Skip to main content
Log in

On the Kramers-Kronig relations

  • Original Contribution
  • Published:
Rheologica Acta Aims and scope Submit manuscript

Abstract

We provide a new derivation of the Kramers-Kronig relations on the basis of the Sokhotski-Plemelj equation with detailed mathematical justifications. The relations hold for a causal function, whose Fourier transform is regular (holomorphic) and square-integrable. This implies analyticity in the lower complex plane and a Fourier transform that vanishes at the high-frequency limit. In viscoelasticity, we show that the complex and frequency-dependent modulus describing the stiffness does not satisfy the relation but the modulus minus its high-frequency value does it. This is due to the fact that despite its causality, the modulus is not square-integrable due to a non-null instantaneous response. The relations are obtained in addition for the wave velocity and attenuation factor. The Zener, Maxwell, and Kelvin-Voigt viscoelastic models illustrate these properties. We verify the Kramers-Kronig relations on experimental data of sound attenuation in seabottoms sediments.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2

Similar content being viewed by others

References

  • Bracewell R (1965) The Fourier transform and its applications. McGraw-Hill Book Co

  • Carcione JM (1999) On energy definition in electromagnetism: an analogy with viscoelasticity. J Acous Soc Am 105(2):626–632

    Article  Google Scholar 

  • Carcione JM (2014) Wave fields in real media. Theory and numerical simulation of wave propagation in anisotropic, anelastic, porous and electromagnetic media. Elsevier. (Third edition, extended and revised)

  • Frisch M (2009) ‘The most sacred tenet’? Causal reasoning in physics. Br J Philos Sci 60(3):459–474

    Article  Google Scholar 

  • Golden JM, Graham GAC (1988) Boundary value problems in linear viscoelasticity. Springer, Berlin

    Book  Google Scholar 

  • Kramers HA (1927) La diffusion de la lumiere par les atomes. Atti Congr Intern Fisica, Como 2:545–557

    Google Scholar 

  • Kronig R de L (1926) On the theory of the dispersion of X-rays. J Opt Soc Am 12:547–557

    Article  Google Scholar 

  • Labuda C, Labuda I (2014) On the mathematics underlying dispersion relations. Eur Phys J H, https://doi.org/10.1140/epjh/e2014-50021-1

  • Liu HP, Anderson DL, Kanamori H (1976) Velocity dispersion due to anelasticity; implications for seismology and mantle composition. Geophys J Roy Astr Soc 47:41–58

    Article  Google Scholar 

  • Nussenzveig HM (1960) Causality an dispersion relations for fixed momentum transfer. Physics 26:209–229

    Google Scholar 

  • Nussenzveig HM (1972). In: Bellman R (ed) Causality and dispersion relations, mathematics in science and engineering, vol 95. Academic Press, New York

  • Plemelj J (1908) Riemannsche Funktionenscharen mit gegebener Monodromiegruppe. Monatshefte für Mathematik und Physik 19:211–246

    Article  Google Scholar 

  • Pritz T (1999) Verification of local Kramers-Kronig relations for complex modulus by means of fractional derivative model. J Sound Vib 228:1145–1165

    Article  Google Scholar 

  • Sokhotskii YW (1873) On definite integrals and functions used in series expansions. St. Petersburg

  • Stastna J, De Kee D, Powley MB (1985) Complex viscosity as a generalized response function. J Rheol 29:457–469

    Article  Google Scholar 

  • Toksöz MN, Johnston DH (eds) (1981) Seismic wave attenuation. Geophysical Reprint Series, Tulsa

  • Wang Y (2007) Seismic inverse Q filtering. Blackwell Pub

  • Zener C (1948) Elasticity and anelasticity of metals. University of Chicago, Chicago

    Google Scholar 

  • Zhou J-X, Zhang X-Z, Knobles DP (2009) Low-frequency geoacoustic model for the effective properties of sandy seabottoms. J Acoust Soc Amer 125:2847–2866

    Article  Google Scholar 

Download references

Acknowledgements

This work is supported by the Specially-Appoin- ted Professor Program of Jiangsu Province, China, the Cultivation Program of “111” Plan of China (BC2018019) and the Fundamental Research Funds for the Central Universities, China.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jing Ba.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendix: the Zener and Maxwell models verify the Kramers-Kronig relations

Appendix: the Zener and Maxwell models verify the Kramers-Kronig relations

In this Appendix we derive, first, a very compact unified formulation of Kramers-Kronig relations, using Hilbert transform. Then we show, through detailed ad hoc computations, that the reduced complex modulus of the Zener model verifies this condition.

