Skip to main content
Log in

A discrete element implementation for concrete: particle generation, contact calculations, packing under gravity and modeling material response

  • Original Paper
  • Published:
Granular Matter Aims and scope Submit manuscript

Abstract

A three-dimensional discrete element modelling capability for concrete based on rigid particles of arbitrary shape and size has been developed. The novel particle generation algorithm allows control of particle size, angularity and flakiness. General rigid body kinematics including finite rotations is accounted for, and an explicit time integration algorithm that conserves energy and momentum is implemented. An efficient contact algorithm with several features to increase the efficiency of the contact computations has been developed. This enables the gravity packing problem for arbitrary shaped particles to be solved in reasonable run time. The proposed procedure is used to generate assemblies of concrete specimens of various sizes that are homogeneous and isotropic in the bulk, and can capture the wall effect due to the formwork. The calibrated specimens are seen to be capable of accurately capturing experimentally observed macro stress strain response and failure patterns. The influence of aggregate shape on texture formation in the packed specimen, and on macro strength and failure patterns in hardened concrete, is demonstrated and is seen to be consistent with experimental results.

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
Fig. 19
Fig. 20
Fig. 21
Fig. 22
Fig. 23
Fig. 24
Fig. 25
Fig. 26
Fig. 27
Fig. 28

Similar content being viewed by others

References

  1. Alexander, M.G., Mindess, S., Diamond, S., Qu, L.: Properties of paste–rock interfaces and their influence on composite behaviour. Mater. Struct. 9(28), 497–506 (1995)

    Article  Google Scholar 

  2. Asahina, D., Landis, E.N., Bolander, J.E.: Modeling of phase interfaces during pre-critical crack growth in concrete. Cement Concr. Compos. 33(9), 966–977 (2011)

    Article  Google Scholar 

  3. Azéma, E., Radjai, F., Saussine, G.: Quasistatic rheology, force transmission and fabric properties of a packing of irregular polyhedral particles. Mech. Mater. 41(6), 729–741 (2009)

    Article  Google Scholar 

  4. Azéma, E., Radjai, F., Dubois, F.: Packings of irregular polyhedral particles: strength, structure, and effects of angularity. Phys. Rev. E 87(6), 062203 (2013)

    Article  ADS  Google Scholar 

  5. Bachmann, H., et al.: Vibration problems in structures. Birkhäuser Basel, Berlin (1995)

    Book  Google Scholar 

  6. Beckmann, B., Schicktanz, K., Reischl, D., Curbach, M.: DEM simulation of concrete fracture and crack evolution. Struct. Concr. 13(4), 213–220 (2012)

    Article  Google Scholar 

  7. Bureau of Indian Standards. IS: 383-1970. Specification for Coarse and Fine Aggregates from Natural sources for Concrete, 2nd edn

  8. Bolander, J.E., Saito, S.: Fracture analyses using spring networks with random geometry. Eng. Fract. Mech. 61(5–6), 569–591 (1998)

    Article  Google Scholar 

  9. Brown, N.J., Chen, J.F., Ooi, J.Y.: A bond model for DEM simulations of cementitious materials and deformable structures. Granul. Matter 16(3), 299–311 (2014)

    Article  Google Scholar 

  10. Camborde, F., Mariotti, C., Donze’, F.V.: Numerical study of rock and concrete behaviour by discrete element modelling. Comput. Geotech. 27(4), 225–247 (2000)

    Article  Google Scholar 

  11. Chang, S.W., Chen, C.S.: A non-iterative derivation of the common plane for contact detection of polyhedral blocks. Int. J. Numer. Methods Eng. 74(5), 734–753 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  12. Cho, G.C., Dodds, J., Santamarina, J.C.: Particle shape effects on packing density, stiffness, and strength: natural and crushed sands. J. Geotech. Geoenviron. Eng. 132(5), 591–602 (2006)

    Article  Google Scholar 

  13. Cui, L., O’sullivan, C., O’neill, S.: An analysis of the triaxial apparatus using a mixed boundary three-dimensional discrete element model. Geotechnique 57(10), 831–844 (2007)

    Article  Google Scholar 

  14. Cundall, P.A., Strack, O.D.L.: A discrete numerical model for granular assemblies. Geotechnique 29(1), 47–65 (1979)

    Article  Google Scholar 

  15. Cundall, P.A.: Formulation of a three-dimensional distinct element model—Part I. A scheme to detect and represent contacts in a system composed of many polyhedral blocks. Int. J. Rock Mech. Min. Sci. Geomech. Abstr. 25(3), 107–116 (1988)

