Skip to main content
Log in

A Computational Model for the Thermal Spallation of Crystalline Rocks

  • Original Paper
  • Published:
Rock Mechanics and Rock Engineering Aims and scope Submit manuscript

Abstract

Thermal spallation is the erosion of brittle materials subjected to high heat fluxes, such as a hot fire in a concrete structure. While experimental investigations provide valuable data on the spallation process, they lack generality to be extrapolated to other materials or loading conditions. This paper presents a numerical method for predicting bulk quantities of the spallation process, such as the average recession rate, for crystalline materials. The method is based on direct numerical simulations of the microstructure. The crystal grains and their boundaries are modeled with finite elements and cohesive zones. These elements have anisotropic and temperature-dependent physical properties, a novel level of fidelity and one that accurately captures known crystal phenomena including the \(\alpha\)\(\beta\) quartz transition. These models of the crystalline material and the thermal spallation process are validated against experiment results for Barre granite, and then applied to predict the thermal spallation of Martian basalt.

Highlights

  • This paper presents a high-fidelity model of a process called thermal spallation.

  • Each individual grain within a rock is simulated directly, with material properties that are both temperature- and direction-dependent.

  • This computational model reproduces experimental results for Barre granite.

  • The model is then used to predict thermal spallation of Martian basalt, which is relevant to landing large robot and human missions on Mars.

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
Fig. 16
Fig. 17
Fig. 18

Similar content being viewed by others

Data Availability

The simulation data presented in these sections are available upon request.

References

  • Air Force Civil Engineering Support Agency (2014) Engineering technical letter (ETL) 14-2 (change 1): preventing and repairing concrete deterioration under MV-22 and CV-22 aircraft. Tech. rep., Department of the Air Force, Tyndall AFB, FL. https://nanopdf.com/download/21-november-2014_pdf

  • Anderson OL, Isaak DG (1995) Elastic constants of mantle minerals at high temperature. In: Ahrens TJ (ed) Mineral Physics and crystallographic: a handbook of physical constants. American Geophysical Union, Washington, DC, pp 64–97

    Google Scholar 

  • Austin SA, Robins PJ, Richards MR (1992) Jet-blast temperature-resistant concrete for Harrier aircraft pavements. Struct Eng 70(23):427–432

    Google Scholar 

  • Baird J, Caskey B, Tinianow M et al (1984) GEODYN: a geological formation/drillstring dynamics computer program. In: SPE annual technical conference and exhibition

  • Bass JD (1995) Elasticity of minerals, glasses, and melts. In: Ahrens TJ (ed) Mineral physics and crystallography: a handbook of physical constants 2. American Geophysical Union, Washington, DC, pp 45–63

    Google Scholar 

  • Beentjes I (2018) Dissolution and thermal spallation of Barre granite using pure and chemically enhanced hydrothermal jets. Master’s thesis, Cornell University

  • Beentjes I, Bender JT, Hillson SD, et al (2018) Hydrothermal spallation of Barre granite using supercritical water jets. In: Proceedings, 43rd workshop on geothermal reservoir engineering, Stanford University, Stanford, California

  • Beentjes I, Bender JT, Hawkins AJ et al (2020) Chemical dissolution drilling of Barre granite using a sodium hydroxide enhanced supercritical water jet. Rock Mech Rock Eng 53(2):483–496

    Article  Google Scholar 

  • Bouchard PO, Bay F, Chastel Y (2003) Numerical modelling of crack propagation: automatic remeshing and comparison of different criteria. Comput Methods Appl Mech Eng 192(35–36):3887–3908

    Article  Google Scholar 

  • Bouquet JBP, Rimoli JJ (2015) A length-dependent model for the thermomechanical response of ceramics. J Mech Phys Solids 82:82–96. https://doi.org/10.1016/j.jmps.2015.05.018

    Article  Google Scholar 

  • Bower AF (2009) Applied mechanics of solids. CRC Press, Boca Raton

    Book  Google Scholar 

  • Branner JC (1895) Decomposition of rocks in Brazil. Bull Geol Soc Am 7(1):255–314. https://doi.org/10.1130/GSAB-7-255

