Skip to main content
Log in

On the Problematic Case of Product Approximation in Backus Average

  • Published:
Journal of Elasticity Aims and scope Submit manuscript

Abstract

Elastic anisotropy might be a combined effect of the intrinsic anisotropy and the anisotropy induced by thin-layering. The Backus average, a useful mathematical tool, allows us to describe such an effect quantitatively. The results are meaningful only if the underlying physical assumptions are obeyed, such as static equilibrium of the material. We focus on the only mathematical assumption of the Backus average, namely, product approximation. It states that the average of the product of a varying function with a nearly constant function is approximately equal to the product of the averages of those functions. We analyse particular problematic case for which the aforementioned assumption is inaccurate. Furthermore, we focus on the seismological context. We examine the inaccuracy’s effect on the wave propagation in a homogenous medium—obtained using the Backus average—equivalent to thin layers. Numerical simulations indicate clearly that the product approximation inaccuracy has negligible effect on wave propagation; irrespective of layers’ symmetries. To give the results a practical focus, we show that the problematic case of product approximation is strictly related to the negative Poisson’s ratio of constituents layers. We discuss the laboratory and well-log cases in which such a ratio has been noticed. Upon thorough literature review, it occurs that examples of so-called auxetic materials (media that have negative Poisson’s ratio) are not extremely rare exceptions as thought previously. The investigation and derivation of Poisson’s ratio for materials exhibiting symmetry classes up to monoclinic become a significant part of this paper. In addition to the main objectives, we also show that the averaging of cubic layers results in an equivalent medium with tetragonal (not cubic) symmetry. We present concise formulations of stability conditions for low symmetry classes, such as trigonal, orthotropic, and monoclinic.

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
Fig. 3
Fig. 4
Fig. 5

Similar content being viewed by others

Data Availability

The data that support the findings of this study are available from the author upon reasonable request.

References

  1. Backus, G.E.: Long-wave elastic anisotropy produced by horizontal layering. J. Geophys. Res. 67(11), 4427–4440 (1962)

    Article  ADS  MATH  Google Scholar 

  2. Bakulin, A., Grechka, V.: Effective anisotropy of layered media. Geophysics 68(6), 1817–1821 (2003)

    Article  ADS  Google Scholar 

  3. Baughman, R.H., Stafstrom, S., Cui, C., Dantas, S.O.: Materials with negative compressibilities in one or more dimensions. Science 279(5356), 1522–1524 (1998)

    Article  ADS  Google Scholar 

  4. Beljankin, D.S., Petrov, V.P.: Occurrence of cristobalite in a sedimentary rock. Am. Mineral. 23(3), 153–155 (1938)

    Google Scholar 

  5. Bos, L., Dalton, D.R., Slawinski, M.A., Stanoev, T.: On Backus average for generally anisotropic layers. J. Elast. 127(2), 179–196 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  6. Bos, L., Danek, T., Slawinski, M.A., Stanoev, T.: Statistical and numerical considerations of Backus-average product approximation. J. Elast. 132(1), 141–159 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  7. Calvert, S.E., Burns, R.G., Smith, J.V., Kempe, D.R.C.: Mineralogy of silica phases in deep-sea cherts and porcelanites [and disucssion]. Phil. Trans. R. Soc. Lond. Math. Phys. Sci. 286(1336), 239–252 (1977)

    Article  ADS  Google Scholar 

  8. Carcione, J.M., Kosloff, D., Behle, A.: Long-wave anisotropy in stratified media: a numerical test. Geophysics 56(2), 245–254 (1991)

    Article  ADS  Google Scholar 

  9. Castagna, J.P., Smith, S.W.: Comparison of AVO indicators: a modeling study. Geophysics 59(12), 1849–1855 (1994)

    Article  ADS  Google Scholar 

  10. Coyner, K.B.: Effects of stress, pore pressure, and pore fluids on bulk strain, velocity, and permeability in rocks. PhD thesis, Massachusetts Institute of Technology (1984)

