Skip to main content
Log in

Modeling the Effects of Grain and Porosity Structure on Copper Spall Response

  • Published:
Journal of Dynamic Behavior of Materials Aims and scope Submit manuscript

Abstract

Ongoing efforts to characterize the incipient spallation of copper have increased the experimental fidelity at which the material’s microstructure is measured. Various imaging techniques have allowed for 3D characterization of the structure of inclusions, porosity, and crystallographic grains. This work employs a combined crystal mechanics and porosity model that, for the first time, addresses the influence of three different types of spatial distributions of second phase particles, relative to features in both experimentally-measured and synthetically-generated microstructures. With these features, the model can probe the sensitivity of copper spall response to both grain and porosity structures. The sensitivity of the model to various nucleation and porosity growth parameters is also shown. These sensitivity studies illustrate where increased experimental accuracy can most readily affect modeling results. Likewise, the model’s fidelity captures many of the key features measured experimentally and is a step toward a model capable of predicting and improving the spall resistance in many materials.

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
Fig. 12
Fig. 13
Fig. 14
Fig. 15

Similar content being viewed by others

Notes

  1. Ideally more instances would have been sampled, but due to the computational expense, only three are simulated. The small variation in spall velocity trace results among these instances suggests that more instances would not yield significant variations.

References

  1. Bingert JF, Suter RM, Lind J, Li SF, Pokharel R, Trujillo CP (2014) High-energy diffraction microscopy characterization of spall damage. In: Song B, Casem D, Kimberley J (eds) Dynamic behavior of materials. Conference proceedings of the society for experimental mechanics series, vol 1. Springer, Cham, pp 397–403. https://doi.org/10.1007/978-3-319-00771-7_48

    Google Scholar 

  2. Davison L, Graham RA (1979) Shock compression of solids. Phys Rep 55(4):255–379

    Article  Google Scholar 

  3. Meyers MA, Aimone CT (1983) Dynamic fracture (spalling) of metals. Prog Mater Sci 28(1):1–96

    Article  Google Scholar 

  4. Curran D, Seaman L, Shockey D (1987) Dynamic failure of solids. Phys Rep 147(5–6):253–388

    Article  Google Scholar 

  5. Fensin S, Escobedo J, Gray G III, Patterson B, Trujillo C, Cerreta E (2014) Dynamic damage nucleation and evolution in multiphase materials. J Appl Phys 115(20):203516

    Article  Google Scholar 

  6. Smith CS (1958) Metallographic studies of metals after explosive shock. Trans Met Soc AIME 212

  7. Smith J (1963) Three low-pressure spall thresholds in copper. In: Symposium on dynamic behavior of materials. ASTM materials science series - 5. ASTM International. https://doi.org/10.1520/STP42034S

  8. McQueen R, Marsh S (1962) Ultimate yield strength of copper. J Appl Phys 33(2):654–665

    Article  Google Scholar 

  9. Brown A, Pham Q, Peralta P, Luo S, Patterson B, Byler D, Koskelo A, Xiao X (2015) Correlations between spall damage mode preference and microstructure in shocked polycrystalline copper: a 3-D X-ray tomography study. J Dyn Behav Mater 1(4):388–396

    Article  Google Scholar 

  10. Chen T, Jiang Z, Peng H, He H, Wang L, Wang Y (2015) Effect of grain size on the spall fracture behaviour of pure copper under plate-impact loading. Strain 51(3):190–197

    Article  Google Scholar 

  11. Escobedo J, Dennis-Koller D, Cerreta E, Patterson B, Bronkhorst C, Hansen B, Tonks D, Lebensohn R (2011) Effects of grain size and boundary structure on the dynamic tensile response of copper. J Appl Phys 110(3):033513

    Article  Google Scholar 

  12. Escobedo J, Cerreta E, Dennis-Koller D (2014a) Effect of crystalline structure on intergranular failure during shock loading. JOM 66(1):156–164

    Article  Google Scholar 

  13. Peralta P, DiGiacomo S, Hashemian S, Luo SN, Paisley D, Dickerson R, Loomis E, Byler D, McClellan K, D’Armas H (2009) Characterization of incipient spall damage in shocked copper multicrystals. Int J Damage Mech 18(4):393–413

    Article  Google Scholar 

  14. Brown AD, Wayne L, Pham Q, Krishnan K, Peralta P, Luo SN, Patterson BM, Greenfield S, Byler D, McClellan KJ (2015) Microstructural effects on damage nucleation in shock-loaded polycrystalline copper. Metall Mater Trans A 46(10):4539–4547

    Article  Google Scholar 

  15. Escobedo JP, Cerreta EK, Dennis-Koller D, Trujillo CP, Bronkhorst CA (2013) Influence of boundary structure and near neighbor crystallographic orientation on the dynamic damage evolution during shock loading. Philos Mag 93(7):833–846

    Article  Google Scholar 

  16. Yang Y, Zhi-qiang P, Xing-zhi C, Zhao-liang G, Tie-gang T, Hai-bo H, Qing-ming Z (2016) Spall behaviors of high purity copper under sweeping detonation. Mater Sci Eng A 651:636–645

    Article  Google Scholar 

  17. Minich RW, Cazamias JU, Kumar M, Schwartz AJ (2004) Effect of microstructural length scales on spall behavior of copper. Metall Mater Trans A 35(9):2663–2673

