Skip to main content
Log in

Nonlinear mechanics of thermoelastic accretion

  • Published:
Zeitschrift für angewandte Mathematik und Physik Aims and scope Submit manuscript

Abstract

In this paper, we formulate a theory for the coupling of accretion mechanics and thermoelasticity. We present an analytical formulation of the thermoelastic accretion of an infinite cylinder and of a two-dimensional block. We develop numerical schemes for the solution of these two problems, and numerically calculate residual stresses and observe a strong dependence of the final mechanical state on the parameters of the accretion process. This suggests the possibility to predict and control thermal accretion processes of soft materials by manipulating thermal parameters, and therefore, to realize additively manufactured soft objects with the desired characteristics and performances.

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
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11

Similar content being viewed by others

Notes

  1. We denote by \({\varvec{T}}^{\star }\) the dual of the \((^1_1)\)-rank tensor \({\varvec{T}}\,\): operating on a 1-form \(\varvec{\lambda }\), it contracts its upper index with \(\varvec{\lambda }\), i.e., \({\varvec{T}}^{\star } \varvec{\lambda }=T^B{}_A\lambda _B{{\mathrm {d}}} X^A\). It should not be confused with the adjoint operator \({}^{{\mathsf {T}}}\), cf. footnote 3.

  2. Rank-1 convexity allowed Sozio and Yavari [49] to use an energetic argument to show that the absence of in-layer deformation and the vanishing of out-of-layer stress (tractions) imply \(\bar{{\varvec{F}}}={\varvec{Q}}\). Note that being an isometry, the accretion tensor \({\varvec{Q}}\) represents an undeformed state.

  3. We denote the adjoint of \({\varvec{F}}\) by \({\varvec{F}}^{{\mathsf {T}}}\) and it is defined such that \({\varvec{g}}({\varvec{F}}{\varvec{W}},{\varvec{w}})={\varvec{G}}({\varvec{W}},{\varvec{F}}^{{\mathsf {T}}}{\varvec{w}})\) for any pair \(({\varvec{W}},{\varvec{w}})\in T_X{\mathcal {B}}_t\times T_{\varphi _t(X)}{\mathcal {S}}\). In components, \((F^{{\mathsf {T}}})^A{}_a=g_{ab}F^b{}_BG^{AB}\).

  4. In Appendix D we discuss the derivation of this inequality.

  5. Note that \({\varvec{G}}^{\sharp } {\mathrm {dT}} = {\text {Grad}}T\).

  6. There are many other choices that will result in the same stress calculation. As a matter of fact, for isotropic solids the geometric theory suggests that the material body is represented by a class of infinitely many isometric Riemannian manifolds. The anisotropic case is slightly more complicated as one needs to look at the symmetry group for the constitutive equation, but there is still some arbitrariness. This was discussed in detail in [48, 50].

  7. If the inner boundary is subject to a traction (pressure) \(p_i\), the condition \(\sigma ^{rr}(S(t),t)=-p_i(t)\) gives the following equation

    $$\begin{aligned} \int \limits _{R_i}^{S(t)} \frac{\mu (T(\xi ,t))}{{\bar{r}}(\xi )} \left[ 1-\frac{{\bar{r}}^4(\xi ) e^{4\omega (T(\xi ,t))} }{r^4(\xi ,t) }\right] {{\mathrm {d}}}\xi = p_i(t). \end{aligned}$$
    (34)
  8. Note that \(\sigma ^{\theta \theta }(S(t),t)=0\) is not assumed; it is anticipated by virtue of \({\varvec{Q}} = \bar{{\varvec{F}}}\).

  9. It is assumed that throughout the whole accretion process the configuration of the accretion surface is given by points (xf(xt) ), for some time-dependent f, so that the vertical addition of material on the whole upper boundary will always be possible.

  10. Unlike the one-dimensional example of the axi-symmetric infinite cylinder, it is not possible to define s(t) as the deformation is not uniform along the width of the block.