    Article  Google Scholar 

  • Brown JM, Angel RJ, Ross NL (2016) Elasticity of plagioclase feldspars. J Geophys Res Solid Earth 121(2):663–675

    Article  Google Scholar 

  • Cashman KV, Marsh BD (1988) Crystal size distribution (CSD) in rocks and the kinetics and dynamics of crystallization II: Makaopuhi lava lake. Contrib Mineral Petrol 99(3):292–305. https://doi.org/10.1007/BF00375363

    Article  Google Scholar 

  • Clayton JD (2005) Dynamic plasticity and fracture in high density polycrystals: constitutive modeling and numerical simulation. J Mech Phys Solids 53(2):261–301

    Article  Google Scholar 

  • Coleman NT, Leroux FH, Cady JG (1963) Biotite–hydrobiotite–vermiculite in soils. Nature 198:409–410. https://doi.org/10.1038/198409c0

    Article  Google Scholar 

  • Dai F, Xia K, Nasseri MHB (2013) Micromechanical model for the rate dependence of the fracture toughness anisotropy of Barre granite. Int J Rock Mech Min Sci 63:113–121

    Article  Google Scholar 

  • Espinosa HD, Zavattieri PD (2003) A grain level model for the study of failure initiation and evolution in polycrystalline brittle materials. Part II: numerical examples. Mech Mater 35(3–6):365–394

    Article  Google Scholar 

  • Fei Y (1995) Thermal expansion. In: Ahrens T (ed) Mineral physics and crystallography: a handbook of physical constants 2. American Geophysical Union, Washington, DC, pp 29–44. https://doi.org/10.1029/rf002p0029

    Chapter  Google Scholar 

  • Finger LW, Ohashi Y (1976) The thermal expansion of diopside to 800 degrees C and a refinement of the crystal structure at 700 degrees C. Am Mineral 61(3–4):303–310

    Google Scholar 

  • Fujino K, Sasaki S, Takeuchi Y et al (1981) X-ray determination of electron distributions in forsterite, fayalite and tephroite. Acta Crystallogr Sect B Struct Crystallogr Cryst Chem 37(3):513–518

    Article  Google Scholar 

  • Gawin D, Pesavento F, Schrefler B (2006) Towards prediction of the thermal spalling risk through a multi-phase porous media model of concrete. Comput Methods Appl Mech Eng 195(41–43):5707–5729

    Article  Google Scholar 

  • Goldsmith W, Sackman JL, Ewerts C (1976) Static and dynamic fracture strength of Barre granite. Int J Rock Mech Min Sci Geomech Abstr 13(11):303–309. https://doi.org/10.1016/0148-9062(76)91829-5

    Article  Google Scholar 

  • Goodrich CA (2003) Petrogenesis of olivine-phyric shergottites Sayh al Uhaymir 005 and Elephant Moraine A79001 lithology A. Geochim Cosmochim Acta 67(19):3735–3772. https://doi.org/10.1016/S0016-7037(03)00171-6

    Article  Google Scholar 

  • Hart KA, Rimoli JJ (2020a) Generation of statistically representative microstructures with direct grain geometry control. Comput Methods Appl Mech Eng. https://doi.org/10.1016/j.cma.2020.113242

    Article  Google Scholar 

  • Hart KA, Rimoli JJ (2020b) MicroStructPy: a statistical microstructure mesh generator in Python. SoftwareX 12:100595. https://doi.org/10.1016/j.softx.2020.100595

    Article  Google Scholar 

  • Haussühl S (1993) Thermoelastic properties of beryl, topaz, diaspore, sanidine and periclase. Zeitschrift fur Kristallographie Crystall Mater 204(1–2):67–76. https://doi.org/10.1524/zkri.1993.204.Part-1.67

    Article  Google Scholar 

  • Hidnert P, Dickson G (1945) Some physical properties of mica. J Res Natl Bureau Stand. https://doi.org/10.6028/jres.035.014

    Article  Google Scholar 

  • Hironaka MC, Malvar LJ (1998) Jet exhaust damaged concrete. Concr Int 20(10):32–35

    Google Scholar 

  • Höfer M, RF (2002) Heat transfer in quartz, orthoclase, and sanidine at elevated temperature. Phys Chem Miner 29(9):571–584. https://doi.org/10.1007/s00269-002-0277-z