  11. Damby, D.E., Llewellin, E.W., Horwell, C.J., Williamson, B.J., Najorka, J., Cressey, G., Carpenter, M.: The \(\alpha \)\(\beta \) phase transition in volcanic cristobalite. J. Appl. Crystallogr. 47(4), 1205–1215 (2014)

    Article  Google Scholar 

  12. Dvorkin, J., Moos, D., Packwood, J.L., Nur, A.M.: Identifying patchy saturation from well logs. Geophysics 64(6), 1756–1759 (1999)

    Article  ADS  Google Scholar 

  13. Dziewoński, A.M., Anderson, D.L.: Preliminary reference Earth model. Phys. Earth Planet. Inter. 25(4), 297–356 (1981)

    Article  ADS  Google Scholar 

  14. Emery, D.J., Stewart, R.R.: Using VP/VS to explore for sandstone reservoirs: well log and synthetic seismograms from the Jeanne d’Arc basin, offshore Newfoundland. CREWES Res. Rep. 18(1), 1–20 (2006)

    Google Scholar 

  15. Fay, M., Larter, S., Bennett, B., Snowdon, L., Fowler, M.: Petroleum Systems and Quantitative Oil Chemometric Models for Heavy Oils in Alberta and Saskatchewan to Characterize Oil-Source Correlations. GeoConvention 2012:Vision (2012)

    Google Scholar 

  16. Fomel, S., Sava, P., Vlad, I., Liu, Y., Bashkardin, V.: Madagascar: open-source software project for multidimensional data analysis and reproducible computational experiments. J. Open Res. Softw. 1(1), e8 (2013)

    Article  Google Scholar 

  17. Freund, D.: Ultrasonic compressional and shear velocities in dry clastic rocks as a function of porosity, clay content, and confining pressure. Geophys. J. Int. 108(1), 125–135 (1992)

    Article  ADS  Google Scholar 

  18. Goodway, B.: AVO and Lamé constants for rock parameterization and fluid detection. CSEG Rec. 26(6), 39–60 (2001)

    Google Scholar 

  19. Gregory, A.R.: Fluid saturation effect on dynamic elastic properties of sedimentary rocks. Geophysics 41(5), 895–921 (1976)

    Article  ADS  Google Scholar 

  20. Grima, J.N., Gatt, R., Zammit, V., Williams, J.J., Evans, K.E., Alderson, A., Walton, R.I.: Natrolite: a zeolite with negative Poisson’s ratios. J. Appl. Phys. 101(8), 086102 (2007)

    Article  ADS  Google Scholar 

  21. Grima, J.N., Jackson, R., Alderson, A., Evans, K.E.: Do zeolites have negative Poisson’s ratios? Adv. Mater. 12(24), 1912–1918 (2000)

    Article  Google Scholar 

  22. Han, D.-H.: Effects of porosity and clay content on acoustic properties of sandstones and unconsolidated sediments. PhD thesis, Stanford University (1986)

  23. Hay, R.L.: Geologic occurrence of zeolites and some associated minerals. Stud. Surf. Sci. Catal. 28(1), 35–40 (1986)

    Article  Google Scholar 

  24. Hayes, B.J.R., Christopher, J.R., Rosenthal, L., Los, G., McKercher, B., Minken, D., Tremblay, Y.M., Fennel, J.: Cretaceous Mannville Group of the Western Canada Sedimentary Basin; in Geological Atlas of the Western Canada Sedimentary Basin. Canadian Society of Petroleum Geologists and Alberta Research Council (1994). 19 (URL: https://ags.aer.ca/publications/chapter-19-Cretaceous-mannville-group.htm)

  25. Helbig, K.: Anisotropy and dispersion in periodically layered media. Geophysics 49(4), 364–373 (1984)

    Article  ADS  Google Scholar 

  26. Helbig, K.: Foundations of Anisotropy for Exploration Seismics, 1st edn. Pergamon, Elmsford (1994)

    Google Scholar 

  27. Hommand-Etienne, F., Houpert, R.: Thermally induced microcracking in granites: characterization and analysis. Int. J. Rock Mech. Min. Sci. Geomech. Abstr. 26(2), 125–134 (1989)

