Skip to main content
Materials logoLink to Materials
. 2024 Mar 27;17(7):1527. doi: 10.3390/ma17071527

Sampling Points-Independent Identification of the Fractional Maxwell Model of Viscoelastic Materials Based on Stress Relaxation Experiment Data

Anna Stankiewicz 1
Editor: Angelo Morro1
PMCID: PMC11012696  PMID: 38612043

Abstract

Considerable development has been observed in the area of applying fractional-order rheological models to describe the viscoelastic properties of miscellaneous materials in the last few decades together with the increasingly stronger adoption of fractional calculus. The fractional Maxwell model is the best-known non-integer-order rheological model. A weighted least-square approximation problem of the relaxation modulus by the fractional Maxwell model is considered when only the time measurements of the relaxation modulus corrupted by additive noises are accessible for identification. This study was dedicated to the determination of the model, optimal in the sense of the integral square weighted model quality index, which does not depend on the particular sampling points applied in the stress relaxation experiment. It is proved that even when the real description of the material relaxation modulus is entirely unknown, the optimal fractional Maxwell model parameters can be recovered from the relaxation modulus measurements recorded for sampling time points selected randomly according to respective randomization. The identified model is a strongly consistent estimate of the desired optimal model. The exponential convergence rate is demonstrated both by the stochastic convergence analysis and by the numerical studies. A simple scheme for the optimal model identification is given. Numerical studies are presented for the materials described by the short relaxation times of the unimodal Gauss-like relaxation spectrum and the long relaxation times of the Baumgaertel, Schausberger and Winter spectrum. These studies have shown that the appropriate randomization introduced in the selection of sampling points guarantees that the sequence of the optimal fractional Maxwell model parameters asymptotically converge to parameters independent of these sampling points. The robustness of the identified model to the measurement disturbances was demonstrated by analytical analysis and numerical studies.

Keywords: viscoelasticity, linear relaxation modulus, fractional Maxwell model, stress relaxation test, experiment randomization, differentiable Lipchitz models

1. Introduction

For several decades, fractional-order rheological models have been used to describe, analyze and improve the viscoelastic properties of different materials. In addition to theoretical research dedicated to fractional-order rheological models [1,2,3,4,5], hundreds of studies have been conducted on the applicability of such models for specific materials to describe their mechanical properties. The applicability of such models to the description of different polymers is well known, for example, poly-isobutylene [4], polyurea and PET [6], shape memory polymers [7], amorphous polymers [8] and flax fiber-reinforced polymer [9]. Fractional viscoelastic models are also used for modeling laminated glass beams in the pre-crack state under explosive loads [10]; stress relaxation behavior of glassy polymers [11]; description of fiber-reinforced rubber concrete [12]; viscoelastic modeling of modified asphalt mastics [13]; and modeling rate-dependent nonlinear behaviors of rubber polymers [14]. The modeling and simulation of viscoelastic foods, for example, food gums [15], carrot root [16], fish burger baking [17], is another field of application of rheological fractional models. Due to the non-integer order of the operations of integration and differentiation, the fractional-order models have improved flexibility and better adjustment to material characteristics, both in the time and frequency domains, compared to those of the classic integer-order models.

Although over the last several decades different fractional differential models have been proposed for modeling the viscoelastic processes in materials, the fractional Maxwell model (FMM) is the best known [4,5]. The relaxation modulus of the FMM, described by the product of Mittag-Leffler and inverse power functions, allows for the modeling of a very wide range of stress relaxation processes in materials. Describing the rheological properties of polymers by the FMM [18,19] is well known. However, the FMM was also applied, for example, for computational modeling and analysis on the damping and vibrational behaviors of viscoelastic composite structures [20], viscoelastic flow in a circular pipe [21], effect of temperature on the dynamic properties of mixed surfactant adsorbed layers at the water/hexane interface [22,23] and constitutive equations of the Mn-Cu damping alloy [24]. Fractional viscoelasticity described by the Maxwell model turned out to model both exponential and non-exponential relaxation phenomena in real materials.

Different identification methods for the recovery of the parameters of the non-integral-order models, including the FMM, from both static [16,25,26,27,28] and dynamic [12,29,30,31] experiments data have been proposed so far. It is known that different identification methods in association with different experiment plans result in different identification data yield models, which may differ [32]. Generally, the identification result, i.e., the chosen model, is influenced by the three entries that are necessary for model identification: the set of models from which the best model is chosen, the rule for the optimal model selection and the measurement data obtained in the experiment [32,33]. For the selected class of models, here, the set of the fractional Maxwell models, the identified model depends on the identification rule and the experiment data. The model parameters are usually determined by guaranteeing the “best-possible” fit to the measurements. Therefore, parameters of the optimal model are dependent on the measure applied for evaluating the “best” [32]. The mean-square approximation error is the predominant selection of the model quality measure, which results in a standard least-squares identification task. For the selected identification index, the model identified is usually dependent, sometimes even very strongly, on the experiment data. This is the case with FMM identification methods known in the literature [12,16,25,26,27,28,29,30,31]. This paper deals with the problem of the FMM identification using measurement data from the stress relaxation test. Therefore, the sampling instants used in the experiment and discrete-time measurements of the relaxation modulus compose the set of the experiment data. To build the optimal fractional Maxwell model whose parameters do not depend on sampling instants applied in the stress relaxation test is the aim of this paper.

In the previous paper [33], the problem of the least-squares approximation of the relaxation modulus has been considered for an assumed wide class of relaxation modulus models. Models being continuous, differentiable and Lipschitz continuous with respect to the parameters have been assumed. The main results in [33] refer to the models that are determined asymptotically, when the number of measurements tend to infinity. Whenever some applicability conditions concerning the chosen class of models are satisfied, the asymptotically optimal FMM parameters can be determined using the measurement data obtained for sampling instants selected randomly due to the appropriate randomization, even when the true relaxation modulus description is completely unknown. For the exponential Maxwell and the exponential stretched Kohlrausch–Williams–Watts models, the applicability conditions are satisfied [33]. It should be noted that the concept of identification being measurement point-independent comes from the Ljung paper [34] and the paper of [35], in which the optimal identification problems for dynamic and static systems have been considered.

In this paper, the concept of introducing an appropriate randomization for the selection of sampling instants at which the measurements of the relaxation modulus are recorded is applied for the fractional Maxwell model identification. Following [33], the real material description is completely unknown and only the measurement data of the relaxation modulus are available for model identification. Identification consists of determining the FMM that solves the problem of an optimal least-squares approximation of a real relaxation modulus. The complicated form of the relaxation modulus of the FMM (the product of Mittag-Leffler and inverse power functions) implies that the applicability of the sampling points-independent identification for FMM identification is not obvious. It is known that the relaxation modulus of the FMM is continuous and differentiable with respect to its four parameters [36]. However, the satisfaction of the Lipschitz continuous property with the bounded Lipschitz constant is proved in this paper for the first time, to guarantee the applicability of the experiment randomization concept.

A complete identification scheme leading to the strongly consistent estimate of the optimal model was specified. Assuming that the measurements are corrupted by additive disturbances, the stochastic-type analysis of the model convergence was carried out, and the exponential rate of convergence was demonstrated both analytically and by numerical studies. For materials described by the unimodal Gauss-like spectrum of relaxation used to describe the rheological properties of the materials [37,38,39] and by the Baumgaertel, Schausberger and Winter (BSW) spectrum [40,41] successfully applied for modeling the polymers [42,43], based on the simulation experiments, both the asymptotic properties and noise robustness of the algorithm were numerically studied. To improve the clarity of this article, the proof of the new FMM Lipschitz property is moved to Appendix A. The tables with the results of the numerical studies are given in Appendix B.

2. Materials and Methods

2.1. Material

A linear viscoelastic material subjected to small deformations for which the uniaxial, non-aging and isotropic stress–strain equation is given by a Boltzmann superposition equation [44]

ςt=tGtτdετdτdτ (1)

is considered, where ςt and εt are, respectively, the stress and strain and Gt denotes the linear relaxation modulus. By Equation (1), the stress ςt at time t depends on the earlier history of the strain rate described by the first-order derivative dετdτ via the kernel given by the relaxation modulus Gt.

The modulus Gt is the stress induced in the material described by constitutive Equation (1) by the unit step strain εt imposed. It is assumed for the studied material that the mathematical description of the modulus Gt is completely unknown. However, the real relaxation modulus Gt is accessible by measurement with a certain accuracy for an arbitrary time tT. Here, T=t0,T with the initial time t0>0 and T.

We make the following assumption [33]:

Assumption 1.

The relaxation modulus Gt of the material is bounded on T, i.e., suptTGtM<.

2.2. Fractional Maxwell Model

Constitutive equation of the fractional order Maxwell model is as follows [2,4,45]:

τrαβdαβςtdtαβ+ςt=Geτrαdαεtdtα, (2)

where Ge denotes the elastic modulus, τr means the relaxation time, α and β are non-integer positive orders of fractional derivatives of the strain εt and stress ςt, respectively. In this paper, dαdtαf(x)=Dtαf(x) means the fractional derivative operator in the sense of Caputo’s of a function f(x) of non-integer-order α with respect to variable t and with a starting point at t=0, which is defined by [1,4]

Dtαft=1Γnα0tt1nα1dndtnftdt,

where n1<α<n and Γn is Euler’s gamma function [1] (Equation (A.1.1)).

The FMM (2) can be considered as a generalization of the classic viscoelastic Maxwell model being the series connection of the ideal spring with a dashpot (see Figure 1a) described by a differential equation of the first order [44,46]:

dςtdt+1τrςt=Gedεtdt, (3)

with the elastic modulus Ge of the spring, the relaxation time τr=η/Ge, where η means the viscosity of the dashpot.

Figure 1.

Figure 1

Viscoelastic models: (a) classic Maxwell model; (b) fractional Scott-Blair model of an order α; (c) fractional Maxwell model; elastic modulus Ge, Ge1, Ge2, viscosity η, relaxation times τr, τ1, τ2.

A series connection (see Figure 1c), analogical to the classic Maxwell model, of two elementary fractional Scott-Blair elements Ge1,τr1,α and Ge2,τr2,β, both described by the fractional differential equation of the general form [2,4,45]

ςt=Geτrαdαεtdtα, (4)

with the parameters Ge,τr,α (see Figure 1b), yields the FMM described by Equation (2), where the parameters Ge1,τr1,α and Ge2,τr2,β uniquely determine the parameters E and τr of the FMM (2); for details, see [16]. The four parameters Ge,τr,α,β of the FMM (2), compared with only two parameters Ge,η, or equivalently Ge,τr of the classic Maxwell model (3), are important for the improvement in the FMM accuracy and flexibility.

The uniaxial stress response of the FMM (2) imposed by the unit step strain εt, i.e., the time-dependent relaxation modulus Gt, for an arbitrary 0<β<α1 is given by the formula [2,4,5]:

Gt=GetτrβEαβ,1βtτrαβ, (5)

where Eκ,μx is the generalized two-parameter Mittag-Leffler function defined by series being convergent in the whole z-complex plane [1,2]:

Eκ,μx=n=0xnΓκn+μ. (6)

Further, for the description of the FMM identification task, relaxation modulus model (5) is denoted as

GMt,g=GetτrβEαβ,1βtτrαβ, (7)

to emphasize the dependence on a four-element vector of model parameters

g=αβGeτrT, (8)

where the subscript ‘M’ means the model.

For the special case α=β, the FMM (2) reduces to the Scott-Blair model (compare (4))

2ςt=Geτrαdαεtdtα, (9)

and the relaxation modulus is described by

GMt,g=Ge2Γ1αtτrα. (10)

Let us consider the following set of the FMM admissible model parameters:

G=g:β0 βα1; GeminGeGemax;τrminτrτrmax (11)

where β0>0 is an arbitrarily small positive number and the maximal and minimal values of elastic modulus Ge and relaxation time τr follow from the a priori knowledge concerning the material under investigation and are such that Gemin>0 and τrmin>t0. G is a compact subset of the four-dimensional real space R4.

