Skip to main content
Log in

Predicting the dislocation nucleation rate as a function of temperature and stress

  • Invited Feature Paper
  • Published:
Journal of Materials Research Aims and scope Submit manuscript

Abstract

Predicting the dislocation nucleation rate as a function of temperature and stress is crucial for understanding the plastic deformation of nanoscale crystalline materials. However, the limited time scale of molecular dynamics simulations makes it very difficult to predict the dislocation nucleation rate at experimentally relevant conditions. We recently develop an approach to predict the dislocation nucleation rate based on the Becker–Döring theory of nucleation and umbrella sampling simulations. The results reveal very large activation entropies, which originated from the anharmonic effects, that can alter the nucleation rate by many orders of magnitude. Here we discuss the thermodynamics and algorithms underlying these calculations in greater detail. In particular, we prove that the activation Helmholtz free energy equals the activation Gibbs free energy in the thermodynamic limit and explain the large difference in the activation entropies in the constant stress and constant strain ensembles. We also discuss the origin of the large activation entropies for dislocation nucleation, along with previous theoretical estimates of the activation entropy.

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.

Similar content being viewed by others

References

  1. X. Li, Y. Wei, L. Lu, K. Lu, and H. Gao: Dislocation nucleation governed softening and maximum strength in nano-twinned metals. Nature 464, 877 (2010).

    Article  CAS  Google Scholar 

  2. J. Li: The mechanics and physics of defect nucleation. MRS Bull. 32, 151 (2007).

    Article  CAS  Google Scholar 

  3. T. Zhu, J. Li, S. Ogata, and S. Yip: Mechanics of ultra-strength materials. MRS Bull. 34, 167 (2009).

    Article  CAS  Google Scholar 

  4. J. Li, K.J. Van Vliet, T. Zhu, S. Yip, and S. Suresh: Atomistic mechanisms governing elastic limit and incipient plasticity in crystals. Nature 418, 307 (2002).

    Article  CAS  Google Scholar 

  5. C.A. Schuh, J.K. Mason, and A.C. Lund: Quantitative insight into dislocation nucleation from high-temperature nanoindentation experiments. Nature Mater. 4, 617 (2005).

    Article  CAS  Google Scholar 

  6. P. Schall, I. Cohen, D.A. Weitz, and F. Spaepen: Visualizing dislocation nucleation by indenting colloidal crystals. Nature 440, 319 (2006).

    Article  CAS  Google Scholar 

  7. W.D. Nix, J.R. Greer, G. Feng, and E.T. Lilleodden: Deformation at the nanometer and micrometer length scales: Effects of strain gradients and dislocation starvation. Thin Solid Films 515, 315 (2007).

    Article  CAS  Google Scholar 

  8. S. Izumi, H. Ohta, C. Takahashi, T. Suzuki, and H. Saka: Shuffle-set dislocation nucleation in semiconductor silicon device. Philos. Mag. Lett. 90, 707 (2010).

    Article  CAS  Google Scholar 

  9. G. Xu, A.S. Argon, and M. Ortiz: Critical configurations for dislocation nucleation from crack tips. Philos. Mag. A 75, 341 (1997).

    Article  CAS  Google Scholar 

  10. F. Frank: Symposium on the Plastic Deformation of Crystalline Solids; Carnegie Institute of Technology and Office of Naval Research, Pittsburgh, PA, 1950; p. 89.

    Google Scholar 

  11. S. Aubry, K. Kang, S. Ryu, and W. Cai: Energy barrier for homogeneous dislocation nucleation: Comparing atomistic and continuum models. Scripta Mater. 64, 1043 (2011).

    Article  CAS  Google Scholar 

  12. M.A. Tschopp, D.E. Spearot, and D.L. McDowell: Atomistic simulations of homogeneous dislocation nucleation in single crystal copper. Model. Simul. Mater. Sci. Eng. 15, 693 (2007).

    Article  CAS  Google Scholar 

  13. E.M. Bringa, K. Rosolankova, R.E. Rudd, B.A. Remington, J.S. Wark, M. Duchaineau, D.H. Kalantar, J. Hawreliak, and J. Belak: Shock deformation of face-centered-cubic metals on subnanosecond timescales. Nat. Mater. 5, 805 (2006).

    Article  CAS  Google Scholar 

  14. T. Zhu, J. Li, A. Samanta, A. Leach, and K. Gall: Temperature and strain-rate dependence of surface dislocation nucleation. Phys. Rev. Lett. 100, 025502 (2008).

    Article  CAS  Google Scholar 

  15. P. Hanggi, P. Talkner, and M. Borkovec: Reaction-rate theory: Fifty years after Kramers. Rev. Mod. Phys. 62, 251 (1990).

    Article  Google Scholar 

  16. R. Becker and W. Döring: The kinetic treatment of nuclear formation in supersaturated vapors. Ann. Phys. (Weinheim) 24, 719 (1935).