    Article  Google Scholar 

  28. Ji, S., Li, L., Motra, H.B., Wuttke, F., Sun, S., Michibayashi, K., Salisbury, M.H.: Poisson’s ratio and auxetic properties of natural rocks. J. Geophys. Res., Solid Earth 123(2), 1161–1185 (2018)

    Article  ADS  Google Scholar 

  29. Ji, S., Sun, S., Wang, Q., Marcotte, D.: Lamé parameters of common rocks in the Earth’s crust and upper mantle. J. Geophys. Res. Solid Earth 115(B6), 1–15 (2010)

    Article  Google Scholar 

  30. Ji, S., Wang, Q., Li, L.: Seismic velocities, Poisson’s ratios and potential auxetic behavior of volcanic rocks. Tectonophysics 766(1), 270–282 (2019)

    Article  ADS  Google Scholar 

  31. Jizba, D.: Mechanical and acoustical properties of sandstones and shales. PhD thesis, Stanford University (1991)

  32. Kudela, I., Stanoev, T.: On possible issues of Backus average (2018). arXiv:1804.01917 [physics.geo-ph]

  33. Lakes, R.S.: Negative-Poisson’s-ratio materials: axetic solids. Annu. Rev. Mater. Res. 47(1), 63–81 (2017)

    Article  Google Scholar 

  34. Liner, C., Fei, T.: The Backus number. Lead. Edge 26(4), 420–426 (2007)

    Article  Google Scholar 

  35. Mavko, G., Jizba, D.: The relation between seismic P- and S-wave velocity dispersion in saturated rocks. Geophysics 59(1), 87–92 (1994)

    Article  ADS  Google Scholar 

  36. Mavko, G., Mukerji, T., Dvorkin, J.: The Rock Physics Handbook, 2nd edn. Cambridge University Press, Cambridge (2009)

    Book  Google Scholar 

  37. McKnight, R.E.A., Moxon, T., Buckley, A., Taylor, P.A., Darling, T.W., Carpenter, M.A.: Grain size dependence of elastic anomalies accompanying the \(\alpha \)\(\beta \) phase transition in polycrystalline quartz. J. Phys. Condens. Matter 20(7), 075229 (2008)

    Article  ADS  Google Scholar 

  38. Mizota, C., Toh, N., Matsuhisa, Y.: Origin of cristobalite in soils derived from volcanic ash in temperate and tropical regions. Geoderma 39(4), 323–330 (1987)

    Article  ADS  Google Scholar 

  39. Mouchat, F., Coudert, F.X.: Necessary and sufficient elastic stability conditions in various crystal systems. Phys. Rev. B 90(22), 1–4 (2014)

    Google Scholar 

  40. Nur, A., Simmons, G.: The effect of saturation on velocity in low porosity rocks. Earth Planet. Sci. Lett. 7(2), 183–193 (1969)

    Article  ADS  Google Scholar 

  41. Sanchez-Valle, C., Lethbridge, Z.A.D., Sinogeikin, S.V., Williams, J.J., Walton, R.I., Evans, K.E., Bass, J.D.: Negative Poisson’s ratios in siliceous zeolite MFI-silicalite. J. Chem. Phys. 128(18), 184503 (2008)

    Article  ADS  Google Scholar 

  42. Schoenberg, M., Muir, F.: A calculus for finely layered anisotropic media. Geophysics 54(5), 581–589 (1989)

    Article  ADS  Google Scholar 

  43. Slawinski, M.A.: Waves and Rays in Elastic Continua, 3rd edn. World Scientific, Singapore (2015)

    Book  MATH  Google Scholar 

  44. Slawinski, M.A.: Waves and Rays in Seismology: Answers to Unasked Questions, 2nd edn. World Scientific, Singapore (2018)

    Book  MATH  Google Scholar 

  45. Yeganeh-Haeri, A., Weidner, D.J., Parise, J.B.: Elasticity of \(\alpha \)-cristobalite: a silicon dioxide with a negative Poisson’s ratio. Science 257(5070), 650–652 (1992)