    Article  Google Scholar 

  18. Escobedo J, Cerreta E, Dennis-Koller D, Patterson B, Bronkhorst C (2014b) Influence of shock loading kinetics on the spall response of copper. In: Journal of Physics: Conference Series, IOP Publishing, vol 500, p 112023

  19. Fensin S, Escobedo-Diaz J, Brandl C, Cerreta E, Gray G, Germann T, Valone S (2014) Effect of loading direction on grain boundary failure under shock loading. Acta Mater 64:113–122

    Article  Google Scholar 

  20. Koller D, Hixson R, Gray G III, Rigg P, Addessio L, Cerreta E, Maestas J, Yablinsky C (2005) Influence of shock-wave profile shape on dynamically induced damage in high-purity copper. J Appl Phys 98(10):103,518

    Article  Google Scholar 

  21. Barton N, Rhee M, Li S, Bernier J, Kumar M, Lind J, Bingert J (2014) Using high energy diffraction microscopy to assess a model for microstructural sensitivity in spall response. J Phys 500:112007

    Google Scholar 

  22. Krishnan K, Brown A, Wayne L, Vo J, Opie S, Lim H, Peralta P, Luo SN, Byler D, McClellan KJ (2015) Three-dimensional characterization and modeling of microstructural weak links for spall damage in fcc metals. Metall Mater Trans A 46(10):4527–4538

    Article  Google Scholar 

  23. Lieberman EJ, Lebensohn RA, Menasche DB, Bronkhorst CA, Rollett AD (2016) Microstructural effects on damage evolution in shocked copper polycrystals. Acta Mater 116:270–280

    Article  Google Scholar 

  24. Menasche DB, Lind J, Li SF, Kenesei P, Bingert JF, Lienert U, Suter RM (2016) Shock induced damage in copper: a before and after, three-dimensional study. J Appl Phys 119(15):154902

    Article  Google Scholar 

  25. Shockey DA, Seaman L, Curran DR (1973) The influence of microstructural features on dynamic fracture. In: Rohde RW, Butcher BM, Holland JR, Karnes CH (eds) Metallurgical effects at high strain rates. Springer, Boston, MA, pp 473–499. https://doi.org/10.1007/978-1-4615-8696-8_27

    Chapter  Google Scholar 

  26. Becker R, LeBlanc M, Cazamias J (2007) Characterization of recompressed spall in copper gas gun targets. J Appl Phys 102(9):093,512

    Article  Google Scholar 

  27. Christy S, Pak HR, Meyers M (1985) Effect of metallurgical parameters on dynamic fracture by spalling of copper. In: International conference on metallurgical applications of shock-wave and high-strain-rate phenomena (EXPLOMET85), pp 835–863

  28. Ahn D, Sofronis P, Kumar M, Belak J, Minich R (2007) Void growth by dislocation-loop emission. J Appl Phys 101(6):063514

    Article  Google Scholar 

  29. Essmann U, Mughrabi H (1979) Annihilation of dislocations during tensile and cyclic deformation and limits of dislocation densities. Philos Mag A 40(6):731–756

    Article  Google Scholar 

  30. Li S, Chu R, Hou J, Wang Z (1998) Microcrack initiation and dislocation structures in an aluminium bicrystal under cyclic loading. Philos Mag A 77(4):1081–1092

    Article  Google Scholar 

  31. Reina C, Marian J, Ortiz M (2011) Nanovoid nucleation by vacancy aggregation and vacancy-cluster coarsening in high-purity metallic single crystals. Phys Rev B 84(10):104117

    Article  Google Scholar 

  32. Wilkerson J, Ramesh K (2014) A dynamic void growth model governed by dislocation kinetics. J Mech Phys Solids 70:262–280

    Article  Google Scholar 

  33. Wilkerson J, Ramesh K (2016) A closed-form criterion for dislocation emission in nano-porous materials under arbitrary thermomechanical loading. J Mech Phys Solids 86:94–116

    Article  Google Scholar 

  34. Belytschko T, Liu WK, Moran B, Elkhodary K (2013) Nonlinear finite elements for continua and structures. Wiley, Chicheste

    Google Scholar 

  35. Ren B, Li S, Qian J, Zeng X (2011) Meshfree simulations of spall fracture. Comput Methods Appl Mech Eng 200(5):797–811

    Article  Google Scholar 

  36. Lebensohn R, Idiart M, Castañeda PP, Vincent PG (2011) Dilatational viscoplasticity of polycrystalline solids with intergranular cavities. Philos Mag 91(22):3038–3067

    Article  Google Scholar 

  37. Lebensohn RA, Escobedo JP, Cerreta EK, Dennis-Koller D, Bronkhorst CA, Bingert JF (2013) Modeling void growth in polycrystalline materials. Acta Mater 61(18):6918–6932

    Article  Google Scholar 

  38. Belak J (1998) On the nucleation and growth of voids at high strain-rates. J Comput Aided Mater Des 5(2):193–206

    Article  Google Scholar 

  39. Clayton J (2005) Modeling dynamic plasticity and spall fracture in high density polycrystalline alloys. Int J Solids Struct 42(16):4613–4640

    Article  Google Scholar 

  40. Clayton JD (2006) Plasticity and spall in high density polycrystals: modeling and simulation. In: AIP conference proceedings, vol 845. AIP, pp 311–314