  11. AY is grateful to Prof. Marshall Slemrod for a discussion that helped us find and correct this mistake.

  12. These simple relations were written incorrectly in Eq. (62) of [43].

  13. Note that in addition to the last two relations in Eq. (62), Eq. (66) in [43] is incorrect as well, and the final form of the energy balance in Eq. (69) must be changed to read \(\rho T\dot{{\mathcal {N}}}=\rho R-{\text {Div}}{\varvec{H}}+\rho T\frac{\partial {\mathcal {N}}}{\partial {\mathbf {G}}}\!:\!\dot{{\mathbf {G}}}\). Fortunately, nothing else in [43] was affected by this mistake.

References

  1. Abi-Akl, R., Abeyaratne, R., Cohen, T.: Kinetics of surface growth with coupled diffusion and the emergence of a universal growth path. Proc. R. Soc. A 475(2221), 20180465 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  2. Bikas, H., Stavropoulos, P., Chryssolouris, G.: Additive manufacturing methods and modelling approaches: a critical review. Int. J. Adv. Manuf. Technol. 83(1–4), 389–405 (2016)

    Article  Google Scholar 

  3. Chadwick, P.: Thermo-mechanics of rubberlike materials. Phil. Trans. R. Soc. Lond. Ser. A Math. Phys. Sci. 276(1260), 371–403 (1974)

    MATH  Google Scholar 

  4. Dadbakhsh, S., Hao, L., Sewell, N.: Effect of selective laser melting layout on the quality of stainless steel parts. Rapid Prototyp. J. 18(3), 241–249 (2012)

    Article  Google Scholar 

  5. de La Batut, B., Fergani, O., Brotan, V., Bambach, M., El Mansouri, M.: Analytical and numerical temperature prediction in direct metal deposition of Ti6Al4V. J. Manuf. Mater. Process. 1(1), 3 (2017)

    Google Scholar 

  6. Denlinger, E.R., Michaleris, P.: Mitigation of distortion in large additive manufacturing parts. Proc. Inst. Mech. Eng. Part B J. Eng. Manuf. 231(6), 983–993 (2017)

    Article  Google Scholar 

  7. Denlinger, E.R., Heigel, J.C., Michaleris, P.: Residual stress and distortion modeling of electron beam direct manufacturing Ti–6Al–4V. Proc. Inst. Mech. Eng. Part B J. Eng. Manuf. 229(10), 1803–1813 (2015)

    Article  Google Scholar 

  8. Dillon Jr., O.W.: A nonlinear thermoelasticity theory. J. Mech. Phys. Solids 10(2), 123–131 (1962)

    Article  MathSciNet  MATH  Google Scholar 

  9. Epstein, M., Maugin, G.A.: Thermomechanics of volumetric growth in uniform bodies. Int. J. Plast 16(7–8), 951–978 (2000)

    Article  MATH  Google Scholar 

  10. Faghih Shojaei, M., Ansari, R.: Variational differential quadrature: a technique to simplify numerical analysis of structures. Appl. Math. Model. 49, 705–738 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  11. Faghih Shojaei, M., Yavari, A.: Compatible-strain mixed finite element methods for incompressible nonlinear elasticity. J. Comput. Phys. 361, 247–279 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  12. Faghih Shojaei, M., Yavari, A.: Compatible-strain mixed finite element methods for 3D compressible and incompressible nonlinear elasticity. Comput. Methods Appl. Mech. Eng. 357, 112610 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  13. Fergani, O., Berto, F., Welo, T., Liang, S.: Analytical modelling of residual stress in additive manufacturing. Fatigue Fract. Eng. Mater. Struct. 40, 971–978 (2017)

    Article  Google Scholar 

  14. Frazier, W.E.: Direct digital manufacturing of metallic components: vision and roadmap. In: 21st Annual International Solid Freeform Fabrication Symposium, Austin, TX, Aug, pp. 9–11 (2010)

  15. Frazier, W.E.: Metal additive manufacturing: a review. J. Mater. Eng. Perform. 23(6), 1917–1928 (2014)

    Article  Google Scholar 

  16. Ganghoffer, J.-F., Goda, I.: A combined accretion and surface growth model in the framework of irreversible thermodynamics. Int. J. Eng. Sci. 127, 53–79 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  17. Garikipati, K., Arruda, E.M., Grosh, K., Narayanan, H., Calve, S.: A continuum treatment of growth in biological tissue: the coupling of mass transport and mechanics. J. Mech. Phys. Solids 52(7), 1595–1625 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  18. Ghosh, S., Choi, J.: Three-dimensional transient finite element analysis for residual stresses in the laser aided direct metal/material deposition process. J. Laser Appl. 17(3), 144–158 (2005)