    CAS  Google Scholar 

  17. H. Eyring: The activated complex in chemical reactions. J. Chem. Phys. 3, 107 (1935).

    Article  CAS  Google Scholar 

  18. G.H. Vineyard: Frequency factors and isotope effects in solid state rate processes. J. Phys. Chem. Solids 3, 121 (1957).

    Article  CAS  Google Scholar 

  19. H. Jónsson, G. Mills, and K.W. Jacobsen: Nudged elastic band method for finding minimum energy paths of transitions. In Classical and Quantum Dynamics in Condensed Phase Simulations; B.J. Berne, G. Ciccotti, and D.F. Coker, Eds.; World Scientific: New York; 1998; pp. 385–404.

  20. A.F. Voter: Introduction to the kinetic Monte Carlo Metho; Springer: New York; 2007.

    Google Scholar 

  21. C. Jin, W. Ren, and Y. Xiang: Computing transition rates of thermally activated events in dislocation dynamics. Script. Mater. 62, 206 (2010).

    Article  CAS  Google Scholar 

  22. D. Chandler: Introduction to Modern Statistical Mechanics; Oxford University Press: Oxford, UK; 1987.

    Google Scholar 

  23. W. E. Ren and E. Vanden-Eijnden: String method for the study of rare events. Phys. Rev. B 66, 052301 (2002).

    Google Scholar 

  24. W. E. Ren and E. Vanden-Eijnden: Finite temperature string method for the study of rare events. J. Phys. Chem. B 109, 6688 (2005).

    Article  CAS  Google Scholar 

  25. S. Ryu, K. Kang, and W. Cai: Entropic effect on the rate of dislocation nucleation. Proc. Natl. Acad. Sci. USA 108, 5174 (2011).

    Article  CAS  Google Scholar 

  26. D. Frenkel and B. Smit: Understanding Molecular Simulation: From Algorithms to Applications, Academic Press: San Diego, CA; 2002.

    Google Scholar 

  27. U.F. Kocks, A.S. Argon, and M.F. Ashby: Thermodynamics and kinetics of slip. Prog. Mater. Sci. 19, 1 (1975).

    Article  Google Scholar 

  28. J.F. Nye: physical Properties of Crystals: Their Representation by Tensors and Matrices; Clarendon Press: New York; 1957.

    Google Scholar 

  29. H. Xiao, O.T. Bruhns, and A. Meyers: Logarithmic strain, logarithmic spin and logarithmic rate. Acta Mech. 124, 89105 (1997).

    Article  Google Scholar 

  30. R. Bechmann, A.D. Ballato, and T.J. Lukaszek: Higher-order temperature coefficients of the elastic stiffinesses and complinaces of alpha-quartz. In Proceedings of the Institute of Radio Engineers, Vol. 50; 1962; p. 1812.

    Google Scholar 

  31. E. Whalley: Use of volumes of activation for determining reaction mechanisms. In Advances in Physical Organic Chemistry; V. Gold, Ed.; Academic Press: London; 1964; pp. 93–162.

  32. M.L. Tonnet and E. Whalley: Effect of pressure on the alkaline hydrolysis of ethyl acetate in acetone–water solutions. Parameters of activation at constant volume. Can. J. Chem. 53, 3414 (1975).

    Article  CAS  Google Scholar 

  33. W.C. Overton Jr. and J. Gaffney: Temperature variation of the elastic constants of cubic elements. I. Copper. Phys. Rev. 98, 969 (1955).

    Article  CAS  Google Scholar 

  34. J.W. Cahn and F.R.N Nabarro: Thermal activation under shear. Philos. Mag. A 81, 1409 (2001).

    Article  CAS  Google Scholar 

  35. A.H. Cottrell: Thermally activated plastic glide. Philos. Mag. Lett. 82, 65 (2002).

    Article  CAS  Google Scholar 

  36. Y. Mishin, M.R. Sorensen, and A.F. Voter: Calculation of point-defect entropy in metals. Philos. Mag. A 81, 2591 (2001).

    Article  CAS  Google Scholar 

  37. D. Gupta, Ed.: Diffusion Processes in Advanced Technological Materials; Springer: New York; 2002; p.140.

    Google Scholar 

  38. C. Kittel: Introduction to Solid State Physics, 8th ed.; Wiley: New York; 2004.

    Google Scholar 

  39. A.S. Argon, R.D. Andrews, J.A. Godrick, and W. Whitney: Plastic deformation bands in glassy polystyrene. J. Appl. Phys. 39, 1899 (1968).

    Article  CAS  Google Scholar 

  40. R.J. DiMelfi, W.D. Nix, D.M. Barnett, J.H. Holbrook, and G.M. Pound: An analysis of the entropy of thermally activated dislocation motion based on the theory of thermoelasticity. Phys. Status Solidi B 75, 573 (1976).