A.1 Compact formulation of Kramers-Kronig relations

As we have seen in the main text, Kramers-Kronig relations for a complex-valued function f of a real variable ω may be expressed as follows:

$$ \left\{ \begin{array}{llllllllllll} \text{Re} f &= \mathcal{H} (\text{Im} f ) \\ \text{Im} f &= - \mathcal{H} (\text{Re} f ) \end{array} \right. $$
(47)

where symbol \( \mathcal {H } \) denotes Hilbert transform.

Summing (1)1 and (1)2 multiplied by the imaginary unit i, and rearranging, we obtain the following:

$$ f = - \mathrm{i} \mathcal{H } (f ) $$
(48)

where we have used the linearity of Hilbert transform.

Conversely, taking real and imaginary parts of Eq. 48, we obtain (47). Therefore, the single condition (48) is equivalent to Kramers-Kronig relations.

A.2 Zener model

We recall that the complex modulus of the Zener model may be expressed as follows:

$$ M(\omega) = M_{\infty} - \frac{ M_{\infty} - M_{0} }{ 1 + \mathrm{i} \tau \omega } $$
(49)

Now, we show that the reduced complex modulus

$$ \overline{M}(\omega) = M(\omega) - M_{\infty} = - \frac{ M_{\infty} - M_{0} }{ 1+ \mathrm{i} \tau \omega } $$
(50)

verifies formulation (48) of Kramers-Kronig relations, i.e.,

$$ \overline{M}(\omega) = (M_{\infty} - M_{0} ) \mathrm{i} \mathcal{H} \left( \frac{ 1 }{ 1 + \mathrm{i} \tau \omega } \right) $$
(51)

The Hilbert transform of a function f(ω) is given by the following:

$$\begin{array}{@{}rcl@{}} \mathcal{H} f(\omega ) &=& \frac{1}{\pi} \mathcal{P} \int_{-\infty}^{\infty} g(\omega, \omega^{\prime}) \mathrm{d} \omega^{\prime} = \frac{1}{\pi} \lim\limits_{L \rightarrow \infty} \mathcal{P} \int_{-L}^{L} g(\omega, \omega^{\prime}) \mathrm{d} \omega^{\prime} \\ &=& \frac{1}{\pi} \lim\limits_{L \rightarrow \infty} \lim\limits_{\epsilon \rightarrow 0^{+}} \left( \int_{-L}^{\omega - \epsilon} g(\omega , \omega^{\prime} ) \mathrm{d} \omega^{\prime} + \int_{\omega + \epsilon}^{L} g(\omega , \omega^{\prime} ) \mathrm{d} \omega^{\prime} \right) \\ &=& \frac{1}{\pi} \lim\limits_{L \rightarrow \infty} \lim\limits_{\epsilon \rightarrow 0^{+}} [ G(\omega, \omega - \epsilon ) - G(\omega, -L )\\&&+ G(\omega, L ) - G(\omega, \omega + \epsilon )] \end{array} $$
(52)

where

$$ g(\omega, \omega^{\prime}) = \frac{ f(\omega^{\prime}) }{ \omega - \omega^{\prime} } $$
(53)

and G is a primitive of the integrand in Eq. 52, i.e.,

$$ G(\omega , \omega^{\prime} ) = \int g(\omega, \omega^{\prime} ) \mathrm{d} \omega^{\prime} $$
(54)

Therefore, the Hilbert transform in Eq. 51 is given by the following:

$$ \mathcal{H} \left( \frac{ 1 }{ 1 + \mathrm{i} \tau \omega } \right) = \frac{1}{ {\tau \omega - \text i} } $$
(55)

as shown in the Lemma below.

