Skip to main content
Log in

About the relations between finite strain in non-cubic crystals and the related phenomenological P-V Equation of State

  • Original papers
  • Published:
Physics and Chemistry of Minerals Aims and scope Submit manuscript

Abstract

The notion of scalar strain in minerals is crucial for the formulation of P-V equations of state (EoS). A scalar strain, μ, holding for any crystal symmetry has been derived by a rigorous and general approach, and then used to develop the related phenomenological P-V EoS. μ, which depends on V and the trace of the G* G0 matrix, can be split into two components, M and Φ, where the former takes values close to those of the scalar strain according to Birch. M, providing the main contribution (often larger than 80%) to μ, is appropriate for the formulation of an EoS as ∂M/∂V behaves regularly in the limit of an unstrained configuration. The phenomenological EoS based on M shows the same dependence on the elastic parameters (bulk modulus and derivatives versus pressure) of the usual Birch-Murnaghan EoS, and yields comparable results. Slight deviations occur for low symmetry minerals. This work is meant to contribute (1) to shed light on the relationships between scalar strain and related P-V EoS‘s, and (2) to provide a most general EoS which includes, as a particular case, the Birch-Murnaghan model and explains why this latter is reliable for crystal symmetry other than the cubic one, for which it was originally derived.

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

Similar content being viewed by others

References

  • Aizu K (1970) Determination of the state parameters and formulation of spontaneous strain for ferroelastics. J Phys Soc Japan 28:706–716

    CAS  Google Scholar 

  • Angel R (2000) Equations of state. Rev Mineral Geochem 41:35–59

    Google Scholar 

  • Birch F (1947) Finite elastic strain in cubic crystals. Phys Rev 71:809–824

    Article  CAS  Google Scholar 

  • Birch F (1986) Equation of state and thermodynamic parameters of NaCl to 300 kbar in the high-temperature domain. J Geophys Res 91:4949–4954

    Google Scholar 

  • Catti M (1985) Calculation of elastic constants by the method of crystal static deformation. Acta Cryst A41:494–500

    Article  Google Scholar 

  • Davies GF (1973) Invariant finite strain measures in elasticity and lattice dynamics. J Phys Solids 34:841–845

    CAS  Google Scholar 

  • Pavese A (1998) Thermoelastic and structural properties of forsterite as a function of P and T: a computer simulation study, by semi-classical potentials and quasi-harmonic approximation. Phys Chem Min 26:44–54

    Article  CAS  Google Scholar 

  • Pavese A (2002) Pressure volume temperature equations of state: a comparative study based on numerical simulations. Phys Chem Min 29:43–51

    Article  CAS  Google Scholar 

  • Pavese A, Diella V, Levy D, Hanfland M (2001) Synchrotron X-ray powder diffraction study of natural P2/n omphacites at high pressure conditions. Phys Chem Min 28:9–16

    Article  CAS  Google Scholar 

  • Pavese A, Levy D, Curetti N, Diella V, Fumagalli P, Sani A (2003) Equation of state and compressibility of pholgopite by in situ high pressure X-ray powder diffraction experiments. Eur J Miner 15:455–463

    Google Scholar 

  • Poirier JP and Tarantola A (1998) A logarithmic equation of state. Phys Earth Planet Int 109:1–8

    Article  Google Scholar 

  • Stacey FD, Brennan BJ, Irvine RD (1981) Finite strain theories and comparisons with seismological data. Geophys Surv 4:189–232

    Google Scholar 

  • Wallace DC (1972) Thermodynamics of crystals. Wiley, New York

    Google Scholar 

  • Weaver JS (1976) Application of finite strain theory to non-cubic crystals. J Phys Chem Solids 37:711–718

    Article  CAS  Google Scholar 

  • Zhang L, Ahsbahs H, Hafner S, Kutoglu A (1997) Single-crystal compression and crystal structure of clinopyroxene up to 10 GPa. Am Miner 82:245–258

    Google Scholar 

Download references

Acknowledgements

The author is grateful to Mauro Gemmi for critical reading of the manuscript before submission. Mauro Prencipe and Ross Angel, who acted as referees for Physics and Chemistry of Minerals, are acknowledged for their suggestions and corrections which enhanced the quality of the manuscript.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Alessandro Pavese.