  41. Vogler T, Clayton J (2008) Heterogeneous deformation and spall of an extruded tungsten alloy: plate impact experiments and crystal plasticity modeling. J Mech Phys Solids 56(2):297–335

    Article  Google Scholar 

  42. Chen Q, Wang J, Liu K (2010) Improved CE/SE scheme with particle level set method for numerical simulation of spall fracture due to high-velocity impact. J Comput Phys 229(19):7503–7519

    Article  Google Scholar 

  43. Barton NR, Winter NW, Reaugh JE (2009) Defect evolution and pore collapse in crystalline energetic materials. Model Simul Mater Sci Eng 17(3):035,003

    Article  Google Scholar 

  44. Nguyen T, Luscher D, Wilkerson J (2017) A dislocation-based crystal plasticity framework for dynamic ductile failure of single crystals. J Mech Phys Solids 108:1–29

    Article  Google Scholar 

  45. Li S, Lind J, Hefferan C, Pokharel R, Lienert U, Rollett A, Suter R (2012) Three-dimensional plastic response in polycrystalline copper via near-field high-energy x-ray diffraction microscopy. J Appl Crystallogr 45(6):1098–1108

    Article  Google Scholar 

  46. Pokharel R, Lind J, Li SF, Kenesei P, Lebensohn RA, Suter RM, Rollett AD (2015) In-situ observation of bulk 3D grain evolution during plastic deformation in polycrystalline Cu. Int J Plast 67:217–234

    Article  Google Scholar 

  47. Marin E, Dawson P (1998) Elastoplastic finite element analyses of metal deformations using polycrystal constitutive models. Comput Methods Appl Mech Eng 165(1):23–41

    Article  Google Scholar 

  48. Cocks A, Ashby M (1980) Intergranular fracture during power-law creep under multiaxial stresses. Met Sci 14(8–9):395–402

    Article  Google Scholar 

  49. Asaro RJ (1983) Crystal plasticity. J Appl Mech 50(4b):921–934

    Article  Google Scholar 

  50. Kweon S (2012) Damage at negative triaxiality. Eur J Mech A 31(1):203–212

    Article  Google Scholar 

  51. Bammann D, Aifantis E (1989) A damage model for ductile metals. Nucl Eng Des 116(3):355–362

    Article  Google Scholar 

  52. Becker R (2002) Ring fragmentation predictions using the Gurson model with material stability conditions as failure criteria. Int J Solids Struct 39(13–14):3555–3580

    Article  Google Scholar 

  53. Chu C, Needleman A (1980) Void nucleation effects in biaxially stretched sheets. J Eng Mater Technol 102(3):249–256

    Article  Google Scholar 

  54. Saje M, Pan J, Needleman A (1982) Void nucleation effects on shear localization in porous plastic solids. Int J Fract 19(3):163–182

    Article  Google Scholar 

  55. Noble C, Anderson A, Barton N, Bramwell J, Capps A, Chang M, Chou J, Dawson D, Diana E, Dunn T, Faux D, Fisher A, Greene P, Heinz I, Kanarska Y, Khairallah S, Liu B, Margraf J, Nichols A, Nourgaliev R, Puso M, Reus J, Robinson P, Shestakov A, Solberg J, Taller D, Tsuji P, White C, White J (2017) ALE3D: an arbitrary Lagrangian-Eulerian multi-physics code. Technical Report LLNL-TR-732040, Lawrence Livermore National Laboratory

  56. Van Leer B (1974) Towards the ultimate conservative difference scheme. II. Monotonicity and conservation combined in a second-order scheme. J Comput Phys 14(4):361–370

    Article  Google Scholar 

  57. Barton N, Rhee M, Brown J (2018) Modeling the constitutive response of tantalum across experimental platforms. In: AIP conference proceedings, vol 1979. AIP Publishing

  58. Gray G III, Hull L, Faulkner J, Briggs M, Cerreta E, Addessio F, Bourne N (2009) The effects of shockwave profile shape and shock obliquity on spallation: kinetic and stress-state effects on damage evolution. In: AIP conference proceedings, vol 1195. AIP, pp 1097–1102

  59. Landron C, Bouaziz O, Maire E, Adrien J (2010) Characterization and modeling of void nucleation by interface decohesion in dual phase steels. Scr Mater 63(10):973–976

    Article  Google Scholar 

  60. Maire E, Bouaziz O, Di Michiel M, Verdu C (2008) Initiation and growth of damage in a dual-phase steel observed by x-ray microtomography. Acta Mater 56(18):4954–4964

    Article  Google Scholar 

  61. Pushkareva M, Adrien J, Maire E, Segurado J, Llorca J, Weck A (2016) Three-dimensional investigation of grain orientation effects on void growth in commercially pure titanium. Mater Sci Eng A 671:221–232

    Article  Google Scholar 

  62. Gray G III, Livescu V, Rigg P, Trujillo C, Cady C, Chen S, Carpenter J, Lienert T, Fensin S (2017) Structure/property (constitutive and spallation response) of additively manufactured 316L stainless steel. Acta Mater 138:140–149

    Article  Google Scholar 

  63. Barton NR, Knap J, Arsenlis A, Becker R, Hornung RD, Jefferson DR (2008) Embedded polycrystal plasticity and adaptive sampling. Int J Plast 24(2):242–266

    Article  Google Scholar 

  64. Moore JA, Barton NR, Florando J, Mulay R, Kumar M (2017) Crystal plasticity modeling of \(\beta\) phase deformation in ti-6al-4v. Model Simul Mater Sci Eng 25(7):075,007