    Article  CAS  Google Scholar 

  41. R.J. DiMelfi, W.D. Nix, D.M. Barnett, and G.M. Pound: The equivalence of two methods for computing the activation entropy for dislocation motion. Acta Mater. 28, 231 (1980).

    Article  Google Scholar 

  42. G. Kemeny and B. Rosenberg: Compensation law in thermodynamics and thermal death. Nature 243, 400 (1973).

    Article  Google Scholar 

  43. A. Yelon, M. Movagha, and H.M. Branz: Origin and consequences of the compensation (Meyer–Neldel) law. Phys. Rev. B 46, 12243 (1992).

    Article  Google Scholar 

  44. M. Born: Thermodynamics of crystals and melting. J. Chem. Phys. 7, 591 (1939).

    Article  CAS  Google Scholar 

  45. S. Brochard, P. Hirel, L. Pizzagalli, and J. Godet: Elastic limit for surface step dislocation nucleation in face-centered cubic metals: Temperature and step height dependence. Acta Mater. 58, 4182 (2010).

    Article  CAS  Google Scholar 

  46. M. Khantha, D.P. Pope and V. Vitek: Dislocation screening and the brittle-to-ductile transition: A Kosterlitz–Thouless type instability. Phys. Rev. Lett. 74, 684 (1994).

    Article  Google Scholar 

  47. Y. Mishin, M.J. Mehl, D.A. Papaconstantopoulos, A.F. Voter, and J.D. Kress: Structural stability and lattice defects in copper: Ab initio, tight-binding, and embedded-atom calculations. Phys. Rev. B 63, 224106 (2001).

    Article  CAS  Google Scholar 

  48. J.P. Hirth and J. Lothe: Theory of Dislocations; Krieger: New York; 1992.

    Google Scholar 

  49. M. Parrinello and A. Rahman: Crystal structure and pair potentials: A molecular dynamics study. Phys. Rev. Lett. 45, 1196 (1980).

    Article  CAS  Google Scholar 

  50. H.C. Anderson: Molecular dynamics at constant pressure and/or temperature. J. Chem. Phys. 72, 2384 (1980).

    Article  Google Scholar 

  51. W.G. Hoover: Canonical dynamics: Equilibrium phase space distribution. Phys. Rev. A 31, 1695 (1985).

    Article  CAS  Google Scholar 

  52. T. Zhu, J. Li, K.J. Van Vliet, S. Ogata, S. Yip, and S. Suresh: Predictive modeling of nanoindentation-induced homogeneous dislocation nucleation in copper. J. Mech. Phys. Sol. 52, 691 (2004).

    Article  CAS  Google Scholar 

  53. A.H.W Ngan, L. Zuo, and P.C. Wo: Size dependence and stochastic nature of yield strength of micron-sized crystals: A case study on Ni3Al. Proc. Royal Soc. A 462, 1661 (2006).

    Article  CAS  Google Scholar 

  54. S. Ryu and W. Cai: Validity of classical nucleation theory for Ising models. Phys. Rev. E 81, 030601(R) (2010).

    Article  CAS  Google Scholar 

  55. T. Zhu, J. Li, A. Samanta, H.G. Kim, and S. Suresh: Interfacial plasticity governs strain rate sensitivity and ductility in nanostructured metals. Proc. Natl. Acad. Sci. USA 104, 3031 (2007).

    Article  CAS  Google Scholar 

  56. S.M. Foiles: Evaluation of harmonic methods for calculating the free energy of defects in solids. Phys. Rev. B 49, 14930 (1994).

    Article  CAS  Google Scholar 

  57. M. de Koning, C.R. Miranda, and A. Antonelli: Atomistic prediction of equilibrium vacancy concentraions in Ni3Al. Phys. Rev. B 66, 104110 (2002).

    Article  CAS  Google Scholar 

  58. S. Ryu and W. Cai: Comparison of thermal properties predicted by interatomic potential models. Model. Simul. Mater. Sci. Eng. 16, 085005 (2008).

    Article  CAS  Google Scholar 

  59. A. Yelon, B. Movaghar, and R.S. Crandall: Multi-exitation entropy: Its role in thermodynamics and kinetics. Rep. Prog. Phys. 69, 1145 (2006).

    Article  CAS  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Wei Cai.

Additional information

This paper has been selected as an Invited Feature Paper.

