FIG 3 - uploaded by Reiner Zorn
Content may be subject to copyright.
͑ a ͒ Correspondence between Cole–Davidson exponent ␥ and Kohlrausch exponent ␤ resulting from equating the second logarithmic moment ͑ continuous curve ͒ . For comparison the ‘‘least-squares’’ correspondence established by Lindsey and Patterson ͑ Ref. 18 ͒ ͑ dashed curve ͒ and the correspondence from asymptotic behavior ͑ dotted curve ͒ . ͑ b ͒ The performance of the three approximations demonstrated by comparing ⑀ Љ ͑␻͒ for ␤ ϭ 1/2. The 

͑ a ͒ Correspondence between Cole–Davidson exponent ␥ and Kohlrausch exponent ␤ resulting from equating the second logarithmic moment ͑ continuous curve ͒ . For comparison the ‘‘least-squares’’ correspondence established by Lindsey and Patterson ͑ Ref. 18 ͒ ͑ dashed curve ͒ and the correspondence from asymptotic behavior ͑ dotted curve ͒ . ͑ b ͒ The performance of the three approximations demonstrated by comparing ⑀ Љ ͑␻͒ for ␤ ϭ 1/2. The 

Source publication
Article
Full-text available
In this paper a novel way to quantify "nonexponential" relaxations is described. So far, this has been done in two ways: (1) by fitting empirical functions with a small number of parameters, (2) by calculation of the underlying distribution function g(ln tau) of (exponential) relaxations using regularization methods. The method described here is in...

Contexts in source publication

Context 1
... this method the logarithmic moments of g (ln ) can be calculated for the Kohlrausch function 1 ͑ row 11 of Table I ͒ . For ␤ ϭ 1/2 the logarithmic moments are the same as for the Rajagopal–Ngai distribution. This is because the latter is constructed to be the exact distribution of relaxation times of the Kohlrausch function in the special case ␤ ϭ 1/2 and a reasonable approximation for other values of ␤ . 14 For the most commonly used empirical relaxation functions with one adjustable parameter, the Kohlrausch, Cole– Davidson, and Cole–Cole expressions, Fig. 1 shows the first and second logarithmic moments. As expected the logarithmic width of the corresponding relaxation time distribution increases with decreasing parameter ␤ , ␣ , or ␥ . If the parameters have the value one all three functions reduce to the Debye/exponential case; thus the width of the distribution vanishes. If the parameters approach zero the width diverges as ␴ ln ␶ ϳ␤ Ϫ 1 , ␣ Ϫ 1 , ␥ Ϫ 1 . For the asymmetric ones among these functions the absolute value of the skewness increases with increasing parameter ␤ or ␥ ; they even diverge as ͉ ␥ 1 ln ␶ ͉ ϳ (1 Ϫ ␤ ) Ϫ 1/2 , (1 Ϫ ␥ ) Ϫ 1/2 . This is a counterintuitive consequence of normalizing the third moment by the cubed width in definition ͑ 10 ͒ . Figure 2 shows a graphical representation of the results for the two-parameter Havriliak-Negami function. As abscissa an alternative definition of the skewness, ␮ 3 ln ␶ / ␴ ln 2 ␶ , ͑ 23 ͒ has been chosen to avoid the divergence of definition ͑ 10 ͒ in the limit ␣ → 1. It can be seen that the Havriliak–Negami function is able to represent any width of the relaxation but only negative skewness ͑ as has been pointed out by Havriliak and Havriliak 16 this deficiency can be resolved by allowing ␥ Ͼ 1 with the weaker restriction ␣␥ Ͻ 1 but only a few authors actually use this generalization ͒ in a certain range delimited by the Cole–Davidson case labeled by ␣ ϭ 1 in the plot. From