    Article  ADS  Google Scholar 

  46. Zaitsev, V.Y., Radostin, A.V., Pasternak, E., Dyskin, A.: Extracting real-crack properties from non-linear elastic behaviour of rocks: abundance of cracks with dominating normal compliance and rocks with negative Poisson ratios. Nonlinear Process. Geophys. 24(3), 543–551 (2017)

    Article  ADS  Google Scholar 

Download references

Acknowledgements

We wish to thank Michael A. Slawinski and David Dalton for their suggestions. Also, we acknowledge Reviewers’ comments and the proofreading of David Dalton. The research was done in the context of The Geomechanics Project partially supported by the Natural Sciences and Engineering Research Council of Canada, grant 202259. The author has no conflict of interests to declare.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Filip P. Adamus.

Additional information

Publisher’s Note

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

Appendices

Appendix A: Backus Average for Anisotropic Layers

Let us write the strain-stress relations in two dimensions (\(x_{3}x_{1}\)–plane), namely,

$$ \sigma _{11}=C_{11}\varepsilon _{11}+C_{13}\varepsilon _{33}\,, $$
(61)
$$ \sigma _{33}=C_{13}\varepsilon _{11}+C_{33}\varepsilon _{33}\,, $$
(62)
$$ \sigma _{13}=2C_{55}\varepsilon _{13}\,, $$
(63)

which are the relations valid for the monoclinic, orthotropic, tetragonal, and TI symmetry class. Upon a rearrangement, we get

$$ \sigma _{11}=\left (C_{11}-\frac{C_{13}^{2}}{C_{33}}\right ) \varepsilon _{11}+\left (\frac{C_{13}}{C_{33}}\right )\sigma _{33}\,, $$
(64)
$$ \varepsilon _{33}=-\left (\frac{C_{13}}{C_{33}}\right )\varepsilon _{11} +\left (\frac{1}{C_{33}}\right )\sigma _{33}\,, $$
(65)
$$ \frac{\partial {u_{1}}}{\partial {x_{3}}}=\left (\frac{1}{C_{55}} \right )\sigma _{13}-\frac{\partial {u_{3}}}{\partial {x_{1}}}\,. $$
(66)

Let us treat the above equations as the stress-strain relations that correspond to many individual constituents that we want to average. To perform the averaging process, we use the three following properties: the average of the sum is a sum of the average, the average of the derivative is a derivative of the average, and, finally, the product approximation. We obtain

$$ \overline{\sigma _{11}}=\left [ \overline{\left (C_{11}-\frac{C_{13}^{2}}{C_{33}}\right )}+ \overline{\left (\frac{C_{13}}{C_{33}}\right )}^{2} \overline{\left (\frac{1}{C_{33}}\right )}^{-1}\right ] \overline{\varepsilon _{11}}+ \overline{\left (\frac{C_{13}}{C_{33}}\right )}\, \overline{\left (\frac{1}{C_{33}}\right )}^{-1} \overline{\varepsilon _{33}}\,, $$
(67)
$$ \overline{\sigma _{33}}= \overline{\left (\frac{C_{13}}{C_{33}}\right )}\, \overline{\left (\frac{1}{C_{33}}\right )}^{-1} \overline{\varepsilon _{11}}+ \overline{\left (\frac{1}{C_{33}}\right )}^{-1} \overline{\varepsilon _{33}}\,, $$
(68)
$$ \overline{\sigma _{13}}=\overline{\left (\frac{1}{C_{55}}\right )}^{-1}2 \,\overline{\varepsilon _{13}}\,. $$
(69)

Comparing equations (67)–(69) with equations (61)–(63), we see that the equivalent elasticity parameters are equal to

$$ C_{11}^{\mathrm{{\overline{\mathit{eq}}}}}= \overline{\left (C_{11}-\frac{C_{13}^{2}}{C_{33}}\right )}+ \overline{\left (\frac{C_{13}}{C_{33}}\right )}^{2} \overline{\left (\frac{1}{C_{33}}\right )}^{-1}\,, $$
(70)
$$ C_{13}^{\mathrm{{\overline{eq}}}}= \overline{\left (\frac{C_{13}}{C_{33}}\right )}\, \overline{\left (\frac{1}{C_{33}}\right )}^{-1} \,, $$
(71)
$$ C_{33}^{\mathrm{{\overline{\mathit{eq}}}}}= \overline{\left (\frac{1}{C_{33}}\right )}^{-1}\,, $$
(72)
$$ C_{55}^{\mathrm{{\overline{\mathit{eq}}}}}= \overline{\left (\frac{1}{C_{55}}\right )}^{-1}\,, $$
(73)