Appendices

Appendix A: Equality of Critical Sizes \(n_c^{{\sigma }}\) and \(n_c^{{\gamma }}\)

Suppose that the Gibbs free energy G(n,σ,T)is maximized at \(n_c^{{\sigma }}\), then

$${\left. {\frac{{\partial G\left( {n,{{\sigma }},T} \right)}}{{\partial n}}} \right|_{{{\sigma }},n = n_c^{{\sigma }}}} = 0\quad.$$
((A1))

Note that T is held constant throughout this section. Through the Legendre transform, Eq. (5), we have the following property forthe Helmholtz free energy F(n,γ,T):

$$ {{{\left. {\frac{{\partial F\left( {n,{{\gamma }},T} \right)}}{{\partial n}}} \right|}_{{{\gamma }},n = n_{\text{c}}^{{\sigma }}}}} \;\; = \;\; {{{\left. {\frac{\partial }{{\partial n}}} \right|}_{\gamma,n = n_{\text{c}}^{{\sigma }}}}\left[ {G\left( {n,{{\sigma }},T} \right) + {{\sigma \gamma }}V} \right]} \\ {} \;\; = \;\; {{{\left. {\frac{{\partial G\left( {n,{{\sigma }},T} \right)}}{{\partial n}}} \right|}_{{{\sigma }},n = n_c^{{\sigma }}}}} \\ {} \;\; {} \;\; { + {{\left. {\frac{{\partial G\left( {n,{{\sigma }},T} \right)}}{{\partial {{\sigma }}}}} \right|}_{{{\sigma }},n = n_c^{{\sigma }}}}\frac{{\partial {{\sigma }}}}{{\partial n}} + \frac{{\partial {{\sigma }}}}{{\partial n}}{{\gamma }}V} \\ {} \;\; = \;\; {0 - V{{\gamma }}\frac{{\partial {{\sigma }}}}{{\partial n}} + V{{\gamma }}\frac{{\partial {{\sigma }}}}{{\partial n}} = 0\quad.} \\ $$
((A2))

By definition, F(n,γ,T) reaches maximum at \(n = n_c^{{\gamma }}\) at constant γ and T,

$${\left. {\frac{{\partial F\left( {n,{{\gamma }},T} \right)}}{{\partial n}}} \right|_{{{\gamma }},n = n_{\text{c}}^{{\gamma }}}} = 0\quad.$$
((A3))

Therefore, we establish that \(n_c^{{\gamma }} = n_c^{{\sigma }},\) that is, the maxi-mizer \(n_c^{{\sigma }}\) of G(n,σ,T) is also the maximizer \(n_c^{{\gamma }}\) of F(n,γ,T).

Appendix B: Equality of Activation Gibbs and Helmholtz Free Energies

The activation Gibbs free energy is the free energy difference between state 0, a perfect crystal, and state 1, a crystal containing a critical dislocation loop under a same shear stress σ. Because of the plastic shear deformation caused by the dislocation loop, state 1 has a higher strain (γ) than state 0 (γ0). It has been shown that the maximizer \(n_c^{{\sigma }}\) of G(n,σ,T) equals the maximizer of \(n_c^{{\gamma }}\) of F(n,γ,T) when σ equals σ(nc,γ,T), as defined in Eq. (6). Note that we keep σ to be the stress at ncγ, and T. Then at the same σ, but for n = 0, the strain becomes γ0. Hence, the activation Gibbs free energy barrier can be written as

$$ {{G_c}} \;\; = \;\; {G\left( {{n_{\text{c}}},{{\sigma }},T} \right) - G\left( {0,{{\sigma }},T} \right)} \\ {} \;\; = \;\; {F\left( {{n_{\text{c}}},{{\gamma }},T} \right) - {{\sigma \gamma }}V - F\left( {0,{{{\gamma }}_0},T} \right) + {{\sigma }}{{{\gamma }}_0}V\,\,\,.} \\ $$
((B1))

Note that F(nc,γ,T) and F(0,γ0,T) do not correspond to the same strain state, so that their difference is not the activation Helmholtz free energy. To construct the activation Helmholtz free energy, we subtract and add the F(0,γ,T) term in the right hand side,

$$ {{G_c}} \;\; = \;\; {F\left( {{n_{\text{c}}},{{\gamma }},T} \right) - F\left( {0,{{\gamma }},T} \right) + F\left( {0,{{\gamma }},T} \right) - F\left( {0,{{{\gamma }}_0},T} \right)} \\ {} \;\; {} \;\; { - V{{\sigma }}\left( {{{\gamma }} - {{{\gamma }}_0}} \right) \approx {F_c} + {{\left. {\frac{{\partial F}}{{\partial {{\gamma }}}}} \right|}_{{{{\gamma }}_0},V}}\left( {{{\gamma }} - {{{\gamma }}_0}} \right)} \\ {} \;\; {} \;\; { + \frac{1}{2}{{\left. {\frac{{{\partial ^2}F}}{{\partial {{{\gamma }}^2}}}} \right|}_{{{{\gamma }}_0},V}}{{\left( {{{\gamma }} - {{{\gamma }}_0}} \right)}^2}} \\ {} \;\; = \;\; {{F_c} + \frac{1}{2}{{\left. {\frac{{{\partial ^2}F}}{{\partial {{{\gamma }}^2}}}} \right|}_{{{{\gamma }}_0},V}}{{\left( {{{\gamma }} - {{{\gamma }}_0}} \right)}^2}} \\ {} \;\; = \;\; {{F_c} + \frac{1}{2}V{{\left. {\frac{{{\partial ^2}F}}{{\partial {{{\gamma }}^2}}}} \right|}_{{{{\gamma }}_0},V}}{{\left( {{{\gamma }} - {{{\gamma }}_0}} \right)}^2}\quad.} \\ $$
((B2))