Appendices

Appendix 1

Let a j be the jth - basis lattice vector of a crystal; then, the metric tensor is defined by G ik = a i · a k . It is well known that det(G)= V2 and G−1=G*, where G* is the metric tensor of the reciprocal lattice and V the volume of the unit-cell fixed by a1, a2 and a3.

We prove, now, Eq. 6. Matrix R, which transforms ortho-normal co-ordinates into crystallographic co-ordinates, relates a j vectors to a chosen set of orthogonal versors t j as follows:

$$ R_{jk}^{\rm T} {\varvec{a}}_{k}= {\varvec{t}}_{j}$$

and therefore

$$ {\varvec{a}}_{j}= R_{jk}^{{\rm T}-1} {\varvec{t}}_{k}.$$

Taking into account that t j · t i ji , it descends that

$$ \begin{aligned} G_{ik} & = {\varvec{a}}_i \cdot {\varvec{a}}_k \cdot = R_{il}^{T - 1} {\varvec{t}}_l \cdot R_{km}^{T - 1} {\varvec{t}}_m = R_{il}^{T - 1} R_{km}^{T - 1} \delta _{ml} = R_{il}^{T - 1} R_{kl}^{T - 1} = R_{il}^{T - 1} R_{lk}^{ - 1} \\ & = ({\varvec{R}}^{T - 1} {\varvec{R}}^{ - 1} )_{ik} = ({\varvec{R}}^{ - 1T} {\varvec{R}}^{ - 1} )_{ik} . \\ \end{aligned} $$

Appendix 2

In the following we report the Lagrangian counterpart of the equations developed through the text for the Eulerian scheme. For Eq. 10:

$${\text{Tr}}({\mathbf{I}} + 2 {\varvec{\eta}} ) = {\text{ Tr}}({\varvec{G}}*^0{\varvec{G}});$$

for Eq. 11

$$\det ({\mathbf{I}} + 2 {\varvec{\eta}} ) = \det ({\varvec{G}}*^0{\varvec{G}}).$$

Replacing 1 + 2μ L with B, and Tr(G*0 G) with λ L , one has: for Eq. 12

$$3B + 2{\text{ Tr}}({\varvec{\Omega}}_{\text{L}}) = \lambda_{\text{L}};$$

and for Eq. 13

$$ B^3 + 2B^2 Tr({\varvec{\Omega}}_{\text{L}} ) + {\text{o}}({\varvec{\Omega}}_{\text{L}}^2 ) + {\text{o}}({\varvec{\Omega}}_{\text{L}}^2 ) = (V/V_0 )^2 . $$

For Eq. 23

$$ \mu _{\text{L}} = \frac{1} {2}\left( {1 - \frac{{\lambda _{\text{L}} }} {3}} \right) $$

Appendix 3

We aim to calculate the derivatives of μ versus V, assuming

$$ \mu = \frac{1} {2}\left( {\frac{\lambda } {3} - 1} \right) $$

One has ∂μ/∂ V=1/6 (∂λ/∂ V). In the case of the Eulerian strain, which is the one treated in the present work, λ=Σ i,j=1,3 G* ij G0 ji , where G ij and G* ij are the ij-terms of the metric matrices of direct and reciprocal lattice, respectively. We have to analyse ∂ G* ji /∂ V. For the sake of clarity, let us focus on the case i=1 and j=2. Then

$$\frac{\partial G^{*}_{12}}{ \partial V}=\frac{\partial G^{*}_{12}}{\partial V^{*}}\frac{\partial V^{*}}{ \partial V}.$$
(26)