The properties of the two-parameter Mittag-Leffler function and the model (7) have been studied by many authors [1,2,3,4,5]. The function Eκ,μx (6) is completely monotonic on the negative real axis for 0<κ1 and μκ, i.e., the function Eκ,μx is completely monotonic for x>0, Ref. [4] (Equation (E.32)). Whence, since t0>0, by virtue of (6), for any tT, and any gG, we have

Eαβ,1βtτrαβ<Eαβ,1β0=1Γ1β1. (12)

Let us introduce the function [4] (Equation (E.53))

eκ,μx;λ=xμ1Eκ,μλxκ, (13)

which, comparing (7) and (13), enables describing the relaxation modulus GMt,g (7) in compact form as follows

GMt,g=Geτrβeαβ,1βt;τrβα. (14)

The function eκ,μx;λ (13) is known to play a crucial role in many problems of fractional calculus [4] (p. 372) because it has many excellent and useful properties; some of them were used in this paper. The function eκ,μx;λ is completely monotonic for x>0 when 0<κμ1 whenever the parameter λ>0 [4] (p. 373) as the product of two completely monotonic functions, which by (14) implies the complete monotonicity of the relaxation modulus model GMt,g for t>0 whenever 0β<α1. This means, in particular, that for t>0 and gG, such that 0<β<α1, the positive definite model GMt,g (7) monotonically decreases with increasing t>0. Therefore, for any t>0 and any gG, such that 0<β<α1, in view of (12)–(14), we have

GMt,gGeτrβt0βEαβ,1βt0τrαβGemaxm0, (15)

where m0 is defined below by the sequence of inequalities valid for any tT and any gG

tτrβτrmaxt0βτrmaxt0=m0, (16)

where t0>0.

For the case α=β, the relaxation modulus GMt,g (10) is also a completely monotonic function of the time for t>0, which in view of (16) is uniformly bounded for tT and gG by Gemaxm0/2.

Therefore, there exists a positive constant M1=Gemaxm0 such that

suptT,gGGMt,gM1<, (17)

i.e., the modulus GMt,g is uniformly bounded on the set T×G.

Inequality (17) combined with Assumption 1 implies the upper bound

suptT,gGGtGMt,gM+M1<. (18)

The Lipschitz continuity of the model GMt,g with respect to parameter g, which is not obvious, in particular, with respect to non-integer orders of fractional derivatives, is fundamental to guarantee the convergence of the optimal models for the applied here experiment randomization. Therefore, before the identification concept and the respective algorithm are presented, the Lipschitz property of the mapping GMt,g (7) will be proved, as summarized in the following theorems. The quite tedious proofs are moved into Appendix A.1.

2.3. Lipschitz Continuity of FMM with Respect to Model Parameters

Due to the relation between the parameter α and β, let us consider two cases separately when (a) β<α and (b) β=α. Therefore, the set of admissible model parameters G (11) is decomposed on two disjoint subsets:

G1=g:β0 β<α1; GeminEGemax;τrminτrτrmax (19)

and

G2=g:β0 β=α1; GeminEGemax;τrminτrτrmax, (20)

in which the relaxation modulus GMt,g is described by the formulas (7) and (10), respectively. The bounded set G1 is non-closed, i.e., the compactness property of the set G (11) is lost here, while G2 is compact.

The following spectral representation derived in [47]

GMt,g=Geτrαπ0τrvαβsinπβ+sinπατrv2αβ+2τrvαβcosπαβ+1vα1etvdv, (21)

which results from the known spectral representation of the two-parameter Mittag-Leffler function [1] (Theorem 4.18, Equations (4.7.17) and (4.7.15)) and is valid for 0<β<α1, will be used for gG1. Applying the differential approach in Appendix A.1, the next result is proved.

Theorem 1.

Let G1 defined by (19) be the set of the fractional Maxwell model admissible parameters. Then, the relaxation modulus GMt,g (7) of the FMM (2) is continuous and differentiable with respect to g (8) for any time tT and

suptT,gG1gGMt,g2M2<, (22)

where gGMt,g denotes the gradient of the function GMt,g with respect to the vector g; here, ·2 is the Euclidean norm in the space R4.

The above theorem means, in particular, that for an arbitrary small positive β0, the mapping GM:T×G1R defined according to Equation (7) is, uniformly with respect to the time tT, a Lipschitz continuous function of the vector of model parameters g with Lipschitz constant M2.

In the case (b) β=α, for the set of model parameters G2 (20), the FMM (2) is described by the power-law relaxation modulus GMt,g (10) and the absolute boundness of the gradient gGMt,g is resolved by the next result proved in Appendix A.2.

Theorem 2.

Let G2, defined by (20), be the set of the fractional Maxwell model admissible parameters with equal parametersα and β. Then, the relaxation modulus GMt,g (10) of the model (9) is continuous and differentiable with respect to g (8) for any time tT and

suptT,gG2gGMt,g2M3<. (23)

From the proofs of the above theorems, especially from the nonnegative definiteness of the derivatives GMt,gE (A5), GMt,gτr (A6) and the two last elements of the gradient gGMt,g (A53), the following property is derived.

Property 1.

Let G defined by (11) be the set of the FMM (2) admissible parameters. Then, for any fixed time tT, the relaxation modulus GMt,g described by (7) or (10) monotonically increases with increasing parameters Ge and τr and other parameters being fixed, i.e., the greater parameters Ge and τr are, the greater the relaxation modulus GMt,g is for the given tT.

2.4. Relaxation Modulus Measurements

Following [33,35], let T1,,TN be independent random variables with a common probability density function ρt; T is the support of ρt. Then, let Gi=GTi be the related relaxation modulus of the material for i=1,,N. Let G¯i denote their measurements corrupted by additive noise Zi, i.e., G¯i=Gi+Zi, recorded in the stress relaxation experiment [44,46,48].

The two assumptions concerning the measurement noises are taken (compare Assumptions 5 and 6 in [33]) as follows:

Assumption 2.

The measurement noise Zi is a time-independent, i.e., independent of the variables Ti, sequence of independent identically distributed (i.i.d.) random variables with zero mean EZi=0 and a common finite variance EZi2=σ2<.

Assumption 3.

The measurement noises Zi are bounded by δ, i.e., Ziδ< for i=1,,N.

Both the above assumptions and Assumption 1, concerning the real relaxation modulus, are natural in the context of the relaxation modulus identification [33].

2.5. Identification Problem

FMM identification involves selecting from a given class of models defined by (7) and (10) the model that best fits the measurement data. Suppose an identification experiment resulted in a set of measurements G¯Ti=GTi+Zi at the sampling times Tit0>0, i=1,,N. The mean-squares index

QNg=1Ni=1NG¯TiGMTi,g2, (24)

is taken as a measure of the FMM model accuracy. Here, the lower index denotes the number of measurements. Then, the problem of the optimal model identification consists of the solution of the minimization task

mingG  QNg=QNgN, (25)

where gN is the optimal model parameter. Since, due to the continuity of the model GMt,g with respect to the parameter g, the index QNg is a continuous function of g and the set of admissible parameters G (11) is compact, the existence of the solution to the optimization problem (25) immediately results from the Weierstrass theorem about the extreme of continuous function on the compact set [49]. Since the minimum gN can be not unique, let GN denote the set of vectors gN that solve the optimization task (25).

The parameters gN of the identified relaxation modulus model GMt,gN are dependent on the measurement data, in particular, on the sampling instants Ti. To make the model independent of specific sampling instants Ti, the optimal sampling points-independent approximation problem is stated in the following subsection.

2.6. The Optimal FMM

Let us consider the following problem of determining such an FMM that minimizes the global approximation error:

Qg=TGtGMt,g2ρtdt, (26)

where the selected weight function, such that 0ρtM0<, is a density on the set T, i.e., Tρtdt=1.

The integral (26) is absolutely integrable, uniformly on G, both for the bounded or unbounded domain T as the product of a function GtGMt,g2, in view of (18) bounded uniformly for t,gT×G, and absolutely integrable function ρt. Therefore, the integral (26) is well defined for any gG.

The problem of the optimal approximation of the real modulus Gt within the class of the fractional Maxwell models relies on determining the parameter g that minimizes Qg over the set G, i.e., in solving optimization task

mingG Qg=Qg. (27)

Due to continuity of GMt,g with respect to the vector g, the index Qg (5) is a continuous function of g, and thus, the existence of the solution g follows from the previously mentioned Weierstrass theorem concerning the extreme of continuous function on the compact set. Let the set of model parameters g solving (27) be denoted by G. Any parameter gG does not depend on the specific time instants applied in the experiment.

3. Results and Discussion

In this section, the analysis of the asymptotic properties of the identified fractional Maxwell model, when the number of measurements tend to infinity, is conducted. The rate of the convergence of this model to the optimal FMM, which does not depend on the experiment data, is studied. The resulting identification algorithm is outlined. Next, the analytically proven properties of the identification method are verified by numerical simulations and studies. Two example materials are simulated. In the first, the “real” material is described by a unimodal Gauss-like relaxation spectrum [37,38,39] with short relaxation times and the Baumgaertel, Schausberger and Winter (BSW) spectrum [40,41] with long relaxation times. Such models are used to describe the rheological properties of various materials, especially polymers and biopolymers. Based on the noise-corrupted data from the simulated randomized stress relaxation experiment, the optimal FMM models are determined. The asymptotic properties and noise robustness have been studied.

3.1. Convergence

The empirical index QNg (24) can be obtained by the replacement of the integral in Qg (26) with the finite mean sum of squares, which is clear from a practical point of view. For i=1,,N, by Assumption 2, the expected value is

EGTi+ZiGMTi,g2=Qg+σ2,

whence, by (24), the expected value is

EQNg=Qg+σ2. (28)

To investigate the stochastic-type asymptotic properties of the empirical identification task given by (25), some properties derived in [35] will be used. Note, that Assumptions A1–A3 from [35], concerning the compactness of the set of model admissible parameters, continuity, differentiability and Lipshitzness of the model are satisfied here. Taken above, Assumption 2 is identical with Assumption A5 in [35], while property (18) also means that Assumption A4 from [35] is satisfied, i.e., all the assumptions from [35] hold here.

By (28), Property 2 from [35] implies the next result.

Property 2.

When Assumptions 1 and 2 are satisfied, then

supgG  Qg+σ2QNg0  w.p.1  as  N, (29)

where  w.p.1 means “with probability one”.

By (28) and (29), the empirical identification index QNg (24) is arbitrarily close to its expected value, uniformly in g over the set G. In consequence, the model parameter gN solving empirical identification task (25) can be related to the parameter g that solves the sampling points-independent minimization task (27). From the uniform in gG convergence of the index QNg in (29), we conclude immediately the main result of this subsection, c.f., Assertion in [35] or Equation (3.5) in [34].

Property 3.

Assume that Assumptions 1 and 2 hold,  T1,,TN are independently and randomly selected from  T, each according to the distribution with probability density function  ρt . If the solutions to the minimization problems (25) and (27) are unique, then

gNg w.p.1  as  N (30)

and

GMt,gNGMt,g w.p.1  as  N. (31)

for all  tT . If the minimization problems (25) and (27) do not have unique solutions, then for any convergent subsequence of the sequence  gN , where  gNGN ,

gNG w.p.1  as  N (32)

and for any  tT and some  gG , the convergence in (31) holds.

The existence of a convergent subsequence of gN so that the asymptotic property (32) holds results directly from the compactness of G (11). Therefore, under Assumptions 1 and 2, the optimal parameter gN of the FMM is a strongly consistent estimate of some parameter gG.

Since, by Theorems 1 and 2, the model GMt,g is Lipschitz on G uniformly in tT, then the almost-sure convergence of gN to g in (30) implies that, c.f., (Ref. [35]: Remark 2):

suptT  GMt,gNGMt,g0  w.p.1  as  N. (33)

i.e., that GMt,gN is a strongly consistent estimate of the optimal FMM GMt,g, uniformly on T.

Concluding, when Assumptions 1 and 2 are satisfied, the arbitrarily fine approximation of the FMM with the optimal parameter g can be determined (almost everywhere) as the number of measurements N grow enough, even if the real description of the material modulus is fully unknown.