Fig. 2 and the equivalent plots for the Jonscher and Matsumoto–Higasi functions it seems that the combination ‘‘large skewness, small width’’ cannot be attained. But this is not a mathematical property because distributions with any combination of the first three moments can be constructed. It is rather a general experimental outcome than an a priori limitation that such highly skewed distributions do not occur in reality. It is clear that all the empirical functions listed in Table I are distinct. They only coincide if one is a special case of another, e.g., the Cole–Davidson function and the Havriliak–Negami function. Especially, the Kohlrausch function has no equivalent in the frequency domain. Its Fourier transform is only known for special cases where ␤ is a rational number, 17 e.g., for ␤ ϭ 1/2: ⑀ ͑ ⑀ ␻ 0 Ϫ ͒ Ϫ ⑀ ρ ⑀ ρ ϭ ͱ 4 i ␻␶ ␲ K exp ͩ 4 i ␻␶ 1 K ͪ erfc ͩ ͱ 4 i ␻␶ 1 K ͪ . ͑ 24 ͒ This expression clearly cannot be represented exactly by any of the empirical functions in frequency domain. Therefore, considerable effort has been made in the literature to establish approximate relationships between the Kohlrausch function and frequency domain expressions. Starting from the asymptotic properties of the Kohlrausch function that ⑀ Љ ͑␻͒ϳ␻ for low frequencies and ⑀ Љ ( ␻ ) ϳ ␻ Ϫ ␤ for high frequencies the Cole–Davidson function with ␥ ϭ ␤ seems to be the appropriate choice. But Fig. 3 ͑ b ͒ shows that the high- and low-frequency wings have different levels and will not coincide despite having the same slope. Therefore, it may be a better compromise for represent- ing actual data to choose a mediatory value ␥ Ͻ ␤ . Such an attempt was first described by Lindsey and Patterson ͑ LP ͒ . 18 On grounds of a least-squares fit they propose the relation ␤ ϭ ͭ 0.970 0.683 ␥ ␥ ϩ ϩ 0.144 0.316 for for 0.2 0.6 р ␥ р р ␥ 0.6, . ͑ 25 ͒ Using the logarithmic moments derived here an alternative correspondence can be established by imposing that the moments up to the second should coincide. This results in the relation . 26 ͱ 6 ␺ Ј ͑ ␥ ͒ Figure 3 ͑ a ͒ shows a comparison of the correspondences. It can be seen that the one based on the logarithmic moments lies in between the LP formula and the ‘‘asymptotic’’ one. Concerning the agreement the LP correspondence performs slightly better in the peak while the logarithmic moments approach works better on the high frequency side. In both cases the low frequency wing is poorly represented. In order to overcome this deficiency one can use a two- parameter function, namely Havriliak–Negami, as was first described by Alvarez, Alegr ́a, and Colmenero ͑ AAC ͒ . 19 They use a numerical algorithm to invert ͑ 6 ͒ and thus obtain the distribution function of the Kohlrausch relaxation. Sub- sequently, they use this distribution to calculate ⑀ K Љ ( ␻ ) from ͑ 7 ͒ . Finally, the Havriliak–Negami function is determined with the parameters for which ⑀ HN Љ ( ␻ ) fits best to the previ- ously calculated function. From the tabulated values of cor- responding exponent parameters, the following approximate relation to the Kohlrausch ␤ is established for given Havriliak–Negami parameters: ␤ ϭ ͑ ␣␥ ͒ 0.813 . ͑ 27 ͒ For a given Kohlrausch ␤ , the additional relation ␥ ϭ 1 Ϫ 0.8121 ͑ 1 Ϫ ␣ ͒ 0.387 ͑ 28 ͒ together with ͑ 27 ͒ fixes the optimal ␣ , ␥ pair to be used. The characteristic times are related ...
