptotic analysis of the model and identification algorithm properties as the number of measurements tends to infinity;

  • The modification of the proposed approach for smoothing the spectrum model with a non-zero boundary condition for zero relaxation time;

  • The modification for non-zero equilibrium modulus;

  • The recurrent realization of the algorithm.

This method can be applied for any deformation process described by definitional Equation (1), i.e., both for uniaxial deformation, uniaxial stress, and uniaxial stretching, assuming that the relaxation modulus of the respective process is experimentally accessible. The relaxation time spectrum in the respective state (uniaxial deformation, stress, or stretching) is then determined. An appropriate modification of the algorithm can be developed to apply the concept of optimal relaxation spectrum model smoothing for oscillatory shear measurements of the storage and loss moduli. This will be the subject of further research.

Appendix A

Appendix A.1. Proof of Proposition 1

Let us define the functions hiv=etiv, i=1, 2, and consider the function

fNv=i=1Ngihiv, (A1)

where the real parameters gi compose the vector gN=g1gNT. By (A1), for any vector gK, we have

0fNv2dv=0i=1Ngihiv2dv, (A2)

which can be rewritten as

0fNv2dv=i=1Nj=1Ngigj0hivhjvdv,

and next, bearing in mind the right equality in (7), is expressed as

0fNv2dv=i=1Nj=1Ngigjϕij. (A3)

Bearing in mind (8) and (A2), Equation (A3) is the quadratic form

0fNv2dv=0i=1Ngihiv2dv=gNTΦNgN.

Thus, gNTΦNgN0 for an arbitrary vector gN, and gNTΦNgN=0, if and only if i=1Ngihiv=0 for almost all v>0. Since the exponential functions h1v=et1v,,hNv=etNv, i.e., the kernel of the Laplace transformation, are linearly independent, the last equality holds if and only if gi=0 for all i=1,,N, i.e., only if the vector gN=0, which yields the positive definiteness of ΦN. The Gram property of ΦN follows directly from the definitional Formulas (7) and (8); its positive definiteness yields the existence of a unique symmetric positive definite square root matrix ΦN1/2 in view of Theorem 7.2.6 in [72] concerning the positive semidefinite k-th roots of the Hermitian positive semidefinite matrices by lying k=2. Thus, ΦN=ΦN1/2ΦN1/2, and ΦN1/2=ΦN1/2T. By Theorem 4.2.1 in [42], the inverse matrix ΦN1 is a positive definite, too. The non-singularity of the square root ΦN1/2 implies the inverse matrix formula ΦN1=ΦN1/2ΦN1/2; thus, this proposition is proved.

Appendix A.2. Some Matrix Identities

In this appendix, some useful vector-matrix identities are given.

Property A1. 

For arbitrary sampling instants ti>0, i=1,,N, such that   ti+1>ti , the matrix ΦN (8) and vector function φNt, defined by (18), satisfy the equation

φNt1,,φNtNT=φNt1,,φNtN=ΦN. (A4)

Proof. 

Equation (A4) follows directly from (8) and (18). □

Property A2. 

For any  γ>0  and any positive definite symmetric matrix  ΦN,  the following inequality

ΦN+4γIN,N2<ΦN2, (A5)

and equations

IN,NΦNΦN+4γIN,N1=4γΦN+4γIN,N1, (A6)
ΦN+4γIN,N1ΦN1/2=ΦN1/2ΦN+4γIN,N1 (A7)

hold.

Proof. 

Inequality (A5) results directly from the following inequality:

ΦN2<ΦN2+8γΦN+16γ2IN,N=ΦN+4γIN,N2.

The equality

IN,NΦNΦN+4γIN,N1=ΦN+4γIN,NΦNΦN+4γIN,N1

implies identity (A6).