    Article  Google Scholar 

  • Isaak DG, Anderson OL, Goto T et al (1989) Elasticity of single-crystal forsterite measured to 1700 K. J Geophys Res Solid Earth 94(B5):5895–5906. https://doi.org/10.1029/JB094iB05p05895

    Article  Google Scholar 

  • Kalifa P, Menneteau FD, Quenard D (2000) Spalling and pore pressure in HPC at high temperatures. Cement Concr Res 30(12):1915–1927. https://doi.org/10.1016/S0008-8846(00)00384-7

    Article  Google Scholar 

  • Kant MA, von Rohr PR (2016) Minimal required boundary conditions for the thermal spallation process of granitic rocks. Int J Rock Mech Min Sci 84:177–186

    Article  Google Scholar 

  • Kant MA, Becker D, Brkic D et al (2015) Investigation of a novel drilling technology-influence of the surface temperature for hydrothermal spallation drilling. In: Proceedings World Geothermal Congress. Melbourne, Australia, pp 19–25

  • Ki Horai (1971) Thermal conductivity of rock-forming minerals. J Geophys Res 76(5):1278–1308. https://doi.org/10.1029/jb076i005p01278

    Article  Google Scholar 

  • Kosinski JA, Gualtieri JG, Ballato A (1992) Thermal expansion of alpha quartz. IEEE Trans Ultrason Ferroelectr Freq Control 39(2):502–507. https://doi.org/10.1109/freq.1991.145883

    Article  Google Scholar 

  • Kukkonen I, Lindberg A (1998) Thermal properties of rocks at the investigation sites: measured and calculated thermal conductivity, specific heat capacity, and thermal diffusivity. Tech. rep, Geological Survey of Finland, Helsinki, Finland

  • Lakshtanov DL, Sinogeikin SV, Bass JD (2007) High-temperature phase transitions and elasticity of silica polymorphs. Phys Chem Miner 34(1):11–22. https://doi.org/10.1007/s00269-006-0113-y

    Article  Google Scholar 

  • Li BQ, Einstein HH (2017) Comparison of visual and acoustic emission observations in a four point bending experiment on Barre granite. Rock Mech Rock Eng 50(9):2277–2296

    Article  Google Scholar 

  • Lindholm U, Yeakley L, Nagy A (1974) The dynamic strength and fracture properties of dresser basalt. Int J Rock Mech Min Sci Geomech Abstr 11(5):181–191. https://doi.org/10.1016/0148-9062(74)90885-7

    Article  Google Scholar 

  • Mantovani JG, Tamasy GJ, Mueller RP et al (2016) Launch pad in a box. ASCE Earth & Space. Orlando, FL, pp 960–971

  • Marsaglia G (1972) Choosing a point from the surface of a sphere. Ann Math Stat 43(2):645–646

    Article  Google Scholar 

  • McNeil LE, Grimsditch M (1993) Elastic moduli of muscovite mica. J Phys Condens Matter 5(11):1681–1690. https://doi.org/10.1088/0953-8984/5/11/008

    Article  Google Scholar 

  • McSween HY, Wyatt MB, Gellert R et al (2006) Characterization and petrologic interpretation of olivine-rich basalts at Gusev Crater, Mars. J Geophys Res Planets 111(E02S10):1–17. https://doi.org/10.1029/2005JE002477

    Article  Google Scholar 

  • Mindeguia JC, Pimienta P, Noumowé A et al (2010) Temperature, pore pressure and mass variation of concrete subjected to high temperature-experimental and numerical discussion on spalling risk. Cement Concr Res 40(3):477–487

    Article  Google Scholar 

  • Ming DW (2016) private interview

  • Moore DE, Lockner DA (1995) The role of microcracking in shear-fracture propagation in granite. J Struct Geol 7(1):95–114. https://doi.org/10.1016/0191-8141(94)E0018-T

    Article  Google Scholar 

  • Morgan SP, Johnson CA, Einstein HH (2013) Cracking processes in Barre granite: fracture process zones and crack coalescence. Int J Fract 180(2):177–204