Context 2
... this method the logarithmic moments of g (ln ) can be calculated for the Kohlrausch function 1 ͑ row 11 of Table I ͒ . For ␤ ϭ 1/2 the logarithmic moments are the same as for the Rajagopal–Ngai distribution. This is because the latter is constructed to be the exact distribution of relaxation times of the Kohlrausch function in the special case ␤ ϭ 1/2 and a reasonable approximation for other values of ␤ . 14 For the most commonly used empirical relaxation functions with one adjustable parameter, the Kohlrausch, Cole– Davidson, and Cole–Cole expressions, Fig. 1 shows the first and second logarithmic moments. As expected the logarithmic width of the corresponding relaxation time distribution increases with decreasing parameter ␤ , ␣ , or ␥ . If the parameters have the value one all three functions reduce to the Debye/exponential case; thus the width of the distribution vanishes. If the parameters approach zero the width diverges as ␴ ln ␶ ϳ␤ Ϫ 1 , ␣ Ϫ 1 , ␥ Ϫ 1 . For the asymmetric ones among these functions the absolute value of the skewness increases with increasing parameter ␤ or ␥ ; they even diverge as ͉ ␥ 1 ln ␶ ͉ ϳ (1 Ϫ ␤ ) Ϫ 1/2 , (1 Ϫ ␥ ) Ϫ 1/2 . This is a counterintuitive consequence of normalizing the third moment by the cubed width in definition ͑ 10 ͒ . Figure 2 shows a graphical representation of the results for the two-parameter Havriliak-Negami function. As abscissa an alternative definition of the skewness, ␮ 3 ln ␶ / ␴ ln 2 ␶ , ͑ 23 ͒ has been chosen to avoid the divergence of definition ͑ 10 ͒ in the limit ␣ → 1. It can be seen that the Havriliak–Negami function is able to represent any width of the relaxation but only negative skewness ͑ as has been pointed out by Havriliak and Havriliak 16 this deficiency can be resolved by allowing ␥ Ͼ 1 with the weaker restriction ␣␥ Ͻ 1 but only a few authors actually use this generalization ͒ in a certain range delimited by the Cole–Davidson case labeled by ␣ ϭ 1 in the plot. From Fig. 2 and the equivalent plots for the Jonscher and Matsumoto–Higasi functions it seems that the combination ‘‘large skewness, small width’’ cannot be attained. But this is not a mathematical property because distributions with any combination of the first three moments can be constructed. It is rather a general experimental outcome than an a priori limitation that such highly skewed distributions do not occur in reality. It is clear that all the empirical functions listed in Table I are distinct. They only coincide if one is a special case of another, e.g., the Cole–Davidson function and the Havriliak–Negami function. Especially, the Kohlrausch function has no equivalent in the frequency domain. Its Fourier transform is only known for special cases where ␤ is a rational number, 17 e.g., for ␤ ϭ 1/2: ⑀ ͑ ⑀ ␻ 0 Ϫ ͒ Ϫ ⑀ ρ ⑀ ρ ϭ ͱ 4 i ␻␶ ␲ K exp ͩ 4 i ␻␶ 1 K ͪ erfc ͩ ͱ 4 i ␻␶ 1 K ͪ . ͑ 24 ͒ This expression clearly cannot be represented exactly by any of the empirical functions in frequency domain. Therefore, considerable effort has been made in the literature to establish approximate relationships between the Kohlrausch function and frequency domain expressions. Starting from the asymptotic properties of the Kohlrausch function that ⑀ Љ ͑␻͒ϳ␻ for low frequencies and ⑀ Љ ( ␻ ) ϳ ␻ Ϫ ␤ for high frequencies the Cole–Davidson function with ␥ ϭ ␤ seems to be the appropriate choice. But Fig. 3 ͑ b ͒ shows that the high- and low-frequency wings have different levels and will not coincide despite having the same slope. Therefore, it may be a better compromise for represent- ing actual data to choose a mediatory value ␥ Ͻ ␤ . Such an attempt was first described by Lindsey and Patterson ͑ LP ͒ . 18 On grounds of a least-squares fit they propose the relation ␤ ϭ ͭ 0.970 0.683 ␥ ␥ ϩ ϩ 0.144 0.316 for for 0.2 0.6 р ␥ р р ␥ 0.6, . ͑ 25 ͒ Using the logarithmic moments derived here an alternative correspondence can be established by imposing that the moments up to the second should coincide. This results in the relation . 