and the resulting medium is either monoclinic, orthotropic, tetragonal, or TI. If layers have cubic symmetry, then \(C_{33}=C_{11}\). In such a case, \(C_{33}^{\mathrm{{\overline{\mathit{eq}}}}}\neq C_{11}^{\mathrm{{\overline{\mathit{eq}}}}}\), which means that the equivalent medium is not cubic. To understand what is the symmetry class of the medium equivalent to cubic layers, we need to derive the analogous equivalent parameters, but for a 3D case. Upon an analogous procedure, shown above, we get

$$ C_{11}^{\mathrm{{\overline{\mathit{eq}}}}}= \overline{\left (C_{11}-\frac{C_{13}^{2}}{C_{11}}\right )}+ \overline{\left (\frac{C_{13}}{C_{11}}\right )}^{2} \overline{\left (\frac{1}{C_{11}}\right )}^{-1}\,, $$
(74)
$$ C_{12}^{\mathrm{{\overline{\mathit{eq}}}}}= \overline{\left (C_{13}-\frac{C_{13}^{2}}{C_{11}}\right )}+ \overline{\left (\frac{C_{13}}{C_{11}}\right )}^{2} \overline{\left (\frac{1}{C_{11}}\right )}^{-1}\,, $$
(75)
$$ C_{13}^{\mathrm{{\overline{\mathit{eq}}}}}= \overline{\left (\frac{C_{13}}{C_{11}}\right )}\, \overline{\left (\frac{1}{C_{11}}\right )}^{-1} \,, $$
(76)
$$ C_{33}^{\mathrm{{\overline{eq}}}}= \overline{\left (\frac{1}{C_{11}}\right )}^{-1}\,, $$
(77)
$$ C_{55}^{\mathrm{{\overline{\mathit{eq}}}}}= \overline{\left (\frac{1}{C_{55}}\right )}^{-1}\,, $$
(78)
$$ C_{66}^{\mathrm{{\overline{\mathit{eq}}}}}=\overline{C_{55}}\,, $$
(79)

where \(C_{11}^{\mathrm{{\overline{\mathit{eq}}}}}=C_{22}^{\mathrm{{\overline{\mathit{eq}}}}}\), \(C_{13}^{\mathrm{{\overline{\mathit{eq}}}}}=C_{23}^{\mathrm{{\overline{\mathit{eq}}}}}\), and \(C_{55}^{\mathrm{{\overline{\mathit{eq}}}}}=C_{44}^{\mathrm{{\overline{\mathit{eq}}}}}\). The equivalent medium has six independent elasticity parameters and exhibits the tetragonal symmetry class.

Appendix B: Backus Procedure for a Trigonal Tensor

First, we write the stress-strain relations in a trigonal medium (expressed in a natural coordinate system) as

$$ \sigma _{11}=C_{11}\varepsilon _{11}+C_{12}\varepsilon _{22}+C_{13} \varepsilon _{33}+C_{15}\frac{\partial u_{1}}{\partial x_{3}}+C_{15} \frac{\partial u_{3}}{\partial x_{1}}\,, $$
(80)
$$ \sigma _{22}=C_{12}\varepsilon _{11}+C_{11}\varepsilon _{22}+C_{13} \varepsilon _{33}-C_{15}\frac{\partial u_{1}}{\partial x_{3}}-C_{15} \frac{\partial u_{3}}{\partial x_{1}}\,, $$
(81)
$$ \sigma _{33}=C_{13}\varepsilon _{11}+C_{13}\varepsilon _{22}+C_{33} \varepsilon _{33}\,, $$
(82)
$$ \sigma _{23}=C_{44}\frac{\partial {u_{2}}}{\partial {x_{3}}}+C_{44} \frac{\partial {u_{3}}}{\partial {x_{2}}}-2C_{15}\varepsilon _{12}\,, $$
(83)
$$ \sigma _{13}=C_{44}\frac{\partial {u_{1}}}{\partial {x_{3}}}+C_{44} \frac{\partial {u_{3}}}{\partial {x_{1}}}+C_{15}\varepsilon _{11}-C_{15} \varepsilon _{22}\,, $$
(84)
$$ \sigma _{12}=(C_{11}-C_{12})\varepsilon _{12}-C_{15} \frac{\partial {u_{2}}}{\partial {x_{3}}}-C_{15} \frac{\partial {u_{3}}}{\partial {x_{2}}}\,. $$
(85)