Note that \({{\gamma }}V\,\, = \,\,\, - \partial G\left( {{n_{\text{c}}},{{\sigma,}}T} \right)/\partial {{\sigma }}\) and \({{{\gamma }}_0}V\,\, = \,\,\, - \partial G\left( {{n_{\text{c}}},{{\sigma,}}T} \right)/\partial {{\sigma }}\). Then \(\left( {{{\gamma }}V\,\, - \,\,{{{\gamma }}_0}V} \right)\) is equivalent to \( - \left( {{\partial \mathord{\left/ {\vphantom {\partial {\partial {{\sigma }}}}} \right. } {\partial {{\sigma }}}}} \right)\left( {G\left( {{n_{\text{c}}},{{\sigma,}}T} \right) - G\left( {0,{{\sigma,}}T} \right)} \right) = - \left( {\partial {G_{\text{c}}}/\partial {{\sigma }}} \right) \equiv {\Omega _c}\),that is, the activation volume. By plugging \(\left( {{{\gamma }} - {{{\gamma }}_0}} \right){\Omega _c}/V\) into the equation, we have

$$ {{G_c}} \;\; = \;\; {{F_c} + \frac{1}{2}\frac{1}{V}{{\left. {\frac{{\partial {{\sigma }}}}{{\partial {{\gamma }}}}} \right|}_{{{{\gamma }}_0},V}}{{\left( {{\Omega ^*}} \right)}^2} + O\left( {{V^{ - 2}}} \right)} \\ {} \;\; = \;\; {{F_c} + O\left( {{V^{ - 1}}} \right)\quad.} \\ $$
((B3))

In the thermodynamics limit \(\left( {V \to \infty } \right)\), we have Gc = Fc. Hence, the nucleation rate does not depend on whether the crystal is subjected to constant stress or constant strain loading. The equality allows us to compute the activation Gibbs free energy\({G_c}\left( {{{\sigma,}}T} \right)\) by combining the activation Helmholtz free energy Fc(γ,T) and the stress–strain relations of the perfect crystal shown in Fig. 2(b) and (d).

Appendix C: Physical Interpretation of Activation Entropy Difference ΔSc

It is well known that entropy is a thermodynamic state variable that is independent of the ensemble of choice, that is, \(S\left( {n,{{\gamma }},T} \right) \equiv \partial F{\left. {{{\left( {n,{{\gamma }},T} \right)} \mathord{\left/ {\vphantom {{\left( {n,{{\gamma }},T} \right)} {\partial T}}} \right. } {\partial T}}} \right|_{n{\text{,}\gamma }}}\) and \(S\left( {n,{{\sigma }},T} \right) \equiv \partial G{\left. {{{\left( {n,{{\sigma }},T} \right)} \mathord{\left/ {\vphantom {{\left( {n,{{\sigma }},T} \right)} {\partial T}}} \right. } {\partial T}}} \right|_{n{\text{,}\sigma }}}\) equal to each other as long as \({{\sigma }} = {V^{ - 1}}{\left. {{{\partial F} \mathord{\left/ {\vphantom {{\partial F} {\partial {{\gamma }}}}} \right. } {\partial {{\gamma }}}}} \right|_{n,T}}\). At the same time, the activation entropy is just the entropy difference between the activated state and the metastable state, that is, \({S_c}\left( {{{\gamma,}}T} \right) = S\left( {{n_c},{{\gamma,}}T} \right) - S\left( {0,{{\gamma }},T} \right)\) and \({S_c}\left( {{{\sigma,}}T} \right) = S\left( {{n_c},{{\sigma,}}T} \right) - S\left( {0,{{\sigma }},T} \right)\). If the entropies in two ensembles can equal each other, it may seem puzzling how the activation entropies can be different.