26 ͱ 6 ␺ Ј ͑ ␥ ͒ Figure 3 ͑ a ͒ shows a comparison of the correspondences. It can be seen that the one based on the logarithmic moments lies in between the LP formula and the ‘‘asymptotic’’ one. Concerning the agreement the LP correspondence performs slightly better in the peak while the logarithmic moments approach works better on the high frequency side. In both cases the low frequency wing is poorly represented. In order to overcome this deficiency one can use a two- parameter function, namely Havriliak–Negami, as was first described by Alvarez, Alegr ́a, and Colmenero ͑ AAC ͒ . 19 They use a numerical algorithm to invert ͑ 6 ͒ and thus obtain the distribution function of the Kohlrausch relaxation. Sub- sequently, they use this distribution to calculate ⑀ K Љ ( ␻ ) from ͑ 7 ͒ . Finally, the Havriliak–Negami function is determined with the parameters for which ⑀ HN Љ ( ␻ ) fits best to the previ- ously calculated function. From the tabulated values of cor- responding exponent parameters, the following approximate relation to the Kohlrausch ␤ is established for given Havriliak–Negami parameters: ␤ ϭ ͑ ␣␥ ͒ 0.813 . ͑ 27 ͒ For a given Kohlrausch ␤ , the additional relation ␥ ϭ 1 Ϫ 0.8121 ͑ 1 Ϫ ␣ ͒ 0.387 ͑ 28 ͒ together with ͑ 27 ͒ fixes the optimal ␣ , ␥ pair to be used. The characteristic times are related ...

Similar publications

Article
Full-text available
For irreducible smooth representations $\Pi$ of $\mathrm{GSp}(4,k)$ over a non-archimedean local field $k$, Piatetskii-Shapiro and Soudry have constructed an $L$-factor depending on the choice of a Bessel model. It factorizes into a regular part and an exceptional part. We determine the regular part for the case of split Bessel models.
Article
Full-text available
The time series of geophysical data are chaotic and, on the other hand, extremely noisy. Thus, though there are a number of advanced methods of chaotic time series prediction, the improvement of geophysical data is crucial to succeed. Mainly, it is connected with low determinism of such time series. The improvement procedure, we are about to repres...
Article
Full-text available
This paper solves the inpainting problem of single depth images. depth images are regarded as natural images without texture. Because of the sparsity property of natural images and the textureless property of depth images, we propose a similar group-based sparse model with sparse gradient regularization. For one thing, the similar group-based spars...
Conference Paper
Full-text available
We present a method for generating approximate 2D and 3D floor plans derived from 3D point clouds. The plans are approximate boundary representations of built indoor structures. The algorithm slices the 3D point cloud, combines concave primary boundary shape detection and regularization algorithms as well as k-means clustering for detection of seco...
Article
Full-text available
In [20], we derived representation formulae for spatially periodic solutions to the generalized, inviscid Proudman-Johnson equation and studied their regularity for several classes of initial data. The purpose of this paper is to extend these results to larger classes of functions including those having arbitrary local curvature near particular poi...

Citations

... Both the fitted parameters and the computed total correlation time τ G shown in Table II obtained for various system sizes do not indicate any system size dependence, which is in accordance with the finding of Celebi et al. 53 who noticed that the finite size correction for the rotational diffusion scales with the inverse box volume and is, therefore, much smaller than the one for translational diffusion. Moreover, we would like to point out that the fitted exponents shown in Table II with β ≈ 0.9 indicate a rather narrow distribution of relaxation times with a distribution width στ with στ ≈ τ K , 55 suggesting a shape of the corresponding spectral densities close to that of a Lorentzian. Next, we want to compute the frequency-dependent spectral densities and thus the frequency-dependent relaxation rates. ...