    Article  Google Scholar 

  65. Horstemeyer M, Lathrop J, Gokhale A, Dighe M (2000) Modeling stress state dependent damage evolution in a cast Al-Si-Mg aluminum alloy. Theor Appl Fract Mech 33(1):31–47

    Article  Google Scholar 

  66. Follansbee P, Kocks U (1988) A constitutive description of the deformation of copper based on the use of the mechanical threshold stress as an internal state variable. Acta Metall 36(1):81–93

    Article  Google Scholar 

  67. Arsenlis A, Barton N, Becker R, Rudd R (2006) Generalized in situ adaptive tabulation for constitutive model evaluation in plasticity. Comput Methods Appl Mech Eng 196(1–3):1–13

    Article  Google Scholar 

  68. Follansbee P (1986) High-strain-rate deformation of fcc metals and alloys. Metall Appl Shock Wave High Strain Rate Phenom 1(1):451–479

    Google Scholar 

  69. Hiratani M, Nadgorny E (2001) Combined model of dislocation motion with thermally activated and drag-dependent stages. Acta Mater 49(20):4337–4346

    Article  Google Scholar 

  70. Meyers MA (1994) Dynamic behavior of materials. Wiley, New York

    Book  Google Scholar 

  71. Estrin Y, Mecking H (1984) A unified phenomenological description of work hardening and creep based on one-parameter models. Acta Metall 32(1):57–70

    Article  Google Scholar 

  72. Mecking H, Kocks U (1981) Kinetics of flow and strain-hardening. Acta Metall 29(11):1865–1875

    Article  Google Scholar 

  73. Benson DJ, Nellis W (1994) Dynamic compaction of copper powder: computation and experiment. Appl Phys Lett 65(4):418–420

    Article  Google Scholar 

  74. Drumheller DS (1998) Introduction to wave propagation in nonlinear fluids and solids. Cambridge University Press, Cambridge

    Book  Google Scholar 

  75. Moore JA (2018) A degradation function consistent with Cocks-Ashby porosity kinetics. Int J Fract 209:231–234

    Article  Google Scholar 

  76. Casals O, Očenášek J, Alcala J (2007) Crystal plasticity finite element simulations of pyramidal indentation in copper single crystals. Acta Mater 55(1):55–68

    Article  Google Scholar 

  77. Kelly A, Groves GW, Kidd P (2000) Crystallography and crystal defects. Wiley, Chichester

    Google Scholar 

  78. Tvergaard V, Needleman A (1984) Analysis of the cup-cone fracture in a round tensile bar. Acta Metall 32(1):157–169

    Article  Google Scholar 

  79. Carreker R Jr, Hibbard W Jr (1953) Tensile deformation of high-purity copper as a function of temperature, strain rate, and grain size. Acta Metall 1(6):654–663

    Article  Google Scholar 

  80. Dao M, Lu L, Shen Y, Suresh S (2006) Strength, strain-rate sensitivity and ductility of copper with nanoscale twins. Acta Mater 54(20):5421–5432

    Article  Google Scholar 

Download references

Acknowledgements

The authors wish to thank Armand Beaudoin, Richard Becker, John Bingert, Christopher Hefferan, Mukul Kumar, Jonathan Lind, James Stölken, and Robert Suter for their valuable discussions and help with this work. This work was performed under the auspices of the U.S. Department of Energy by Lawrence Livermore National Laboratory under Contract DE-AC52-07NA27344 (LLNL-JRNL-751104-DRAFT). The work is supported in part by the Joint DOD/DOE Munitions Technology Development Program.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to John A. Moore.

Appendices

Appendix 1

Kinematics

The material kinematics are based on a multiplicative decomposition of the deformation gradient

$$\begin{aligned} \pmb {\text {F}} = \pmb {\text {V}} \cdot \pmb {\text {R}} \cdot \pmb {\text {F}}_p \cdot \pmb {\text {F}}_d, \end{aligned}$$
(8)

where \(\pmb {\text {R}}\) is the lattice rotation, and \(\pmb {\text {V}}\) is the thermoelastic lattice stretch. The plastic deformation gradient due to dislocation slip is described by \(\pmb {\text {F}}_p\), and \(\pmb {\text {F}}_d\) captures the volume change from porosity (i.e., damage). This multiplicative decomposition results in several intermediate configurations \({\bar{{\mathscr {B}}}}\), \(\hat{{\mathscr {B}}}\), \(\tilde{{\mathscr {B}}}\) (similar to those in [47]) between a reference configuration \({\mathscr {B}}_o\) and current (deformed) configuration \({\mathscr {B}}\) as is illustrated in Fig. 16.

