Skip to main content
Erschienen in: Archive of Applied Mechanics 3/2018

Open Access 08.11.2017 | Original

Elastic stress–strain analysis of an infinite cylindrical inclusion with eigenstrain

verfasst von: F. D. Fischer, G. A. Zickler, J. Svoboda

Erschienen in: Archive of Applied Mechanics | Ausgabe 3/2018

Aktivieren Sie unsere intelligente Suche, um passende Fachinhalte oder Patente zu finden.

search-config
loading …

Abstract

The ensemble of interstitial atoms as C attracted to a dislocation is well established as “Cottrell cloud” phenomenon. The deposition of the interstitial atoms in octahedral or tetrahedral positions in a bcc lattice may yield a remarkable internal stress state according to their anisotropic misfit eigenstrains. The stress fields of the atoms may then lead to a significant change in the stress field, e.g., around a dislocation. In such a case, the interstitial atoms are situated near the dislocation core in cylindrical volume elements along the dislocation line. As the occupancy of the sites in each cylinder by interstitial atoms can be considered as constant, also the eigenstrain state tensor is constant in the cylinder. A complete set of analytical expressions for the eigenstress state inside and outside of the cylinder is presented. The resultant stress field is then given by superposition, which allows also the determination of the interaction energy between the Cottrell cloud and the dislocation.

1 Introduction

The problem of interaction between dislocations and interstitial atoms as carbon or hydrogen is of increasing interest with respect to a detailed understanding of Cottrell clouds [1], and their kinetics [2]. To calculate the main physical quantity, i.e., the interaction energy between the stress field of a distinct dislocation and of the interstitial atoms, one can use the superposition of the elastic fields generated by the dislocation and by eigenstrains due to deposition of the atoms in interstitial positions. Here the pioneering work by Cochardt et al. [3] can be mentioned, who explained in detail the interaction of a dislocation with a single interstitial atom in the bcc lattice; see also the later Krempasky et al. [4]. There exist three types of anisotropic octahedral positions for interstitial atoms being shorter in one of the three main crystallographic \(\langle 100 \rangle \) directions in the bcc lattice. Thus, deposition of an atom in one of the octahedral position leads to a remarkably anisotropic eigenstrain in the region of the deposited atom, which can be considered as a spherical inclusion of the volume \(\omega \) of a substitutional atom with a given eigenstrain tensor \({\varvec{\upvarepsilon }}_i^\omega \) depending on the type i, \(i=1,2,3\), of the octahedral position. This problem of interaction was dealt with also by Friedel [5], Sect. 15.4., in recent papers by Cahn [6, 7], Mishin and Cahn [8] and Cai et al. [9]. In most studies, the anisotropic eigenstrain is replaced by a volume misfit, which is approved for interstitial atoms in fcc lattice or for substitutional atoms in any cubic lattice. In such a case, a rather simple analytical solution for the stress field inside and outside of the inclusion exists. Experimental work, see Wilde et al. [10], and atomistic simulations, Veiga et al. [11], Waseda et al. [12], has shown that an arrangement of interstitial atoms may exist leading to a significant overall stress field stemming from eigenstrains of atoms deposited in the three types of octahedral interstitial positions in the bcc lattice. Since the interaction energy of atoms with the stress field of a distinct dislocation depends significantly on the type of the octahedral interstitial positions, also the local site fraction of atoms in various types of octahedral interstitial positions is significantly different in thermodynamic equilibrium. As a result, the eigenstrains by deposition of interstitial atoms can significantly change the deformation field generated by the dislocation.
The treatment of the interaction with a cloud of interstitial atoms can be, e.g., performed for an edge dislocation, embedded in an infinite elastic lattice, with the dislocation line coinciding with the z-axis. Then the cloud of interstitial atoms around the dislocation is independent of z coordinate. The distribution of the interstitial atoms can then be described by a set of cylinders of infinite length along the z-axis having small cross sections and possessing in their whole volume constant site fractions of atoms in all three types of interstitial positions. Then the eigenstrain in each cylinder can be calculated and considered as given.
The goal of this paper is to check existing analytical relations for the stress state due to anisotropic eigenstrain state in an infinite cylinder embedded in infinite lattice and to provide a fully analytical solution for the stress field inside and outside of the cylinder. These results will provide the base for a thermodynamic kinetic model treating interaction of interstitial atoms with a moving dislocation.

2 Problem description

According to the recent atomistic simulations [11, 12], we assume a cylindrical inclusion as a representative volume element with the radius R in the xy-plane normal to the z-axis, which may coincide with the dislocation line. We consider a homogeneous eigenstrain state with the in-plane components \({\varepsilon _x^{\text {eig}}}\), \({\varepsilon _y^{\text {eig}}}\), \({\varepsilon _{\textit{xy}}^{\text {eig}}}\) and the component \({\varepsilon _z^{\text {eig}}}\) in the z-direction acting in the inclusion and no eigenstrain in the surrounding matrix. For the sake of completeness, it should be mentioned that the role of eigenstrains, \({\varepsilon _{\textit{xz}}^{\text {eig}}}\), \({\varepsilon _{\textit{yz}}^{\text {eig}}}\) can be dealt with extra by a simple use of Hooke’s law. All the eigenstrain components may have different values.
The cylindrical inclusion is restricted with respect to its expansion in z-direction by the surrounding material. Therefore, we assume, as simplest representation of the constraint in z-direction, a plane strain state with zero total strain in z-direction \(({\varepsilon _z \equiv 0})\). A recent atomistic study [8] has shown that the assumption of isotropic material is acceptable for cubic crystals. The isotropic material constants as Young’s modulus E and Poisson’s ratio \(\nu \) are assumed to be spatially constant in both the inclusion and the surrounding matrix; see Fig. 1.
From pioneering works by Eshelby [13, 14], it follows that a homogeneous eigenstrain in an ellipsoidal inclusion induces a homogeneous stress field in the inclusion. If one approximates the cylinder by a needle-like spheroid with R for axes a and b and an infinite length of axis c, one may utilize the solutions with normal and shear eigenstrains in Mura’s book [15], Chpt. 2 there, or the results worked out in Fischer and Böhm [16]. However, it must be kept in mind that the interstitials interact due to their exterior stress field with the adjacent interstitials and the dislocations. Therefore, the exterior stress field is of immediate relevance. The according mathematical framework in context with the Eshelby concept is rather extensive and somewhat demanding; see also Chpt. 2 in Mura [15] and the treatment by Li et al. [17, 18] for inclusions embedded in a finite domain. Our intention to provide easy-to-handle relations for the stress field of a cloud forced us to look for a straightforward solution of our particular problem. Since we have also to consider a shear eigenstrain in the cylinder, the analytical concept by Markenscoff and Dundurs [19], dealing also with a shear eigenstrain in an annulus and the corresponding comment by Shodja and Korshidi [20] concerning the disappearance of the singularities of stresses, motivated us to follow an exact analytical solution.
Since we assume an infinite domain, we expect rather easy-to-handle equations for the stress and deformation state utilizing the Airy function in polar coordinates. Here we refer to the pioneering work by Michell [21] who provided a set of solutions of the biharmonic equation in polar coordinates already in 1899; see also the book by Mal and Singh [22], Sect. 7.5, the book by Barber [23], Sect. 8.4, and the book by Bower [24], Sect. 5.2.3. Such an approach is in accordance with that applied in classical works on circular holes in disks (plates), e.g., by Kirsch [25], more than a century ago, by Bickley [26] and Sen [27].