Article
We present a computational framework for reliably determining the frequency-dependent intermolecular and intramolecular nuclear magnetic resonance (NMR) dipole–dipole relaxation rates of spin 1/2 nuclei from Molecular Dynamics (MD) simulations. This approach avoids the alterations caused by the well-known finite-size effects of translational diffusion. Moreover, a procedure is derived to control and correct for effects caused by fixed distance-sampling cutoffs and periodic boundary conditions. By construction, this approach is capable of accurately predicting the correct low-frequency scaling behavior of the intermolecular NMR dipole–dipole relaxation rate and thus allows for the reliable calculation of the frequency-dependent relaxation rate over many orders of magnitude. Our approach is based on the utilization of the theory of Hwang and Freed for the intermolecular dipole–dipole correlation function and its corresponding spectral density [L.-P. Hwang and J. H. Freed, J. Chem. Phys. 63, 4017–4025 (1975)] and its combination with data from MD simulations. The deviations from the Hwang and Freed theory caused by periodic boundary conditions and sampling distance cutoffs are quantified by means of random walker Monte Carlo simulations. An expression based on the Hwang and Freed theory is also suggested for correcting those effects. As a proof of principle, our approach is demonstrated by computing the frequency-dependent intermolecular and intramolecular dipolar NMR relaxation rates of 1H nuclei in liquid water at 273 and 298 K based on the simulations of the TIP4P/2005 model. Our calculations are suggesting that the intermolecular contribution to the 1H NMR relaxation rate of the TIP4P/2005 model in the extreme narrowing limit has previously been substantially underestimated.
... First, the finite-difference error in estimating the reaction coefficient will be quantified. The finite-difference derivative approximation to the reaction coefficient in Eq. (23) will always overestimate the exact analytical solution by a small offset since the odd derivatives of a decay function are always negative [21], whereby P (t), P (t) < 0 in the Taylor expamsion: ...
Article
A reaction limited by standard diffusion is simulated stochastically to illustrate how the continuous time random walk (CTRW) formalism can be implemented with minimum statistical error. A step-by-step simulation of the diffusive random walk in one dimension reveals the fraction of surviving reactants P(t) as a function of time, and the time-dependent unimolecular reaction rate coefficient K(t). Accuracy is confirmed by comparing the time-dependent simulation to results from the analytical master equation, and the asymptotic solution to that of Fickian diffusion. An early transient feature is shown to arise from higher spatial harmonics in the Fourier distribution of walkers between reaction sites. Statistical ‘shot’ noise in the simulation is quantified along with the offset error due to the discrete time derivative, and an optimal simulation time interval \(\Delta t_0\) is derived to achieve minimal error in the finite time-difference estimation of the reaction rate. The number of walkers necessary to achieve a given error tolerance is derived, and \(W = 10^7\) walkers is shown to achieve an accuracy of \(\pm 0.2\%\) when the survival probability reaches \(P(t) \sim \frac{1}{3}\). The stochastic method presented here serves as an intuitive basis for understanding the CTRW formalism, and can be generalized to model anomalous diffusion-limited reactions to prespecified precision in regimes where the governing wait-time distributions have no analytical solution.
... (eqn (8) 15 ), e Àhln[t]i (eqn (9) 26 ). ...
... Previously we used numerically-approximated ESDs for the Generalised Debye model. 10 Logarithmic moments have been shown to be an appropriate metric for comparing different distributions, 26 ...
... 31 It may also occur that the distribution of rates in a sample is significantly skewed, and the waveform data are not well represented by a symmetric Generalised Debye distribution: one alternative is to use the Havriliak-Negami model 7 which also has a known expectation value and variance. 26 For compounds that are required to be modelled like this then we expect that e hln[t]i will provide the best measure to compare relaxation times in both cases. ...
