Log in

Modeling electrical resistivity and particle fluxes with multiply broken power-law distributions

  • Regular Article
  • Published:
The European Physical Journal Plus Aims and scope Submit manuscript

Abstract

Multiply broken power-law densities are introduced to model empirical data sets extending over several logarithmic decades. Two examples demonstrating the wide range of applicability of these distributions are discussed. First, the temperature dependence of the electrical resistivity of metals (Cu, Au, Cr, Al, W, Fe, Ni) is inferred by nonlinear least-squares regression covering the solid phase up to the melting point. The regressed broken power laws are compared with high- and low-temperature scaling predictions obtained from electron–phonon and electron–electron scattering. In the intermediate temperature range, inflection points arise in Log–Log plots of the electrical resistivity, unaccounted for by the Bloch–Grüneisen theory. In the second example, the energy evolution of cosmic-ray electron and positron fluxes is analyzed in terms of multiply broken and exponentially cut power-law densities, based on recently obtained number counts. Index functions quantifying the spectral variation and survival functions (complementary cumulative distributions) of the particle and energy fluxes are calculated from the regressed densities.

Graphic abstract

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

Access this article

Subscribe and save

Springer+ Basic
EUR 32.99 /Month
  • Get 10 units per month
  • Download Article/Chapter or Ebook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
Subscribe now

Buy Now

Price includes VAT (Thailand)

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16
Fig. 17
Fig. 18
Fig. 19
Fig. 20
Fig. 21
Fig. 22
Fig. 23
Fig. 24
Fig. 25
Fig. 26