For the Gram matrix ΦN=ΦN1/2ΦN1/2, we have

ΦN1/2ΦN+4γIN,N=ΦN+4γIN,NΦN1/2, (A8)

whence

ΦN=ΦN1/2ΦN1/2=ΦN+4γIN,NΦN1/2ΦN+4γIN,N1ΦN1/2,

and finally

ΦN+4γIN,N1ΦN=ΦN1/2ΦN+4γIN,N1ΦN1/2. (A9)

From (A8), after multiplying by ΦN+4γIN,N1 on the right and on the left, we obtain identity (A7). □

Appendix A.3. Proof of Theorem 2

A dual approach will be used to find the saddle point of the modified Lagrange Functional (25). First, the function HMτ minimizing the Functional (25) is found, i.e., the primary optimization task

minHM·L20,LmHM,λN=LmH¯M,λN, (A10)

is solved. Next, the dual function is maximized, and the vector of optimal Lagrange multipliers is found.

The necessary and sufficient optimality condition for the primary problem (A10) takes the form

2H¯Mτi=1Nλieti/ττ=0,

from which we obtain

H¯Mτ=12i=1Nλieti/ττ, (A11)

which, substituted into the modified Lagrange functional (25), yields the dual function

LD,mλN=LmH¯M,λN,

given by

LD,mλN=i=1NλiG¯ti14j=1Nλj0etj/ττeti/ττdτγi=1Nλi2. (A12)

By definition (7) and (8) of the matrix ΦN, the structure of the vectors G¯N (10) and λN, the dual function (A12) can be expressed in compact form as

LD,mλN=λNTG¯N14λNTΦNλNγλNTλN,

which, in view of Proposition 1, means that for any parameter γ>0 the dual function is strictly the concave function of λN with the gradient

dLD,mλNdλN=G¯N12ΦNλN2γλN.

Therefore, the unique solution of the dual problem is given by the formula

λNγ=2ΦN+4γIN,N1G¯N. (A13)

By (A11), and bearing in mind the notations (13) and (14), the corresponding relaxation spectrum model is expressed as follows:

H¯Nγτ=12i=1Nλiγeti/ττ=12λNγThNτ. (A14)

The pair H¯Nγ,λNγ is a unique saddle point of the Lagrange functional (25). The substitution of λNγ (A13) into (A14) yields Formula (26).

By (1) and the left equality in (A11), bearing in mind (16), for any t>0 we have

G¯Nγt=0H¯Nγττet/τdτ=12i=1Nλiγ0eti+t/ττ2dτ=12i=1Nλiγφit,

with the functions φit given by (17), which, through (18), can be expressed as

G¯Nγt=12λNγTφNt,

and, when combined with (A13), implies Formula (27). The result is proved.

Appendix A.4. Derivation of the Inequality (28)

By (A13), we have

λNγTλNγ=4G¯NTΦN+4γIN,N2G¯N, (A15)

while for the Lagrange multiplier λN (11) the respective value is

λNTλN=G¯NTΦN2G¯N. (A16)

Therefore, by virtue of (A15), (A16), and property (A5), for the optimal vectors of the Lagrange multipliers λN (11) and λNγ (A13), the inequality (28) holds.

Appendix A.5. Proof of Proposition 2

The following differential properties which hold for the arbitrary differentiable matrix functions Ax and Bx [42] (Equations (P2.1.2a) and (P2.1.2b)):

xAxBx=AxxBx+AxBxx, (A17)
xAx1=Ax1AxxAx1, (A18)

assuming that matrix Ax is invertible, will be used.