3.2. Exponential Rate of Convergence

Analyzing the convergence in (30) and (32), the question immediately arises of how fast gN tends to some gG as N grows large. As in [35], the distance between the model parameters gN and g will be evaluated by means of the integral identification index Qg (26), i.e., in the sense of the difference QgQgN. For this purpose, it will be checked how fast, for a given small ε>0, the probability PQgQgNε tends to zero, as N increases. From the well-known Hoeffding’s inequality [50], the upper bound of this probability can be derived, analogous to inequality (15) in [35] or inequality (22) in [33] (for details, see Appendix A.1 in [33]):

PQgQgNε2expNε28M^2, (34)

for any ε>0, where

M^=2M+M12+σ2+δ2+2M+M1δ, (35)

with the constants M and M1 defined in Assumption 1 and Equation (17), respectively, the noises’ variance σ2 and upper bound δ are introduced by Assumptions 2 and 3.

The inequality (34) describes the influence of the number of measurements N and the noises’ ”strength” on the rate of convergence. For ε being fixed, the bounds for PQgQgNε decrease exponentially to zero as N increases. The convergence rate is the higher, the lower is M^ (35). In particular, a quick inspection of (35) shows that for stronger measurement noises, the rate of convergence is reduced. Larger δ and σ yield a greater decrease in the rate. This is as expected, since with large disturbances, the measurements are not very adequate. Simultaneously, the larger M+M1, i.e., in view of the estimation (18), the greater the discrepancy between the real modulus and the FMM, the worse the convergence.

3.3. Identification Algorithm

In view of the convergence properties (30), (31) the computation of the parameter gN approximation the parameter g of the optimal FMM requires the next steps:

  1. Select randomly from the set T the sampling times t1,,tN, choosing each ti independently, according to the probability distribution of the density ρt defined given by the weight function in the integral Qg (26).

  2. Conduct the stress relaxation experiment [44,46,48], measure and store the measurements G¯i of the relaxation modulus for the selected time instants ti, i=1,,N.

  3. Solve the identification optimization task (25) and compute the identified model parameter gN.

  4. Put N¯=N and gN¯=gN. To extend the set of experiment data, select new NN¯.

  5. Repeat Steps 1–3 for a new N, that is, randomly choose new sampling times, conduct the rheological experiment once more for a new sample of the material and determine the next gN.

  6. Examine if gN¯gN2<ε, where ε is a small positive number, to check if gN¯ is an adequate approximation of g. If yes, stop the scheme and take gN¯ as the approximate value of g. Otherwise, go again to Step 4.

Remark 1.

A less restrictive testing regarding whether  QN¯gN¯QNgN<ε  holds can be used as an alternative for the stopping rule from Step 6. Both types of stopping rules are commonly used in numerical optimization techniques.

3.4. Numerical Studies

The results of the numerical studies are concerned with the asymptotic properties of the determined optimal FMM and the influence of the measurement noises on this model. Apart from the theoretical analysis above, these simulation studies make it possible to show the respectability and effectiveness of the method developed for FMM identification.

Firstly, it is assumed that the rheological properties of the material are described by the Gaussian-like distribution of the relaxation spectrum, which were used to represent the viscoelastic properties of numerous materials, e.g., polyacrylamide gels [48], native starch gels [38], glass [39], poly(methyl methacrylate) [37], polyethylene [51] and carboxymethylcellulose (CMC) [52]. The spectra of various biopolymers determined by many researchers are Gaussian in nature, for example, cold gel-like emulsions stabilized with bovine gelatin [53], fresh egg white-hydrocolloids [52], some (wheat, potato, corn and banana) native starch gels [38], xanthan gum water solution [52] and wood [54,55].

Next, it is assumed that the material is modeled by the Baumgaertel, Schausberger and Winter (BSW) spectrum [40,41], which was used to describe the viscoelasticity of various polymers; for example, polydisperse polymer melts [42,43], polymethylmethacrylate (PMMA) [56], polybutadiene (PBD) [56] and polymer pelts [57].

The “real” material and the FMM model were simulated in Matlab R2023b, The Mathworks, Inc., Natick, MA, USA. Functions MLFFIT2 [58] and MLF [59], provided by Podlubny, were used for the FMM simulation and numerical solution of the optimal identification tasks.

3.5. Material I

Consider the material whose relaxation spectrum is described by the unimodal Gauss-like distribution:

Hτ=ϑe1τm2/q/τ,

where the parameters are as follows [60]: ϑ=31520 Pa·s, m=0.0912s1 and q=3.25×103s2. The related relaxation modulus is [60]

Gt=πq2ϑ e14t2qmterfc12tqmq, (36)

where the complementary error function erfcx is given by [4] (Equation (C.2))

erfcx=2π  xez2dz.

Following [47], for numerical simulations, the time interval T=0,200 seconds is chosen. Hence, the weighting function in Qg (26) is ρt=1200s1. The elements of the optimal parameter vector g solving the measurement-independent optimization task (27) are given in Table 1.

Table 1.

The components α, β, Ge and τr of the FMM parameter g solving the optimal identification problem (27) and the optimal integral quadratic indices Qg defined by (27) for the “real” relaxation modulus Gt (36).

Qg kPa2 α β Ge kPa τrs
5.2054279 × 10−4 0.920029 1.469033 × 10−2 3.086723 12.949456

The N sampling instants ti for the simulated stress relaxation test were selected randomly according to the uniform distribution on T. A normal distribution with zero mean value and variance σ2 was applied to the random independent generation of the additive measurement noises zi. In the noise robustness analysis, the standard deviations σ=2,5,8 Pa were used. In the analysis of the model asymptotic properties, for any σ numbers of measurements, NN have been applied, where N=50;100;200;500;1000;2000;5000;7000;10,000;12,000;15,000.

3.5.1. Asymptotic Properties

Then, for every pair N,σ, the optimal parameter gN was determined through solving the minimization task (25). The elements of the vectors gN, the mean square indices QNgN and integral QgN indices, and the relative percentage errors of the approximation of the measurement-independent parameter g, defined as

ERR=gNg22/g22·100%, (37)

are given in Table A1, Table A2 and Table A3 for the three standard deviations of the noises. The model approximation error was also estimated via the relative mean error defined as (compare (24))

QNrelg=1Ni=1NG¯TiGMTi,g2G¯Ti2. (38)

The optimal model parameters gN as the functions of the number of measurements N are illustrated by Figure 2 for the noises of σ=2,5,8 Pa. In any subplot, the values of the related parameters of the sampling points-independent model g are depicted by horizontal purple lines. The asymptotic properties are also illustrated by Figure 3 juxtaposing the empirical index mean-square index QNgN, Equation (24), and the integral quadratic sampling instants-independent index QgN, Equation (26), as the functions of N with the index Qg, marked with horizontal lines. In Figure 2 and Figure 3, a logarithmic scale is applied for the horizontal axes. These plots confirm the asymptotic properties of the proposed identification algorithm. The convergence of gN to the parameter g is directly translated into the convergence of QgN into Qg, especially for N5000. The values of the index QNgN for N=50, small compared to those for N100 (see Table A1, Table A2 and Table A3), result from the good fit of the FMM, whose four parameters are optimally selected in problem (25), to only 50 measurement points. For more measurement points, such a good fit is, generally, impossible whenever the real characteristic does not depend on the pre-assumed class of models. A comparison of Figure 2 and Figure 3b with Figure 3a shows that the impact of stronger noises on the values of the empirical index QNgN is much stronger than the impact of the noises on the values of the FMM parameter gN and, consequently, also on the integral index QgN, which does not directly depend on the measurement noises. Given Equation (28), this property is natural and fully justified.

Figure 2.

Figure 2

Dependence of the parameters of the FMM approximating the “real” relaxation modulus (36): (a) αN; (b) βN; (c) GeN; and (d) τrN on the number of measurements N for disturbances σ=2,5,8 Pa; the horizontal purple lines are related to the optimal parameters α, β, Ge and τr independent on the sampling instants used in the rheological experiment.

Figure 3.

Figure 3

The indices of the “real” relaxation modulus (36) approximation by the FMM: (a) the mean-square empirical index QNgN, Equation (24), (b) the integral quadratic sampling instants-independent index QgN, Equation (26), as the functions of the number of measurements N and noises σ=2,5,8 Pa; the horizontal purple lines correspond to the optimal integral index Qg defined in Equation (27).

The quality of the real modulus Gt approximation by the FMM is illustrated in Figure 4, where the measurements G¯i of the real modulus Gt fitted by the optimal model GMt,gN are plotted for the N=100 and N=10,000 measurements and the strongest disturbances; σ=8 Pa. Although, for the N=100 measurements, the models GMt,gN and GMt,g differ slightly (see small subplot), for the N=10,000 measurements, they are practically identical, which is confirmed by the values of ERR (37) equaling 0.52% for N=100 and equaling only 5.58 × 10−4 % for N=10,000 (see Table A3). Even for the strongest noises, the relative errors ERR (37) of the parameters g and gN discrepancy is smaller than 0.002% for N200. This almost excellent fitting of the experiment data by the model GMt,gN is confirmed by the values of the relative square model approximation index QNrelgN (38), which for N200 and the weakest noises does not exceed 0.015%, while for the strongest noises, it does not exceed 0.28%. For the noises considered, the values of the model fit indices QNgN (24) and QNrelgN (38) and the integral quadratic index QgN (26) indicate an excellent fit of the model to the experiment data and the fast convergence of gN to g as N tends to infinity; compare Table A1, Table A2 and Table A3.

Figure 4.

Figure 4

The measurements G¯i (red points) of the “real” relaxation modulus (36) and optimal FMM models: sampling points-independent GMt,g and empirical GMt,gN for N measurements and normal distribution noises with the standard deviation σ=8 Pa: (a) N=100; (b) N=10,000.

3.5.2. Noise Robustness

To examine the effect of the measurement noises, for every pair N,σ, the simulated experiment was repeated n=50 times. In each experiment repetition, the measurement noises zi were generated independently and randomly with a normal distribution, with a zero mean value and variance σ2.

Having in mind the definition of the index QNrelg (38), for the n-element sample, the mean relative relaxation modulus approximation error was determined as follows:

ERRQNrel=1nj=1nQNrelgN,j, (39)

for any pair N,σ, where the vector of the optimal FMM parameters gN,j was computed for j-th experiment repetition, j=1,,n.

For the true relaxation modulus approximation, the mean optimal integral error

ERRQ=1nj=1nQgN,j (40)

was determined for every pair N,σ.

The distance between the vector gN,j and the measurement-independent vector g for the n element sample was estimated by the mean relative error defined as follows (compare ERR (37)):

MERR=1nj=1ngN,jg22/g22·100%. (41)

The indices ERRQNrela (39) and ERRQ (40), as the functions of N and σ, are depicted in the bar in Figure 5, while the index MERR (41) is shown in Figure 6.

Figure 5.

Figure 5

Dependence of the mean indices of the “real” relaxation modulus Gt (36) optimal approximation by the FMM: (a) relative empirical error ERRQNrel (39), (b) integral error ERRQ (40) on the number of measurements N and the noises’ standard deviations σ.

Figure 6.

Figure 6

Dependence of the mean relative error MERR (41) between the optimal parameters gN and g of the FMM approximating the “real” relaxation modulus Gt (36) on the number of measurements N and the noises’ standard deviation σ.

From Figure 5b, it is seen that for N>2000, the number of measurements do not essentially affect the integral index ERRQ, either for weak or strong noises, while both the empirical index ERRQNrel and mean relative error MERR decrease exponentially with the increasing number of measurements, which confirms the analytical analysis performed above. The MERR index is of order 0.55% for N=100, it does not exceed 103 % for N1000 and is smaller than 5×105 % even for the strongest disturbances. This, practically, means determining the sampling points-independent parameter g. The algorithm ensures the very good quality of the measurement approximation even for large noises. The values of the relative relaxation modulus approximation error ERRQNrel, which due to the “real” modulus model difference is lower bounded by 3.191×104%, already for N100 measurements do not exceed 0.35%, and for N1000, fall below 0.028%. The course of the mean integral sampling points-independent index ERRQ (40) as the function of N indicates the asymptotic independence of the model from the sampling points, especially for N5000.