    Article  Google Scholar 

  19. Gibson, I., Rosen, D., Stucker, B.: Additive Manufacturing Technologies: 3D Printing, Rapid Prototyping, and Direct Digital Manufacturing. Springer, New York (2014)

    Google Scholar 

  20. Goriely, A.: The Mathematics and Mechanics of Biological Growth, vol. 45. Springer, New York (2017)

    Book  MATH  Google Scholar 

  21. Gu, D., Meiners, W., Wissenbach, K., Poprawe, R.: Laser additive manufacturing of metallic components: materials, processes and mechanisms. Int. Mater. Rev. 57(3), 133–164 (2012)

    Article  Google Scholar 

  22. Hodge, N., Ferencz, R., Solberg, J.: Implementation of a thermomechanical model for the simulation of selective laser melting. Comput. Mech. 54(1), 33–51 (2014)

    Article  MathSciNet  Google Scholar 

  23. Hodge, N., Ferencz, R., Vignes, R.: Experimental comparison of residual stresses for a thermomechanical model for the simulation of selective laser melting. Addit. Manuf. 12, 159–168 (2016)

    Google Scholar 

  24. Holzapfel, G., Simo, J.: Entropy elasticity of isotropic rubber-like solids at finite strains. Comput. Methods Appl. Mech. Eng. 132(1), 17–44 (1996)

    Article  MathSciNet  MATH  Google Scholar 

  25. Horn, T.J., Harrysson, O.L.: Overview of current additive manufacturing technologies and selected applications. Sci. Prog. 95(3), 255–282 (2012)

    Article  Google Scholar 

  26. Hu, H., Argyropoulos, S.A.: Mathematical modelling of solidification and melting: a review. Modell. Simul. Mater. Sci. Eng. 4(4), 371 (1996)

    Article  Google Scholar 

  27. Huang, S.H., Liu, P., Mokasdar, A., Hou, L.: Additive manufacturing and its societal impact: a literature review. Int. J. Adv. Manuf. Technol. 67, 1–13 (2013)

    Article  Google Scholar 

  28. Klingbeil, N.W., Beuth, J.L., Chin, R., Amon, C.: Residual stress-induced warping in direct metal solid freeform fabrication. Int. J. Mech. Sci. 44(1), 57–77 (2002)

    Article  MATH  Google Scholar 

  29. Levy, G.N., Schindel, R., Kruth, J.-P.: Rapid manufacturing and rapid tooling with layer manufacturing (lm) technologies, state of the art and future perspectives. CIRP Ann. Manuf. Technol. 52(2), 589–609 (2003)

    Article  Google Scholar 

  30. Li, C., Fu, C., Guo, Y., Fang, F.: A multiscale modeling approach for fast prediction of part distortion in selective laser melting. J. Mater. Process. Technol. 229, 703–712 (2016)

    Article  Google Scholar 

  31. Li, Y., Gu, D.: Thermal behavior during selective laser melting of commercially pure titanium powder: numerical simulation and experimental study. Addit. Manuf. 1, 99–109 (2014)

    Google Scholar 

  32. Loh, L.-E., Chua, C.-K., Yeong, W.-Y., Song, J., Mapar, M., Sing, S.-L., Liu, Z.-H., Zhang, D.-Q.: Numerical investigation and an effective modelling on the selective laser melting (SLM) process with aluminium alloy 6061. Int. J. Heat Mass Transf. 80, 288–300 (2015)

    Article  Google Scholar 

  33. Marsden, J., Hughes, T.: Mathematical Foundations of Elasticity. Dover Civil and Mechanical Engineering Series. Dover, Mineola (1983)

    Google Scholar 

  34. Matsumoto, M., Shiomi, M., Osakada, K., Abe, F.: Finite element analysis of single layer forming on metallic powder bed in rapid prototyping by selective laser processing. Int. J. Mach. Tools Manuf. 42(1), 61–67 (2002)

    Article  Google Scholar 

  35. Mercelis, P., Kruth, J.-P.: Residual stresses in selective laser sintering and selective laser melting. Rapid Prototyp. J. 12(5), 254–265 (2006)