    Article  Google Scholar 

  • Nasseri M, Mohanty B, Young R (2006) Fracture toughness measurements and acoustic emission activity in brittle rocks. Pure Appl Geophys 163(5):917–945

    Article  Google Scholar 

  • National Research Council (2012) NASA space technology roadmaps and priorities: restoring NASA’s technological edge and paving the way for a new era in space. The National Academies Press, Washington, DC. https://doi.org/10.17226/13354. https://nap.nationalacademies.org/catalog/13354/nasa-space-technology-roadmaps-and-priorities-restoring-nasas-technological-edge

  • Navrotsky A (1995) Thermodynamic properties of minerals. Mineral physics and crystallography: a handbook of physical constants, AGU Ref Shelf, vol 2, pp 18–28

  • Ohno I, Harada K, Yoshitomi C (2006) Temperature variation of elastic constants of quartz across the \(\alpha\)–$\beta $ transition. Phys Chem Miner 33(1):1–9. https://doi.org/10.1007/s00269-005-0008-3

    Article  Google Scholar 

  • Panagouli O, Margaronis K, Tsotoulidou V (2020) A multiscale model for thermal contact conductance of rough surfaces under low applied pressure. Int J Solids Struct 200–201:106–118

    Article  Google Scholar 

  • Pletcher RH, Tannehill JC, Anderson D (2012) Computational fluid mechanics and heat transfer. CRC Press, Boca Raton

    Google Scholar 

  • Preston FW (1926) The spalling of bricks. J Am Ceram Soc 9(10):654–658. https://doi.org/10.1111/j.1151-2916.1926.tb17115.x

    Article  Google Scholar 

  • Preston FW, White HE (1934) Observations on spalling. J Am Ceram Soc 17(1–12):137–144. https://doi.org/10.1111/j.1151-2916.1934.tb19296.x

    Article  Google Scholar 

  • Pribnow D, Umsonst T (1993) Estimation of thermal conductivity from the mineral composition: influence of fabric and anisotropy. Geophys Res Lett 20(20):2199–2202. https://doi.org/10.1029/93GL02135

    Article  Google Scholar 

  • Rauenzahn RM (1986) Analysis of rock mechanics and gas dynamics of flame-jet thermal spallation drilling. PhD thesis, Massachusetts Institute of Technology, http://hdl.handle.net/1721.1/14884

  • Rauenzahn RM, Tester JW (1991a) Numerical simulation and field testing of flame-jet thermal spallation drilling—1. Model development. Int J Heat Mass Transf 34(3):795–808. https://doi.org/10.1016/0017-9310(91)90126-Y

    Article  Google Scholar 

  • Rauenzahn RM, Tester JW (1991b) Numerical simulation and field testing of flame-jet thermal spallation drilling—2. Experimental verification. Int J Heat Mass Transf 34(3):809–818. https://doi.org/10.1016/0017-9310(91)90127-Z

    Article  Google Scholar 

  • Rimoli J, Ortiz M (2010) A three-dimensional multiscale model of intergranular hydrogen-assisted cracking. Philos Mag 90(21):2939–2963

    Article  Google Scholar 

  • Rimoli J, Gürses E, Ortiz M (2010) Shock-induced subgrain microstructures as possible homogenous sources of hot spots and initiation sites in energetic polycrystals. Phys Rev B 81(1):014112

    Article  Google Scholar 

  • Robertson EC (1988) Thermal properties of rocks. Tech. Rep. 88-441, United States Department of the Interior

  • Robie RA, Hemingway BS, Fisher JR (1978) Thermodynamic properties of minerals and related substances at 298.15K and 1bar pressure and at higher temperatures. Tech. rep., US Government Printing Office

  • Rosenholtz JL, Smith DT (1941) Linear thermal expansion and inversion of quartz, var. rock crystal. Am Mineral 26(6):103–109

    Google Scholar 

  • Saksala T (2018) Numerical modelling of thermal spallation of rock. In: Proceedings of XII Argentine Congress on Computational Mechanics. Mecánica Computacional, Yerba Buena, Argentina, pp 1567–1574

  • Smith M (2009) ABAQUS/standard user’s manual, version 6.9. Dassault Systèmes Simulia Corp, United States