3.6. Material II

Consider the material described by the empirical spectrum of relaxation times τ introduced by Baumgaertel, Schausberger and Winter [40,41],

Hτ=β1ττcρ1+β2ττcρ2eττmax, (42)

which is known to effectively describe polydisperse polymer melts [42,43], with the coefficients [43,47,61] as follows: 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. The spectrum Hτ uniquely defines the relaxation modulus Gt by the following integral [44]:

Gt=0Hττet/τdτ. (43)

Following [47], the time interval T=0,2000 seconds is taken for the experiment simulations; the weighting function is ρt=12000s1. The elements of the optimal parameter vector g, which solve the measurement-independent optimization task (27), are given in Table 2.

Table 2.

The elements α, β, Ge and τr of the FMM parameter g solving the optimal identification task (27) and the optimal integral quadratic indices Qg defined by (27) for the “real” relaxation modulus Gt (42), (43).

Qg MPa2 α β Ge MPa τrs
2.383349 × 10−5 0.736706 8.088257 × 10−2 1.2634125 6.397636 × 103

As previously described, in the simulations, the sampling points ti were randomly selected according to the uniform distribution on T. The standard deviations σ=3,6,8 kPa of the random normally distributed noises zi combined with the number of measurements NN were used for the analysis of the model asymptotic properties.

3.6.1. Asymptotic Properties

For every pair N,σ, the elements of the optimal model parameter gN, the empirical QNgN, QNrelgN and integral QgN indices and the relative percentage errors ERR (37) are given in Table A4, Table A5 and Table A6 in Appendix B. The dependence of the optimal model parameters gN on the number of measurements N for the noises of σ=3,6,8 kPa are illustrated by Figure 7. Figure 8 illustrates the empirical QNgN and integral QgN indices as the functions of N; the value of Qg is marked with purple horizontal lines. These plots confirm the asymptotic properties of the proposed identification algorithm. Figure 8a shows the impact of noises on the values of the empirical index QNgN.

Figure 7.

Figure 7

The parameters of the FMM approximating the relaxation modulus (43) of material described by the BSW relaxation spectrum (42): (a) αN; (b) βN; (c) GeN; and (d) τrN as the functions of the number of measurements N for noises σ=3,6,8 kPa; the horizontal purple lines correspond to optimal model parameters α, β, Ge and τr being independent on the sampling instants used in the experiment.

Figure 8.

Figure 8

The indices of the BSW relaxation modulus (42), (43) approximation by the FMM: (a) the mean-square empirical index QNgN, Equation (24), (b) the integral quadratic sampling instants-independent index QgN, Equation (26), as the functions of the number of measurements N for the noises σ=2,5,8 Pa; the horizontal purple lines correspond to the optimal integral index Qg defined in Equation (27).

The approximation of the real modulus Gt by the FMM is illustrated in Figure 9, where the measurements G¯i of the real modulus Gt along with optimal models GMt,gN and GMt,g are plotted for the N=100 and N=10,000 measurements and the strongest noises σ=8 Pa. However, for N=100, the model parameter error ERR=5.35%, while for N=10,000, we have ERR=0.15%; both for N=100 and N=10,000, the models GMt,gN and GMt,g differ slightly and the respective empirical indices are QNrelgN=3.33 × 10−4 % and QNrelgN=2.0 × 10−7%, respectively.

Figure 9.

Figure 9

The measurements G¯i (red points) of the real relaxation modulus (43) of the material described by the BSW spectrum (42) and the fractional Maxwell optimal models: sampling points-independent GMt,g and empirical GMt,gN for N measurements and additive random normally distributed noises with standard deviation =8 Pa and zero mean value: (a) N=100; (b) N=10,000.

3.6.2. Noise Robustness

For every pair N,σ, the simulated experiment was repeated n=50 times. The mean relative relaxation modulus approximation error ERRQNrel (39), the mean optimal integral error ERRQ (40) and the mean relative error of the parameter g approximation MERR (41) were determined. The indices ERRQNrela and ERRQ are depicted in Figure 10 as the functions of N and σ. Figure 11 illustrates the dependence of the index MERR on N and σ.

Figure 10.

Figure 10

Dependence of the mean indices of the “real” BSW relaxation modulus (42), (43) approximation by the FMM: (a) the mean relative empirical error ERRQNrel (39), (b) the mean optimal sampling points-independent integral error ERRQ (40) on the number of measurements N and the noises’ standard deviation σ.

Figure 11.

Figure 11

Dependence of the mean relative error MERR (41) between the parameters g and gN of the FMM approximating the “real” BSW relaxation modulus (42), (43) on the number of measurements N and the noises’ standard deviation σ.

The mean integral error ERRQ for N12,000 does not depend essentially on the number of measurements, either for small or large noises (see Figure 10b), while both the empirical index ERRQNrel and mean relative error MERR decrease exponentially with the increasing number of measurements, the MERR for N7000. For N1000, the MERR index does not exceed 1.01%, for N7000 it does not exceed 0.22%, while for N10,000, it falls below 0.05% even for the strongest disturbances. The globally optimal parameter g was determined. As is seen from Figure 9, the algorithm practically ensures an excellent quality of the relaxation modulus approximation even for the strongest noises. The values of the relative relaxation modulus approximation error ERRQNrel, already for the N100 measurements, do not exceed 3.3 × 10−4% and for N1000, fall below 8.3 × 10−6%. From the course of the mean integral sampling points-independent index ERRQ (40), as the function of N, we can conclude that the model is practically independent on the sampling instants for N12,000, independently on the measurement noises. The above combined with the close to zero values of ERRQNrel means the determining of the globally optimal model with the parameter g. In conclusion, the courses of both the index ERRQNrel (38), and the indices MERR (41) and ERRQ (40) as the functions of N, indicate the asymptotic independence of the model from the sampling points for a sufficiently large number of measurements.

4. Conclusions

The fractional Maxwell model allows for the modeling of a very wide range of stress relaxation processes in materials. The goal of the FMM identification is, generally, not to achieve a true description of the genuine relaxation modulus, but one that is “optimally accurate” in the assumed sense of the square weighted approximation error and does not depend on the particular sampling instants used in the stress relaxation experiment. The stochastic-type analytical analysis and numerical studies demonstrated that, despite the fact that the real description of the relaxation modulus is completely unknown, an arbitrarily exact approximation of the sampling points-independent optimal FMM can be identified based on the relaxation modulus data sampled randomly, according to respective randomization, when the number of the measurements applied in the experiment appropriately grow large. The four parameters of the approximate FMM are strongly consistent estimates of the parameters of the sampling points-independent model minimizing the integral square approximation error. The resulting identification scheme is simple and useful in application. It requires only the a priori, before the experiment is performed, independent random choice of the time instants at which the relaxation modulus is recorded from the assumed set according to a stationary rule.

Although this article is about modeling the relaxation modulus, the proposed identification scheme can also be successfully applied to the identification of the fractional-order models of creep compliance using the measurements obtained in the retardation test, whenever the respective set of sampling instants is open to manipulation during experimental data collecting. Therefore, the applicability of the identification asymptotically independent of the time instants used in the rheological experiment, to other fractional-order models determination, in particular, Kelvin–Voight, Zener and anti-Zener models, can be the subject of future research.

Appendix A

Appendix A.1. Proof of Theorem 1

Note, firstly, that the assumption gG1 means, in particular, that 0<β0 β<α1. The assumption tT, by definition of the set T=t0,T, where t0>0, implies inequality t>0.

Differentiability and whence continuity of the relaxation modulus model GMt,g (7) with respect to parameter Ge is obvious. The two-parameter Mitteg-Leffler function Eκ,μx (6) is known to be differentiable and continuous with respect to real argument x and parameters κ and μ [36]. Therefore, by (13) and (14), the differentiability of the function GMt,g with respect to the positive relaxation time parameter τr and parameters α and β directly results. To show that the condition (22) is satisfied, it is enough to prove that the partial derivatives of GMt,g with respect to the four model parameters are bounded uniformly on T×G1.

To prove the boundness of the partial derivatives with respect to the relaxation time parameter τr and the orders α and β of the stress and strain derivatives, the spectral representation (21) of the FMM will be applied together with the property concerning the absolute boundness, uniform on the set T×G1, of some definite integrals, which results from the following known [62] property concerning the absolute integrability of the product of absolutely integrable and bounded functions.

Property A1

([62]). If the function fx is absolutely integrable in the interval a, and the function gx is bounded in a,, then the product fxgx is absolutely integrable in a,.

Let us consider definite integral

I0t,g=0r0v,t,gf0v,t,gdv, (A1)

where the function f0v,t,gR+×T×G0R is absolutely integrable with respect to the variable of integration v in R+, uniformly on the set T×G0, where R+=0,, G0Rk, k1 and the function r0v,t,g:R+×T×G0R is absolutely bounded, uniformly on R+×T×G0. The first assumption means that there exists a positive constant m¯ such that

0f0v,t,gdvm¯<, (A2)

for any tT and any gG0, while the second assumption yields

r0v,t,gm̿<, (A3)

for any v,t,gR+×T×G0. From Property A1, the convergence of the integral

0r0v,t,gf0v,t,gdv, (A4)

and, in consequence, of the integral I0t,g (A1) for any tT and any gG0 follows. In view of (A3) and (A4), we have

I0t,g0r0v,t,gf0v,t,gdvm̿0f0v,t,gdvm̿m¯=M¯.

Therefore, the next result holds.

Property A2.

If the function f0v,t,g is absolutely integrable with respect to v in the interval R+=0,, uniformly on the set T×G0 of the rest arguments t,g, and the function r0v,t,g is bounded uniformly on R+×T×G0, then the product r0v,t,gf0v,t,g is absolutely integrable in R+ for any t,gT×G0 and the integral I0t,g (A1) is absolutely bounded uniformly on T×G0.

Below, the proof is divided into four parts related to the four model parameters.

Appendix A.1.1. Uniform on T×G1 Boundness of the FMM Derivative with Respect to Ge

From (7), we have

GMt,gGe=tτrβEαβ,1βtτrαβ, (A5)

whence, by (12) and (16), the uniform boundness of the above derivative on the set T×G1 follows.

Appendix A.1.2. Uniform on T×G1 Boundness of the FMM Derivative with Respect to τr

By (7) and (13), we can express GMt,g as follows

GMt,g=Geeαβ,1βtτr;1.

Whence, the partial derivative with respect to the relaxation time, in the respective notation, is given by

GMt,gτr=Ge1tτr2ddxeαβ,1βx;1x=tτr, (A6)

where, due to the complete monotonicity of the function eαβ,1βx;1, the negative derivative ddxeαβ,1βx;1 monotonically increases to zero for x>0.

From (A5) and (A6), the nonnegative definiteness of GMt,gGe and GMt,gτr, being positive for any gG1 and t<, follows; therefore, Property 1 is formulated.

To examine the asymptotic properties of GMt,gτr as t and as tt0, let us express (A6), applying the known differentiation formula [4] (Equation (E.55))

eκ,μx;λ=ddxeκ,μ+1x;λ,

in the form

GMt,gτr=Ge1tτr2eαβ,βtτr;1,

or, having in mind definition (13), directly in terms of the Mittag-Leffler function

GMt,gτr=1GeτrtτrβEαβ,βtτrαβ. (A7)

The following asymptotic approximation [63] (Equation (12)), see also [4] (Equation (E.30)):

Eκ,μtκn=11n1tκnΓμκn,

which holds for t, applied to (A7), yields

GMt,gτr1GeτrtτrβtτrαβΓα+n=21n1tτrαβnΓβαβn,

whence, for large times, especially for tτrαβ1, we obtain the asymptotic long-time approximation

GMt,gτr1GeτrΓαtτrα,

where means “approximately equal”. Therefore, derivative GMt,gτr tends to zero as t for any admissible parameter gG1.