The resolution of this apparent paradox is that under the constant applied stress, the nucleation of a dislocation loop causes a strain increase. Let γ be the strain at the state defined by n = nc, σ, and T, and γ0 be the strain at the state defined by n = 0, σ, and T, then γ > γ0. Hence, we have \(S\left( {{n_c},{{\sigma,}}T} \right) = S\left( {{n_c},{{\gamma,}}T} \right)\) and \(S\left( {{n_c},{{\sigma,}}T} \right) = S\left( {{n_c},{{\gamma,}}T} \right)\), but \(S\left( {0,{{\gamma,}}T} \right) \ne S\left( {0,{{{\gamma }}_0}{\text{,}}T} \right)\).

$$ {{S_c}\left( {{{\sigma }},T} \right)} \;\; = \;\; {S\left( {{n_c}{{,\sigma }},T} \right) - S\left( {0{{,\sigma }},T} \right)} \\ {} \;\; = \;\; {S\left( {{n_c}{{,\gamma }},T} \right) - S\left( {0{\text{,}}{{{\gamma }}_0},T} \right)} \\ {} \;\; = \;\; {S\left( {{n_c}{{,\gamma }},T} \right) - S\left( {0{{,\gamma }},T} \right)} \\ {} \;\; {} \;\; { + S\left( {0{{,\gamma }},T} \right) - S\left( {0{\text{,}}{{{\gamma }}_0},T} \right)} \\ {} \;\; = \;\; {{S_c}\left( {{{\gamma }},T} \right) + S\left( {0{{,\gamma }},T} \right) - S\left( {0{\text{,}}{{{\gamma }}_0},T} \right)\quad.} \\ $$
((C1))

This shows that the activation entropy difference \(\Delta {S_c} \equiv {S_c}\left( {{\sigma }} \right) - {S_c}\left( {{\gamma }} \right)\) equals to \(S\left( {0,{{\gamma,}}T} \right) - S\left( {0,{{{\gamma }}_0}{\text{,}}T} \right)\), which is the entropy difference of the perfect crystal at two slightly different strains.

In the limit of \(V \to \infty \), because we expect \(\left( {{{\gamma }} - {{{\gamma }}_0}} \right) \to 0\), we might reach a false conclusion that \(\Delta {S_c} = \left( {S\left( {{\text{0,}\gamma },T} \right) - S\left( {{\text{0,}}{{{\gamma }}_0},T} \right)} \right) \to 0\). Instead, the correct behavior in the thermodynamic limit can be obtained by expanding ΔSc in a Taylor series,

$$ {s\left( {{\gamma }} \right) - S\left( {{{{\gamma }}_0}} \right)} \;\; = \;\; {\frac{{\partial S}}{{\partial {{\gamma }}}}\left( {{{\gamma }} - {{{\gamma }}_0}} \right) + \cdots } \\ {} \;\; = \;\; {{{\left. { - \frac{{\partial {{\sigma }}}}{{\partial T}}} \right|}_{{{\gamma,}}V}}V\left( {{{\gamma }} - {{{\gamma }}_0}} \right) + \cdots \quad.} \\ $$
((C2))

where the Maxwell relationship \({\left. {{{\partial S} \mathord{\left/ {\vphantom {{\partial S} {\partial {{\gamma }}}}} \right. } {\partial {{\gamma }}}}} \right|_T} = - V{\left. {{{\partial {{\sigma }}} \mathord{\left/ {\vphantom {{\partial {{\sigma }}} {\partial T}}} \right. } {\partial T}}} \right|_{{{\gamma }},V}}\) is used. The term (γV–γ0V) equals the activation volume Ωc and can be interpreted as plastic strain γpl due to the formation of a dislocation loop times the volume of the crystal, that is,

$$\left( {{{\gamma }}V - {{{\gamma }}_0}V} \right) = {\Omega _c} = {{{\gamma }}^<PublisherLocation></PublisherLocation>}V = b{A_{\text{c}}}\quad,$$
((C3))

where b is the magnitude of the Burgers vector and Ac is the area of the critical dislocation loop. Using the relation \(\left( {{{\gamma }} - {{{\gamma }}_0}} \right) = {{{\Omega _c}} \mathord{\left/ {\vphantom {{{\Omega _c}} V}} \right. } V}\), we have

$$\Delta {S_c} = S\left( {{\gamma }} \right) - S\left( {{{{\gamma }}_0}} \right) = - {\left. {\frac{{\partial {{\sigma }}}}{{\partial T}}} \right|_{{{\gamma,}}V}}{\Omega _c} + O\left( {{V^{ - 1}}} \right)\quad,$$
((C4))

which is exactly the same as Eq. (24).

A similar expression has been obtained for the difference between point defect formation entropies under constant pressure (Sp) and under constant volume (Sv),39 with

$${S_{\text{p}}} - {S_{\text{v}}} = {{\beta }}B{V_{{\text{rel}}}}\quad,$$
((C5))