3 Theory and solution concept

We introduce cylindrical coordinates \(r,\vartheta ,z\) and the according displacements uvw. The kinematic relations between the total strain components \(\varepsilon _r ,\varepsilon _\vartheta , \varepsilon _z\) and the shear angles \(\gamma _{r\vartheta }, \gamma _{rz}, \gamma _{\vartheta z}\) in terms of uvw and the equilibrium equations can be taken from any continuum mechanics textbook. Hooke’s law links the stress components \(\sigma _r, \sigma _\vartheta , \sigma _z, \sigma _{r\vartheta }, \sigma _{rz}, \sigma _{\vartheta z}\) to the elastic strain contributions. Furthermore, we use the following abbreviations
$$\begin{aligned} \cos \vartheta =c, \quad \sin \vartheta =s, \quad \cos 2\vartheta =c_2 , \quad \sin 2\vartheta =s_2 . \end{aligned}$$
(1)

3.1 Solution for the area eigenstrain \({\varepsilon ^{\text {eig}}}\)

Although the solution for this problem is known, for the sake of completeness we start the stress calculation with an in-plane (area) eigenstrain \({\varepsilon ^{\text {eig}}}\) together with an eigenstrain \({\varepsilon _z^{\text {eig}}}\) in z-direction keeping in mind that \(\varepsilon _z \equiv 0\), which yields
$$\begin{aligned} \sigma _z =-E{\varepsilon _z^{\text {eig}}} +\nu ({\sigma _r +\sigma _\vartheta }). \end{aligned}$$
(2)
Expressing the equilibrium equation only in u(r), since \(v=0\) and \(w=0\), one finds after some analysis
$$\begin{aligned} u=Ar+B/r. \end{aligned}$$
(3)
The integration constants A, B can be calculated from the two contact conditions
$$\begin{aligned} r=R: u |_\mathrm{i} = u |_\mathrm{o},\quad r=R: \sigma _r|_\mathrm{i} = \sigma _r |_\mathrm{o}. \end{aligned}$$
(4)
The subscripts “i” and “o” stand for inside and outside of the inclusion, resp. Finally, we have the following solution for the nonzero stress components as
$$\begin{aligned} 0\le & {} r\le R \qquad \sigma _r=\sigma _\vartheta =-\frac{E \left( {{\varepsilon ^{\text {eig}}}+\nu \varepsilon _z^{\text {eig}}}\right) }{2({1-\nu ^{2}})},\quad \sigma _z=-\frac{E\left( {\nu \varepsilon ^{\text {eig}}+\varepsilon _z^{\text {eig}}}\right) }{({1-\nu ^{2}})}, \end{aligned}$$
(5.1)
$$\begin{aligned} r> & {} R \qquad \qquad \sigma _r=\frac{-E \left( {\varepsilon ^{\text {eig}}+\nu \varepsilon _z^{\text {eig}}}\right) }{2({1-\nu ^{2}})}\frac{R^{2}}{r^{2}}, \quad \sigma _\vartheta =-\sigma _r,\quad \sigma _z=-E\varepsilon _z^{\text {eig}} . \end{aligned}$$
(5.2)
Note that only a deviatoric in-plane stress state exists for \(r\ge R\) and \(\varepsilon _z^{\text {eig}} =0\). This fact is denoted as “Bitter–Crum” theorem; for details, see Fratzl and Penrose [28].