By (A17), we obtain

xAx2=xAx1Ax1=Ax1xAx1+Ax1Ax1x, (A19)

from which, including (A18), we immediately obtain the next useful differential formula:

xAx2=Ax1AxxAx2Ax2AxxAx1. (A20)

From (31), it follows that

γεNTεN=32γG¯NTΦN+4γIN,N2G¯N+16γ2G¯NTγΦN+4γIN,N2G¯N. (A21)

Since, by virtue of (A20),

γΦN+4γIN,N2=8ΦN+4γIN,N3,

Equation (A21) takes the form

γεNTεN=32γG¯NTΦN+4γIN,N2G¯N16·8·γ2G¯NTΦN+4γIN,N3G¯N,

and can be rewritten as follows

γεNTεN=32γG¯NTΦN+4γIN,N3ΦN+4γIN,N4γIN,NG¯N,

from which the next formula is directly obtained

γεNTεN=32γG¯NTΦN+4γIN,N3ΦNG¯N. (A22)

Therefore, using (A9), we obtain

ΦN+4γIN,N3ΦN=ΦN+4γIN,N2ΦN1/2ΦN+4γIN,N1ΦN1/2,

whence, through the double use of identity (A7), we find

ΦN+4γIN,N3ΦN=ΦN+4γIN,N1ΦN+4γIN,N1ΦN1/2ΦN+4γIN,N1ΦN1/2=ΦN+4γIN,N1ΦN1/2ΦN+4γIN,N2ΦN1/2=ΦN1/2ΦN+4γIN,N3ΦN1/2. (A23)

Combining (A22) and (A23) yields

γεNTεN=32γG¯NTΦN1/2ΦN+4γIN,N3ΦN1/2G¯N, (A24)

from which the positive definiteness of the first derivative follows in view of the positive definiteness of the matrix ΦN+4γIN,N3.

From (A24), the second derivative is obtained

2γ2εNTεN=32G¯NTΦN1/2ΦN+4γIN,N3ΦN1/2G¯N+32γG¯NTΦN1/2γΦN+4γIN,N3ΦN1/2G¯N. (A25)

Applying (A17) again to the second term in (A25) gives

2γ2εNTεN=32G¯NTΦN1/2ΦN+4γIN,N3ΦN1/2G¯N+32γG¯NTΦN1/2γΦN+4γIN,N1ΦN+4γIN,N2ΦN1/2G¯N+32γG¯NTΦN1/2ΦN+4γIN,N1γΦN+4γIN,N2ΦN1/2G¯N,

whence, using differential Formulas (A18) and (A20), we obtain

2γ2εNTεN=32G¯NTΦN1/2ΦN+4γIN,N3ΦN1/2G¯N32·4·γG¯NTΦN1/2ΦN+4γIN,N2ΦN+4γIN,N2ΦN1/2G¯N32·8·γG¯NTΦN1/2ΦN+4γIN,N1ΦN+4γIN,N3ΦN1/2G¯N,

and next

2γ2εNTεN=32G¯NTΦN1/2ΦN+4γIN,N3ΦN1/2G¯N32·12·γG¯NTΦN1/2ΦN+4γIN,N4ΦN1/2ΦN1/2G¯N,

which can be finally expressed as the quadratic form

2γ2εNTεN=32G¯NTΦN1/2ΦN+4γIN,N2ΦN8γIN,NΦN+4γIN,N2ΦN1/2G¯N.

The matrix ΦN+4γIN,N2 is a positive definite. The conditions concerning convexity and concativity of the square model error εNTεN result directly from the conditions of positive or negative definiteness of the matrix ΦN8γIN,N. The proposition is proved.

Appendix A.6. Derivation of Formulas (34) and (35)

By (33) and differential Formulas (A19) and (A18), for any γ>0, we have

γINγ=4G¯NTΦN1/2ΦN+4γIN,N2ΦN+4γIN,N1ΦN1/2G¯N4G¯NTΦN1/2ΦN+4γIN,N1ΦN+4γIN,N2ΦN1/2G¯N,

which yields (34).

Therefore, using (A17), we obtain

2γ2INγ=8G¯NTΦN1/2γΦN+4γIN,N2ΦN+4γIN,N1ΦN1/2G¯N8G¯NTΦN1/2ΦN+4γIN,N2γΦN+4γIN,N1ΦN1/2G¯N.