Fig. 16
figure 16

The kinematics of the multiplicative decomposition of the deformation gradient

The velocity gradient is decomposed into a symmetric (\(\pmb {\text {D}}\)) and skew-symmetric (\(\pmb {\text {W}}\)) part as

$$\begin{aligned} \pmb {\text {L}} = \text {sym}(\pmb {\text {L}}) + \text {skew}(\pmb {\text {L}})= \pmb {\text {D}}+ \pmb {\text {W}}. \end{aligned}$$
(9)

The velocity gradient is

$$\begin{aligned} \pmb {\text {L}} = \dot{\pmb {\text {F}}} \cdot \pmb {\text {F}}^{-1}. \end{aligned}$$
(10)

Combining Eqs. 8 and 10 gives

$$\begin{aligned} \pmb {\text {L}} = \dot{\pmb {\text {V}}}\cdot \pmb {\text {V}}^{-1} + \pmb {\text {V}}\cdot \left[ \pmb {\text {W}}_R + \pmb {\text {R}} \cdot \left( \hat{\pmb {\text {L}}}_p + \pmb {\text {F}}_p \cdot {\bar{\pmb {\text {L}}}}_d \cdot \pmb {\text {F}}_p^{-1} \right) \cdot \pmb {\text {R}}^{T} \right] \cdot \pmb {\text {V}}^{-1}, \end{aligned}$$
(11)

where \(\hat{\pmb {\text {L}}}_p \equiv \dot{\pmb {\text {F}}}_p\cdot \pmb {\text {F}}_p^{-1}\) and \({\bar{\pmb {\text {L}}}}_p \equiv \dot{\pmb {\text {F}}}_d\cdot \pmb {\text {F}}_d^{-1}\) represent deformation from slip and porosity growth respectively and \(\pmb {\text {W}}_R \equiv \dot{\pmb {\text {R}}}\cdot \pmb {\text {R}}^T\) accounts for rigid rotation.

Dislocation Slip Kinematics

The decomposition of \(\hat{\pmb {\text {L}}}_p\) into symmetric (\(\hat{\pmb {\text {D}}}_p\)) and skew-symmetric (\(\hat{\pmb {\text {W}}}_p\)) parts gives

$$\begin{aligned} \hat{\pmb {\text {L}}}_p = \text {sym}(\hat{\pmb {\text {L}}}_p) + \text {skew}(\hat{\pmb {\text {L}}}_p)= \hat{\pmb {\text {D}}} _p+ \hat{\pmb {\text {W}}}_p. \end{aligned}$$
(12)

The governing equations for the material response concerning slip are then

$$\begin{aligned} \pmb {\text {R}}^T \cdot \pmb {\text {D}}^\prime \cdot \pmb {\text {R}}= \frac{1}{a} \dot{\tilde{\pmb {\text {V}}}}^\prime + \hat{\pmb {\text {D}}} _p, \end{aligned}$$
(13)
$$\begin{aligned} \pmb {\text {R}}^T \cdot \pmb {\text {W}}\cdot \pmb {\text {R}}= \pmb {\text {R}}^T \cdot \pmb {\text {W}}_R \cdot \pmb {\text {R}} + \hat{\pmb {\text {W}}} _p - \frac{1}{2a^2} \left( \dot{\tilde{\pmb {\text {V}}}}^\prime \cdot \tilde{\pmb {\text {V}}}^\prime - \tilde{\pmb {\text {V}}}^\prime \cdot \dot{\tilde{\pmb {\text {V}}}}^\prime \right) , \end{aligned}$$
(14)

where \((\bullet )^\prime\) denotes the deviator and \(\tilde{\pmb {\text {V}}} \equiv \pmb {\text {R}}^T\cdot \pmb {\text {V}}\cdot \pmb {\text {R}}\). These equations are similar to those in [63, 64], where more details on their derivation are given.

Following the assumptions of [49], \(\hat{\pmb {\text {D}}} _p\) and \(\hat{\pmb {\text {W}}}_p\) are determined from

$$\begin{aligned} \hat{\pmb {\text {L}}}_p= {\dot{\pmb {\text {F}}}}_p \cdot \pmb {\text {F}}_p^{-1}= \sum _{\alpha =1}^{N} {\dot{\gamma }}^{\alpha } \; \hat{\pmb {\text {s}}} ^{\alpha } \otimes \hat{\pmb {\text {m}}} ^{\alpha } , \end{aligned}$$
(15)

where N is the number of slip systems, \({\dot{\gamma }}\) is the shear slip rate, \(\hat{\pmb {\text {s}}}^{\alpha }\) is the slip direction, and \(\hat{\pmb {\text {m}}}^{\alpha }\) is the slip plane normal, all for a slip system \(\alpha\). In Eqs. 13 and 14 it is assumed that that the deviatoric part of the elastic strain is small (a reasonable assumption for metal alloys) such that the elastic stretch can be written as

$$\begin{aligned} \pmb {\text {V}} \equiv a \left( \pmb {\text {I}} + \pmb {\epsilon }^* \right) , \end{aligned}$$
(16)

where \(\pmb {\text {I}}\) is the identity tensor, \(\pmb {\epsilon }^*\) is defined as a deviatoric tensor, and a is a variable related to volume change. Equations 13 and 14 are solved for fundamental degrees-of-freedom \(\pmb {\epsilon }^*\), \(\pmb {\text {W}}_\mathrm {R}\) where \(\pmb {\text {D}}\) and \(\pmb {\text {W}}\) are known from the finite element method.