    Article  Google Scholar 

  36. Michaleris, P.: Modeling metal deposition in heat transfer analyses of additive manufacturing processes. Finite Elem. Anal. Des. 86, 51–60 (2014)

    Article  Google Scholar 

  37. Ogden, R.: Large deformation isotropic elasticity: on the correlation of theory and experiment for compressible rubberlike solids. Proc. R. Soc. Lond. A 328(1575), 567–583 (1972a)

    Article  MATH  Google Scholar 

  38. Ogden, R.: Large deformation isotropic elasticity: on the correlation of theory and experiment for incompressible rubberlike solids. Proc. R. Soc. Lond. A Math. Phys. Sci. 326(1567), 565–584 (1972b)

    MATH  Google Scholar 

  39. Ogden, R.: On the thermoelastic modeling of rubberlike solids. J. Therm. Stress. 15(4), 533–557 (1992)

    Article  MathSciNet  Google Scholar 

  40. Ozakin, A., Yavari, A.: A geometric theory of thermal stresses. J. Math. Phys. 51(3), 032902 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  41. Richmond, O., Tien, R.: Theory of thermal stresses and air-gap formation during the early stages of solidification in a rectangular mold. J. Mech. Phys. Solids 19(5), 273–284 (1971)

    Article  Google Scholar 

  42. Rifkin, J.: The Third Industrial Revolution: How Lateral Power Is Transforming Energy, the Economy, and the World. St. Martin’s Press, New York (2011)

    Google Scholar 

  43. Sadik, S., Yavari, A.: Geometric nonlinear thermoelasticity and the time evolution of thermal stresses. Math. Mech. Solids 22(7), 1546–1587 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  44. Schwerdtfeger, K., Sato, M., Tacke, K.-H.: Stress formation in solidifying bodies. Solidification in a round continuous casting mold. Metall. Mater. Trans. B 29(5), 1057–1068 (1998)

    Article  Google Scholar 

  45. Shamsaei, N., Yadollahi, A., Bian, L., Thompson, S.M.: An overview of direct laser deposition for additive manufacturing; part ii: mechanical behavior, process parameter optimization and control. Addit. Manuf. 8, 12–35 (2015)

    Article  Google Scholar 

  46. Shiomi, M., Osakada, K., Nakamura, K., Yamashita, T., Abe, F.: Residual stress within metallic model made by selective laser melting process. CIRP Ann. Manuf. Technol. 53(1), 195–198 (2004)

    Article  Google Scholar 

  47. Shu, C.: Differential Quadrature and Its Application in Engineering. Springer, New York (2012)

    Google Scholar 

  48. Sozio, F., Yavari, A.: Nonlinear mechanics of surface growth for cylindrical and spherical elastic bodies. J. Mech. Phys. Solids 98, 12–48 (2017)

    Article  MathSciNet  Google Scholar 

  49. Sozio, F., Yavari, A.: Nonlinear mechanics of accretion. J. Nonlinear Sci. 29(4), 1813–1863 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  50. Sozio, F., Yavari, A.: Riemannian and Euclidean material structures in anelasticty. Math. Mech. Solids (2019). https://doi.org/10.1177/1081286519884719

    Article  Google Scholar 

  51. Stefan, J.: Über einge problem der theoric der wärmeleitung. Sitzber. Wien Akad. Mat. Nat.r 98, 173–484 (1989)

    Google Scholar 

  52. Swain, D., Gupta, A.: Biological growth in bodies with incoherent interfaces. Proc. R. Soc. A Math. Phys. Eng. Sci. 474(2209), 20170716 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  53. Thompson, S.M., Bian, L., Shamsaei, N., Yadollahi, A.: An overview of direct laser deposition for additive manufacturing; part i: transport phenomena, modeling and diagnostics. Addit. Manuf. 8, 36–62 (2015)

    Article  Google Scholar 

  54. Tomassetti, G., Cohen, T., Abeyaratne, R.: Steady accretion of an elastic body on a hard spherical surface and the notion of a four-dimensional reference space. J. Mech. Phys. Solids 96, 333–352 (2016)