Article
Full-text available
The use of magnetisation decay measurements to characterise very slow relaxation of the magnetisation in single-molecule magnets is becoming increasingly prevalent as relaxation times move to longer timescales outside of the AC susceptibility range. However, experimental limitations and a poor understanding of the distribution underlying the stretched exponential function, commonly used to model the data, may be leading to misinterpretation of the results. Herein we develop guidelines on the experimental design, data fitting, and analysis required to accurately interpret magnetisation decay measurements. Various measures of the magnetic relaxation rate extracted from magnetisation decay measurements of [Dy(Dtp)2][Al{OC(CF3)3}4] previously characterised by Evans et al., fitted using combinations of fixing or freely fitting different parameters, are compared to those obtained using the innovative square-wave "waveform" technique of Hilgar et al. The waveform technique is comparable to AC susceptometry for measurement of relaxation rates on long timescales. The most reliable measure of the relaxation time for magnetisation decays is found to be the average logarithmic relaxation time, e〈ln[τ]〉, obtained via a fit of the decay trace using a stretched exponential function, where the initial and equilibrium magnetisation are fixed to first measured point and target values respectively. This new definition causes the largest differences to traditional approaches in the presence of large distributions or relaxation rates, with differences up to 50% with β = 0.45, and hence could have a significant impact on the chemical interpretation of magnetic relaxation rates. A necessary step in progressing towards chemical control of magnetic relaxation is the accurate determination of relaxation times, and such large variations in experimental measures stress the need for consistency in fitting and interpretation of magnetisation decays.
... 7), ⟨τ⟩ −1 (Eq. 8 15 ), e −⟨ln [τ]⟩ (Eq. 9 26 ). ...
... 31 It may also occur that the distribution of rates in a sample is significantly skewed, and the waveform data are not well represented by a symmetric Generalised Debye distribution: one alternative is to use the Havriliak-Negami model 7 which also has a known expectation value and variance. 26 For compounds that are required to be modelled like this then we expect that e ⟨ln [τ]⟩ will provide the best measure to compare relaxation times in both cases. ...
Preprint
Full-text available
The use of magnetisation decay measurements to characterise very slow relaxation of the magnetisation in single-molecule magnets are becoming increasingly prevalent as relaxation times move to longer timescales outside of the AC susceptibility range. However, experimental limitations and a poor understanding of the distribution underlying the stretched exponential function, commonly used to model the data, may be leading to misinterpretation of the results. Herein we develop guidelines on the experimental design, data fitting, and analysis required to accurately interpret magnetisation decay measurements. Different measures of the magnetic relaxation rate extracted from magnetisation decay measurements of [Dy(Dtp)2][Al{OC(CF3)3}4], fitted using different combinations of fixing or freely fitting different parameters, are compared to those obtained using the innovative square-wave “waveform” technique of Hilgar et al, which is comparable to AC susceptometry for measurement of relaxation rates on long timescales. The most reliable measure of the relaxation time for magnetisation decays is found to be the the average logarithmic relaxation time, $e^{⟨ln [τ]⟩}$, obtained via a fit of the decay trace using a stretched exponential function, where the initial and equilibrium magnetisation are fixed to first measured point and target values respectively. This new definition has causes the largest differences to traditional approaches in the presence of large distributions or relaxation rates, with differences up to 50% with β = 0.45, and hence could have a significant impact on the chemical interpretation of magnetic relaxation rates. A necessary step in progressing towards chemical control of magnetic relaxation is the accurate determination of relaxation times, and such large variations in experimental measures stress the need for consistency in fitting and interpretation of magnetisation decays.
... We will use this convention from now on as it is the usual choice in the literature. Both equations (23) and (27) lead to the same real and imaginary parts of equations (24) (equivalent to the dielectric case given in equation (26)). The Debye model fails to model short times (and thus high frequencies) and violates the sum-rule that ∞ 0 ωχ (ω) dω should remain finite, so modifications sometimes need to be considered if very high-frequency studies are carried out [17] which, for example, can be done using time-domain terahertz spectroscopy [18]. ...
... In the case of a single relaxation time τ c , g(τ ) = δ(τ − τ c ) and the ideal Debye model with τ = τ c is recovered as applicable to systems such as superparamagnets [21] and single-molecule magnets [1]. Various forms of distributions of relaxation times have been considered and are listed in reference [27], though in that work the distribution of relaxation times f (ln(τ )) is defined in terms of the logarithm of the relaxation time, so that the integral in equation (34) is written as d ln τ f (ln τ )/(1 + iωτ ) (though the two forms can easily be related using d ln ...