Porosity Kinematics

Porosity evolution is assumed to account for volume change only and is governed by the Cocks–Ashby porosity kinetics in Eq. 3. Porosity evolution is connected to the crystal mechanics via the trace of Equation 11. Since \(\pmb {\text {F}}_p\) and \(\pmb {\text {W}}_R\) do not contribute to volume change this gives

$$\begin{aligned} \mathrm {tr} \left( \pmb {\text {D}}\right) = \mathrm {tr} \left( {\dot{\pmb {\text {V}}}}\cdot \pmb {\text {V}}^{-1} \right) + \mathrm {tr} \left( {\dot{\pmb {\text {F}}}_d}\cdot \pmb {\text {F}}_d^{-1} \right) , \end{aligned}$$
(17)

which for simplicity (in implementation) is written as

$$\begin{aligned} \mathrm {tr} \left( \pmb {\text {D}}\right) = \mathrm {tr} \left( \pmb {\text {D}}_V\right) + \mathrm {tr} \left( {\bar{\pmb {\text {D}}}}_d\right) , \end{aligned}$$
(18)

where \(\pmb {\text {D}}_V = {\dot{\pmb {\text {V}}}}\cdot \pmb {\text {V}}^{-1}\) and \({\bar{\pmb {\text {D}}}}_d\) is the symmetric part of \({\bar{\pmb {\text {L}}}}_d\). Using the assumption in [65], the deformation due to porosity change is \(\pmb {\text {F}}_d = \varPhi \pmb {\text {I}}\). Applying the definition \(J_d = \mathrm {det} (\pmb {\text {F}}_d)\) then gives \(\varPhi = J_d^{1/3}\) and \(\pmb {\text {F}}_d = J_d^{1/3}\pmb {\text {I}}\). Differentiating gives

$$\begin{aligned} \dot{\pmb {\text {F}}}_d = \frac{1}{3}J_d^{-2/3}\dot{J}_d \pmb {\text {I}}, \end{aligned}$$
(19)

resulting in

$$\begin{aligned} \bar{\pmb {\text {D}}}_d = \text {sym} (\dot{\pmb {\text {F}}_d}\cdot \pmb {\text {F}}_d^{-1}) = \frac{1}{3} \frac{{\dot{J}}_d}{J_d} \pmb {\text {I}}. \end{aligned}$$
(20)

Thus, \(\text {tr}({\bar{\pmb {\text {D}}}}_d) = {\dot{J}}_d/J_d\) and rearranging shows that this an ordinary differential equation (ODE) of the form

$$\begin{aligned} {\dot{J}}_d = \mathrm {tr}({\bar{\pmb {\text {D}}}}_d) J_d. \end{aligned}$$
(21)

Similarly, applying the definition \(J_V = \mathrm {det} (\pmb {\text {V}})\) and the approximation that \(\mathrm {det} (\pmb {\text {V}}) = a^3\), results in \(\pmb {\text {V}} = J_V^{1/3}(\pmb {\text {I}} + \pmb {\epsilon }^*)\). Differentiating then gives

$$\begin{aligned} \dot{\pmb {\text {V}}} = \frac{1}{3}J_V^{-2/3}{\dot{J}}_V (\pmb {\text {I}} + \pmb {\epsilon }^*). \end{aligned}$$
(22)

Using the Cayley-Hamilton theorem for the inverse of a \(3 \times 3\) matrix and neglecting higher-order term gives \(\pmb {\text {V}}^{-1} = \frac{1}{a} \left( \pmb {\text {I}} - \pmb {\epsilon }^*\right)\) and

$$\begin{aligned} \pmb {\text {D}}_V = {\dot{\pmb {\text {V}}}} \cdot \pmb {\text {V}}^{-1} = \frac{1}{3} \frac{\dot{J}_V}{J_V}\pmb {\text {I}}. \end{aligned}$$
(23)

Following the same logic used to arrive at Eq. 21 gives

$$\begin{aligned} {\dot{J}}_V = \mathrm {tr}(\pmb {\text {D}}_V) J_V. \end{aligned}$$
(24)

Solving the ODEs in 21 and 24 numerically gives

$$\begin{aligned} J_d= & {} \exp \left( \text {tr}({\bar{\pmb {\text {D}}}}_d) \varDelta t\right) J_d^o, \end{aligned}$$
(25)
$$\begin{aligned} J_V= & {} \exp \left( \text {tr}(\pmb {\text {D}}_V) \varDelta t\right) J_V^o. \end{aligned}$$
(26)

Where \(J_d\) and \(J_V\) are at the current step and \(J_d^o\) and \(J_V^o\) are at the previous step and \(\varDelta t\) is the current time increment. In implementing porosity evolution, \(\text {tr}({\bar{\pmb {\text {D}}}}_d)\) is considered as unknown (i.e., degree-of-freedom). The value of \(\text {tr}\left( \pmb {\text {D}}_V \right)\) is then determined from Eq. 17 and \(J_V\) is used to determine stress, while \(J_d\) is used to determine the porosity. Following the derivation by Bammann and Aifantis [51] and accounting for nucleated porosity (\(f_n\)) gives

$$\begin{aligned} J_d = \frac{1 - f_o}{1 - (f - f_n)}, \end{aligned}$$
(27)

where f is the total porosity and \(f_o\) is the initial porosity (which in this work is always zero). The porosity is then calculated from Eq. 27. The pressure and porosity are used to evaluate Eq. 3. Using Eq. 18, the pressure (based on \(J_V\)) and porosity (based on \(J_d\)) are determined from \(\mathrm {tr} \left( {\bar{\pmb {\text {D}}}}_d\right)\) which is iterated upon until Eq. 3 is satisfied.