    Article  MathSciNet  Google Scholar 

  55. Truesdell, C., Noll, W.: The Non-Linear Field Theories of Mechanics, 3rd edn. Springer, New York (2004)

    Book  MATH  Google Scholar 

  56. Truskinovsky, L., Zurlo, G.: Nonlinear elasticity of incompatible surface growth. Phys. Rev. E 99(5), 053001 (2019)

    Article  MathSciNet  Google Scholar 

  57. Viskanta, R.: Heat transfer during melting and solidification of metals. ASME Trans. J. Heat Transf. 110, 1205–1219 (1988)

    Article  Google Scholar 

  58. Weller, C., Kleer, R., Piller, F.T.: Economic implications of 3d printing: market structure models in light of additive manufacturing revisited. Int. J. Prod. Econ. 164, 43–56 (2015)

    Article  Google Scholar 

  59. Yavari, A.: A geometric theory of growth mechanics. J. Nonlinear Sci. 20(6), 781–830 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  60. Zaeh, M.F., Branner, G.: Investigations on residual stresses and deformations in selective laser melting. Prod. Eng. Res. Dev. 4(1), 35–45 (2010)

    Article  Google Scholar 

Download references

Acknowledgements

This research was supported by NSF-Grant Nos. CMMI 1130856 and ARO W911NF-18-1-0003.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Arash Yavari.

Additional information

Publisher's Note

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

Appendices

A A nonlinear thermoelastic constitutive model

In this appendix, we present a thermoelastic model following the works of Chadwick [3], Ogden [37,38,39], and Holzapfel and Simo [24]. See [43] for more details. For a homogeneous isotropic solid, we denote by \(\kappa _0\), \(\mu _0\) and \(\beta _0\), the bulk modulus, the shear modulus, and the volumetric coefficient of thermal expansion at \(T_0\), respectively, with \(\beta (X,T) = \frac{\partial }{\partial T}{\text {tr}} \varvec{\omega }(X,T)\), while in the isotropic case \(\beta (X,T) = 3\frac{\partial }{\partial T}\omega (X,T) = 3 \alpha (X,T)\) in 3D, and \(\beta (X,T) = 2\frac{\partial }{\partial T}\omega (X,T) = 2 \alpha (X,T)\) in 2D. We consider the following constitutive model

$$\begin{aligned} \bar{{\mathcal {W}}}(T,\tilde{I},J)= \frac{\mu _0}{2}\frac{T}{T_0} (\tilde{I}-3)+\frac{\kappa _0}{2}\frac{T}{T_0}(J-1)^2 -{\kappa _0}\beta _0(J-1)\left( T-T_0\right) -\rho \int \limits _{T_0}^Tc_E(\eta )\frac{T-\eta }{\eta } {{\mathrm {d}}}\eta , \end{aligned}$$
(51)

where \(\tilde{I}=J^{-2/3}I\), \(I={\text {tr}}{\varvec{C}}\), \({J} = \sqrt{\det {\varvec{C}}}\), and \(c_E\) is the specific heat capacity at constant strain. In the incompressible case, we have the constraint \(J=1\) associated with the pressure field p as the Lagrange multiplier and the constitutive model transforms to read

$$\begin{aligned} \bar{{\mathcal {W}}}(T,{I})=\frac{\mu _0}{2}\frac{T}{T_0}({I}-3)-\rho \int \limits _{T_0}^Tc_E(\eta )\frac{T-\eta }{\eta } {{\mathrm {d}}}\eta , \end{aligned}$$
(52)

plus the Lagrange multiplier part \(p(J-1)^2\). Note that the shear modulus is linear in temperature, i.e.,

$$\begin{aligned} \mu (T)=\mu _0 \frac{T}{T_0} . \end{aligned}$$
(53)

The simplest thermoelastic model for the expansion coefficients is given by a constant \(\alpha \), viz.

$$\begin{aligned} \alpha (T)= \alpha _0,~~~~\omega (T) = \omega _0 \frac{T}{T_0},~~~~\omega _0=\alpha _0 T_0 . \end{aligned}$$
(54)