3.2 Solution concept for \({\varepsilon _x^{\text {eig}}} \ne 0\) and \(\varepsilon _y^{\text {eig}} \ne 0\)

First we calculate the stress state for \(\varepsilon _x^{\text {eig}} \ne 0\) only and transform the according eigenstrain tensor in polar coordinates as
$$\begin{aligned} {\varvec{\upvarepsilon }}_x^{\text {eig}} =\frac{\varepsilon _x^{\text {eig}}}{2} ({\mathbf{I}+\mathbf{P}}), \quad \mathbf{P}=\left[ {\begin{array}{cc} c_2 &{} -s_2 \\ -s_2 &{} -c_2 \\ \end{array}} \right] , \end{aligned}$$
(6)
with I as unit tensor and the abbreviations by Eq. (1). The first part with I of the r.h.s. of Eq. (6) represents a homogeneous eigenstrain with the solutions of Sect. 3.1 for \(\varepsilon ^{\text {eig}}={\varepsilon _x^{\text {eig}}}/2\) and \(\varepsilon _z^{\text {eig}} =0\). The second part with \(\mathbf{P}\) represents a tensor varying with the angle \(2\vartheta \). Here we select a proper biharmonic Airy stress function \(\phi (r,\vartheta )\), inspired by the solution of the “inclusion problem,” by Mal and Singh [22], Example 7.5–8, as
$$\begin{aligned} {\phi }= ({Ar^{2}+Cr^{4}}) c_2 \quad \hbox {for} \quad 0\le r\le R, \quad {\phi }= ({B/r^{2}+D})c_2 \quad \hbox {for} \quad r\ge R. \end{aligned}$$
(7)
We find the nonzero stress components by the following operation, see [21], Sect. 5.2.3 there,
$$\begin{aligned} \sigma _r= & {} 1/r {\partial \phi }/{\partial r}+1/{r^{2}}{\partial ^{2}\phi }/{\partial \vartheta ^{2}}, \nonumber \\ \sigma _\vartheta= & {} {\partial ^{2}\phi }/{\partial r^{2}},\quad \sigma _{r\vartheta } ={-\partial }/{\partial r} ({1/r\;{\partial \phi }/{\partial \vartheta }}), \end{aligned}$$
(8)
yielding
$$\begin{aligned} 0\le r\le R \qquad \sigma _r= & {} -2Ac_2 ,\quad \sigma _\vartheta = \left( {2A+12Cr^{2}}\right) c_2 , \nonumber \\ \sigma _{r\vartheta }= & {} \left( {2A+6Cr^{2}}\right) s_2; \end{aligned}$$
(9.1)
$$\begin{aligned} r\ge R \qquad \qquad \sigma _r= & {} \left( {-{6B}/{r^{4}}-{4D}/{r^{2}}}\right) c_2, \quad \sigma _\vartheta = \left( {{6B}/{r^{4}}}\right) c_2,\nonumber \\ \sigma _{r\vartheta }= & {} \left( {{-6B}/{r^{4}}-{2D}/{r^{2}}}\right) s_2 . \end{aligned}$$
(9.2)
The integration constants ABCD can be calculated from four contact conditions as
$$\begin{aligned} r=R: \qquad u |_\mathrm{i} = u |_\mathrm{o} ,\; \vartheta |_\mathrm{i} = \vartheta |_\mathrm{o} ,\quad \; {\sigma _r} |_\mathrm{i} = {\sigma _r } |_\mathrm{o} ,\quad \; {\sigma _{r\vartheta }} |_\mathrm{i} = {\sigma _{r\vartheta }} |_\mathrm{o} . \end{aligned}$$
(10)
According to Eshelby [13], the stress state must be homogeneous in the inclusion with the consequence that C must become zero. Furthermore, the contact conditions enforce the calculation of \(u (r, \vartheta )\) and \(v (r,\vartheta )\) via the integration of the kinematic equations for \(\varepsilon _r\) and \(\varepsilon _\vartheta \), combined with Hooke’s law (including \(\varepsilon _z =0\) yielding \(\sigma _z =\nu ({\sigma _r +\sigma _\vartheta }))\), as
$$\begin{aligned} E\varepsilon _r= & {} {E\partial u}/{\partial r}=\sigma _r -\nu \sigma _\vartheta -\nu ^{2} ({\sigma _r +\sigma _\vartheta }) +E\varepsilon _x^{\text {eig}} {c_2 }/2, \end{aligned}$$
(11.1)
$$\begin{aligned} E\varepsilon _\vartheta= & {} E{ ({{\partial v}/{\partial \vartheta }+u})}/r = ( {-\nu \sigma _r +\sigma _\vartheta -\nu ^{2} ({\sigma _r +\sigma _\vartheta })}) - E\varepsilon _x^{\text {eig}} {c_2 }/2. \end{aligned}$$
(11.2)
The integration of Eqs. (11) involves two functions, namely \(f_u (\vartheta )\) in \(u ({r,\vartheta })\) and \(g_v (r)\) in \(v ({r,\vartheta })\). These two functions can be found by employing \(E\varepsilon _{r\vartheta }\) with Eu and Ev from above as
$$\begin{aligned} E\varepsilon _{r\vartheta } = E{\left( {\frac{1}{r}\frac{\partial u}{\partial \vartheta }+\frac{\partial v}{\partial r}-\frac{v}{r}}\right) }\Big /2 = ({1+\nu })\sigma _{r\vartheta } -E\varepsilon _x^{\text {eig}} {s_2 }/2. \end{aligned}$$
(12)
Inserting now u together with \(f_u (\vartheta )\) and v together with \(g_v (r)\) in Eq. (12) shows that both \(f_u (\vartheta )\) and \(g_r (r)\) can be interpreted as “rigid” body motions and, therefore, can be skipped.
To summarize, the calculation of the coefficients \(A,\, B,\, C,\, D\) and, consequently, the stresses makes a lot of algebraic operations necessary, which were left to mathematica (https://​www.​wolfram.​com/​mathematica/​) but result indeed in \(C=0\), see above, and rather simple equations for the stresses. Using the abbreviation \(\tilde{E}={E\varepsilon _x^{\text {eig}}}/ {({8 ({1-\nu ^{2}})})}\) and denoting the stress terms corresponding to P in Eq. (6) with a subscript “P” yield
$$\begin{aligned} 0\le r\le R: \qquad \sigma _r^{\mathbf{P}}= & {} -\tilde{E}c_2 ,\quad \sigma _\vartheta ^{\mathbf{P}} =\tilde{E}c_2 ,\nonumber \\ \sigma _{r\vartheta }^{\mathbf{P}}= & {} \tilde{E}s_2 ,\quad \sigma _z^{\mathbf{P}} =0; \end{aligned}$$
(13.1)
$$\begin{aligned} r\ge R:\qquad \qquad \sigma _r^{\mathbf{P}}= & {} -\tilde{E}\left( {{4R^{2}}/{r^{2}}-{3R^{4}}/{r^{4}}} \right) c_2 , \quad \sigma _\vartheta ^{\mathbf{P}} =-\tilde{E}\left( {{3R^{4}}/{r^{4}}} \right) c_2 ,\nonumber \\ \sigma _{r\vartheta }^{\mathbf{P}}= & {} \tilde{E}\left( {{3R^{4}}/{r^{4}}-{2R^{2}}/{r^{2}}} \right) s_2 , \quad \sigma _z^{\mathbf{P}} =-4\nu \tilde{E}{R^{2}}/{r^{2}}. \end{aligned}$$
(13.2)
Let us check now the stress state inside the inclusion with respect to the xy system. This means that stresses \(\sigma _r^{\mathbf{P}}, \sigma _\vartheta ^{\mathbf{P}}, \sigma _{r\vartheta }^{\mathbf{P}}\) must be transformed into the xy system. The components \(\sigma _x^{\mathbf{P}}, \sigma _y^{\mathbf{P}}, \sigma _{\textit{xy}}^{\mathbf{P}}\) follow after some algebra
$$\begin{aligned} 0\le r\le R:\sigma _x^{\mathbf{P}} =- \tilde{E},\quad \sigma _y^{\mathbf{P}} = \tilde{E}, \quad \sigma _{\textit{xy}}^{\mathbf{P}} =0. \end{aligned}$$
(13.3)
This result is in accordance with the Eshelby [13] theorem, stating that the stress state must be homogeneous in the inclusion for a homogeneous eigenstrain.
The problem of calculating the stress state to an eigenstrain field \(\varepsilon _y^{\text {eig}}\) only is easily solved by rotation of the total configuration by \(\pi /2\) in relation to the eigenstrain field \(\varepsilon _x^{\text {eig}}\), yielding, in analogy to Eq. (6),
$$\begin{aligned} {\varvec{\upvarepsilon }}_y^{\text {eig}} = \frac{\varepsilon _y^{\text {eig}} }{2} ({\mathbf{I}-\mathbf{P}}). \end{aligned}$$
(14)
The superposition of \({\varvec{\upvarepsilon }}_x^{\text {eig}}\) and \({\varvec{\upvarepsilon }}_y^{\text {eig}}\) yields for the homogeneous eigenstrain \(\varepsilon ^{\text {eig}}={({\varepsilon _x^{\text {eig}} + \varepsilon _y^{\text {eig}}})}/2\) and for the contribution due to P the eigenstrain difference \(\Delta \varepsilon ^{\text {eig}}=\varepsilon _x^{\text {eig}} -\varepsilon _y^{\text {eig}}\) and with the abbreviations \(\tilde{E}^{I}={E ( {\varepsilon ^{\text {eig}}+\nu \varepsilon _z^{\text {eig}}})}/{8 ({1-\nu ^{2}} )} \quad \tilde{E}_\varepsilon ^{\mathbf{P}} ={E\Delta \varepsilon ^{\text {eig}}}/{8 ({1-\nu ^{2}})}\),
$$\begin{aligned} 0\le r\le R:\quad \sigma _r= & {} -4\tilde{E}^{\mathbf{I}}- \tilde{E}_\varepsilon ^{\mathbf{P}} c_2 ,\quad \sigma _\vartheta =-4\tilde{E}^{\mathbf{I}}+\tilde{E}_\varepsilon ^{\mathbf{P}} c_2,\nonumber \\ \sigma _{r\vartheta }= & {} \tilde{E}_\varepsilon ^{\mathbf{P}} s_2 ,\quad \sigma _z =-E\frac{\left( {\nu \varepsilon ^{\text {eig}}+\varepsilon _z^{\text {eig}} } \right) }{\left( {1-\nu ^{2}} \right) }; \end{aligned}$$
(15.1)
$$\begin{aligned} r\ge R:\quad \sigma _r= & {} -4\tilde{E}^{\mathbf{I}}{R^{2}}/{r^{2}}-\tilde{E}_\varepsilon ^{\mathbf{P}} \left( {{4R^{2}}/{r^{2}}-{3R^{4}}/{r^{4}}} \right) c_2 , \quad \sigma _\vartheta =4\tilde{E}^{\mathbf{I}}{R^{2}}/{r^{2}}-3\tilde{E}_\varepsilon ^{\mathbf{P}} \left( {{R^{4}}/{r^{4}}} \right) c_2 ,\nonumber \\ \sigma _{r\vartheta }= & {} \tilde{E}_\varepsilon ^{\mathbf{P}} \left( {{3R^{4}}/{r^{4}}-{2R^{2}}/{r^{2}}} \right) s_2 , \quad \sigma _z =-4\nu \tilde{E}_\varepsilon ^{\mathbf{P}} \left( {{R^{2}}/{r^{2}}} \right) c_2. \end{aligned}$$
(15.2)