To estimate the value of GMt,gτr (A7) for t=t0, the series representation

GMt,gτr=Geτrtτrβn=01n+1tτrαβnΓαβnβ

resulting directly from (6) and (A7), is used. The first summand of the series is positive, while the next elements are positive or negative, depending on the index n and the relation between parameters α and β. Since

tτrβn=01n+1tτrαβnΓαβnβtτrβn=0tτrαβnΓαβnβ, (A8)

and the argument of the gamma function is such that

αβnββ>1,

which in view of the monotonicity of the gamma function implies

ΓαβnβΓxmin=0.8856032>0.8856, (A9)

where xmin1.4616321 is the real nonnegative argument at which a minimum of the function Γx occurs [64]. In view of (A8) and (A9), having in mind the nonnegative definiteness of GMt,gτr, we obtain the following estimation

GMt,gτr<Ge1.1292τrtτrβn=0tτrαβn,

which, by the inequalities τrminτrτrmax, for t=t0, implies the next estimation

GMt,gτrt=t0<Ge1.1292τrminτrmaxτrminβt0τrminβn=0t0τrminαβn.

By the assumption t0<τrmin, the above estimation can be rewritten in compact form as

GMt,gτrt=t0<Ge1.1292τrminτrmaxτrminβt0τrminβ1t0τrminαβ. (A10)

Since, for an arbitrary β0 β<α1, we have

t0τrminβ1t0τrminαβ=1t0τrminβt0τrminα<1t0τrminβ,

inequality (A10) for any gG1 implies

GMt,gτrt=t0<Ge1.1292τrminτrmaxt0β,

which means that

GMt,gτrt=t0<Gemax 1.1292τrminτrmaxt0=Gemax 1.1292τrminm0,

i.e., the derivative for t=t0 is bounded, uniformly on the set G1, where positive parameter m0 is defined in Equation (16).

Since the continuous function GMt,gτr of the time t is bounded for t=t0 and for t for any fixed gG1, derivative GMt,gτr as a function of the time is bounded both for the bounded and not-bounded set T. However, due to the non-compactness of the set G1, from which α=β is excluded, the uniform on T×G1 boundness of GMt,gτr is not obvious. Therefore, it should be examined if the maximum of GMt,gτr (with respect to the time) is bounded, as αβ+. To this end, an alternative to (A6) and (A7), the representation of GMt,gτr is derived based on the spectral representation given by Equation (21).

Differentiation of (21) on both sides with respect to τr yields

GMt,gτr=Geατrα1π0rv,gvα1etvdv+Geτrαπ0αβτrvαβ1sinπβτrv2αβ+2τrvαβcosπαβ+1vαetvdvGeτrαπ02αβτrv2αβ1+2αβτrvαβ1cosπαβτrv2αβ+2τrvαβcosπαβ+1rv,gvαetvdv, (A11)

where the function

rv,g=τrvαβsinπβ+sinπατrv2αβ+2τrvαβcosπαβ+1, (A12)

by (21), is such that

GMt,g=Geτrαπ0rv,gvα1etvdv. (A13)

Whence, introducing the notations

r1v,g=τrvαβsinπβτrv2αβ+2τrvαβcosπαβ+1, (A14)
r2v,g=τrv2αβ+τrvαβcosπαβτrv2αβ+2τrvαβcosπαβ+1, (A15)

Equation (A11) can be rewritten as

GMt,gτr=ατrGMt,g+αβGeτrα1π0r1v,gvα1etvdv2αβGeτrα1π0r2v,grv,gvα1etvdv,

or in a more compact form as a linear combination:

GMt,gτr=ατrGMt,g+αβGeτrα1πI1t,g2αβGeτrα1πI2t,g, (A16)

where the integrals:

I1t,g=0r1v,gvα1etvdv, (A17)
I2t,g=0 r2v,grv,gvα1etvdv (A18)

The denominator

qv,g=τrv2αβ+2τrvαβcosπαβ+1 (A19)

of the fractions rv,g, r1v,g and r2v,g is positive for any v0, whenever αβ1, i.e., for any admissible parameter gG1, which satisfies the following inequalities

0<αβ1β1β0<1.

For αβ+, the denominator qv,gτrvαβ+121. By (A13), (17) and nonnegative definiteness of rv,g (A12) on the set R+×G1, we have

0rv,gvα1etvdv=0rv,gvα1etvdv=πGeτrαGMt,gπGeτrαM1

for any t,gT×G1, i.e., the function rv,gvα1etv as the function of the variable v is absolutely integrable, uniformly on the set T×G1, with the constant m¯ (compare (A2)) given by

0rv,gvα1etvdvm¯=πGeminγ2M1, (A20)

where the parameter

0<γ2=minβ0α1τrminα<. (A21)

In view of Property A2, bearing in mind inequality (17), or (15), to prove the absolute uniform boundness of the derivative GMt,gτr (A16) on the set T×G1, it is enough to demonstrate that the integrals I1t,g (A17) and I2t,g (A18) are convergent and absolutely bounded, uniformly on the set T×G1. For this purpose, we express the two integrals as definite integrals of the product of some absolutely integrable function and a bounded function.

The continuous, nonnegative definite for any v,gR+×G1, function r1v,g (A14) is equal to zero for v=0, tends to zero for v and takes the maximal value for v=1/τr, independently on the values of α and β. Whence, for any v,gR+×G1, the inequalities hold

r1v,gsinπβ2cosπαβ+212cosπ1β0+2=122cosπβ0=m1<, (A22)

i.e., r1v,g is absolutely bounded uniformly on R+×G1. By the following notable integral [4] (Equation (A.21):

0vα1etvdv=0vα1etvdv=Γαtα, (A23)

which holds for any α>0 and t>0, vα1etv is the absolutely integrable function of the variable v0 for any tT. Recalling the definitions of the sets T and G1 (19), and the monotonicity of the gamma function Γα for 0<α1, we immediately obtain the estimation

0vα1etvdvΓβ0γ1, (A24)

valid for any tT and any gG1, where

0<γ1=minβ0α1t0α<. (A25)

Therefore, according to Property A2, the integral I1t,g (A17) is convergent for any t>0 and any gG1 and absolutely bounded by the upper bound equal to m1Γβ0/γ1, uniformly on the set T×G1.

It is easy to check that continuous function r2v,g (A15) is equal to zero for v=0 and tends to 1, as v, independently on the values of α and β from the set G1, i.e., for α>β. Function r2v,g can be expressed as

r2v,g=1τrvαβcosπαβqv,g1qv,g,

with qv,g defined by (A19), where the absolute value of the second summand takes the maximal value for v=1/τr, while the third summand takes the maximal value whenever τrvαβ=cosπαβ, if cosπαβ<0, and for v=0 in the opposite case. Therefore, for any v,gR+×G1, the next estimation holds

r2v,g1+cosπαβ2cosπαβ+2+mα,β, (A26)

where

mα,β=11cos2παβ      if cosπαβ<0 1                           if cosπαβ0. (A27)

Since, for gG1, the inequality α>β holds, for αβ+, we have cosπαβ1 and mα,β1. Simultaneously, if α1 and ββ0, then mα,β11cos2πβ0>1, whenever β0<12 and mα,β1 for β012. Therefore, for any v,gR+×G1, by (A26) and (A27), we have

r2v,g1+122cosπβ0+11cos2πβ0=1+m1+m2<, (A28)

where

m2=11cos2πβ0<, (A29)

and according to Property A2, the integral I2t,g (A18) is convergent for any tT and any gG1 and absolutely bounded uniformly on T×G1, with the upper bound π1+m1+m2Geminγ2M1 resulting from (A20) and (A28).

Combining the absolute boundness of the three summands of the right-hand side of (A16), uniform on the set T×G1, the respective uniform boundness of GMt,gτr is proved.

Appendix A.1.3. Uniform on T×G1 Boundness of the FMM Derivative with Respect to β

Differentiation of Equation (21) on both sides with respect to β yields

GMt,gβ=Geτrαπ0lnτrvτrvαβsinπβ+πτrvαβcosπβτrv2αβ+2τrvαβcosπαβ+1vα1etvdv+2Geτrαπ0rv,glnτrvτrv2αβ+lnτrvτrvαβcosπαβπτrvαβsinπαβτrv2αβ+2τrvαβcosπαβ+1vα1etvdv

where rv,g is given by (A12), which using the notations r1v,g (A14) and r2v,g (A15) and introducing functions

r3v,g=τrvαβcosπβτrv2αβ+2τrvαβcosπαβ+1, (A30)
r4v,g=τrvαβsinπαβτrv2αβ+2τrvαβcosπαβ+1, (A31)

can be rewritten in compact form as a linear combination:

GMt,gβ=GeτrαI3t,g2GeτrαI4t,gGeτrαπI5t,g+2GeτrαπI6t,g, (A32)

of four integrals:

I3t,g=0r3v,gvα1etvdv, (A33)
I4t,g=0 r4v,grv,gvα1etvdv, (A34)
I5t,g=0r1v,glnτrvvα1etvdv, (A35)
I6t,g=0r2v,grv,glnτrvvα1etvdv. (A36)

To prove the absolute boundness of the derivative GMt,gβ (A32), uniform on the set T×G1, it is enough to demonstrate that the four above integrals are convergent and absolutely bounded, uniformly on T×G1.

For any v,gR+×G1, the continuous function r3v,g (A30) satisfies the following inequalities (compare (A14) and (A22))

r3v,gcosπβ2cosπαβ+212cosπ1β0+2=122cosπβ0=m1, (A37)

where m1 is defined in (A22), i.e., r3v,g is absolutely bounded, uniformly on R+×G1, which combined with the absolutely integrability of the function vα1etv for any tT and any gG1, according to Property A2, implies the convergence of the integral I3t,g (A33) for any tT and any gG1 and its absolute boundness by the upper bound m1Γβ0/γ1 resulting from and (A24), (A25) and (A37), valid uniformly on the set T×G1.

The nonnegative function r4v,g (A31) is for any v,gR+×G1 bounded by

r4v,gsinπαβ2cosπαβ+212cosπ1β0+2=122cosπβ0=m1, (A38)

similarly to r1v,g (A14) and r3v,g (A30), which combined with the absolutely integrability of rv,gvα1etv for any gG1 and any tT, implies the convergence of the integral I4t,g (A34) for any t,gT×G1 and the absolute boundness of I4t,g by the upper bound π m1Geminγ2M1, derived from (A20) and (A38), uniformly on the set T×G1.

To demonstrate the convergence and uniform boundness of the integrals I5t,g (A35) and I6t,g (A36), let us express them as follows

I5t,g=0r12v,gr5v,t,gvα1e12tvdv, (A39)
I6t,g=0r22v,gr5v,t,grv,gvα1e12tvdv. (A40)

where, by (A14) and (A15)

r12v,g=sinπβτrv2αβ+2τrvαβcosπαβ+1, (A41)
r22v,g=τrvαβ+cosπαβτrv2αβ+2τrvαβcosπαβ+1, (A42)

and nonnegative continuous function

r5v,t,g=τrvαβlnτrve12tv=lnτrvτrvαβe12tv. (A43)

Since (compare (A13)),

GMt2,g=Geτrαπ0rv,gvα1e12tvdv,

the nonnegative function rv,gvα1e12tv is absolutely integrable for any t,gT×G1 with the upper bound πGeminγ2M1; compare (A20). In turn, by (A23), vα1e12tv is an absolutely integrable function of the variable v0, uniformly on the set T×G1, with the upper bound 2Γβ0/γ1; compare (A24) and (A25).

Functions r12v,g (A41) and r22v,g (A42) are absolutely bounded uniformly on R+×G1, since the following estimations hold for any v,gR+×G1:

r12v,g11cos2πβ0=m2, (A44)
r22v,g122cosπβ0+11cos2πβ0=m1+m2<, (A45)

where constants m1 and m2 are defined in (A22) and (A29), respectively.

To examine the properties of r5v,t,g (A43), the asymptotic properties as v0+ and v are studied. This function is expressed as

r5v,t,g=lnτrvτrvαβe12tv,

where the nominator tends to and the denominator tends to +, as the variable v0+. Therefore, by applying the L’Hospital’s rule, in view of α>β>0, we obtain

limv0+r5v,t,g=limv0+1τrαβ+12t τrvτrvαβe12tv=0.

Since τrvαβlnτrv tends to +, while e12tv tends to zero, as the variable v tends to infinity, using the L’Hospital’s rule double times to the right expression in (A43), we have

limvr5v,t,g=limvτr+αβlnτrvτr12te12tvτrv1αβ,

and next

limvr5v,t,g=limvτr+αβlnτrvτr12te12tvτrv1αβ=limvαβ12t12tv+1α+βτrv1α+βe12tv=0+.