Preprint
Full-text available
The experimental technique of a.c. susceptibility can be used as a probe of magnetic dynamics in a wide variety of systems. Its use is restricted to the low-frequency regime and thus is sensitive to relatively slow processes. Rather than measuring the dynamics of single spins, a.c. susceptibility can be used to probe the dynamics of collective objects, such as domain walls in ferromagnets or vortex matter in superconductors. In some frustrated systems, such as spin glasses, the complex interactions lead to substantial spectral weight of fluctuations in the low-frequency regime, and thus a.c. susceptibility can play a unique role. We review the theory underlying the technique and magnetic dynamics more generally and give applications of a.c. susceptibility to a wide variety of experimental situations.
... Moleculat dynamics simulation studies [5] have recently shown that the stretching of the relaxation cannot be simply assigned to the superposition of spatially distributed heterogeneities, but already exists on a very local scale. Obtaining a model-free distribution of relaxation times from experimental data is a difficult task and it has been proposed to describe the distribution of relaxation times by means of logarithmic moments that quantify the characteristic time of the relaxation, its width or stretching, and its asymmetry [6]. Molecular mobility has often been characterized in computer simulations using different approaches to evaluate the distribution of relaxation times in equilibrium states [7,8]. ...
... The width of the distribution, or stretching parameter σ logτ , has also been evaluated as the square root of the variance of the distribution of logarithmic relaxation times, its asymmetry γ logτ as the skewness of the distribution g(logτ), and its kurtosis. These parameters are also good descriptors to compare with the experimental-wide distributions of relaxation times, regardless of the models (see [6]). ...
Article
Full-text available
Glass transition processes have often been explained in terms of wide distributions of relaxation times. By means of a simple stochastic model we here show how dynamic heterogeneity is the key to the emergence of the glass transition. A non-Markovian model representing a small open region of the amorphous material was previously shown to reproduce the time and thermal characteristic behavior of supercooled liquids and glasses. Due to the interaction of the open regions with their environment, the temperature dependence of the equilibrium relaxation times differs from the featureless behavior of the relaxation times of closed regions, whose static disorder does not lead to a glass transition, even with wider distributions of equilibrium relaxation times. The dynamic heterogeneity of the open region produces a glass transition between two different regimes: a faster-than-Arrhenius and non-diverging growth of the supercooled liquid relaxation times and an average Arrhenius behavior of the ideal glass. The Kovacs’ expansion gap was studied by evaluating the nonequilibrium distribution of relaxation times after the temperature quenches.
... Different statistical parameters have to be calculated to compare the shape of the spectra and quantify the changes with polymer dosage and temperature. Considering the normalized form for the spectra, the α th moment is calculated using the following equation as defined in Tschoegl (1989), Zorn (2002), and Bhattacharjee et al. (2011). ...
Article
Full-text available
Modification using elastomeric thermoplastic polymers is commonly adopted to improve the high-temperature performance of paving bitumen. The performance of modified bitumen classified under the same grade is highly variable depending on the type of base bitumen, the polymer architecture, and its dosage. The current specification parameters are insensitive to such variability. Identification of a suitable set of parameters that can quantify the changes in rheological properties due to various interaction mechanisms of bitumen with modifier thus becomes necessary. In this study, the base bitumen obtained using two different processes, namely air rectification and component blending, are considered. Though the same grade of bitumen produced using both processes is considered, the material compositions are different, and this necessitated the use of different polymer architectures (diblock and triblock SBS) for the two binders. Three different dosages are used for each modifier. A stress relaxation experiment is conducted, and the peak modulus and stress relaxation time are determined. In addition, the continuous relaxation spectrum and the associated parameters are computed. The base bitumen and the polymer architecture of the corresponding polymer influenced the stress relaxation response substantially. These factors also influenced the response of the material captured using the relaxation spectrum and exhibited interesting insights regarding the influence of temperature.