B A remark on the material metric

Equation (2) is a generalization of the isotropic case that appeared in [40]. In section 2.1 of [43], due to a mis-manipulation of the musical operators for raising and lowering indices, the representation of the material metric using the \(({}^{1}_{1})\)-tensor \(\varvec{\omega }\) was mistakenly presented as

$$\begin{aligned} {\varvec{G}}(X,T)= {\varvec{G}}_0(X) e^{2\varvec{\omega }(X,T)}. \end{aligned}$$
(55)

Indeed, this representation may violate the symmetry requirement for the Riemannian metric \({\varvec{G}}\), while the representation (2) ensures its symmetry. In what follow, we provide a correction of the proof appearing in section 2.1 of [43]. The manifold \(({\mathcal {B}},{\varvec{G}}_0)\)—which corresponds to the stress-free temperature field \(T_0=T_0(X)\)—is flat. Hence, there exists a local coordinate chart \(\{Y^A\}\) in which

$$\begin{aligned} {\varvec{G}}_0=\delta _{AB}{\mathrm {dY}}^A\otimes {\mathrm {dY}}^B. \end{aligned}$$
(56)

Following Ozakin and Yavari [40], the temperature-dependent material metric can be written as

$$\begin{aligned} {\varvec{G}}(X,T)=\sum _K e^{2\omega _K(X,T)}{\mathrm {dY}}^K\otimes {\mathrm {dY}}^K, \end{aligned}$$
(57)

where \(\{\omega _{A}\}_{A=1,2,3}\) describes the thermal expansion properties of the material such that \(\omega _K\) is related to the thermal expansion coefficient \(\alpha _K\) in the direction \(\frac{\partial }{\partial Y^K}\) by

$$\begin{aligned} \alpha _K(X,T)=\frac{\partial \omega _K}{\partial T}(X,T), \end{aligned}$$
(58)

and \(\omega _K(X,T_0)=0\). Let the change of basis between \(\{Y^A\}_{A=1,2,3}\) and some arbitrary local coordinate chart \(\{X^A\}_{A=1,2,3}\) be written as

$$\begin{aligned} {\mathrm {dY}}^K=A^K{}_J{\mathrm {dX}}^J, \end{aligned}$$
(59)

which also reads as

$$\begin{aligned} \frac{\partial }{\partial Y^K}=(A^{-1})^I{}_K\frac{\partial }{\partial X^I}. \end{aligned}$$
(60)

Then it follows that

$$\begin{aligned} {\varvec{G}}=\left( \sum _K e^{2\omega _K} A^K{}_I A^K{}_J\right) {\mathrm {dX}}^I\otimes {\mathrm {dX}}^J. \end{aligned}$$
(61)

Let \(\varvec{\omega }\) be the \(({}^{1}_{1})\)-tensor

$$\begin{aligned} \varvec{\omega }=\sum _K \omega _K \frac{\partial }{\partial Y^K}\otimes {\mathrm {dY}}^K = \sum _K (A^{-1})^I{}_K {\omega _K} A^K{}_J \frac{\partial }{\partial X^I} \otimes {\mathrm {dX}}^J. \end{aligned}$$
(62)

However, in \(\{ X^A\}\), one has

$$\begin{aligned} {\varvec{G}}_0= \delta _{AB} A^A{}_I A^B{}_J {\mathrm {dX}}^I\otimes {\mathrm {dX}}^J. \end{aligned}$$
(63)

Hence, the material metric transforms to

$$\begin{aligned} {\varvec{G}}(X,T)= e^{\omega ^L{}_I} (G_0)_{LM}~\!e^{\omega ^M{}_J} {\mathrm {dX}}^I\otimes {\mathrm {dX}}^J, \end{aligned}$$
(64)

where \((G_0)_{LM}\) are the components of \({\varvec{G}}_0\) in \(\{ X^A\}\). This is the coordinate representation of (2). Note that even though the general representation for the material metric used in [43] had this symmetry fallacy, the results of the paper and the examples were not affected and remain valid.