  • Smyth JR, McCormick TC (1995) Crystallographic data for minerals. In: Ahrens TJ (ed) Mineral physics and crystallography: a handbook of physical constants 2. American Geophysical Union, Washington, DC, pp 1–17. https://doi.org/10.1029/RF002

    Chapter  Google Scholar 

  • Song I, Suh M, Woo YK et al (2004) Determination of the elastic modulus set of foliated rocks from ultrasonic velocity measurements. Eng Geol 72(3–4):293–308. https://doi.org/10.1016/j.enggeo.2003.10.003

    Article  Google Scholar 

  • Sumino Y, Nishizawa O, Goto T et al (1977) Temperature variation of elastic constants of single-crystal forsterite between \(-\,190^\circ\) and 400$^\circ $C. J Phys Earth 25(4):377–392

    Article  Google Scholar 

  • Suzuki I (1975) Thermal expansion of periclase and olivine, and their anharmonic properties. J Phys Earth. https://doi.org/10.4294/jpe1952.23.145

    Article  Google Scholar 

  • Tijssens MG, Sluys BL, van der Giessen E (2000) Numerical simulation of quasi-brittle fracture using damaging cohesive surfaces. Eur J Mech A/Solids 19(5):761–779

    Article  Google Scholar 

  • Tutti F, Dubrovinsky LS, Nygren M (2000) High-temperature study and thermal expansion of phlogopite. Phys Chem Miner 27(9):599–603. https://doi.org/10.1007/s002690000098

    Article  Google Scholar 

  • Waeselmann N, Brown JM, Angel RJ et al (2016) The elastic tensor of monoclinic alkali feldspars. Am Mineral 101(5):1228–1231

    Article  Google Scholar 

  • Wahid SM, Madhusudana C (2000) Gap conductance in contact heat transfer. Int J Heat Mass Transf 43(24):4483–4487

    Article  Google Scholar 

  • Walsh SD, Lomov IN (2013) Micromechanical modeling of thermal spallation in granitic rock. Int J Heat Mass Transf 65:366–373. https://doi.org/10.1016/j.ijheatmasstransfer.2013.05.043

    Article  Google Scholar 

  • Walsh SD, Vogler D (2020) Simulating electropulse fracture of granitic rock. Int J Rock Mech Min Sci 128:104238. https://doi.org/10.1016/j.ijrmms.2020.104238

    Article  Google Scholar 

  • Wang Z, Palmer TA, Beese AM (2016) Effect of processing parameters on microstructure and tensile properties of austenitic stainless steel 304L made by directed energy deposition additive manufacturing. Acta Mater 110:226–235. https://doi.org/10.1016/j.actamat.2016.03.019

    Article  Google Scholar 

  • Wilkinson MA (1989) Computational modeling of the gas-phase transport phenomena and experimental investigation of surface temperatures during flame-jet thermal spallation drilling. PhD thesis, Massachusetts Institute of Technology. http://hdl.handle.net/1721.1/14031

  • Wilkinson MA, Tester JW (1993) Experimental measurement of surface temperatures during flame-jet induced thermal spallation. Rock Mech Rock Eng 26(1):29–62

    Article  Google Scholar 

  • Williams RE, Potter RM, Miska S (1996) Experiments in thermal spallation of various rocks. J Energy Resour Technol 118(1):2–8

    Article  Google Scholar 

  • Xia K, Nasseri M, Mohanty B et al (2008) Effects of microstructures on dynamic compression of Barre granite. Int J Rock Mech Min Sci 45(6):879–887

    Article  Google Scholar 

  • Yu QL, Ranjith PG, Liu HY et al (2015) A mesostructure-based damage model for thermal cracking analysis and application in granite at elevated temperatures. Rock Mech Rock Eng 48:2263–2282

    Article  Google Scholar 

  • Zhao Z, Radovitzky R, Cuitino A (2004) A study of surface roughening in FCC metals using direct numerical simulation. Acta Mater 52(20):5791–5804

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Julian J. Rimoli.

Additional information

Publisher's Note

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

Appendix A: Numerical Solution Process

Appendix A: Numerical Solution Process

1.1 A.1: 2D Plane Strain Solution

The model presented above is fully 3D; however, modern computers are limited such that a statistically representative volume element would be computationally intractable. The model is instead simplified to 2D plane strain, which is common practice in crack propagation models (Bouchard et al. 2003). The entire process described in this section is extensible to 3D, when computing power becomes available to model an RVE with a statistically significant number of grains.