Similar content being viewed by others

References

  1. M.P.H. Stumpf, M.A. Porter, Science 335, 665 (2012)

    Article  ADS  MathSciNet  Google Scholar 

  2. M.E.J. Newman, Contemp. Phys. 46, 323 (2005)

    Article  ADS  Google Scholar 

  3. A. Clauset, C.R. Shalizi, M.E.J. Newman, SIAM Rev. 51, 661 (2009)

    Article  ADS  MathSciNet  Google Scholar 

  4. X. Gabaix, Annu. Rev. Econ. 1, 255 (2009)

    Article  Google Scholar 

  5. X. Gabaix, Q. J. Econ. 114, 739 (1999)

    Article  Google Scholar 

  6. E. Puppin, Eur. Phys. J. Plus 135, 420 (2020)

    Article  Google Scholar 

  7. C.D. Lai, D. Murthy, M. **e, in Springer Handbook of Engineering Statistics, H. Pham, ed. (Springer, London, 2006)

  8. R. Tomaschitz, Physica A 483, 438 (2017)

    Article  ADS  MathSciNet  Google Scholar 

  9. M. Dalla Via, C. Bianca, I. El Abassi, A. Darcherif, Eur. Phys. J. Plus 135, 198 (2020)

  10. D. Ahlborn, Eur. Phys. J. Plus 135, 568 (2020)

    Article  Google Scholar 

  11. J.C. Phillips, Rep. Prog. Phys. 59, 1133 (1996)

    Article  ADS  Google Scholar 

  12. D. Apitz, P.M. Johansen, J. Appl. Phys. 97, 063507 (2005)

  13. M.N. Berberan-Santos, E.N. Bodunov, B. Valeur, Chem. Phys. 315, 171 (2005)

    Article  Google Scholar 

  14. E. Limpert, W.A. Stahel, M. Abbt, Bioscience 51, 341 (2001)

    Article  Google Scholar 

  15. M. Mitzenmacher, Internet Math. 1, 226 (2004)

    Article  MathSciNet  Google Scholar 

  16. J. Eeckhout, Am. Econ. Rev. 94, 1429 (2004)

    Article  Google Scholar 

  17. J. Eeckhout, Am. Econ. Rev. 99, 1676 (2009)

    Article  Google Scholar 

  18. M. Levy, Am. Econ. Rev. 99, 1672 (2009)

    Article  Google Scholar 

  19. M. Bee, M. Riccaboni, S. Schiavo, Econ. Lett. 120, 232 (2013)

    Article  Google Scholar 

  20. R. Tomaschitz, Physica A 541, 123188 (2020)

  21. R. Tomaschitz, Appl. Phys. A 126, 102 (2020)

    Article  ADS  Google Scholar 

  22. R.A. Matula, J. Phys. Chem. Ref. Data 8, 1147 (1979)

    Article  ADS  Google Scholar 

  23. T.K. Chu, C.Y. Ho, CINDAS Rep. 60 (1982), https://ia800104.us.archive.org/28/items/DTIC_ADA129078/DTIC_ADA129078.pdf

  24. P.D. Desai, H.M. James, C.Y. Ho, J. Phys. Chem. Ref. Data 13, 1131 (1984)

    Article  ADS  Google Scholar 

  25. P.D. Desai, T.K. Chu, H.M. James, C.Y. Ho, J. Phys. Chem. Ref. Data 13, 1069 (1984)

    Article  ADS  Google Scholar 

  26. J. Yang, in Thermal Conductivity, T.M. Tritt, ed. (Springer, New York, 2004)

  27. C. Uher, in Thermal Conductivity, T.M. Tritt, ed. (Springer, New York, 2004)

  28. M. Aguilar et al., Phys. Rev. Lett. 122, 041102 (2019)

  29. M. Aguilar et al., Phys. Rev. Lett. 122, 101101 (2019)

  30. O. Adriani et al., Phys. Rev. Lett. 120, 261102 (2018)

  31. R. Tomaschitz, Physica B 593, 412243 (2020)

  32. R. Tomaschitz, J. Phys. Chem. Solids 152, 109773 (2021)

  33. N.L. Johnson, S. Kotz, N. Balakrishnan, Continuous Univariate Distributions, vol. 1, 2nd ed. (Wiley, New York, 1994)

  34. J.B. McDonald, Econometrica 52, 647 (1984)

    Article  Google Scholar 

  35. J.B. McDonald, Y.J. Xu, J. Econometrics 66, 133 (1995)

    Article  Google Scholar 

  36. J. Luckstead, S. Devadoss, D. Danforth, Physica A 474, 237 (2017)

    Article  ADS  MathSciNet  Google Scholar 

  37. I. Băncescu, L. Chivu, V. Preda, M. Puente-Ajovín, A. Ramos, Physica A 526, 121017 (2019)

  38. R. Tomaschitz, Phys. Lett. A 393, 127185 (2021)

  39. C. Bianca, M. Menale, Eur. Phys. J. Plus 134, 143 (2019)

    Article  Google Scholar 

  40. C. Bianca, M. Menale, Eur. Phys. J. Plus 136, 243 (2021)

    Article  Google Scholar 

  41. R. Tomaschitz, Physica A 451, 456 (2016)

    Article  ADS  MathSciNet  Google Scholar 

  42. S.W. Barwick et al., Astrophys. J. 498, 779 (1998)

    Article  ADS  Google Scholar 

  43. M.A. DuVernois et al., Astrophys. J. 559, 296 (2001)

    Article  ADS  Google Scholar 

  44. M. Boezio et al., Adv. Space Res. 27, 669 (2001)

    Article  ADS  Google Scholar 

  45. C. Grimani et al., Astron. Astrophys. 392, 287 (2002)

    Article  ADS  Google Scholar 

  46. M. Aguilar et al., Phys. Lett. B 646, 145 (2007)

    Article  ADS  Google Scholar 

  47. M. Ackermann et al., Phys. Rev. Lett. 108, 011103 (2012)

  48. O. Adriani et al., Phys. Rev. Lett. 111, 081102 (2013)

  49. J. Alcaraz et al., Phys. Lett. B 484, 10 (2000)

    Article  ADS  Google Scholar 

  50. J.J. Beatty et al., Phys. Rev. Lett. 93, 241102 (2004)

  51. J. Chang et al., Nature 456, 362 (2008)

    Article  ADS  Google Scholar 

  52. F. Aharonian et al., Phys. Rev. Lett. 101, 261104 (2008)

  53. F. Aharonian et al., Astron. Astrophys. 508, 561 (2008)

    Article  ADS  Google Scholar 

  54. O. Adriani et al., Phys. Rev. Lett. 106, 201101 (2011)

  55. G. Ambrosi et al., Nature 552, 63 (2017)

    Article  ADS  Google Scholar 

  56. S. Abdollahi et al., Phys. Rev. D 95, 082007 (2017)

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Roman Tomaschitz.

Appendix A: Log–Log and Lin-Log coordinates, power-law slopes, lognormals, and Index functions

Appendix A: Log–Log and Lin-Log coordinates, power-law slopes, lognormals, and Index functions

A positive function \(y = f(x)\) in linear coordinates \(x,y \ge 0\) is transformed into logarithmic coordinates (denoted by a hat) \(x = 10^{{\hat{x}}}\), \(y = 10^{{\hat{y}}}\) by

$$ \hat{y} = {\text{Log}}(f(10^{{\hat{x}}} )) = \frac{{\log (f({\text{e}}^{{\hat{x}\log 10}} ))}}{\log 10}, $$
(A.1)

where \({\text{Log(}}x) = \log x/\log 10\) denotes the decadic logarithm. Conversely, a function \(\hat{y} = g(\hat{x})\) in Log–Log representation is converted via \(\hat{x} = {\text{Log(}}x)\), \(\hat{y} = {\text{Log(}}y)\) into linear coordinates \(x,y \ge 0\) by

$$ y = 10^{{g({\text{Log}}(x))}} = \exp \left[ {g(\frac{\log x}{{\log 10}})\log 10} \right]. $$
(A.2)