C Christoffel symbols for the accreting cylinder

First note that denoting with \((\Gamma _0)^A{}_{BC}\) the Christoffel symbols relative to \({\varvec{G}}_0\) as in (6), one obtains

$$\begin{aligned} \Gamma ^A{}_{BC} = (\Gamma _0)^A{}_{BC} + (G_0)^{AD} [ (G_0)_{DB} \omega _{,C} + (G_0)_{DC} \omega _{,B} - (G_0)_{BC} \omega _{,D} ] . \end{aligned}$$
(65)

Therefore, as \({\varvec{G}}_0^{\sharp }:{\varvec{G}}_0={\text {dim}}{\mathcal {B}}\), one has

$$\begin{aligned} \Gamma ^B{}_{BC} = (\Gamma _0)^B{}_{BC} + ({\text {dim}}{\mathcal {B}})~\omega _{,C} . \end{aligned}$$
(66)

Under the assumption of a uniform material one has \(\omega _{,C} = \frac{\partial \omega }{\partial T} \,T_{,C} = \alpha \,T_{,C}\), i.e., \({\mathrm {d}}\omega = \alpha \,{\mathrm {d}}T\), so one writes

$$\begin{aligned} \Gamma ^B{}_{BC} = (\Gamma _0)^B{}_{BC} + ({\text {dim}}{\mathcal {B}}) \alpha \,T_{,C} . \end{aligned}$$
(67)

Now we compute the Christoffel symbols for the material metric \({\varvec{G}}\) of (25). The only nonzero Christoffel symbols are

$$\begin{aligned} \begin{aligned} \Gamma ^R{}_{RR}&= T_{,R}\alpha (T),\\ \Gamma ^R{}_{\Theta \Theta }&= -{\bar{r}}(R) \left[ {\bar{r}}'(R) + {\bar{r}}(R) T_{,R} \alpha (T) \right] ,\\ \Gamma ^\Theta {}_{R\Theta }&= \frac{{\bar{r}}'(R)}{{\bar{r}}(R)} + T_{,R}\alpha (T). \end{aligned} \end{aligned}$$
(68)

Now one can compute the material divergence \({\text {Div}}[K(T) {\varvec{G}}^\sharp {\mathrm {dT}}]\) for the heat equation as

$$\begin{aligned} \begin{aligned} {\text {Div}}[K(T) {\varvec{G}}^\sharp {\mathrm {dT}}]&= \left[ K(T) G^{AB} T_{,B} \right] _{|A}\\&= \left[ K(T) G^{AB}\right] _{|A} T_{,B} + K(T) G^{AB} \left[ T_{,B}\right] _{|A}\\&= \left[ K(T)\right] _{|A} G^{AB} T_{,B} + K(T) G^{AB} \left[ T_{,AB} - \Gamma ^C{}_{AB} T_{,C}\right] \\&= \frac{{\mathrm {dK}}}{{\mathrm {dT}}} T_{,R}{}^2 G^{RR} + K(T) \left[ T_{,RR}G^{RR} -\left( \Gamma ^R{}_{RR}G^{RR}+\Gamma ^R{}_{\Theta \Theta }G^{\Theta \Theta } \right) T_{,R} \right] \\&= \left[ \frac{{\mathrm {dK}}}{{\mathrm {dT}}} T_{,R}{}^2 + \frac{K(T)}{{\bar{r}}(R)}\frac{\partial }{\partial R}\left( {\bar{r}}(R)\frac{\partial T}{\partial R}\right) \right] e^{-2\omega (T)}. \end{aligned} \end{aligned}$$
(69)

D The reduced form of the Clausius–Duhem inequality in thermoelastic accretion

In this appendix, we discuss the restrictions that the second law of thermodynamics imposes on constitutive equations. In particular, we correct a mistake in [43],Footnote 11 which fortunately did not affect any of the results or conclusions of that work. The localized form of the Clausius-Duhem inequality reads

$$\begin{aligned} \rho \dot{{\mathcal {N}}}\ge \rho \frac{R}{T}-\text {Div}\left( \frac{{\varvec{H}}}{T}\right) +\rho \frac{\partial {\mathcal {N}}}{\partial {\varvec{G}}}\!:\!\dot{{\varvec{G}}}, \end{aligned}$$
(70)

where \({\mathcal {N}}={\mathcal {N}}(X,T,{\varvec{C}}^\flat ,{\varvec{G}})\) is the specific entropy. Expanding Eq. (70) and multiplying by \(T>0\) one obtains