3.3 Solution concept for \(\varepsilon _{\textit{xy}}^{\text {eig}} \)

We complete the results by calculating the stress state according to the shear eigenstrain \(\varepsilon _{\textit{xy}}^{\text {eig}}\) (or the shear angle \(\gamma _{\textit{xy}}^{\text {eig}} =2\varepsilon _{\textit{xy}}^{\text {eig}}\)) only and consider a coordinate system \(x'{-}y'{-}z\) rotated by an angle of \(\pi {/}4\) in relation to the master xyz coordinate system with the eigenstrains \(\varepsilon _{{x}'}^{\text {eig}} =\varepsilon _{\textit{xy}}^{\text {eig}}\) and \(\varepsilon _{{y}'}^{\text {eig}} =-\varepsilon _{\textit{xy}}^{\text {eig}}\). Then we can apply the solution concept for \(\varepsilon _x^{\text {eig}}\) and \(\varepsilon _y^{\text {eig}}\), however, in the \(x'{-}y'{-}z\) coordinate system with \(\varepsilon ^{\text {eig}}=0\), \(\Delta \varepsilon ^{\text {eig}} = \gamma _{\textit{xy}}^{\text {eig}}\) and \(\varepsilon _z^{\text {eig}} =0\). Using the abbreviation \(\tilde{E}_\gamma ^{\mathbf{P}} ={E\varepsilon _{\textit{xy}}^{\text {eig}}}/ {4 ({1-\nu ^{2}})}\) and noting that \({\vartheta }' = \vartheta -\pi /4\), the relations in the master coordinate system read as
$$\begin{aligned} 0\le r\le R \qquad \sigma _r= & {} -\tilde{E}_\gamma ^{\mathbf{P}} s_2 ,\quad \sigma _\vartheta =\tilde{E}_\gamma ^{\mathbf{P}} s_2 ,\nonumber \\ \sigma _{r\vartheta }= & {} -\tilde{E}_\gamma ^{\mathbf{P}} c_2 ,\quad \sigma _z =0; \end{aligned}$$
(15.1)
$$\begin{aligned} R<r \quad \qquad \sigma _r= & {} -\tilde{E}_\gamma ^{\mathbf{P}} \left( {{4R^{2}}/{r^{2}}-{3R^{4}}/{r^{4}}}\right) s_2 ,\quad \sigma _\vartheta =-3\tilde{E}_\gamma ^{\mathbf{P}} \left( {{R^{4}}/{r^{4}}}\right) s_2 ,\nonumber \\ \sigma _{r\vartheta }= & {} -\tilde{E}_\gamma ^{\mathbf{P}} \left( {{3R^{4}}/{r^{4}}-{2R^{2}}/{r^{2}}}\right) c_2 ,\quad \sigma _z =-4v\tilde{E}_\gamma ^{\mathbf{P}} \left( R^{2}/r^{2}\right) s_2 . \end{aligned}$$
(15.2)