Next, applying Formulas (A18) and (A19), we find:

2γ2INγ=8·4G¯NTΦN1/2ΦN+4γIN,N2ΦN+4γIN,N2ΦN1/2G¯N+8·4G¯NTΦN1/2ΦN+4γIN,N1ΦN+4γIN,N2ΦN+4γIN,N1ΦN1/2G¯N+8·4G¯NTΦN1/2ΦN+4γIN,N2ΦN+4γIN,N2ΦN1/2G¯N,

whence

2γ2INγ=8·12G¯NTΦN1/2ΦN+4γIN,N4ΦN1/2G¯N,

i.e., Formula (35) follows.

Appendix B

Table A1.

For Example I, the square model error εNTεN (31), the smoothness index INγ=0H¯Nγτ2dτ, Equation (32), and the noise robustness index QNγ (38) for N measurements of the relaxation modulus corrupted by normally distributed additive independent noises with zero mean value and the standard deviation σ=4Pa; parameter γ is introduced in the modified Lagrange Functional (25).

γ [s1] Index N=20 N=50 N=100 N=500 N=1000
5 × 10−7 εNTεN kPa2 5.43595 × 10−4 1.00519 × 10−3 1.84977 × 10−3 7.70001 × 10−3 1.55419 × 10−2
INγ kPa2·s 83.336529 70.77716 71.747521 73.185807 83.497189
QNγ kPa2·s 6.384582 6.648187 5.829650 5.484982 13.108415
1 × 10−6 εNTεN kPa2 5.63927 × 10−4 1.01733 × 10−3 1.85720 × 10−3 7.707668 × 10−3 1.556163 × 10−2
INγ kPa2·s 76.168804 66.385898 69.099367 70.438599 76.296796
QNγ kPa2·s 3.564357 3.317502 3.031949 2.621451 5.193645
5 × 10−6 εNTεN kPa2 6.61796 × 10−4 1.04504 × 10−3 1.887541 × 10−3 7.733015 × 10−3 1.56079 × 10−2
INγ kPa2·s 65.608941 63.014540 65.816002 67.562022 70.882789
QNγ kPa2·s 0.851793 1.172124 0.914936 0.414335 0.754181
1 × 10−5 εNTεN kPa2 7.48551 × 10−4 1.06649 × 10−3 1.92315 × 10−3 7.75582 × 10−3 1.56467 × 10−2
INγ kPa2·s 62.549842 62.274208 64.589149 66.768457 69.528702
QNγ kPa2·s 0.410221 0.739890 0.596089 0.238231 0.379671

Table A2.

For Example I, the square model error εNTεN (31), the smoothness index INγ=0H¯Nγτ2dτ, Equation (32), and the noise robustness index QNγ (38) for N measurements of the relaxation modulus corrupted by normally distributed additive independent noises with zero mean value and the standard deviation σ=6Pa; parameter γ is introduced in the modified Lagrange functional (25).

γ [s1] Index N=20 N=50 N=100 N=500 N=1000
5 × 10−7 εNTεN kPa2 1.21470 × 10−3 2.25864 × 10−3 4.16166 × 10−3 1.73258 × 10−2 3.49726 × 10−2
INγ kPa2·s 96.874904 77.140666 77.6860944 79.138260 100.199781
QNγ kPa2·s 14.365310 14.958420 13.116713 12.341209 29.493933
1 × 10−6 εNTεN kPa2 1.24932 × 10−3 2.28259 × 10−3 4.17777 × 10−3 1.73428 × 10−2 3.50181 × 10−2
INγ kPa2·s 84.636611 68.436067 71.912153 73.025494 83.586342
QNγ kPa2·s 8.019804 7.464379 6.821885 5.898264 11.685702
5 × 10−6 εNTεN kPa2 1.39704 × 10−3 2.32565 × 10−3 4.22678 × 10−3 1.73873 × 10−2 3.51075 × 10−2
INγ kPa2·s 68.346247 62.895476 66.290649 67.641705 72.583587
QNγ kPa2·s 1.916535 2.637279 2.058606 0.932255 1.696907
1 × 10−5 εNTεN kPa2 1.51741 × 10−3 2.34999 × 10−3 4.27376 × 10−3 1.74152 × 10−2 3.516484 × 10−2
INγ kPa2·s 64.090823 62.048757 64.664280 66.661361 70.566065
QNγ kPa2·s 0.922997 1.664754 1.341202 0.536019 0.854259