1.2 A.2: Staggered One-Way Coupling

General solutions to the system of equations defined in Sect. 2.2 cannot be found analytically, so instead they are computed using direct numerical simulation (DNS). The material is discretized into finite elements, with unstructured triangular elements inside the grains and cohesive elements at the grain boundaries.

The equations presented in Sect. 2.2 show that the mechanical PDE is strongly influenced by temperature, but the thermal PDE is weakly influenced by displacement. In addition, mechanical wave propagation in igneous rocks is orders of magnitude higher than temperature propagation, so changes in stress occur nearly instantaneously compared to changes in temperature. Staggering the solution of these two equations avoids using an unnecessary quasi-static approach to solving the mechanical equations. Time is advanced through the thermal simulation, then a static mechanical analysis is performed at the end of each thermal step. A Crank–Nicolson scheme advances time in the thermal simulation, while the mechanical analysis is solving \(Ku=F\), as in standard FEM (Pletcher et al. 2012).

The simulations are performed by Abaqus, with a Python manager that prepares input files and processes output files from each time step. The flowchart for the manager is shown in Fig. 19. The two “step” blocks are calls to Abaqus, with the other blocks performed by the manager.

Fig. 19
figure 19

Staggered thermomechanical simulation of thermal spallation

1.3 A.3: Initialization

At the beginning of the simulation, mesh and job files are created from defined parameters for the run. These parameters include the magnitude of the heat flux, size of the domain, and the properties of the constituent materials. The first step is to run MicroStructPy (Hart and Rimoli 2020a, b), to create a domain with the same composition and grain size distributions as the material being simulated. Cohesive elements are inserted between elements at grain interfaces. A material section is created for each grain, and the material orientations of these sections are sampled from a uniform random distribution:

$$\begin{aligned}{} & {} q' \sim Z + Z i + Z j + Z k \end{aligned}$$
(24)
$$\begin{aligned}{} & {} q = \frac{q'}{||q'||} \end{aligned}$$
(25)

To uniformly sample orientations, the components of a quaternion are sampled from the standard normal distribution, Eq. 24, then the quaternion is normalized to have unit magnitude, Eq. 25 (Marsaglia 1972). An undirected graph of the elements is also generated, which is used in the mechanical post-processing step.

1.4 A.4: Time Advancement in Thermal Analysis

1.4.1 A.4.1: Pre-processor

The thermal pre-processor updates the thermal mesh and heat flux boundary condition. If it is the first step, \(i=0\), then there are no updates from the initial mesh and boundary conditions. The nodes, elements, and contact pairs in the thermal mesh are updated with the outputs from the \(i-1\) mechanical step. Nodal coordinates are updated with the displacements computed by the mechanical step. Elements are removed from the thermal mesh based on the mechanical post-processor output. When an entire grain is removed, then the contact pairs associated with it are also removed. The thermal job is also updated to reflect changes to the initial and boundary conditions. The initial temperature field of step i is set to the final temperature field of step \(i-1\).

1.4.2 A.4.2: Thermal Step

The thermal step takes the job file created by the thermal pre-processor and runs it in Abaqus. The temperature jump conditions in Eq. 6 are implemented with a gap conductance interaction between contact pairs. Since the nodal coordinates were updated by the pre-processor, gaps that are associated with failed cohesive elements do not conduct heat (\(\kappa =0\)).

The boundary conditions of the thermal model are adiabatic on all sides except the side with an applied heat flux. For example if heat is applied to the \(+z\) face of the domain, then the \(\pm x\), \(\pm y\), and \(-z\) faces are all adiabatic. Since many applications are nearly infinite in the \(-z\) direction, additional grains must be included in the simulation to avoid numerical artifacts in the simulation results. For the granite simulations below, this ghost layer is four times the average grain size.

The Crank–Nicolson algorithm is used to advance the temperature states in time. It is an unconditionally stable, implicit scheme. The temperature at each node is advanced from timestep i to \(i+1\) using central differences at timestep \(i+1\) to approximate the gradient in Eq. 1. For grid points on the ends of the mesh, a one-sided finite difference is used instead.