4 Representative examples and discussion

We demonstrate two examples in a dimension-free form. Assuming \(E \varepsilon _x^{\text {eig}} =8 ({1-\nu ^{2}})\), we demonstrate \(\sigma _x\) and \(\sigma _y\) along the x-axis in Fig. 2a and \(\sigma _{r\vartheta }\) along the radius r for \(\vartheta =45^{\circ }\) in Fig. 2b. Assuming \(E\varepsilon _{\textit{xy}}^{\text {eig}} =4 ({1-\nu ^{2}})\) we demonstrate \(\sigma _r\) and \(\sigma _\vartheta \) along the radius r for \(\vartheta =45^{\circ }\) in Fig. 3a and \(\sigma _{\textit{xy}}\) along the x-axis in Fig. 3b. All curves are checked by a finite element study with ABAQUS (http://​www.​3ds.​com/​de/​produkte-und-services/​simulia/​produkte/​abaqus/​).
It is interesting to note that the stresses outside of the inclusion, dominated by a \(({R/r})^{2}\) term, decay to nearly zero over a remarkably long distance of approximately 10R.
For practical application of the results, we present for each kind of eigenstrain (i.e., \(\varepsilon _x^{\text {eig}}, \varepsilon _y^{\text {eig}}, \varepsilon _{\textit{xy}}^{\text {eig}}, \varepsilon _z^{\text {eig}}\)) the stress fields in a Cartesian coordinate system in “Appendix.”

Acknowledgements

Open access funding provided by Montanuniversität Leoben. Financial support by the Austrian Federal Government (in particular from the Bundesministerium für Verkehr, Innovation and Technologie and the Bundesministerium für Wirtschaft und Arbeit) and the Styrian Provincial Government, represented by Österreichische Forschungsförderungsgesellschaft mbH and by Steirische Wirtschaftsförderungsgesellschaft mbH, within the research activities of the K2 Competence Centre on “Integrated Research in Materials, Processing and Product Engineering,” operated by the Materials Center Leoben Forschung GmbH in the framework of the Austrian COMET Competence Centre Programme, Projects A1.23 and A2.32, is gratefully acknowledged. J.S. gratefully acknowledges the financial support by the Czech Science Foundation in the frame of the Project 17-01641S.
Open AccessThis article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://​creativecommons.​org/​licenses/​by/​4.​0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.
Anhänge