$$\begin{aligned} \rho T \left( \frac{\partial {\mathcal {N}}}{\partial T}{\dot{T}} +\frac{\partial {\mathcal {N}}}{\partial {\varvec{C}}^{\flat }}\!:\!\dot{{\varvec{C}}}^{\flat } \right) \ge \rho R-\text {Div}{\varvec{H}} +\frac{1}{T}\left<{\mathrm {d}}T,{\varvec{H}}\right> , \end{aligned}$$
(71)

where in a local coordinate chart \(\{X^A\}\), the 1-form \({\mathrm {d}}T\) has the representation \({\mathrm {d}}T=\frac{\partial T}{\partial X^A}dX^A\). The specific free energy function has the form \(\Psi =\Psi (X,T,{\varvec{C}}^\flat ,{\varvec{G}})\). The internal energy is defined as the Legendre transform of the free energy with respect to the conjugate variables T and \({\mathcal {N}}\), i.e., \({\mathcal {E}}=T{\mathcal {N}}+\Psi \), and hence, \({\mathcal {E}}={\mathcal {E}}(X,{\mathcal {N}},{\varvec{C}}^\flat ,{\varvec{G}})\,\). ThereforeFootnote 12

$$\begin{aligned} \frac{\partial {\mathcal {E}}}{\partial {\varvec{G}}}=T\frac{\partial {\mathcal {N}}}{\partial {\varvec{G}}} +\frac{\partial \Psi }{\partial {\varvec{G}}} ,\quad \frac{\partial {\mathcal {E}}}{\partial {\varvec{C}}^\flat }=T\frac{\partial {\mathcal {N}}}{\partial {\varvec{C}}^\flat } +\frac{\partial \Psi }{\partial {\varvec{C}}^\flat }. \end{aligned}$$
(72)

The localized balance of energy reads [43]

$$\begin{aligned} \rho \dot{{\mathcal {E}}}={\varvec{S}}\!:\!{\varvec{D}}-\text {Div}\,{\varvec{Q}}+\rho R +\rho \frac{\partial {\mathcal {E}}}{\partial {\varvec{G}}}\!:\!\dot{{\varvec{G}}}. \end{aligned}$$
(73)

This can be rewritten in terms of the specific entropy as

$$\begin{aligned} \rho \left( {\mathcal {N}}+\frac{\partial \Psi }{\partial T}\right) {\dot{T}}+\rho T \frac{\partial {\mathcal {N}}}{\partial T} {\dot{T}} +\rho T \frac{\partial {\mathcal {N}}}{\partial {\varvec{C}}^\flat }\!:\!\dot{{\varvec{C}}}^{\flat } +\left( \rho \frac{\partial \Psi }{\partial {\varvec{C}}^\flat }-\frac{1}{2}{\mathbf {S}} \right) \!:\!\dot{{\varvec{C}}}^{\flat } =\rho R-\text {Div}\,{\varvec{H}}. \end{aligned}$$
(74)

Substituting (74) into (71) one obtains

$$\begin{aligned} \rho \left( {\mathcal {N}}+\frac{\partial \Psi }{\partial T}\right) {\dot{T}} +\left( \rho \frac{\partial \Psi }{\partial {\varvec{C}}^\flat } -\frac{1}{2}{\mathbf {S}} \right) \!:\!\dot{{\varvec{C}}}^{\flat } +\frac{1}{T}\left<{\mathrm {d}}T,{\varvec{H}}\right> \le 0. \end{aligned}$$
(75)

This inequality must hold for all deformations \(\varphi \) and metrics \({\varvec{G}}\,\). ThereforeFootnote 13

$$\begin{aligned} {\mathcal {N}}=-\frac{\partial \Psi }{\partial T},~~~{\mathbf {S}}=2\rho \frac{\partial \Psi }{\partial {\varvec{C}}^\flat } , ~~~ \left<{\mathrm {d}}T,{\varvec{H}}\right> \le 0 . \end{aligned}$$
(76)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Sozio, F., Faghih Shojaei, M., Sadik, S. et al. Nonlinear mechanics of thermoelastic accretion. Z. Angew. Math. Phys. 71, 87 (2020). https://doi.org/10.1007/s00033-020-01309-5

Download citation

  • Received:

  • Revised:

  • Published:

  • DOI: https://doi.org/10.1007/s00033-020-01309-5

Keywords

Mathematics Subject Classification

Navigation