We can directly rewrite equations (82)–(84) in a manner that the nearly-constant stresses and strains are on the right-hand side, whereas the sole varying function of displacements is on the left-hand side. We get,

$$ \varepsilon _{33}=\sigma _{33} \underbrace{ \left (\frac{1}{C_{33}}\right ) }_{\text{$g_{1}$}}- \underbrace{\left (\frac{C_{13}}{C_{33}}\right )}_{\text{$g_{2}$}} \varepsilon _{11}-\underbrace{\left (\frac{C_{13}}{C_{33}}\right )}_{\text{$g_{3}$}}\varepsilon _{22}\,, $$
(86)
$$ \frac{\partial {u_{2}}}{\partial {x_{3}}}=\sigma _{23} \underbrace{\left (\frac{1}{C_{44}}\right )}_{\text{$g_{4}$}}- \frac{\partial {u_{3}}}{\partial {x_{2}}}- \underbrace{\left (\frac{C_{15}}{C_{44}}\right )}_{\text{$g_{t}$}}2 \varepsilon _{12}\,, $$
(87)
$$ \frac{\partial {u_{1}}}{\partial {x_{3}}}=\sigma _{13} \underbrace{\left (\frac{1}{C_{44}}\right )}_{\text{$g_{5}$}}- \frac{\partial {u_{3}}}{\partial {x_{1}}}-\left (\frac{C_{15}}{C_{44}} \right )\varepsilon _{11}+\left (\frac{C_{15}}{C_{44}}\right ) \varepsilon _{22}\,. $$
(88)

Now, we insert the right-hand side of equation (86) and (88) into equations (80) and (81). Also, we insert the right-hand side of (87) into (85). Upon simple calculations, we obtain

$$ \sigma _{11}= \sigma _{33}\left (\frac{C_{13}}{C_{33}}\right ) + \sigma _{13}\left (\frac{C_{15}}{C_{44}}\right ) + \underbrace{\left (C_{11}-\frac{C_{13}^{2}}{C_{33}}-\frac{C_{15}^{2}}{C_{44}}\right )}_{\text{$g_{6}$}}\varepsilon _{11} + \underbrace{\left (C_{12}-\frac{C_{13}^{2}}{C_{33}}+\frac{C_{15}^{2}}{C_{44}}\right )}_{\text{$g_{7}$}}\varepsilon _{22} \,, $$
(89)
$$ \sigma _{22}= \sigma _{33}\left (\frac{C_{13}}{C_{33}}\right ) - \sigma _{13}\left (\frac{C_{15}}{C_{44}}\right ) + \left (C_{12}- \frac{C_{13}^{2}}{C_{33}}+\frac{C_{15}^{2}}{C_{44}}\right ) \varepsilon _{11} + \underbrace{\left (C_{11}-\frac{C_{13}^{2}}{C_{33}}-\frac{C_{15}^{2}}{C_{44}}\right )}_{\text{$g_{8}$}}\varepsilon _{22} \,, $$
(90)
$$ \sigma _{12}= -\sigma _{23}\left (\frac{C_{15}}{C_{44}}\right ) - \underbrace{\left (\frac{C_{11}-C_{12}}{2}-\frac{C_{15}^{2}}{C_{44}}\right )}_{\text{$g_{9}$}}2\varepsilon _{12} \,. $$
(91)

Terms in parentheses in equations (86)–(91) correspond to various \(g\); we denote them as \(g_{i}\) or \(g_{t}\). We notice that in case of trigonal symmetry, \(g_{2}=g_{3}\), \(g_{4}=g_{5}\), and \(g_{6}=g_{8}\).