Appendix: Eigenstress fields in a Cartesian coordinate system

With the abbreviations in Eq. (1), we have
$$\begin{aligned} \sigma _x= & {} c^{2}\sigma _r +s^{2}\sigma _\vartheta -s_2 \sigma _{r\vartheta } , \quad \sigma _y =s^{2}\sigma _r +c^{2}\sigma _\vartheta -s_2 \sigma _{r\vartheta }\nonumber \\ \end{aligned}$$
(A.1)
$$\begin{aligned} \sigma _{\textit{xy}}= & {} sc ({\sigma _r -\sigma _\vartheta }) +c_2 \sigma _{r\vartheta } , \quad \sigma _z =-E\varepsilon _z^{\text {eig}} +\nu ({\sigma _r +\sigma _\vartheta }), \end{aligned}$$
(A.2)
with \(c=x/r\), \(s=y/r\) and \(r^{2}=x^{2}+y^{2}\) and \(\bar{{x}}=x/R\), \(\bar{{y}}=y/R\) and \(\bar{{r}}=r/R\).
Expressing the relations for the individual cases of eigenstrain yields:
$$\begin{aligned} \varepsilon _x^{\text {eig}} \ne 0 \end{aligned}$$
Inside
$$\begin{aligned} \sigma _x =-\frac{3E\varepsilon _x^{\text {eig}} }{8\left( {1-\nu ^{2}}\right) }, \quad \sigma _y =-\frac{E\varepsilon _x^{\text {eig}} }{8\left( {1-\nu ^{2}}\right) }, \quad \sigma _z =-\frac{Ev\varepsilon _x^{\text {eig}} }{2\left( {1-\nu ^{2}}\right) }. \end{aligned}$$
Outside
$$\begin{aligned} \sigma _x= & {} -\frac{E\varepsilon _x^{\text {eig}} \left[ {2\bar{{r}}^{4}\left( {\bar{{x}}^{2}-\bar{{y}}^{2}}\right) + 4\bar{{r}}^{2}\bar{{x}}^{2}\left( {\bar{{x}}^{2}-3\bar{{y}}^{2}}\right) -3 \left( {\bar{{x}}^{4}-6\bar{{x}}^{2}\bar{{y}}^{2} +\bar{{y}}^{4}}\right) }\right] }{8\left( {1-\nu ^{2}}\right) \bar{{r}}^{8}},\\ \sigma _y= & {} \frac{E\varepsilon _x^{\text {eig}} \left[ {2\bar{{r}}^{4} \left( {\bar{{x}}^{2}-\bar{{y}}^{2}}\right) -4\bar{{r}}^{2}\bar{{y}}^{2} \left( {3\bar{{x}}^{2}-\bar{{y}}^{2}}\right) -3 \left( {\bar{{x}}^{4}- 6\bar{{x}}^{2}\bar{{y}}^{2}+\bar{{y}}^{4}}\right) }\right] }{8\left( {1-\nu ^{2}}\right) \bar{{r}}^{8}},\\ \sigma _{\textit{xy}}= & {} -\frac{E\varepsilon _x^{\text {eig}} \bar{{x}}\bar{{y}} \left[ {\bar{{r}}^{4}+\left( {2\bar{{r}}^{2}-3}\right) \left( {\bar{{x}}^{2}-\bar{{y}}^{2}}\right) }\right] }{2\left( {1-\nu ^{2}}\right) \bar{{r}}^{8}}, \quad \sigma _z =-\frac{Ev\varepsilon _x^{\text {eig}} \left( {\bar{{x}}^{2}-\bar{{y}}^{2}}\right) }{2\left( {1-\nu ^{2}}\right) \bar{{r}}^{4}}. \end{aligned}$$
$$\begin{aligned} \varepsilon _y^{\text {eig}}\ne & {} 0 \end{aligned}$$
Inside
$$\begin{aligned} \sigma _x =-\frac{E\varepsilon _y^{\text {eig}} }{8\left( {1-\nu ^{2}}\right) }, \quad \sigma _y =-\frac{3E\varepsilon _y^{\text {eig}} }{8\left( {1-\nu ^{2}}\right) }, \quad \sigma _z =-\frac{Ev\varepsilon _y^{\text {eig}} }{2\left( {1-\nu ^{2}}\right) }. \end{aligned}$$
Outside
$$\begin{aligned} \sigma _x= & {} -\frac{E\varepsilon _y^{\text {eig}} \left[ {2\bar{{r}}^{4}\left( {\bar{{x}}^{2}-\bar{{y}}^{2}}\right) - 4\bar{{r}}^{2}\bar{{x}}^{2} \left( {\bar{{x}}^{2}-3\bar{{y}}^{2}}\right) + 3 \left( {\bar{{x}}^{4}-6\bar{{x}}^{2}\bar{{y}}^{2}+\bar{{y}}^{4}}\right) } \right] }{8\left( {1-\nu ^{2}}\right) \bar{{r}}^{8}},\\ \sigma _y= & {} \frac{E\varepsilon _y^{\text {eig}} \left[ {2\bar{{r}}^{4}\left( {\bar{{x}}^{2}-\bar{{y}}^{2}}\right) + 4\bar{{r}}^{2}\bar{{y}}^{2}\left( {3\bar{{x}}^{2}-\bar{{y}}^{2}}\right) +3 \left( {\bar{{x}}^{4}-6\bar{{x}}^{2}\bar{{y}}^{2}+\bar{{y}}^{4}}\right) } \right] }{8\left( {1-\nu ^{2}}\right) \bar{{r}}^{8}},\\ \sigma _{\textit{xy}}= & {} -\frac{E\varepsilon _y^{\text {eig}} \bar{{x}}\bar{{y}} \left[ {\bar{{r}}^{4}-\left( {2\bar{{r}}^{2}-3}\right) \left( {\bar{{x}}^{2}- \bar{{y}}^{2}}\right) }\right] }{2\left( {1-\nu ^{2}}\right) \bar{{r}}^{8}}, \quad \sigma _z =-\frac{Ev\varepsilon _y^{\text {eig}} \left( {\bar{{x}}^{2}-\bar{{y}}^{2}}\right) }{2\left( {1-\nu ^{2}}\right) \bar{{r}}^{4}}. \end{aligned}$$
$$\begin{aligned} \varepsilon _{\textit{xy}}^{\text {eig}}\ne & {} 0 \end{aligned}$$
Inside
$$\begin{aligned} \sigma _x =\sigma _y =0,\quad \sigma _{\textit{xy}} =-\frac{E\varepsilon _{\textit{xy}}^{\text {eig}} }{4\left( {1-\nu ^{2}}\right) }, \end{aligned}$$
Outside
$$\begin{aligned} \sigma _x= & {} -\frac{E\varepsilon _{\textit{xy}}^{\text {eig}} \bar{{x}}\bar{{y}} \left[ {\bar{{r}}^{2} \left( {3\bar{{x}}^{2}-\bar{{y}}^{2}}\right) - 3\left( {\bar{{x}}^{2}-\bar{{y}}^{2}}\right) }\right] }{\left( {1-\nu ^{2}}\right) \bar{{r}}^{8}}, \quad \sigma _y =\frac{E\varepsilon _{\textit{xy}}^{\text {eig}} \bar{{x}}\bar{{y}} \left[ {\bar{{r}}^{2} \left( {\bar{{x}}^{2}-3\bar{{y}}^{2}}\right) -3\left( {\bar{{x}}^{2}-\bar{{y}}^{2}}\right) }\right] }{\left( {1-\nu ^{2}}\right) \bar{{r}}^{8}},\\ \sigma _{\textit{xy}}= & {} \frac{E\varepsilon _{\textit{xy}}^{\text {eig}} \left( {2\bar{{r}}^{2}-3}\right) \left( {\bar{{x}}^{4}-6\bar{{x}}^{2}\bar{{y}}^{2}+\bar{{y}}^{4}}\right) }{4\left( {1-\nu ^{2}}\right) \bar{{r}}^{8}}, \quad \sigma _z =-\frac{2Ev\varepsilon _{\textit{xy}}^{\text {eig}} \left( {\bar{{x}}\bar{{y}}}\right) }{\left( {1-\nu ^{2}}\right) \bar{{r}}^{4}}. \end{aligned}$$
$$\begin{aligned} \varepsilon _z^{\text {eig}}\ne & {} 0 \end{aligned}$$
Inside