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

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.Ferry J.D. Viscoelastic Properties of Polymers. 3rd ed. John Wiley & Sons; New York, NY, USA: 1980. [Google Scholar]
  • 2.Mead D.W. Numerical interconversion of linear viscoelastic material functions. J. Rheol. 1994;38:1769–1795. doi: 10.1122/1.550526. [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.Hajikarimi P., Moghadas Nejad F. Chapter 6—Interconversion of constitutive viscoelastic functions. In: Hajikarimi P., Nejad F.M., editors. Applications of Viscoelasticity. Elsevier; Amsterdam, The Netherlands: 2021. pp. 107–139. [DOI] [Google Scholar]
  • 5.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]
  • 6.Chang F.L., Hu B., 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]
  • 7.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]
  • 8.Demidov A.V., Makarov A.G. Spectral Simulation of Performance Processes of Polymeric Textile Materals. Fibre Chem. 2023;54:222–226. doi: 10.1007/s10692-023-10381-2. [DOI] [Google Scholar]
  • 9.Macey H.H. On the Application of Laplace Pairs to the Analysis of Relaxation Curves. J. Sci. Instrum. 1948;25:251. doi: 10.1088/0950-7671/25/7/323. [DOI] [Google Scholar]
  • 10.Sips R. Mechanical behavior of viscoelastic substances. J. Polym. Sci. 1950;5:69–89. doi: 10.1002/pol.1950.120050103. [DOI] [Google Scholar]
  • 11.Yamamoto R. Stress relaxation property of the cell wall and auxin-induced cell elongation. J. Plant Res. 1996;109:75–84. doi: 10.1007/BF02344291. [DOI] [Google Scholar]
  • 12.Widder D.V. An Introduction to Transformation Theory. Academic Press; Cambridge, MA, USA: 1971. [Google Scholar]
  • 13.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]
  • 14.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]
  • 15.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]
  • 16.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]
  • 17.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]
  • 18.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]
  • 19.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]
  • 20.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]
  • 21.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]
  • 22.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]
  • 23.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]
  • 24.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]
  • 25.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]
  • 26.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]
  • 27.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]
  • 28.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]
  • 29.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]
  • 30.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]
  • 31.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]
  • 32.Kontogiorgos V. Calculation of Relaxation Spectra from Stress Relaxation Measurements. In: Elnashar M., editor. Biopolymers. Sciyo; Rijeka, Croatia: 2010. [DOI] [Google Scholar]
  • 33.Bachman G., Narici L. Functional Analysis. Academic Press; New York, NY, USA: 1966. [Google Scholar]
  • 34.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]
  • 35.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;13:13040629. doi: 10.3390/polym13040629. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Malkin A.I.A., Malkin A.Y., Isayev A.I. Rheology: Concepts, Methods and Applications. ChemTec; Deerfield Beach, FL, USA: 2006. [Google Scholar]
  • 37.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]
  • 38.Martinez J.M., Boukamel A., Méo S., Lejeunes S. Statistical approach for a hyper-visco-plastic model for filled rubber: Experimental characterization and numerical modeling. Eur. J. Mech.-A/Solids. 2011;30:1028–1039. doi: 10.1016/j.euromechsol.2011.06.013. [DOI] [Google Scholar]
  • 39.Ljung L. System Identification: Theory for the User. 1st ed. Prentice Hall; Englewood Cliffs, NJ, USA: 1999. [Google Scholar]