where \({{\beta }} \equiv {V^{ - 1}}{\left. {\left( {{{\partial V} \mathord{\left/ {\vphantom {{\partial V} {\partial T}}} \right. } {\partial T}}} \right)} \right|_{p = 0}}\) is the thermal expansion factor at zero hydrostatic pressure p, B is the isothermal bulk modulus, and Vrel is the relaxation volume of the defect. In Cu, the value of SpSv is estimated to be–1.7 kB for a vacancy and 13.7 kB for an interstitial.36 Comparing Eq. (C5) with Eq. (C4), we note that relaxation volume Vrel for point defects corresponds to the activation volume X for dislocation nucleation and that ßß corresponds to the \({\left( {{{\partial {{\sigma }}} \mathord{\left/ {\vphantom {{\partial {{\sigma }}} {\partial T}}} \right. } {\partial T}}} \right)}\) term. The similarity between these two equations is because they both express the entropy difference between two states, and the choice of the two states depends on whether the stress or the strain is kept constant when the defect is introduced. In contrast, there are also some differences between the physics expressed by these two equations. First, thermal expansion plays a prominent role in Eq. (C5) because it focuses on hydrostatic stress and strain effects. In comparison, thermal expansion does not play a role in Eq. (C4) because it focuses on shear stress and strain effects. Second, the formation entropy of a point defect is the entropy difference between two metastable states and governs equilibrium properties, for example, the density of vacancies at thermal equilibrium. In comparison, the activation entropy is the entropy difference between a saddle (i.e., unstable) state and a metastable state and governs kinetics, such as the dislocation nucleation rate. In addition, the saddle state (i.e., the size of the critical nucleus) depends on stress and temperature, whereas such complexity does not arise in the formation entropy of point defects.

Appendix D: Approximation of \({S_c}\left( {{\sigma }} \right)\)

We now introduce a series of simplifying approximations to estimate the magnitude of \({S_c}\left( {{\sigma }} \right)\) in the low temperature, low stress limit. In the temperature range of 0–300 K, the activation entropy is found to be insensitive to temperature. Starting from Eqs. (24) and (38), we have

$${S_c}\left( {{\sigma }} \right) = {S_c}\left( {{\gamma }} \right) + \Delta {S_c} \approx - \frac{{{E_c}\left( {{\gamma }} \right)}}{{{{\mu }}\left( 0 \right)}}\frac{{\partial {{\mu }}}}{{\partial T}} - {\Omega _c}{\left. {\frac{{\partial {{\sigma }}}}{{\partial T}}} \right|_{{\gamma }}}\quad.$$
((D1))

If we assume the crystal is linear elastic, that is, σ = μγ, then

$${S_c}\left( {{\sigma }} \right) \approx - \frac{{{H_c}\left( {{\sigma }} \right) + {\Omega _c}\left( {{\sigma }} \right)\cdot{{\sigma }}}}{{{{\mu }}\left( 0 \right)}}\frac{{\partial {{\mu }}}}{{\partial T}}\quad.$$
((D2))

Similar expressions can be obtained for normal (com-pressive) loading by replacing γ with e and replacing μ by the Young’s modulus.

To gain more intuition, we note that in the limit of σ →0, the line tension model estimates that \({H_c}\left( {{\sigma }} \right) \propto {{{\sigma }}^{ - 1}}\). In addition, in the limit of \(T \to 0,{\Omega _c}\left( {{\sigma }} \right) \approx {{\partial {H_c}} \mathord{\left/ {\vphantom {{\partial {H_c}} {\partial {{\sigma }}}}} \right. } {\partial {{\sigma }}}}\). Under these conditions, \({\Omega _c}\left( {{\sigma }} \right) \times {{\sigma }} \approx {H_c}\left( {{\sigma }} \right)\), so that

$${S_c}\left( {{\sigma }} \right) \approx - \frac{{2{H_c}\left( {{\sigma }} \right)}}{{{{\mu }}\left( 0 \right)}}\frac{{\partial {{\mu }}}}{{\partial T}}\quad.$$
((D3))

Comparing Eq. (D3) with Eq. (38), we have

$$\frac{{{S_c}\left( {{\sigma }} \right)}}{{{H_c}\left( {{\sigma }} \right)}} \approx 2\frac{{{S_c}\left( {{\gamma }} \right)}}{{{{\text{E}}_c}\left( {{\gamma }} \right)}}\quad.$$
((D4))

This trend is qualitatively observed in heterogeneous nucleation, when comparing Fig. 8(b) and (d), and is less clear in homogeneous nucleation, when comparing Fig. 8(a) and (c). This is probably because the stress-strain relationship is more nonlinearin the case ofhomogeneous nucleation.