Substituting (55) into (51) and rearranging, one obtains an identity, which proves that the reduced complex modulus \( \overline {M}(\omega )\) verifies Kramers-Kronig relations.

Lemma

The Hilbert transform of functionf(ω) = 1/(1 + iτω) isgiven by the following:

$$ \mathcal{H} \left( \frac{ 1 }{ 1 + \mathrm{i} \tau \omega } \right) = \frac{1}{ {\tau \omega - \text i} } $$
(56)

Proof We rely on identity (52), where now

$$ g(\omega , \omega^{\prime} ) = \frac{ 1 }{ (\omega - \omega^{\prime}) (1+ \mathrm{i} \tau \omega^{\prime}) } $$
(57)

and hence its primitive is

$$ G(\omega , \omega^{\prime} ) = \frac{ 2 \tan^{-1}(\tau \omega^{\prime}) + 2 \mathrm{i} \ln(\omega^{\prime} - \omega) - \mathrm{i} \ln[1 + (\tau \omega^{\prime})^{2} ] }{ 2 (\tau \omega - \mathrm{i}) } $$
(58)

as can be checked by differentiating with respect to \( \omega ^{\prime } \). Substituting (58) into (52), one obtains (56). Computations are cumbersome, but straightforward: indeed two divergent terms appear, \( \ln (\epsilon ) \) and \( \ln (- \epsilon ) \), but they are present only in the combination \( \ln (\epsilon ) - \ln (- \epsilon ) = \ln (-1) = \mathrm {i} \pi \).

A.3 Maxwell model

The Maxwell model is obtained from Fig. 1 by removing the parallel spring. Its reduced complex modulus is as follows:

$$ \overline{M}(\omega) = M(\omega) - M_{\infty} = - \frac{ M_{\infty} }{ 1+ \mathrm{i} \tau \omega } $$
(59)

(e.g., Carcione 2014). From Eqs. 48 and 56, we have the following:

$$ \overline{M}(\omega) = - \mathrm{i} \mathcal{H } (\overline{M} ) = \mathrm{i} M_{\infty} \mathcal{H} \left( \frac{ 1 }{ 1 + \mathrm{i} \tau \omega } \right) = \frac{\mathrm{i} M_{\infty}}{\tau \omega - \text i} = \overline{M} (\omega) , $$
(60)

so we have shown that the Maxwell model satisfies the relations.

A.4 Kelvin-Voigt model

The Kelvin-Voigt model is obtained from Fig. 1 by removing the series spring. Its complex modulus is as follows:

$$ M(\omega) = M_{0} (1+ \mathrm{i} \tau \omega ) $$
(61)

(e.g., Carcione 2014). \(M_{\infty } = \infty \) in this case. We may consider \(\overline {M} = \mathrm {i} \tau \omega M_{0}\), since the Hilbert transform of a constant (M0) is zero. This is the complex modulus of the dashpot.

The primitive function is as follows:

$$ G(\omega , \omega^{\prime} ) = - \mathrm{i} \tau M_{0} [\omega^{\prime}+\omega \ln (\omega^{\prime}-\omega)] . $$
(62)

It can be easily seen that the calculation (52) diverges, and therefore the Kelvin-Voigt solid does not satisfy the relations. Another explanation may be found by considering the relaxation rate \(\dot \psi = M_{0} \delta + M_{0} \tau \delta \), which is the inverse Fourier transform of the complex modulus. One immeditely sees that it is causal, but both of its terms are singular distributions. Thus, there is no way, because of the lack of regularity, to introduce a “reduced complex modulus” as in the previous cases. Equivalently, one may notice that both terms in the complex modulus M = M0 + iωτM0 are not square-integrable.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Carcione, J.M., Cavallini, F., Ba, J. et al. On the Kramers-Kronig relations. Rheol Acta 58, 21–28 (2019). https://doi.org/10.1007/s00397-018-1119-3

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00397-018-1119-3

Keywords

Navigation