1.4.3 A.4.3: Post-processor

The thermal step results are post-processed after the step completes. The final temperature of each node is used as the initial condition in thermal step \(i+1\) and in the thermal strain for mechanical step i, which follows immediately after thermal step i.

1.5 A.5: Static Mechanical Analysis with Temperature Change

1.5.1 A.5.1: Mechanical Pre-processor

Similar to the thermal pre-processor, the mechanical pre-processor updates the mechanical mesh and job file. If \(i=0\), no updates are needed and the initial mesh and job are used in the mechanical step. The job file is updating by creating a *RESTART from the \(i-1\) mechanical step. Displacement, stress, and damage fields calculated for step \(i-1\) are loaded as initial conditions for step i. The initial temperature field for the mechanical job is taken from the end of step \(i-1\) and the final temperature field is from the end of step i. These temperatures set the material properties of the grains and their difference results in a thermal strain, \(\alpha _{kl} (T-T_0)\). This strain is applied in addition to the strains from the end of step \(i-1\).

Since the mechanical steps use the *RESTART option, a new mesh does not need to be created for each step. When spalls are removed from the mesh, *MODEL CHANGE steps are included to remove the elements and contact pairs associated with those spalls.

1.5.2 A.5.2: Mechanical Step

The mechanical step is a static finite element analysis that takes place instantaneously, at the end of step i. The domain is traction-free on the surface where heat is applied, and fixed displacement on the other boundaries. For example, the \(\pm x\) faces are fixed in x, \(\pm y\) faces are fixed in y, and \(-z\) face is fixed in z. The only load applied is the \(\Delta T\) at each node in the mesh, representing the change in temperature from the beginning of step i to the end.

The static finite element procedure solves the linear system of equations \(Ku=F\). Cohesive elements create additional degrees of freedom and force balancing equations compared to a model with only continuum elements. In this case, additional rows would be inserted into K, with stiffness defined in Eq. 8. Iteration on the static analysis is required, since the displacement field determines the damage state of the cohesive elements, the damage changes the stiffnesses in matrix K, and that matrix is used to solve for the displacements.

As in the thermal step, the \(-z\) boundary condition is artificial, imposed to limit the domain to a tractable size. To avoid numerical artifacts in the simulation results, a ghost layer of grains is added to the \(-z\) side of the domain. This ghost layer can be multiple grains wide, with the exact number of grains depending on their thermal diffusivity. The grain interfaces in this model are represented by cohesive elements and contact pairs. Abaqus cohesive elements do not enforce contact on their own, so contact is enforced separately with a pressure–overclosure relationship and no softening.

If there are spalls to be removed, *MODEL CHANGE steps are performed before the static finite element analysis. These steps remove the spalls identified at the end of job \(i-1\) before performing the job i static step.

1.6 A.6: Spall Identification and Removal

The results from the mechanical step are used to determine if a spall has formed and should be removed from the material. First, the cohesive elements with damage \(D\ge 0.8\) are removed from an undirected graph of the mesh elements. This value was chosen as a compromise between accuracy on the dissipated cohesive energy and the ability of DNS models containing damaged cohesive elements to converge. Allowing the damage to reach the value of 1 results in singular stiffness matrices that halt the simulation process. In our experience, the remaining energy under the curve in the TSL is vanishingly small after \(D=0.8\) , having minimal effect on the energetics of the problem while allowing the models to converge. It may be the case that not all of the cohesive elements surrounding a spall have failed, for example a single element with \(D < 0.8\). This is a numerical feature of DNS, rather than a realistic one, so an additional step is taken to identify and remove these cohesive elements.

Fig. 20
figure 20

Identification of under-damaged cohesive elements using shortest paths

As shown in Fig. 20, under-damaged cohesive elements are identified by finding the shortest paths between the continuum elements on each side of the fully damaged elements. These shortest paths are computed using the undirected graph, where the vertices of the graph are the elements and the edges are weighted by the distances between element centroids. Inspecting Fig. 20 visually, it is clear which cohesive elements should be removed. The shortest paths approach gives the job manager the ability to algorithmically identify these elements. If the under-damaged cohesive elements represent less than one-third of the total area of cohesive elements removed, then the spall is considered to be topologically separated from the main body. This threshold balances fidelity of the spall formation model with reliability of the DNS to converge to a solution. If the spall fully topologically separates from the main body, those elements become unconstrained and can prevent the solver from finding a converged displacement field.