$$\begin{aligned} \sigma _x =-\frac{E\nu \varepsilon _z^{\text {eig}} }{2\left( {1-\nu ^{2}}\right) }, \quad \sigma _y =-\frac{E\nu \varepsilon _z^{\text {eig}} }{2\left( {1-\nu ^{2}}\right) }, \quad \sigma _z =-\frac{E\varepsilon _z^{\text {eig}} }{\left( {1-\nu ^{2}}\right) }. \end{aligned}$$
Outside
$$\begin{aligned} \sigma _x= & {} -\frac{E\nu \varepsilon _z^{\text {eig}} \left( {\bar{{x}}^{2}-\bar{{y}}^{2}}\right) }{2\left( {1-\nu ^{2}}\right) \bar{{r}}^{4}}, \quad \sigma _y =\frac{E\nu \varepsilon _z^{\text {eig}} \left( {\bar{{x}}^{2}-\bar{{y}}^{2}}\right) }{2\left( {1-\nu ^{2}}\right) \bar{{r}}^{4}},\\ \sigma _{\textit{xy}}= & {} -\frac{E\nu \varepsilon _z^{\text {eig}} \bar{{x}}\bar{{y}}}{\left( {1-\nu ^{2}}\right) \bar{{r}}^{4}},\quad \sigma _z =0. \end{aligned}$$
Literatur
1.
Zurück zum Zitat Cottrell, A.H., Bilby, B.A.: Dislocation theory on yielding and strain ageing in iron. Proc. Phys. Soc. Lond. Sect. A 62, 49–62 (1949)CrossRef Cottrell, A.H., Bilby, B.A.: Dislocation theory on yielding and strain ageing in iron. Proc. Phys. Soc. Lond. Sect. A 62, 49–62 (1949)CrossRef
2.
Zurück zum Zitat Svoboda, J., Zickler, G.A., Kozeschnik, E., Fischer, F.D.: Kinetics of interstitials segregation in Cottrell atmospheres and grain boundaries. Philos. Mag. Lett. 95, 458–465 (2015)CrossRef Svoboda, J., Zickler, G.A., Kozeschnik, E., Fischer, F.D.: Kinetics of interstitials segregation in Cottrell atmospheres and grain boundaries. Philos. Mag. Lett. 95, 458–465 (2015)CrossRef
3.
Zurück zum Zitat Cochardt, A.W., Schoeck, G., Wiedersich, H.: Interaction between dislocations and interstitial atoms in body-centered cubic metals. Acta Metall. 3, 533–537 (1955)CrossRef Cochardt, A.W., Schoeck, G., Wiedersich, H.: Interaction between dislocations and interstitial atoms in body-centered cubic metals. Acta Metall. 3, 533–537 (1955)CrossRef
4.
Zurück zum Zitat Krempasky, C., Liedl, U., Werner, E.A.: A note on the diffusion of carbon atoms to dislocations. Comput. Mater. Sci. 38, 90–97 (2006)CrossRef Krempasky, C., Liedl, U., Werner, E.A.: A note on the diffusion of carbon atoms to dislocations. Comput. Mater. Sci. 38, 90–97 (2006)CrossRef
5.
Zurück zum Zitat Friedel, J.: Dislocations, Pergamon Student Editions, vol. 3. Pergamon Press, Oxford (1964) Friedel, J.: Dislocations, Pergamon Student Editions, vol. 3. Pergamon Press, Oxford (1964)
6.
Zurück zum Zitat Cahn, J.W.: Thermodynamic aspects of Cottrell atmospheres. Philos. Mag. 93, 3741–3746 (2013)CrossRef Cahn, J.W.: Thermodynamic aspects of Cottrell atmospheres. Philos. Mag. 93, 3741–3746 (2013)CrossRef
7.
Zurück zum Zitat Cahn, J.W.: Reprise: partial chemical strain dislocations and their role in pinning dislocations to their atmospheres. Philos. Mag. 94, 3170–3176 (2014)CrossRef Cahn, J.W.: Reprise: partial chemical strain dislocations and their role in pinning dislocations to their atmospheres. Philos. Mag. 94, 3170–3176 (2014)CrossRef
8.
Zurück zum Zitat Mishin, Y., Cahn, J.W.: Thermodynamics of Cottrell atmospheres tested by atomistic simulations. Acta Mater. 117, 197–206 (2016)CrossRef Mishin, Y., Cahn, J.W.: Thermodynamics of Cottrell atmospheres tested by atomistic simulations. Acta Mater. 117, 197–206 (2016)CrossRef
9.
Zurück zum Zitat Cai, W., Sills, R.B., Barnett, D.M., Nix, W.D.: Modeling a distribution of point defects as misfitting inclusions in stressed solids. J. Mech. Phys. Solids 66, 154–171 (2014)CrossRefMATH Cai, W., Sills, R.B., Barnett, D.M., Nix, W.D.: Modeling a distribution of point defects as misfitting inclusions in stressed solids. J. Mech. Phys. Solids 66, 154–171 (2014)CrossRefMATH
10.
Zurück zum Zitat Wilde, J., Cerezo, A., Smith, G.D.W.: Three-dimensional atomic-scale mapping of a Cottrell atmosphere around a dislocation in iron. Scripta Mater. 43, 39–48 (2000)CrossRef Wilde, J., Cerezo, A., Smith, G.D.W.: Three-dimensional atomic-scale mapping of a Cottrell atmosphere around a dislocation in iron. Scripta Mater. 43, 39–48 (2000)CrossRef
11.
Zurück zum Zitat Veiga, R.G.A., Perez, M., Becquart, C.S., Comain, C.: Atomistic modeling of carbon Cottrell atmospheres in bcc iron. J. Phys. Condens. Matter 25, 025401 (2013). (7 pp)CrossRef Veiga, R.G.A., Perez, M., Becquart, C.S., Comain, C.: Atomistic modeling of carbon Cottrell atmospheres in bcc iron. J. Phys. Condens. Matter 25, 025401 (2013). (7 pp)CrossRef
12.
Zurück zum Zitat Waseda, O., Veiga, R.G.A., Morthomas, J., Chantrenne, P., Becquart, C.S., Ribeiro, F., Jelea, A., Goldenstein, H., Perez, M.: Formation of carbon Cottrell atmospheres and their effect on the stress field around an edge dislocation. Scripta Mater. 129, 16–19 (2017)CrossRef Waseda, O., Veiga, R.G.A., Morthomas, J., Chantrenne, P., Becquart, C.S., Ribeiro, F., Jelea, A., Goldenstein, H., Perez, M.: Formation of carbon Cottrell atmospheres and their effect on the stress field around an edge dislocation. Scripta Mater. 129, 16–19 (2017)CrossRef
13.