Appendix E: Activation Free Energy Data

Table E.I and Table E.II provide the activation free energy data at all temperature and strain conditions in this study so that interested readers can use them as a benchmark. For homogeneous nucleation, the strain γxy is defined as \({{\Delta x} \mathord{\left/ {\vphantom {{\Delta x} {h_y^0}}} \right. } {h_y^0}}\), where Δx is the displacement of the repeat vector initially in the y direction along the x axis at each pure shear stress condition and \({h_y^0}\) is the height of the cell along the y axis at zero temperature without external loading. Shear stress σxy is determined from the xy component of the average Virial stress. For heterogeneous nucleation, we take only the elastic strain into account. The elastic strain ϵzz at T is defined as \({{\left[ {{L_z}\left( {{{\sigma }},T} \right) - {L_z}\left( {{{\sigma = 0,}}T} \right)} \right]} \mathord{\left/ {\vphantom {{\left[ {{L_z}\left( {{{\sigma }},T} \right) - {L_z}\left( {{{\sigma = 0,}}T} \right)} \right]} {L_z^0}}} \right. } {L_z^0}}\), where \({{L_z}\left( {{{\sigma }},T} \right)}\) is the length of the repeat vector along the z axis;\({{L_z}\left( {{{\sigma = 0,}}T} \right)}\) is the equilibrium length at temperature T under zero stress; and \(L_z^0 = 20{a_0} = 72.3{\text{{\AA}}}\) is the reference length (before relaxation), where a0 is the lattice constant of copper. The compressional stress σzz is defined by \({{{\sigma }}_{zz}} = < {\text{F}} >/{d^2}\), where F> is the axial force computed from the zz component of the average Virial stress and d = 15a0 = 54.225 Å is the reference side length of the nanorod.

TABLE E.I.
figure TabE1

Data for homogeneous nucleation: σxy, (Gpa); Ec \({\tilde E_c}\), and Fc (eV); and \(f_c^ + \left( {{{10}^{14}}{{\text{s}}^{ - 1}}} \right)\).

TABLE. E.II.
figure TabE2

Data for heterogeneous nucleation: σzz (Gpa), Ec and Fc (eV), and \(f_c^ + \left( {{{10}^{14}}{{\text{s}}^{ - 1}}} \right)\).

Appendix F: Activation Volume and Critical Loop Size

The activation volume Ωc is defined as the derivative of activation free energy with stress, that is, \({\Omega _c}\left( {{{\sigma }},T} \right) = - {\left. {{{\partial {G_c}} \mathord{\left/ {\vphantom {{\partial {G_c}} {\partial {{\sigma }}}}} \right. } {\partial {{\sigma }}}}} \right|_T}\), and measures the sensitivity of the nucleation rate to the stress. Physically, it is interpreted as plastic strain associated with the dislocation loop times the volume of the crystal, that is, Ωc = bAc,where Ac is theareaof the critical dislocation loop (see Appendix C). Here, we use our numerical data to test the validity of the latter interpretation.

Because the activation volume measures the sensitivity of Gc(σ,T) to applied stress [see Eq. (10)], it must be proportional to the Schmid factor S in uniaxial loading. Hence, the hypothesis we wish to test is

$${\Omega _c} = {n_c}b{A_{\text{a}}}S\quad,$$
((F1))

where b is the magnitude of the Burgers vector and Aa is the average area each atom occupies on the {111} slip plane. Given that the lattice constant of Cu is a0 = 3.615 Å, we have b = a0/√6 = 1.48 Å and \({A_{\text{a}}} = \sqrt 6 {{a_0^2} \mathord{\left/ {\vphantom {{a_0^2} 4}} \right. } 4} = 5.66{{\text{{\AA}}}^2}\), so that bAa = 8.35 Å3. For the pure shear loading in our homogeneous nucleation case, the Schmid factor S = 1. For the uniaxial loading in our heterogeneous nucleation case, S = 0.471.

Figure E1 plots nc versus Ωc for both homogeneous and heterogeneous dislocation nucleations. In both cases, Ωc appears to be roughly linear with nc as expected from Eq. (F1). For homogeneous nucleations under pure shear [Fig. E.I(a)], the slope of the curves is roughly \({\Omega _c}/{n_{\text{c}}} \approx 10{{\text{{\AA}}}^3}\), close to the expected value of 8.35 Å3. For heterogeneous nucleation under compression, [Fig. E1(a)], the slopes of the curves after correction for the Schmid factor is roug hly \({\Omega _c}/\left( {{n_{\text{c}}}S} \right) \approx 8{{\text{{\AA}}}^3}\), which is similar to the case of homogeneous nucleation. Therefore, our data confirm that the activation volume is proportional to the size of the critical dislocation loop. That \({\Omega _c}/\left( {{n_{\text{c}}}S} \right)\) is somewhat smaller than bAa supports the notion that the Burgers vector of a critical dislocation nucleus is smaller than that of a fully formed dislocation.9,11

FIG. E1.
figure E1

The relation between critical dislocation size nc and the activation volume Ω ≡ −(∂Gc/∂σ). (a) Homogeneous nucleation and (b) heterogeneous nucleation. Circles represent the activation volume obtained from the derivative of Gc with respect to σ. Squares represent the activation volume data multiplied by 1/S, where S is the Schmid factor. Dashed lines are linear fits to the data.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Ryu, S., Kang, K. & Cai, W. Predicting the dislocation nucleation rate as a function of temperature and stress. Journal of Materials Research 26, 2335–2354 (2011). https://doi.org/10.1557/jmr.2011.275

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1557/jmr.2011.275

Navigation