To obtain a tangent line in a double-logarithmic plot, we note \(\hat{y}^{\prime}(\hat{x}) = xf^{\prime}(x)/f(x)\), according to (A.1). The Log–Log tangent at \(\hat{x}_{0} = \log x_{0} /\log 10\), \(\hat{y}_{0} = \log f(x_{0} )/\log 10\), reads \(\hat{y} = \alpha \hat{x} + a\), with slope \(\alpha = x_{0} f^{\prime}(x_{0} )/f(x_{0} )\) and vertical intercept \(a = (\log f(x_{0} ) - \alpha \log x_{0} )/\log 10\). This Log–Log tangent corresponds to the power law \(y = 10^{a} x^{\alpha }\) in linear coordinates, cf. (A.2). As for Log–Log inflection points, we note \(\hat{y}^{\prime\prime}(\hat{x})/\log 10 = (xf^{\prime}(x)/f(x))^{\prime}x\); if \(\hat{x}_{0}\) is an inflection point of a Log–Log plot, the derivative of \(xf^{\prime}(x)/f(x)\) vanishes at \(x_{0} = 10^{{\hat{x}_{0} }}\). Local minima and maxima of \(xf^{\prime}(x)/f(x)\) thus indicate inflection points of \(\hat{y}(\hat{x})\) in (A.1). In general, a Log–Log inflection point does not have a counterpart in linear representation.

Example 1

\(y = Ax^{{\alpha + \beta \log x + \gamma \log^{2} x + \cdots }}\), where the ellipsis in the exponent indicates possible higher powers of \(\log x\), so that \(y^{\prime} = (\alpha + 2\beta \log x + 3\gamma \log^{2} x + \cdots )y/x\). If \(\gamma = 0\), \(\beta < 0\), this is a lognormal distribution. In Log–Log coordinates, we obtain a polynomial,

$$ \hat{y} = \frac{\log A}{{\log 10}} + \alpha \hat{x} + \beta \hat{x}^{2} \log 10 + \gamma \hat{x}^{3} \log^{2} 10 + \cdots . $$
(A.3)

Lognormals are second-order polynomials in Log–Log representation, and a power law (\(\beta = \gamma = 0\)) is Log–Log linear. Thus, the regression of a lognormal distribution just amounts to fitting a second-order polynomial to a Log–Log plot of the data set, and a power law implies linear regression.

Example 2

\(y = Ax^{\alpha } \exp ( - (x/d)^{\delta } )\), with \(A,d,\delta > 0\), so that \(y^{\prime} = (\alpha - \delta (x/d)^{\delta } )y/x\). In Log–Log coordinates, this Weibull distribution reads

$$ \hat{y} = \frac{\log A}{{\log 10}} + \alpha \hat{x} - \frac{{{\text{e}}^{{\delta \hat{x}\log 10}} }}{{d^{\delta } \log 10}}, $$
(A.4)

the power-law factor \(Ax^{\alpha }\) becomes a linear function, and the cutoff factor emerges as subtracted exponential in this representation.

To define Index functionals, we start with a function \(y = f(x)\), \(x > 0\), which reads, in Lin-Log coordinates, \(y = f(10^{{\hat{x}}} )\), the logarithmic coordinate being denoted by a hat, \(\hat{x} = {\text{Log(}}x)\). The function \(y = I_{f} (x) = xf^{\prime}(x)/f(x)\) is the linear representation of the Log–Log slope (Index) of \(f(x)\), see after (A.2). The Lin-Log representation of the Index \(I_{f} (x)\) is \(y = I_{f} (10^{{\hat{x}}} ) = 10^{{\hat{x}}} f^{\prime}(10^{{\hat{x}}} )/f(10^{{\hat{x}}} )\).

For instance, a lognormal \(f(x) = Ax^{\alpha + \beta \log x}\), \(\beta < 0\), has the Index function \(y = I_{f} (10^{{\hat{x}}} ) = \alpha + 2\beta \hat{x}\log 10\). Thus, the Index of a lognormal is linear in Lin-Log representation, and the Index of a power law is constant. Conversely, if the Index \(I_{f} (x)\) is constant or linear (with negative slope) in Lin-Log representation \(y = I_{f} (10^{{\hat{x}}} )\), then \(f(x)\) is a power law or lognormal, cf. Example 1 above, where \(y = I_{f} (x) = \alpha + 2\beta \log x + \cdots\). A Weibull distribution, \(f(x) = Ax^{\alpha } \exp ( - (x/d)^{\delta } )\), has the Index function \(y = I_{f} (x) = \alpha - \delta (x/d)^{\delta } = \alpha - \delta {\text{e}}^{{\delta \hat{x}\log 10}} /d^{\delta }\), cf. Example 2 above.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Tomaschitz, R. Modeling electrical resistivity and particle fluxes with multiply broken power-law distributions. Eur. Phys. J. Plus 136, 629 (2021). https://doi.org/10.1140/epjp/s13360-021-01542-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1140/epjp/s13360-021-01542-5

Navigation