Applying two known inequalities ex<k!k!+xk and [65]

lnxnx1n1, (A46)

being valid for any integer n,k>0 and real x>0, by putting n=1 and k=2 function r5v,t,g (A43) for any v,t,gR+×T×G1 can be estimated by

r5v,t,g2τrvαβτrv12+14tτr2τrv2τrvαβτrv+11+18t0τr2τrv2τrmaxvαβτrmaxv+11+18t0τr2τrv2. (A47)

For τrmaxv<1, the right inequality in (A47) implies

r5v,t,g<21+18t0τr2τrv22, (A48)

while for τrmaxv1, the middle inequality in (A47) yields

r5v,t,gτrvαβ1+τrvαβ218t0τr2218t0τr216t0τrmax2=16m02. (A49)

where positive m0 is defined in Equation (16). Combining (A48) and (A49), we obtain the inequality

r5v,t,g<max16m02,2=m4, (A50)

valid for any v,t,gR+×T×G1, which together with (A44) and (A45) means an absolute boundness of continuous functions r12v,gr5v,t,g and r22v,gr5v,t,g, respectively, and in view of the absolute integrability of vα1e12tv and rv,gvα1e12tv, imply the convergence of the integrals I5t,g (A39) and I6t,g (A40) for any tT and any gG1. The absolute boundness of I5t,g and I6t,g, uniform on the set T×G1, with upper bounds estimations 2Γβ0m2m4γ1 and πM1m1+m2m4Geminγ2, respectively, follows from Property A2. Therefore, the absolute boundness of GMt,gβ (A32), uniform on T×G1, is proved.

Appendix A.1.4. Uniform on T×G1 Boundness of the FMM Derivative with Respect to α

The same integral properties and the spectral representation (21) will be applied to prove the boundness of the partial derivative with respect to parameter α. Differentiation (21) on both sides with respect to α yields

dGMt,gdα=lnτrGMt,g+Geτrαπ0lnτrvτrvαβsinπβ+πcosπατrv2αβ+2τrvαβcosπαβ+1vα1etvdv2Geτrαπ0rv,glnτrvτrv2αβ+lnτrvτrvαβcosπαβπτrvαβsinπαβτrv2αβ+2τrvαβcosπαβ+1vα1etvdv+Geτrαπ0rv,gvα1lnvetvdv, (A51)

where rv,g is described by (A12). Having in mind (A13), recalling the notations r1v,g (A14), qv,g (A19), r2v,g (A15), r4v,g (A31), r5v,t,g (A43) and introducing the integrals

I7t,g=0cosπαqv,gvα1etvdv, (A52)
I8t,g=0 rv,gr5v,t,gvβ1e12tvdv,

we can express dGMt,gdα (A51) in a compact form as

dGMt,gdα=GeτrαπI5t,g+GeτrαI7t,g2GeτrαπI6t,g+2GeτrαI4t,g+GeτrαπI8t,g,

where the integrals I5t,g (A35), I6t,g (A36) and I4t,g (A34) are absolutely bounded uniformly on T×G1. Therefore, only the convergence and boundness of the two new integrals I7t,g and I8t,g must be proved.

The boundness of cosπαqv,g, uniform on R+×G1, with the upper bound

cosπαqv,g11cos2πβ0=m2,

combined with the absolute integrability of vα1etv for any t,gT×G1, yields the convergence of the integral I7t,g (A52) and its absolute boundness, uniform on T×G1.

From (A12), the upper bound of the nonnegative function rv,g follows

rv,g122cosπβ0+11cos2πβ0=m1+m2,

where constants m1 and m2 are defined in (A22) and (A29), respectively; therefore, the absolute boundness of r5v,t,g, uniform on R+×T×G1 (c.f., (A50)), and integrability of vβ1e12tv imply both the convergence and the absolute boundness of I8t,g, uniformly on the set T×G1.

The partial derivatives of the FMM with respect to the four model parameters are proved to be absolutely bounded uniformly on the set T×G1; therefore, the uniform boundness (22) of the gradient gGMt,g follows. The theorem is proved.

Appendix A.2. Proof of Theorem 2

Since, for any tT function, GMt,g (10) is differentiable with respect to g and in this case β=α, the four-element vector of model parameters g (8) is as follows

g=ααGeτrT,

the gradient is given by

gGMt,g=GMt,gαGMt,gα12Γ1αtτrαGeα2Γ1αttτr1αT, (A53)

with the partial derivative

GMt,gα=Ge2Γ1αψ1α+lnτrtτrtα, (A54)

where

ψx=ΓxΓx (A55)

is the digamma (or psi) function [66] defined as the logarithmic derivative of the gamma function [66]:

ψx=ddxlnΓx.

For 0<β0α<1, the nonnegative digamma function ψ1α strictly decreases from finite ψ1β0<0 to , while the positive gamma function Γ1α strictly increases from Γ1β0>0 to +. To evaluate the first summand of GMt,gα (A54), the following result proved by Mező and Hoffman [67] is helpful in providing an infinite product representation of the entire function ψzΓz.

Property A3

([67] (Theorem 2.1)). For all zC

ψzΓz=e2γzk=01zαkez/αk, (A56)

where C is the set of complex numbers, γ is the Euler’s constant and αk are the zeros of the digamma function ψz.

It is known [67] that the zeros αk are real, and all but one are negative; here, α0 is the positive zero, and α1, α2, are the negative ones in decreasing order. The course of ψzΓz for the real 0z1 is illustrated by Figure A1. For α=1, by (A56), the quotient ψ1αΓ1α is as follows ψ0Γ0=1, while for α=β0, the quotient ψ1αΓ1α tends to ψ1Γ1=γ. Therefore, for any t,gT×G2

Eψ1α2Γ1ατrtαGemax2τrmaxt0,

i.e., the first summand of GMt,gα (A54) is bounded.

Figure A1.

Figure A1

The quotient ψzΓz of the psi function ψz (A55) by the gamma function Γz for real argument 0z1.

Using the inequality (A46), the second summand of GMt,gα (A54) can be estimated as follows

Ge2Γ1αlnτrtτrtαmGe2Γ1ατrt1m1τrtα,

where an integer m>0. From the above, for t,gT×G2, the next inequality follows

Ge2Γ1αlnτrtτrtαm Gemax2τrmaxt01m1τrmaxt0,

i.e., the second summand of GMt,gα (A54) is bounded, too.

Since, for 0<β0 α1 gamma function Γ1αΓ1β0>1, the last two elements of the gradient (A53) are nonnegative definite (c.f., Property 1) and bounded by 12τrmaxt0 and Gemax2t0τrmaxt0, respectively, for any t,gT×G2. The theorem is proved.

Appendix B

Appendix B.1. The Results of the Numerical Studies for Material I

Table A1.

The elements αN, βN, GeN and τrN of the FMM parameter vector gN solving identification task (25) for real relaxation modulus (36) of the material described by the unimodal Gauss-like distribution, the mean-square identification indices QNgN, Equation (24), the mean relative square model approximation index QNrelgN, Equation (38), the sampling points-independent integral indices QgN defined by the optimization task (26), and the relative errors ERR (37) of the FMM parameter g approximation for N relaxation modulus measurements independently disturbed by additive, zero mean, normally distributed noises with standard deviation σ=2 kPa.

N QNgN kPa2 QNrelgN QgN kPa2 ERR [%] αN βN GeN kPa τrNs
50 9.295186 × 10−5 2.102093 × 10−3 2.863765 × 10−3 0.527 0.963788 8.644275 × 10−2 2.862573 15.170858
100 5.614324 × 10−4 5.907274 × 10−5 1.350707 × 10−3 1.224 0.965614 7.207174 × 10−2 2.745219 16.225154
200 2.185904 × 10−4 1.109724 × 10−4 4.959376 × 10−3 4.537 × 10−3 0.954725 8.815551 × 10−2 3.065976 14.289093
500 5.559271 × 10−4 5.053629 × 10−5 5.262647 × 10−4 7.740 × 10−3 0.915232 1.261495 × 10−2 3.113879 12.68090
1000 4.594139 × 10−4 1.476206 × 10−4 5.268387 × 10−4 3.104 × 10−4 0.920937 1.341017 × 10−2 3.081285 12.926718
2000 4.668059 × 10−4 3.243456 × 10−6 5.396114 × 10−4 5.315 × 10−2 0.929099 2.375304 × 10−2 3.015563 13.557098
5000 5.289533 × 10−4 3.282759 × 10−7 5.213213 × 10−4 2.876 × 10−4 0.920502 1.445854 × 10−2 3.081489 12.984136
7000 5.224292 × 10−4 3.978418 × 10−6 5.215398 × 10−4 1.654 × 10−4 0.920364 1.419721 × 10−2 3.082754 12.979817
10,000 5.082251 × 10−4 3.739909 × 10−6 5.219846 × 10−4 9.301 × 10−5 0.920327 1.391317 × 10−2 3.083746 12.942422
12,000 5.150186 × 10−4 1.363350 × 10−5 5.208965 × 10−4 1.689 × 10−5 0.920187 1.504288 × 10−2 3.085455 12.942422
15,000 5.247854 × 10−4 3.687418 × 10−6 5.205439 × 10−4 1.938 × 10−8 0.920014 1.472034 × 10−2 3.086680 12.949456

Table A2.

The elements αN, βN, GeN and τrN of the FMM parameter vector gN solving identification task (25) for real relaxation modulus (36) of the material described by the unimodal Gauss-like distribution, the mean-square identification indices QNgN, Equation (24), the mean relative square model approximation index QNrelgN, Equation (38), the sampling points-independent integral indices QgN defined by the optimization task (26), and the relative errors ERR (37) of the FMM parameter g approximation for N measurements independently disturbed by additive, zero mean, normally distributed noises with standard deviation σ=5 kPa.

N QNgN kPa2 QNrelgN QgN kPa2 ERR [%] αN βN GeN kPa τrNs
50 1.246174 × 10−4 3.1182072 × 10−3 2.953417 × 10−3 0.501 0.962059 8.720788 × 10−2 2.868287 15.170858
100 6.042548 × 10−4 9.788569 × 10−3 9.505292 × 10−4 0.542 0.950711 5.528005 × 10−2 2.859441 15.193338
200 2.469407 × 10−4 2.622478 × 10−4 4.888504 × 10−3 3.999 × 10−3 0.954563 8.747687 × 10−2 3.067250 14.289093
500 5.754599 × 10−4 1.136117 × 10−7 5.288353 × 10−4 1.1543 × 10−4 0.917799 1.594113 × 10−2 3.090037 12.837417
1000 4.8724914 × 10−4 6.769619 × 10−4 5.261736 × 10−4 2.129 × 10−4 0.921142 1.374989 × 10−2 3.082219 12.926718
2000 4.8687842 × 10−4 3.040833 × 10−7 5.400918 × 10−4 5.345 × 10−2 0.929234 2.394509 × 10−2 3.015365 13.557098
5000 5.4973567 × 10−4 3.940686 × 10−6 5.211862 × 10−4 3.026 × 10−4 0.920472 1.459525 × 10−2 3.081354 12.984135
7000 5.401956 × 10−4 1.559930 × 10−6 5.214093 × 10−4 1.573 × 10−4 0.920350 1.428889 × 10−2 3.082852 12.979817
10,000 5.261243 × 10−4 2.744246 × 10−6 5.221499 × 10−4 5.516 × 10−4 0.920905 1.446157 × 10−2 3.079473 12.979817
12,000 5.356062 × 10−4 5.782985 × 10−5 5.208668 × 10−4 2.386 × 10−5 0.920122 1.495904 × 10−2 3.085216 12.942422
15,000 5.457918 × 10−4 4.535517 × 10−6 5.205517 × 10−4 2.513 × 10−7 0.919980 1.477095 × 10−2 3.086569 12.949456

Table A3.

The elements αN, βN, GeN and τrN of the FMM parameter vector gN solving identification task (25) for real relaxation modulus (36) of the material described by the unimodal Gauss-like distribution, the mean relative square model approximation index QNrelgN, Equation (38), the sampling points-independent integral indices QgN defined by the optimization task (26), and the relative errors ERR (37) of the parameter g approximation for N measurements independently disturbed by additive, zero mean, normally distributed noises with standard deviation σ=8 kPa.