  • 40.Sufleta Z. Identifiability of time-varying parameters in a class of linear dynamical systems. IEEE Trans. Autom. Control. 1982;27:1114–1116. doi: 10.1109/TAC.1982.1103072. [DOI] [Google Scholar]
  • 41.Zangwill W.I. Nonlinear Programming: A Unified Approach. Prentice-Hall; Saddle River, NJ, USA: 1969. [Google Scholar]
  • 42.Golub G.H., Van Loan C.F. Matrix Computations. Johns Hopkins University Press; Baltimore, MD, USA: 2013. [Google Scholar]
  • 43.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]
  • 44.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]
  • 45.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. [Google Scholar]
  • 46.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]
  • 47.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]
  • 48.Bardet S., Gril J. Modelling the transverse viscoelasticity of green wood using a combination of two parabolic elements. Comptes Rendus Mécanique. 2002;330:549–556. doi: 10.1016/S1631-0721(02)01503-6. [DOI] [Google Scholar]
  • 49.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]
  • 50.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]
  • 51.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]
  • 52.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]
  • 53.Unidad H.J., Goad M.A., Bras A.R., Zamponi M., Faust R., Allgaier J., Pyckhout-Hintzen W., Wischnewski A., Richter D., Fetters L.J. Consequences of Increasing Packing Length on the Dynamics of Polymer Melts. Macromolecules. 2015;48:6638–6645. doi: 10.1021/acs.macromol.5b00341. [DOI] [Google Scholar]
  • 54.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]
  • 55.Stankiewicz A. Sampling Points-Independent Identification of the Fractional Maxwell Model of Viscoelastic Materials Based on Stress Relaxation Experiment Data. Materials. 2024;17:1527. doi: 10.3390/ma17071527. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.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]
  • 57.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]
  • 58.Anderssen R.S., Saiful A., Husain R., Loy J. The Kohlrausch function: Properties and applications. Anziam J. 2003;45:C800–C816. doi: 10.21914/anziamj.v45i0.924. [DOI] [Google Scholar]
  • 59.De Gennes P.-G. Relaxation Anomalies in Linear Polymer Melts. Macromolecules. 2002;35:3785–3786. doi: 10.1021/ma012167y. [DOI] [Google Scholar]
  • 60.Ngai K.L., Roland C.M. Development of cooperativity in the local segmental dynamics of poly(vinylacetate): Synergy of thermodynamics and intermolecular coupling. Polymer. 2002;43:567–573. doi: 10.1016/S1089-3156(01)00011-3. [DOI] [Google Scholar]
  • 61.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]
  • 62.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]
  • 63.Harry Pollard H. The representation of exλ a Laplace integral. Bull. Am. Math. Soc. 1946;52:908–910. doi: 10.1090/S0002-9904-1946-08672-3. [DOI] [Google Scholar]
  • 64.Gorenflo R., Kilbas A.A., Mainardi F., Rogosin S.V. Mittag-Leffler Functions, Related Topics and Applications. Springer; Berlin/Heidelberg, Germany: 2014. [DOI] [Google Scholar]
  • 65.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]
  • 66.Maruoka H., Nishimura A., Ushiki H., Hatada K. Stretched exponential relaxation process of onion structures under various oscillatory shears with analysis using Shannon entropy. Chem. Phys. 2018;513:280–286. doi: 10.1016/j.chemphys.2018.08.020. [DOI] [Google Scholar]
  • 67.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]
  • 68.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]
  • 69.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]
  • 70.Winter H.H. Glass transition as the rheological inverse of gelation. Macromolecules. 2013;46:2425–2432. doi: 10.1021/ma400086v. [DOI] [Google Scholar]
  • 71.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]
  • 72.Horn R.A., Johnson C.R. Matrix Analysis. 2nd ed. Cambridge University Press; Cambridge, NY, USA: 2013. [Google Scholar]

Associated Data

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

Data Availability Statement

Data are contained within the article.


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

RESOURCES