Zurück zum Zitat Eshelby, J.D.: The determination of the elastic field of an inclusion, and related problems. Proc. R. Soc. Lond. A 241, 376–396 (1957)MathSciNetCrossRefMATH Eshelby, J.D.: The determination of the elastic field of an inclusion, and related problems. Proc. R. Soc. Lond. A 241, 376–396 (1957)MathSciNetCrossRefMATH
15.
Zurück zum Zitat Mura, T.: Micromechanics of Defects in Solids, 2nd edn. Martinus Nijhoff Publishers, Dordrecht (1987)CrossRefMATH Mura, T.: Micromechanics of Defects in Solids, 2nd edn. Martinus Nijhoff Publishers, Dordrecht (1987)CrossRefMATH
16.
Zurück zum Zitat Fischer, F.D., Böhm, H.J.: On the role of transformation eigenstrain in the growth or shrinkage of a spheroidal inclusion, with a general eigenstrain state. Acta Mater. 54(151–156), 55 (2006, 2007) Fischer, F.D., Böhm, H.J.: On the role of transformation eigenstrain in the growth or shrinkage of a spheroidal inclusion, with a general eigenstrain state. Acta Mater. 54(151–156), 55 (2006, 2007)
17.
Zurück zum Zitat Li, S., Sauer, R., Wang, G.: A cicular inclusion in a finite domain I. The Dirichlet–Eshelby problem. Acta Mech. 179, 67–90 (2005)CrossRefMATH Li, S., Sauer, R., Wang, G.: A cicular inclusion in a finite domain I. The Dirichlet–Eshelby problem. Acta Mech. 179, 67–90 (2005)CrossRefMATH
18.
Zurück zum Zitat Wang, G., Li, S., Sauer, R.: A circular inclusion in a finite domain II. The Neumann–Eshelby problem. Acta Mech. 179, 91–110 (2005)CrossRef Wang, G., Li, S., Sauer, R.: A circular inclusion in a finite domain II. The Neumann–Eshelby problem. Acta Mech. 179, 91–110 (2005)CrossRef
19.
Zurück zum Zitat Markenscoff, X., Dundurs, J.: Annular inhomogeneities with eigenstrain and interphase modeling. J. Mech. Phys. Solids 64, 468–482 (2014)MathSciNetCrossRef Markenscoff, X., Dundurs, J.: Annular inhomogeneities with eigenstrain and interphase modeling. J. Mech. Phys. Solids 64, 468–482 (2014)MathSciNetCrossRef
20.
Zurück zum Zitat Shodja, H.M., Khorshidi, A.: Comment on “Annular inhomogeneities with eigenstrain and interphase modeling [2014, J. Mech. Phys. Solids 64, 468–482]”. J. Mech. Phys. Solids 73, 1–2 (2014)MathSciNetCrossRef Shodja, H.M., Khorshidi, A.: Comment on “Annular inhomogeneities with eigenstrain and interphase modeling [2014, J. Mech. Phys. Solids 64, 468–482]”. J. Mech. Phys. Solids 73, 1–2 (2014)MathSciNetCrossRef
21.
Zurück zum Zitat Michell, J.H.: On the direct determination of stress in elastic solid, with application to the theory of plates. Proc. Lond. Math. Soc. 31, 100–124 (1899)MathSciNetCrossRefMATH Michell, J.H.: On the direct determination of stress in elastic solid, with application to the theory of plates. Proc. Lond. Math. Soc. 31, 100–124 (1899)MathSciNetCrossRefMATH
22.
Zurück zum Zitat Mal, A.K., Singh, S.J.: Deformation of Elastic Solids. Prentice-Hall, New Jersey (1991)MATH Mal, A.K., Singh, S.J.: Deformation of Elastic Solids. Prentice-Hall, New Jersey (1991)MATH
23.
24.
Zurück zum Zitat Bower, A.F.: Applied Mechanics of Solids. CRC Press, Boca Raton (2009) Bower, A.F.: Applied Mechanics of Solids. CRC Press, Boca Raton (2009)
25.
Zurück zum Zitat Kirsch, G.: Die Theorie der Elastizität und die Bedürfnisse der Festigkeitslehre. Z. VDI 42, 797–807 (1898). (in German) Kirsch, G.: Die Theorie der Elastizität und die Bedürfnisse der Festigkeitslehre. Z. VDI 42, 797–807 (1898). (in German)
26.
Zurück zum Zitat Bickley, W.G.: The distribution of stress round a circular hole in a plate. Philos. Trans. R. Soc. Lond. 227, 383–415 (1928)CrossRef Bickley, W.G.: The distribution of stress round a circular hole in a plate. Philos. Trans. R. Soc. Lond. 227, 383–415 (1928)CrossRef
27.
Zurück zum Zitat Sen, B.: Problems of thin plates with circular holes. Bull. Calcutta Math. Soc. 37, 37–42 (1945)MathSciNetMATH Sen, B.: Problems of thin plates with circular holes. Bull. Calcutta Math. Soc. 37, 37–42 (1945)MathSciNetMATH
28.
Zurück zum Zitat Fratzl, P., Penrose, O., Lebowitz, J.L.: Modelling of phase separation in alloys with coherent elastic misfit. J. Stat. Phys. 95, 1429–1503 (1999)CrossRefMATH Fratzl, P., Penrose, O., Lebowitz, J.L.: Modelling of phase separation in alloys with coherent elastic misfit. J. Stat. Phys. 95, 1429–1503 (1999)CrossRefMATH
Metadaten
Titel
Elastic stress–strain analysis of an infinite cylindrical inclusion with eigenstrain
verfasst von
F. D. Fischer
G. A. Zickler
J. Svoboda
Publikationsdatum
08.11.2017
Verlag
Springer Berlin Heidelberg
Erschienen in
Archive of Applied Mechanics / Ausgabe 3/2018
Print ISSN: 0939-1533
Elektronische ISSN: 1432-0681
DOI
https://doi.org/10.1007/s00419-017-1318-x

Weitere Artikel der Ausgabe 3/2018

Archive of Applied Mechanics 3/2018 Zur Ausgabe

    Marktübersichten

    Die im Laufe eines Jahres in der „adhäsion“ veröffentlichten Marktübersichten helfen Anwendern verschiedenster Branchen, sich einen gezielten Überblick über Lieferantenangebote zu verschaffen.