Slip Kinetics

As discussed in [66], thermally activated glide and dislocation drag are competing mechanisms of slip where the controlling mechanism is strain-rate dependent. Due to the heterogeneity of the microstructure simulated in this work, a range of local strain-rates occur and thus both thermally activated and dislocation drag slip kinetics are modeled. To account for both mechanisms, it is assumed that the time needed for dislocation motion is the sum of the time spent overcoming thermal barriers and the time spent gliding between barriers respectively. Since the slip-rate of a given slip system is inversely proportional to the time a dislocation spends overcoming or moving between various barriers, the total slip-rate is assumed to follow the form

$$\begin{aligned} {\dot{\gamma }}^{\alpha } = \left( \frac{1}{\dot{\gamma }_T^\alpha } + \frac{1}{ \dot{\gamma }_d^\alpha } \right) ^{-1}, \end{aligned}$$
(28)

where \({\dot{\gamma }}_T^\alpha\) accounts for dislocations overcoming barriers by thermal activation and \({\dot{\gamma }}_d^\alpha\) accounts for the dislocation glide motion before encountering the nearest forest dislocations (i.e., dislocation drag). This drag term (similar to [43, 67,68,69,70]) is given by

$$\begin{aligned} {\dot{\gamma }}_d^\alpha = \left\{ \begin{array}{cl} \left( {\dot{\gamma }}_{do} h \right) \left( 1 - \exp \left( -\frac{(\tau ^{\alpha } - g^{*\alpha })}{D_d^*} \right) \right) &{} {\text{for}} \quad \tau ^\alpha> g^{*\alpha } \\ 0 &{} {\text{for}} \quad \tau ^\alpha \le g^{*\alpha } \end{array}, \right. \end{aligned}$$
(29)

where \(D_d\) is the drag stress, \({\dot{\gamma }}_{do}\) is a reference slip-rate between obstacles, and \(g^\alpha\) is the strength of slip system \(\alpha\), and star superscripts denoted degraded strength values. The variable h is the current dislocation density normalized by a reference dislocation density: \(1\, {\upmu \text{m}}^{-2}\). The initial value of normalized dislocation density is denoted by \(h_o\). The thermally activated term [43, 66] is given by

$$\begin{aligned} {\dot{\gamma }}_T^\alpha = \frac{{\dot{\gamma }}_{To}}{\sqrt{h}} \left[ \exp \left( -\frac{\varDelta G^\alpha (\tau ^\alpha )}{kT}\right) - \exp \left( -\frac{\varDelta G^\alpha (-\tau ^\alpha )}{kT}\right) \right] , \end{aligned}$$
(30)

where k is the Boltzmann constant, T is the temperature, \({\dot{\gamma }}_{To}\) is a reference thermal activation slip-rate. The free energy \(\varDelta G\) is given by

$$\begin{aligned} \varDelta G^\alpha (\tau ^\alpha ) = {\mathcal {G}}_o b^3 \mu ^{\alpha } \left[ 1 - \left( \frac{(\tau ^\alpha - \tau _a^*)}{g^{*\alpha }}\right) ^{p} \right] ^{q} \;\;\text{ for} \quad \tau ^\alpha \le g^{*\alpha }+ \tau _a^* , \end{aligned}$$
(31)

where \({\mathcal {G}}_o\) is a normalized activation energy, b is the magnitude of the Burgers vector, \(\mu ^{\alpha }\) is the shear modulus resolved onto the slip system, and p and q are constant parameters. The slip system strength caused by dislocations is given by

$$\begin{aligned} g^\alpha = s \sqrt{h}, \end{aligned}$$
(32)

where s is constant and the slip system strength is assumed to be the same for all systems. For dislocation density evolution, a Kocks–Mecking [71, 72] type hardening model is used

$$\begin{aligned} {\dot{h}} = ( k_1\sqrt{h} - k_2 h ){\dot{\gamma }} \;\;\text{ where }\;\; k_2= \left( k_{2o}\frac{{\dot{\gamma }}_{k0}}{\dot{\gamma }}\right) ^{1/c} . \end{aligned}$$
(33)

The first term accounts for dislocation multiplication and the second term accounts for dislocation sinks by annihilation and is a function of shearing rate and strain-rate sensitivity, c. The parameter \(k_1\), \(k_2\), \(k_{2o}\), and \({\dot{\gamma }}_{ko}\) are material dependent. The normalized dislocation density h is the final fundamental degrees-of-freedom for the system.

Equation of State

The model uses a Grüneisen equation of state in the form employed in [73]

$$\begin{aligned} p = \frac{\rho _o C^2 {\mathscr {U}} \left[ 1 + \left( 1 - \varGamma _o/2\right) {\mathscr {U}} - A {\mathscr {U}}^2/2\right] }{\left[ 1 - (S-1){\mathscr {U}} \right] ^2} + \left( \varGamma _o + A {\mathscr {U}}\right) E , \end{aligned}$$
(34)

where \(\rho\) and \(\rho _o\) are current and initial density, respectively C is the bulk wave speed, \(\varGamma _o\) is the Grüneisen parameter, A is a volume correction term, E is internal energy, and

$$\begin{aligned} {\mathscr {U}} = \frac{\rho }{\rho _o} - 1 = J_V^{-1} - 1 . \end{aligned}$$
(35)

The temperature is calculated using

$$\begin{aligned} T = \frac{E -E_c}{\rho C_v} , \end{aligned}$$
(36)

where \(C_v\) is the heat capacity at a constant volume and \(E_c\) represents the cold energy [74]. No heat transfer is modeled.