Appendix C: Proof of Lemma 1

Proof

Let us consider first part of Lemma 1, which states that

if \(g_{2}^{\normalfont {\mathrm{ort}}}<0\) and \(g_{3}^{\normalfont {\mathrm{ort}}}<0\) then \(n_{1}\) and \(n_{2}\) cannot be both positive.

To prove it, let us assume that \(g_{2}^{\mathrm{ort}}<0\) and \(g_{3}^{\mathrm{ort}}<0\). Since, according to stability conditions, \(C_{33}\geq 0\), the above assumption is tantamount to \(C_{13}<0\) and \(C_{23}<0\). Also, assume that \(n_{1}>0\) and \(n_{2}>0\), which is tantamount to

$$ C_{13}C_{22}-C_{12}C_{23}>0 $$
(92)

and

$$ C_{23}C_{11}>C_{12}C_{13}\,, $$
(93)

respectively. From stability conditions, we also know that \(C_{22}\geq 0\). Hence, to satisfy expression (92), \(C_{12}\) must be positive. Therefore, we can rewrite inequality (92) as

$$ \frac{C_{13}C_{22}}{C_{12}}>C_{23}\,. $$
(94)

Now, let us consider inequality (93). Upon inserting some larger value in the place of \(C_{23}\) the inequality will remain true. However, if we insert there the left-hand side of inequality (94), we obtain

$$ C_{11}C_{22}< C_{12}^{2}\,, $$
(95)

which is mathematically correct, but not allowed by the stability condition.

Second part of Lemma 1 states that

if \(g_{2}^{\normalfont {\mathrm{ort}}}<0\) then \(n_{1}>0\) together with \(n_{2}<0\) are not allowed.

To prove it, let us assume that \(g_{2}^{\mathrm{ort}}<0\), which is tantamount to \(C_{13}<0\). Also, assume that \(n_{1}>0\) and \(n_{2}<0\). We get,

$$ \frac{C_{12}C_{23}}{C_{22}}< C_{13} $$
(96)

and

$$ \frac{C_{12}C_{13}}{C_{11}}>C_{23}\,. $$
(97)

Herein, we have invoked the so-called strict stability conditions (the matrix representing the elasticity tensor must be positive definite instead of positive semidefinite), which constitute the fact that \(C_{11}\) and \(C_{22}\) are greater than zero. We insert the left-hand side of inequality (97) in place of \(C_{23}\) inside inequality (96) to again obtain

$$ C_{11}C_{22}< C_{12}^{2}\,. $$
(98)

To prove third part of Lemma 1 stating that

if \(g_{3}^{\normalfont {\mathrm{ort}}}<0\) then \(n_{1}<0\) together with \(n_{2}>0\) are not allowed

we assume \(g_{3}^{\mathrm{ort}}<0\) and consider \(n_{1}<0\) and \(n_{2}>0\). We obtain

$$ \frac{C_{12}C_{23}}{C_{22}}>C_{13} $$
(99)

and

$$ \frac{C_{12}C_{13}}{C_{11}}< C_{23}\,. $$
(100)

If we insert the left-hand side of inequality (99) in place of \(C_{13}\) from inequality (100), which is a mathematically justified operation, we again obtain expression not allowed by stability conditions.

The last part stating that

if either \(g_{2}^{\normalfont {\mathrm{ort}}}>0\) or \(g_{3}^{\normalfont {\mathrm{ort}}}>0\) then \(n_{1}\) and \(n_{2}\) cannot be both negative,

can be proved trivially using same strategy as for the previous parts of Lemma 1. □

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Adamus, F.P. On the Problematic Case of Product Approximation in Backus Average. J Elast 144, 55–80 (2021). https://doi.org/10.1007/s10659-021-09826-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10659-021-09826-8

Mathematics Subject Classification

Keywords

Navigation