Taking into account that G* 12= a* 1 a* 2 cosγ*, it follows

$$\left(\frac{\partial G^{*}_{12}}{\partial a^{*}_{1}}\frac{\partial a^{*}_{1}}{\partial V^{*}}+ \frac{\partial G^{*}_{12}}{\partial a^{*}_{2}}\frac{\partial a^{*}_{2}}{\partial V^{*} }+\frac{\partial G^{*}_{12}}{\partial \gamma^{*}}\frac{\partial \gamma^{*}}{\partial V^{*}}\right)\left(\frac{\partial V^{*}}{\partial V}\right).$$
(27)

One has now to keep in mind that, in the present case, the derivatives of the lattice parameters with respect to V* [i.e. (∂a* 1/∂V*), (∂a* 2/∂V*) and (∂γ*/∂ V*)] must be calculated assuming infinitesimal changes of volume produced by pressure. Hence, for a* 1, one writes

$$\frac{\partial a^{*}_{1}}{\partial V^{*}}=\frac{\partial a^{*}_{1}}{\partial P}\frac{\partial P}{\partial V^{*}}$$

which can be trivially re-cast into

$$ \frac{{a_{\text{1}}^{\text{*}} }} {{V^{\text{*}} }}\frac{{\beta _{a_1^{\text{*}} } }} {{\beta _{V^{\text{*}} } }} $$
(28)

where β X is the compressibility of X. Expressions similar to Eq. 28 hold for a* 2 and γ* . This all implies that Eq. 27 can be formulated as follows

$$ \left[ {\frac{{G_{12}^* }} {{V^* }}\frac{{\beta _{a_1^* } }} {{\beta _{V^* } }} + \frac{{G_{12}^* }} {{V^* }}\frac{{\beta _{a_2^* } }} {{\beta _{V^* } }} + \frac{{\partial G_{12}^* }} {{\partial \gamma ^* }}\frac{{\partial \gamma ^* }} {{\partial P}}\frac{{\partial P}} {{\partial V^* }}} \right]\left( { - \frac{1} {{V^2 }}} \right). $$

(∂γ*/∂ P) is in most cases negligible with respect to its counterparts in terms of axial compression, and therefore one can drop off the contribution depending on γ*. \( {{\beta _{a_i^* } } \mathord{\left/ {\vphantom {{\beta _{a_i^* } } {\beta _{V^* } }}} \right. \kern-\nulldelimiterspace} {\beta _{V^* } }} \) involves the axial compressibility of a* i , and represents a serious hindrance for the development of a P-V EoS. The simplest way to overcame such a difficulty is to drastically approximate \( {{\beta _{a_i^* } } \mathord{\left/ {\vphantom {{\beta _{a_i^* } } {\beta _{V^* } }}} \right. \kern-\nulldelimiterspace} {\beta _{V^* } }} \) with 1/3, which holds rigorously in the case of cubic symmetry only. However, this is not an unrealistic approximation for non-cubic symmetry, too, save the case of structures with a highly anisotropic behaviour under compression. In so doing one obtains

$$\frac{\partial G^{*}_{12}}{\partial V}\approx-\frac{2}{3} \frac{G^{*}_{12}}{V}$$

and therefore

$$\frac{\partial \lambda}{\partial V} \approx -\frac{2}{3}\frac{\lambda}{V}$$

leading finally to

$$\frac{\partial \mu}{\partial V} =-\frac{1}{9}\frac{\lambda}{V} =-\frac{1}{3} \frac{1+2 \mu}{V}.$$

Second and third derivatives are straightforwardly determined by ∂μ /∂ V.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Pavese, A. About the relations between finite strain in non-cubic crystals and the related phenomenological P-V Equation of State. Phys Chem Minerals 32, 269–276 (2005). https://doi.org/10.1007/s00269-005-0465-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00269-005-0465-8

Keywords

Navigation