Stress and Material Degradation

The deviatoric Kirchhoff stress is determined using

$$\begin{aligned} \pmb {\tau }^\prime = \pmb {\mathscr {L}}: \text {ln}(\pmb {\text {V}}^\prime ) , \end{aligned}$$
(37)

where \(\pmb {\mathscr {L}}\) is a fourth-order tensor representing a constant crystallographic elastic moduli. For this work \(\pmb {\mathscr {L}}\) is characterized by three elastic constants \(C_{11}\), \(C_{12}\), and \(C_{33}\) that describe \(\pmb {\mathscr {L}}\) in Voigt notation (see [34]).

The equivalent stress (which is based on the deviator of the Kirchhoff stress, \(\pmb {\tau }\)) is given by \(\tau _{eq} = \sqrt{\frac{3}{2}\pmb {\tau }^\prime :\pmb {\tau }^\prime }\). The Cauchy stress \(\pmb {\sigma }\) is given by

$$\begin{aligned} \pmb {\sigma }=\frac{\pmb {\tau }}{ J_V J_d} . \end{aligned}$$
(38)

The degraded matrix material equivalent strength \(\tau _{eq}^*\) is given by

$$\begin{aligned} \tau ^*_{eq} = w \cdot \tau _{eq}, \end{aligned}$$
(39)

where w is a degradation function given by \(w=1 - \text {tanh}(c_2f)\). All associated crystal slip system strengths are likewise degraded by w and are denoted with stars (e.g., \(g^{*\alpha }\), \(D_r^*\), \(\tau _a^*\), etc.). The degradation function phenomenologically represents a smooth transition from full load carrying capability at zero porosity to a very low loading carrying capability (the value of which is determined by \(c_2\)) at high porosity without the implementation complexity of a micromechanics-based degradation function as in [75].

Implementation

The materials model presented here respectively solves Eqs. 13, 14, 3, 33 for degrees-of-freedom \(\pmb {\epsilon }^*\), \(\pmb {\text {W}}_\mathrm {R}\) , \(\text {tr}({\bar{\pmb {\text {D}}}}_d)\), and h. Equation 13 is solved using a dogleg trust region method. The crystal mechanics and porosity equations are solved in a staggered manner holding \(J_V\) constant (i.e., at its beginning of time step value) during the solution for Eqs. 13, 14 and holding \(\pmb {\epsilon }^*\), \(\pmb {\text {W}}_\mathrm {R}\) constant during the solution of Eq. 3.

Appendix 2

The parameters used in each simulation are shown in Table 2 in Appendix 2. The parameters associated with thermal activation \({\mathcal {G}}_o\), b, p and q are set to the values determined by Follansbee and Kocks [66] for their mechanical threshold stress model of 99.99\(\%\) copper. Arsenlis et al.’s [67] values of resolved shear modulus \(\mu\) and drag stress \(D_r\) for unalloyed copper are also used. The initial temperature is 300 K and the elastic constants \(C_{11}\), \(C_{12}\), and \(C_{44}\) values were reported in Casals et al.’s [76] study of single crystal copper. The 12 face-centered cubic \(\langle 1{\bar{1}}0 \rangle\)/\(\left\{ 111\right\}\) slip systems are also used [77]. The equation of state parameter S, \(\varGamma _o\), and A values are taken from Benson et al. [73].

Table 2 Parameters for copper simulations

The strength (\(\tau _a\), s), kinetics (\({\dot{\gamma }}_{To}\), \({\dot{\gamma }}_{do}\)), and hardening (\(k_1\), \(k_{2o}\), \({\dot{\gamma }}_{ko}\), c, \(h_o\)) parameters are all calibrated to the quasi-static stress–strain response reported by Fensin et al. [5].

The value of the degradation parameter \(c_2\) is estimated as 20. In each simulation the material is considered fully degraded when the degradation function in Eq. 39 is \(w = 0.001\). Using \(c_2 = 20\) and solving the degradation function expression for porosity gives a failure porosity of \(f_{fail} \approx 0.2\). This value lies between the coalescence and failure porosity used by Tvergaard and Needleman [78] of 0.15 and 0.25 respectively and is considered (by the authors) to be an adequate estimate for the purposes of this work.

The unitless value of strain-rate sensitivity exponent n is chosen as 10. At room temperature, n values for copper are generally higher than 10 (see [66, 79, 80]) and a function of strain-rate. Lower strain-rates tend to result in a rate insensitive responses (i.e., high n) and higher strain-rates result in a rate dependent responses (i.e., low n) [66]. For this work, the heterogeneity of the system leads to a spectrum of local strain-rates. Since the region of interest (where spalling occurs) has the highest strain-rates, a low value of n (such as 10) is more appropriate for these regions. While this may introduce inaccuracies for regions away from the spall plane, these inaccuracies (as they pertain to the porosity kinetics in Eq. 3) are alleviated partially through the calibration of the porosity kinetics parameter \(c_1\). The Cocks–Ashby [48] porosity kinetics expression, was originally derived for material creep, where they considered values of n ranging from 1 to 9; therefore, the chosen value of 10 is similar to the values where Cocks–Ashby’s assumptions are considered valid.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Moore, J.A., Li, S.F., Rhee, M. et al. Modeling the Effects of Grain and Porosity Structure on Copper Spall Response. J. dynamic behavior mater. 4, 464–480 (2018). https://doi.org/10.1007/s40870-018-0170-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s40870-018-0170-6

Keywords

Navigation