N QNgN kPa2 QNrelgN QgN kPa2 ERR [%] αN βN GeN kPa τrNs
50 1.823164 × 10−4 4.314509 × 10−3 2.993805 × 10−3 0.504 0.961565 8.769949 × 10−2 2.867641 15.1708579
100 6.647523 × 10−4 0.353591 9.411604 × 10−4 0.517 0.949857 5.428199 × 10−2 2.864822 15.193338
200 2.961081 × 10−4 4.334879 × 10−4 4.686224 × 10−3 4.277 × 10−3 0.954203 8.597868 × 10−2 3.066582 14.289093
500 6.078210 × 10−4 2.946700 × 10−5 5.271033 × 10−4 3.193 × 10−4 0.918084 1.579814 × 10−2 3.092238 12.837417
1000 5.328032 × 10−4 2.789884 × 10−3 5.267889 × 10−4 1.300 × 10−4 0.921325 1.356372 × 10−2 3.083203 12.926718
2000 5.248407 × 10−4 3.026647 × 10−7 5.403471 × 10−4 5.202 × 10−2 0.929297 2.403479 × 10−2 3.016320 13.557098
5000 5.883023 × 10−4 1.431097 × 10−5 5.210845 × 10−4 3.179 × 10−4 0.920442 1.473303 × 10−2 3.081219 12.984135
7000 5.759879 × 10−4 1.780341 × 10−7 5.220610 × 10−4 2.087 × 10−3 0.921658 1.538542 × 10−2 3.072622 13.066791
10,000 5.616302 × 10−4 2.127802 × 10−6 5.220589 × 10−4 5.578 × 10−4 0.920855 1.448935 × 10−2 3.079433 12.979817
12,000 5.739756 × 10−4 2.163690 × 10−4 5.211071 × 10−4 8.574 × 10−4 0.921109 1.594595 × 10−2 3.077685 13.006321
15,000 5.843841 × 10−4 5.116492 × 10−6 5.207537 × 10−4 4.857 × 10−4 0.920823 1.552819 × 10−2 3.079921 13.006321

Appendix B.2. The Results of the Numerical Studies for Material II

Table A4.

For the optimal FMM approximating the relaxation modulus (43) of the material described by the BSW spectrum (42): the elements αN βN, GeN and τrN of the vector gN solving identification task (25), the mean-square identification indices QNgN, Equation (24), the mean relative square model approximation index QNrelgN, Equation (38), the sampling points-independent integral indices QgN defined by the optimization task (26), and the relative errors ERR (37) of the parameter g approximation for N relaxation modulus measurements independently disturbed by additive normally distributed noises with standard deviation σ=3 kPa.

N QNgN MPa2 QNrelgN QgN MPa2 ERR [%] αN βN GeN MPa τrNs
50 1.315712 × 10−5 3.597318 × 10−7 5.638335 × 10−5 2.342 0.677682 6.870627 × 10−2 1.370585 5.418555 × 103
100 1.161929 × 10−5 5.298049 × 10−7 4.939762 × 10−5 4.151 0.656134 6.493121 × 10−2 1.409899 5.094185 × 103
200 1.03161 × 10−5 1.124491 × 10−8 5.287861 × 10−5 4.759 0.65004 6.373051 × 10−2 1.421409 5.002011 × 103
500 9.772230 × 10−6 2.140316 × 10−8 2.966257 × 10−5 0.897 0.686475 7.329572 × 10−2 1.336261 5.791700 × 103
1000 1.216504 × 10−5 1.472521 × 10−8 2.964808 × 10−5 0.847 0.691723 7.332173 × 10−2 1.331643 5.808836 × 103
2000 9.462435 × 10−6 2.397636 × 10−9 3.703687 × 10−5 1.928 0.674709 6.955942 × 10−2 1.364964 5.509183 × 103
5000 3.372717 × 10−5 1.236704 × 10−8 2.439945 × 10−5 0.061 0.750336 8.307524 × 10−2 1.245023 6.555588 × 103
7000 3.499392 × 10−5 3.0361484 × 10−9 2.578132 × 10−5 0.179 0.761153 8.483605 × 10−2 1.231209 6.669047 × 103
10,000 3.974638 × 10−5 1.327619 × 10−9 2.524376 × 10−5 0.136 0.757999 8.429268 × 10−2 1.235341 6.633616 × 103
12,000 3.289041 × 10−5 2.540537 × 10−10 2.384835 × 10−5 1.384 × 10−3 0.735839 8.062010 × 10−2 1.265447 6.373837 × 103
15,000 3.259757 × 10−5 2.789747 × 10−10 2.383509 × 10−5 6.536 × 10−5 0.737489 8.098389 × 10−2 1.262561 6.402808 × 103

Table A5.

For the optimal FMM approximating the relaxation modulus (43) of the material described by the BSW spectrum (42): the elements αN, βN, GeN and τrN of the parameter vector gN, the mean-square identification indices QNgN, Equation (24), the mean relative square model approximation index QNrelgN, Equation (38), the sampling points-independent integral indices QgN defined by the optimization task (26), and the relative errors ERR (37) of the parameter g for N measurements corrupted by the noises with standard deviation σ=6 kPa.

N QNgN MPa2 QNrelgN QgN MPa2 ERR [%] αN βN GeN MPa τrNs
50 5.334837 × 10−5 1.607199 × 10−6 5.383509 × 10−5 1.783 0.691583 7.062629 × 10−2 1.352647 5.543339 × 103
100 4.625811 × 10−5 1.912785 × 10−6 5.226179 × 10−5 4.853 0.652363 6.407151 × 10−2 1.421237 4.988252 × 103
200 4.104462 × 10−5 6.4067158 × 10−8 6.964808 × 10−5 7.841 0.632293 5.905374 × 10−2 1.467987 4.606199 × 103
500 3.664178 × 10−5 7.819189 × 10−8 3.074233 × 10−5 1.039 0.681071 7.285674 × 10−2 1.342891 5.745449 × 103
1000 3.879589 × 10−5 6.729167 × 10−8 2.961505 × 10−5 0.869 0.689812 7.328929 × 10−2 1.333233 5.801105 × 103
2000 3.618881 × 10−5 1.385717 × 10−9 3.604767 × 10−5 1.786 0.675299 6.992477 × 10−2 1.362099 5.542684 × 103
5000 5.999448 × 10−5 3.264396 × 10−8 2.456599 × 10−5 8.109 × 10−2 0.752093 8.335898 × 10−2 1.242558 6.579817 × 103
7000 6.106792 5.065819 × 10−9 2.598612 × 10−5 0.195 0.763258 8.506666 × 10−2 1.229229 6.680137 × 103
10,000 6.592012 × 10−5 1.717204 × 10−9 2.536338 × 10−5 0.144 0.759401 8.444666 × 10−2 1.234044 6.640191 × 103
12,000 5.916358 × 10−5 2.831521 × 10−9 2.384641 × 10−5 5.855 × 10−4 0.737150 8.077941 × 10−2 1.264056 6.382156 × 103
15,000 5.893401 × 10−5 1.984301 × 10−10 2.384002 × 10−5 2.644 × 10−4 0.738283 8.108685 × 10−2 1.261701 6.408038 × 103

Table A6.

For the optimal FMM approximating the relaxation modulus (43) of the material described by the BSW spectrum (42): the elements αN, βN, GeN and τrN of the parameter vector gN, the mean-square identification indices QNgN, Equation (24), the mean relative square model approximation index QNrelgN, Equation (38), the sampling points-independent integral indices QgN defined by the optimization task (26), and the relative errors ERR (37) of the parameter g for N measurements corrupted by the noises with standard deviation σ=8 kPa.

N QNgN MPa2 QNrelgN QgN MPa2 ERR [%] αN βN GeN MPa τrNs
50 9.520744 × 10−5 2.941612 × 10−6 6.384002 × 10−5 1.448 0.701586 7.198997 × 10−2 1.340251 5.627729 × 103
100 8.218302 × 10−5 3.329357 × 10−6 5.485285 × 10−5 5.348 0.649888 6.349515 × 10−2 1.428867 4.918087 × 103
200 7.290485 × 10−5 1.233057 × 10−7 8.330015 × 10−5 10.352 0.620745 5.578770 × 10−2 1.501144 4.339259 × 103
500 6.446603 × 10−5 1.355064 × 10−7 3.166061 × 10−5 1.145 0.677484 7.255754 × 10−2 1.347428 5.713136 × 103
1000 6.646278 × 10−5 1.240244 × 10−7 2.966288 × 10−5 0.885 0.688528 7.326539 × 10−2 1.334319 5.795691 × 103
2000 6.395292 × 10−5 2.673029 × 10−9 3.542823 × 10−5 1.693 0.675699 7.016833 × 10−2 1.360185 5.565162 × 103
5000 8.738813 × 10−5 5.134265 × 10−8 2.469399 × 10−5 9.607 × 10−2 0.753278 8.354811 × 10−2 1.240915 6.595934 × 103
7000 8.835062 × 10−5 6.705427 × 10−9 2.613160 × 10−5 0.205 0.764677 8.522041 × 10−2 1.227913 6.687353 × 103
10,000 9.319389 × 10−5 2.003354 × 10−9 2.544649 × 10−5 0.149 0.760331 8.454841 × 10−2 1.233188 6.644512 × 103
12,000 8.640779 × 10−5 6.119895 × 10−9 2.384921 × 10−5 2.494 × 10−4 0.738014 8.088431 × 10−2 1.263145 6.387532 × 103
15,000 8.625405 × 10−5 1.518245 × 10−10 2.384484 × 10−5 4.544 × 10−4 0.738792 8.115159 × 10−2 1.261157 6.411273 × 103