... This union has brought new formulations for fractional diffusive models, since they render the Riemann-Liouville-Caputo fractional derivatives as individual cases [8,9,24]. In addition, the fractional Prabhakar derivative has been the most complete and sophisticated tool for describing complexity in physical systems such as, in the description of relaxation and response in anomalous dielectrics of Havriliak-Negami type [2,28], in dynamical models of spherical stellar systems [4] and in the fractional Viscoelasticity [10]. For a practical guide to Prabhakar fractional calculus refer to [11]. ...
... In addition, assume that the matrix I − K C T γ ,ρ,α,ω is nonsingular and f k,M is the approximate solution of the problem (1) obtained using Eq. (28). Then ...
... where F is the approximate solution of the linear algebraic system (28). ...
Article
Full-text available
This paper presents a Legendre wavelet spectral method for solving a type of fractional Fredholm integro–differential equations. The fractional derivative is defined in the Caputo–Prabhakar sense. The derivative of Prabhakar consists of an integro–differential operator that has a Mittag–Leffler function with three parameters in the integration kernel, so it generalizes the Riemann–Liouville and Caputo fractional operators. Moreover, it has many applications in several fields of computational physics. We first derive a matrix method to solve linear problems. In this method, the given linear problem is reduced to a linear system of algebraic equations. The detailed convergence analysis of the proposed matrix method is given. An iterative matrix method is then constructed for nonlinear problems. The nonlinear problem is first replaced with a sequence of linear problems by utilizing the quasilinearization technique. Then, this sequence of problems is successively solved using the matrix method. Numerical examples are included to demonstrate the efficiency and accuracy of the proposed methods.
... where ω 0 is the angular turnover frequency and α is a dimensionless number between zero and one [26,128,129]. It is generally established that it is the diversity of relaxation timescales that is responsible for the observed anomalous electric response of tissue environments [130], which is the source of fractional order evolution in our model as well. ...
Preprint
Full-text available
We present a theoretical framework to model the electric response of cell aggregates. We establish a coarse representation for each cell as a combination of membrane and cytoplasm dipole moments. Then we compute the effective conductivity of the resulting system, and thereafter derive a Fokker-Planck partial differential equation that captures the time-dependent evolution of the distribution of induced cellular polarizations in an ensemble of cells. Our model predicts that the polarization density parallel to an applied pulse follows a skewed t-distribution, while the transverse polarization density follows a symmetric t-distribution, which are in accordance with our direct numerical simulations. Furthermore, we report a reduced order model described by a coupled pair of ordinary differential equations that reproduces the average and the variance of induced dipole moments in the aggregate. We extend our proposed formulation by considering fractional order time derivatives that we find necessary to explain anomalous relaxation phenomena observed in experiments as well as our direct numerical simulations. Owing to its time-domain formulation, our framework can be easily used to consider nonlinear membrane effects or intercellular couplings that arise in several scientific, medical and technological applications.
... The separation of -process is clearly seen in the spectra of shrimps after 3 days of starvation and in the group regenerated by 14 days. Although there are no clear peaks visible in the spectra, the dielectric loss curves from the 3-day and regenerated groups have been fitted by Negami dielectric functions [21][22][23] which are given by the equation: ...
Preprint
Full-text available
In this work, dielectric studies on Neocaridina davidi shrimps have been presented. The effect of starvation on dielectric properties such as conductivity or permittivity have been shown. It was found that the onset frequency of electrode polarization depends on starvation period, which is probably related to the cytoplasm viscosity. In the dielectric spectra of shrimps two relaxation processes have been identified i.e. α and β process. The α process is probably related to the counter ion polarization while β process to the mobility of macromolecules present in the body of a shrimp, mainly to the amount of lipids. It was also found that there is a difference in dielectric response between control group and the group regenerated after 14 day starvation period. Basing on dielectric response, one can conclude that the viscosity of cytoplasm of regenerated shrimps is higher and the cells are rich in the lipid droplets when compared to control group.