Topological separation is a necessary, but not sufficient, condition for a spall to be removed from the material. If a grain in the middle of the material were to become topologically separated, it should not be removed, because it has no path away from the main body. To be removable, a spall must also satisfy a geometric condition. As shown in Fig. 21a, the surface attached to the removed cohesive elements has a set of normal vectors, \(n_{1-4}\). The mathematical condition that must be true to remove a spall with a given set of normal vectors, N, is that

$$\begin{aligned} \exists \,u \quad n_i\cdot u \ge 0 \quad \forall \,n_i \in N \end{aligned}$$
(26)

The linear system of inequalities \(n_i \cdot u \ge 0\) ensures that the pull direction, u, would not cause the spall to collide with the remaining material at any of its boundaries. From the expressions in Eq. 26, it is not immediately clear how to test for the existence of a pull direction, u. An equivalent condition to Eq. 26 is that u exists if N can be contained by a hemisphere. For a given u, the set of points on that sphere that have a positive dot product with u is a hemisphere. In Fig. 21b, the normal vectors are all contained in the \(+y\) hemisphere, so this spall can be removed. If, however, there were an \(n_5\) that was mostly in the \(+x\) direction with a small \(-y\) component, the set would still fall within a hemisphere, with a pull direction in the \((+x,+y)\) quadrant. To systematically test for the existence of a pull direction, the convex hull of the set N is taken.

Fig. 21
figure 21

Spall normal vectors and their convex hull

If the origin is contained within this convex hull, then N spans more than a hemisphere and a pull direction does not exist. On the other hand, if the convex hull does not contain the origin, then the set N spans less than a hemisphere, a pull direction u exists, and the spall is geometrically removable. This approach to testing if a spall can be removed is general for any number of dimensions and does not need to be modified for 3D simulations.

For time steps where multiple potential spalls have formed, the geometric conditions are modified if the spalls shared a grain boundary. If the spall in Fig. 20, for example, had a jagged grain boundary between the two halves, then each half would not have a pull direction. The union of the two halves is removable, but not each half individually. To catch for these cases, the algorithm first tests the unions of neighboring spalls, then tests them individually. If an individual spall is removable, the algorithm iterates to check if other spalls can be removed after that individual has been removed. For example, if the \(n_2\) grain boundary extended to divide the spall in Fig. 20 into two individuals, initially the spall on the left would be considered removable and the spall on the right would not. After removing the spall on the left, the spall on the right no longer has that confining surface, so it would be considered removable in the second iteration.

1.7 A.7: Reapplication of the Heat Flux

The heat flux applied to the top of the domain is updated to reflect the changes in nodal coordinates and element removal. The total heat applied to the domain during each thermal step is held constant, so if a surface element stretches or rotates then the heat applied to it must change. To compute the heat flux on each surface element, it is projected onto the plane perpendicular to the direction of the applied heat flux. For example, if heat is applied in the z-direction, each surface element is projected onto the xy plane. Once projected, the job manager checks for overlapping surface elements, then determines which elements to apply the heat flux to using ray-tracing. This ensures that the total heat flux remains constant, and that heat is only applied to those surfaces that are visible to the heat flux. For flame-jet spallation, heat would be applied, wherever the hot gas flows; however, an estimate for that flux value would be inaccurate without CFD.

1.8 A.8: Termination Conditions

After material has been removed from the domain, there is a check to determine if the model should terminate. If the top surface of the domain penetrates the ghost layer, then the model should terminate. This check prevents the domain from becoming small enough that numerical artifacts are introduced into the simulation results.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hart, K.A., Rimoli, J.J. A Computational Model for the Thermal Spallation of Crystalline Rocks. Rock Mech Rock Eng 56, 8235–8254 (2023). https://doi.org/10.1007/s00603-023-03498-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00603-023-03498-7

Keywords

Navigation