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The author declares no conflicts 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.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]
  • 2.Schiessel H., Metzler R., Blumen A., Nonnejunacher T.F. Generalized viscoelastic models: Their fractional equations with solutions. J. Phys. A Math. Gen. 1995;28:6567–6584. doi: 10.1088/0305-4470/28/23/012. [DOI] [Google Scholar]
  • 3.Metzler R., Nonnenmacher T.F. Fractional relaxation processes and fractional rheological models for the description of a class of viscoelastic materials. Int. J. Plast. 2003;19:941–959. doi: 10.1016/S0749-6419(02)00087-6. [DOI] [Google Scholar]
  • 4.Mainardi F. Fractional Calculus and Waves in Linear Viscoelasticity: An Introduction to Mathematical Models. World Scientific; Hackensack, NJ, USA: 2022. [Google Scholar]
  • 5.Mainardi F., Spada G. Creep, Relaxation and Viscosity Properties for Basic Fractional Models in Rheology. Eur. Phys. J. Spec. Top. 2011;193:133–160. doi: 10.1140/epjst/e2011-01387-1. [DOI] [Google Scholar]
  • 6.Meng R., Yin D., Drapaca C.S. A variable order fractional constitutive model of the viscoelastic behavior of polymers. Int. J. Non-Linear Mech. 2019;113:171–177. doi: 10.1016/j.ijnonlinmec.2019.04.002. [DOI] [Google Scholar]
  • 7.Shiru G., Hai P., Mengsen Y., Ziheng S., Dejun L., Jinyong X., Yiying F. New perspective on the creep characteristic of fiber–dependent shape memory polymers: Variable–order fractional constitutive model. J. Mater. Res. Technol. 2023;24:5177–5185. doi: 10.1016/j.jmrt.2023.04.111. [DOI] [Google Scholar]
  • 8.Liang Z., Li J., Zhang X., Kan Q. A viscoelastic-viscoplastic constitutive model and its finite element implementation of amorphous polymers. Polym. Test. 2023;117:107831. doi: 10.1016/j.polymertesting.2022.107831. [DOI] [Google Scholar]
  • 9.Xu B., Blok R., Teuffel P. An investigation of the effect of relative humidity on viscoelastic properties of flax fiber reinforced polymer by fractional-order viscoelastic model. Compos. Commun. 2023;37:101406. doi: 10.1016/j.coco.2022.101406. [DOI] [Google Scholar]
  • 10.Viviani L., Di Paola M., Royer-Carfagni G. Fractional viscoelastic modeling of laminated glass beams in the pre-crack state under explosive loads. Int. J. Solids Struct. 2022;248:111617. doi: 10.1016/j.ijsolstr.2022.111617. [DOI] [Google Scholar]
  • 11.Xiang G., Yin D., Meng R., Cao C. Predictive model for stress relaxation behavior of glassy polymers based on variable-order fractional calculus. Polym. Adv. Technol. 2021;32:703–713. doi: 10.1002/pat.5123. [DOI] [Google Scholar]
  • 12.Popov I.I., Shitikova M.V., Levchenko A.V., Zhukov A.D. Experimental identification of the fractional parameter of the fractional derivative standard linear solid model for fiber-reinforced rubber concrete. Mech. Adv. Mater. Struct. 2023:1–9. doi: 10.1080/15376494.2023.2191600. [DOI] [Google Scholar]
  • 13.Hajikarimi P., Ehsani M., EL Haloui Y., Tehrani F.F., Absi J., Nejad F.M. Fractional viscoelastic modeling of modified asphalt mastics using response surface method. Constr. Build. Mater. 2022;317:125958. doi: 10.1016/j.conbuildmat.2021.125958. [DOI] [Google Scholar]
  • 14.Cai W., Wang P. Rate-dependent fractional constitutive model for nonlinear behaviors of rubber polymers. Eur. J. Mech. A/Solids. 2024;103:105186. doi: 10.1016/j.euromechsol.2023.105186. [DOI] [Google Scholar]
  • 15.David S.A., Katayama A.H. Fractional Order for Food Gums: Modeling and Simulation. Appl. Math. 2013;4:28194. doi: 10.4236/am.2013.42046. [DOI] [Google Scholar]
  • 16.Stankiewicz A. Fractional Maxwell model of viscoelastic biological materials. BIO Web Conf. 2018;10:02032. doi: 10.1051/bioconf/20181002032. [DOI] [Google Scholar]
  • 17.Bainy E.M., Lenzi E.K., Corazza M.L., Lenzi M.K. Mathematical Modeling of Fish Burger Baking Using Fractional Calculus. Therm. Sci. 2017;21:41–50. doi: 10.2298/TSCI160422241B. [DOI] [Google Scholar]
  • 18.Lei D., Liang Y., Xiao R. A fractional model with parallel fractional Maxwell elements for amorphous thermoplastics. Phys. A Stat. Mech. Its Appl. 2018;490:465–475. doi: 10.1016/j.physa.2017.08.037. [DOI] [Google Scholar]
  • 19.Hou T., Chen H. Isothermal physical aging of PEEK and PPS investigated by fractional Maxwell model. Polymer. 2012;53:2509–2518. doi: 10.1016/j.polymer.2012.04.005. [DOI] [Google Scholar]
  • 20.Elmoghazy Y.H., Safaei B., Asmael M., Sahmani S., Zeeshan Q., Qin Z. Computational Modelling and Analysis of Effect of Viscoelastic Materials on Damping and Vibrational Behaviors of Composite Structures—An Extensive Review. Arch. Computat. Methods Eng. 2024:1–52. doi: 10.1007/s11831-023-10057-4. [DOI] [Google Scholar]
  • 21.Gritsenko D., Paoli R. Theoretical Analysis of Fractional Viscoelastic Flow in Circular Pipes: General Solutions. Appl. Sci. 2020;10:9093. doi: 10.3390/app10249093. [DOI] [Google Scholar]
  • 22.Loglio G., Kovalchuk V.I., Bykov A.G., Ferrari M., Krägel J., Liggieri L., Miller R., Noskov B.A., Pandolfini P., Ravera F., et al. Interfacial Dilational Viscoelasticity of Adsorption Layers at the Hydrocarbon/Water Interface: The Fractional Maxwell Model. Colloids Interfaces. 2019;3:66. doi: 10.3390/colloids3040066. [DOI] [Google Scholar]
  • 23.Kovalchuk V.I., Loglio G., Bykov A.G., Ferrari M., Krägel J., Liggieri L., Miller R., Milyaeva O.Y., Noskov B.A., Ravera F., et al. Effect of Temperature on the Dynamic Properties of Mixed Surfactant Adsorbed Layers at the Water/Hexane Interface under Low-Gravity Conditions. Colloids Interfaces. 2020;4:27. doi: 10.3390/colloids4030027. [DOI] [Google Scholar]
  • 24.Mao B., Zhu R., Wang Z., Yang Y., Han X., Zhao Q. A Modified Fractional Maxwell Numerical Model for Constitutive Equation of Mn-Cu Damping Alloy. Materials. 2020;13:2020. doi: 10.3390/ma13092020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Schmidt A., Gaul L. Parameter Identification and FE Implementation of a Viscoelastic Constitutive Equation Using Fractional Derivatives. Proc. Appl. Math. Mech. 2002;1:153–154. doi: 10.1002/1617-7061(200203)1:1&#x0003c;153::AID-PAMM153&#x0003e;3.0.CO;2-J. [DOI] [Google Scholar]
  • 26.Xiao Z., Haitian Y., Yiqian F. Identification of constitutive parameters for fractional viscoelasticity. Commun. Nonlinear Sci. Numer. Simul. 2014;19:311–322. doi: 10.1016/j.cnsns.2013.05.019. [DOI] [Google Scholar]
  • 27.Zhou H.W., Yi H.Y., Mishnaevsky L., Wang R., Duan Z.Q., Chen Q. Deformation analysis of polymers composites: Rheological model involving time-based fractional derivative. Mech. Time-Depend. Mater. 2017;21:151–161. doi: 10.1007/s11043-016-9323-y. [DOI] [Google Scholar]
  • 28.Zhou S., Cao J., Chen Y. Genetic Algorithm-Based Identification of Fractional-Order Systems. Entropy. 2013;15:1624–1642. doi: 10.3390/e15051624. [DOI] [Google Scholar]
  • 29.Lewandowski R., Chorążyczewski B. Identification of the parameters of the Kelvin–Voigt and the Maxwell fractional models, used to modeling of viscoelastic dampers. Comput. Struct. 2010;88:1–17. doi: 10.1016/j.compstruc.2009.09.001. [DOI] [Google Scholar]
  • 30.Beda T., Chevalier Y. New Methods for Identifying Rheological Parameter for Fractional Derivative Modeling of Viscoelastic Behavior. Mech. Time-Depend. Mater. 2004;8:105–118. doi: 10.1023/B:MTDM.0000027671.75739.10. [DOI] [Google Scholar]
  • 31.Yin Y., Yang Z., Shi M. Circular arc rules of complex plane plot for model parameters determination of viscoelastic material. Mech. Time-Depend. Mater. 2021;25:631–643. doi: 10.1007/s11043-020-09465-x. [DOI] [Google Scholar]
  • 32.Ljung L. System Identification: Theory for the User. 1st ed. Prentice Hall; Englewood Cliffs, NJ, USA: 1999. [Google Scholar]
  • 33.Stankiewicz A., Juściński S. How to Make the Stress Relaxation Experiment for Polymers More Informative. Polymers. 2023;15:4605. doi: 10.3390/polym15234605. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Ljung L. Convergence analysis of parametric identification methods. IEEE Trans. Autom. Control. 1978;23:770–783. doi: 10.1109/TAC.1978.1101840. [DOI] [Google Scholar]
  • 35.Hasiewicz Z., Stankiewicz A. On input-dependent system identification by Monte Carlo approach. IEEE Trans. Autom. Control. 1986;31:192. doi: 10.1109/TAC.1986.1104226. Erratum in IEEE Trans. Autom. Control 1985, 30, 480–483. [DOI] [Google Scholar]
  • 36.Apelblat A. Differentiation of the Mittag-Leffler Functions with Respect to Parameters in the Laplace Transform Approach. Mathematics. 2020;8:657. doi: 10.3390/math8050657. [DOI] [Google Scholar]
  • 37.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]
  • 38.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]
  • 39.Wang J., Wang X., Ruan H. On the mechanical β relaxation in glass and its relation to the double-peak phenomenon in impulse excited vibration at high temperatures. J. Non-Cryst. Solids. 2020;533:119939. doi: 10.1016/j.jnoncrysol.2020.119939. [DOI] [Google Scholar]
  • 40.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]
  • 41.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]
  • 42.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]
  • 43.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]
  • 44.Ferry J.D. Viscoelastic Properties of Polymers. 3rd ed. John Wiley & Sons; New York, NY, USA: 1980. [Google Scholar]
  • 45.Gorenflo R., Mainardi F. Fractional Calculus: Integral and Differential Equations of Fractional Order. In: Carpinteri A., Mainardi F., editors. Fractals and Fractional Calculus in Continuum Mechanics. Springer; Wien, Austria: New York, NY, USA: 1997. pp. 223–276. [Google Scholar]
  • 46.Malkin A.I.A., Malkin A.Y., Isayev A.I. Rheology: Concepts, Methods and Applications. ChemTec; Deerfield Beach, FL, USA: 2006. [(accessed on 5 April 2023)]. Available online: https://books.google.pl/books?id=8rGafjhgz-UC. [Google Scholar]
  • 47.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]
  • 48.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]
  • 49.Andrei N. Modern Numerical Nonlinear Optimization. Springer; Cham, Switzerland: 2022. [DOI] [Google Scholar]
  • 50.Hoeffding W. Probability inequalities for sums of bounded random variables. J. Am. Stat. Assoc. 1963;58:13–30. doi: 10.1080/01621459.1963.10500830. [DOI] [Google Scholar]
  • 51.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]
  • 52.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 June 2023)]. Available online: https://onlinelibrary.wiley.com/doi/book/10.1002/9781119108580. [Google Scholar]
  • 53.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]
  • 54.Bardet S., Gril J. Modelling the transverse viscoelasticity of green wood using a combination of two parabolic elements. C. R. Mec. 2002;330:549–556. doi: 10.1016/S1631-0721(02)01503-6. [DOI] [Google Scholar]
  • 55.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]
  • 56.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]
  • 57.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]
  • 58.Podlubny I. Fitting Data Using the Mittag-Leffler Function. MATLAB Central File Exchange. [(accessed on 23 July 2023)]. Available online: https://www.mathworks.com/matlabcentral/fileexchange/32170-fitting-data-using-the-mittag-leffler-function.
  • 59.Podlubny I. Mittag-Leffler Function with Two Parameters. MATLAB Central File Exchange. [(accessed on 23 July 2023)]. Available online: https://www.mathworks.com/matlabcentral/fileexchange/8738-mittag-leffler-function.
  • 60.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]
  • 61.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]
  • 62.Fikhtengol’ts G.M. The Fundamentals of Mathematical Analysis. Volume 2 Pergamon Press; Oxford, UK: London, UK: 2013. (International Series of Monographs in Pure and Applied Mathematics). [Google Scholar]
  • 63.Lemes N.H.T., dos Santos J.P.C., Braga J.P. A generalized Mittag-Leffler function to describe nonexponential chemical effects. Appl. Math. Model. 2016;40:7971–7976. doi: 10.1016/j.apm.2016.04.021. [DOI] [Google Scholar]
  • 64.Deming W., Colcord C. The Minimum in the Gamma Function. Nature. 1935;135:917. doi: 10.1038/135917b0. [DOI] [Google Scholar]
  • 65.Topsøe F. Some bounds for the logarithmic function. In: Cho Y.J., Kim J.K., Dragomir S.S., editors. Inequality Theory and Applications. Nova Science Publishers; New York, NY, USA: 2007. p. 137. [Google Scholar]
  • 66.Alzer H. On Some Inequalities for the Gamma and Psi Functions. Math. Comput. 1997;66:373–389. doi: 10.1090/S0025-5718-97-00807-7. [DOI] [Google Scholar]
  • 67.Mező I., Hoffman M.E. Zeros of the digamma function and its Barnes G-function analogue. Integral Transform. Spec. Funct. 2017;28:846–858. doi: 10.1080/10652469.2017.1376193. [DOI] [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 Materials are provided here courtesy of Multidisciplinary Digital Publishing Institute (MDPI)

RESOURCES