    Article  Google Scholar 

  16. Cusatis, G., Pelessone, D., Mencarelli, A.: Lattice discrete particle model (LDPM) for failure behavior of concrete I: theory. Cement Concr. Compos. 33(9), 881–890 (2011)

    Article  Google Scholar 

  17. Cusatis, G., Mencarelli, A., Pelessone, D., Baylot, J.: Lattice discrete particle model (LDPM) for failure behavior of concrete II: calibration and validation. Cement Concr. Compos. 33(9), 891–905 (2011)

    Article  Google Scholar 

  18. d’Addetta, G.A., Kun, F., Ramm, E.: On the application of a discrete model to the fracture process of cohesive granular materials. Granul. Matter 4(2), 77–90 (2002)

    Article  MATH  Google Scholar 

  19. De Josselin, De Jong, G., Verruijt, A.: Etude photo-elastique d’un empilement de disques. Cah. Groupe Fr. Rheol. 2(1), 73–86 (1969)

    Google Scholar 

  20. Escoda, J., Jeulin, D., Willot, F., Toulemonde, C.: Three-dimensional morphological modelling of concrete using multiscale Poisson polyhedra. J. Microsc. 258(1), 31–48 (2015)

    Article  Google Scholar 

  21. Golub, G.H., Matt, U.V.: Quadratically constrained least squares and quadratic problems. Numer. Math. 59(1), 561–580 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  22. Gyurkó, Z., Bagi, K., Borosnyói, A.: Discreteelement modelling of uniaxial compression test of hardened concrete. J. Silic. Based Compos. Mater. 66(4), 113–119 (2014)

    Google Scholar 

  23. Hart, R., Cundall, P.A., Lemos, J.: Formulation of a three-dimensional distinct element model—Part II. Mechanical calculations for motion and interaction of a system composed of many polyhedral blocks. Int. J. Rock Mech. Min. Sci. Geomech. Abstr. 25(3), 117–125 (1988)

    Article  Google Scholar 

  24. Hentz, S., Daudeville, L., Donzé, F.V.: Identification and validation of a discrete element model for concrete. J. Eng. Mech. 130(6), 709–719 (2004)

    Article  Google Scholar 

  25. Hordijk, D.A.: Local approach to fatigue of concrete. Ph.D Thesis, Delft University of Technology, Delft, The Netherlands (1991)

  26. Huang, H.: Discrete element modeling of railroad ballast using imaging based aggregate morphology characterization. Ph.D. Thesis, University of Illinois at Urbana-Champaign, Urbana, IL (2010)

  27. Jensen, R.P., Edil, T.B., Bosscher, P.J., Plesha, M.E., Kahla, N.B.: Effect of particle shape on interface behavior of DEM-simulated granular materials. Int. J. Geomech. 1(1), 1–19 (2001)

    Article  Google Scholar 

  28. Jivkov, A.P., Yates, J.R.: Elastic behaviour of a regular lattice for meso-scale modelling of solids. Int. J. Solids Struct. 49(22), 3089–3099 (2012)

    Article  Google Scholar 

  29. Kaplan, M.F.: The effects of the properties of coarse aggregates on the workability of concrete. Mag. Concr. Res. 10(29), 63–74 (1958)

    Article  Google Scholar 

  30. Kim, Y., Souza, L.T.: Effects of aggregate angularity on mix design characteristics and pavement performance. Department of Civil Engineering, University of Nebraska-Lincoln, Technical report; MPM-10 (2009)

  31. Kumar, C.N.S., Rao, T.D.: Fracture parameters of high strength concrete Mode II testing. Mag. Concr. Res. 62(3), 157–162 (2010)

    Article  Google Scholar 

  32. Lee, H., Kwon, J., Cho, H.: Simulation of impact breakage of concrete blocks using DGB. Geosyst. Eng. 10(2), 31–36 (2007)

    Article  Google Scholar 

  33. Liao, C.L., Chang, T.P., Young, D.H.: Stress–strain relationship for granular materials based on the hypothesis of best fit. Int. J. Solids Struct. 34(31–32), 4087–4100 (1997)

    Article  MATH  Google Scholar 

  34. Liu, J.X., Deng, S.C., Zhang, J., Liang, N.G.: Lattice type of fracture model for concrete. Theoret. Appl. Fract. Mech. 48(3), 269–284 (2007)

    Article  Google Scholar 

  35. Lu, M., McDowell, G.R.: The importance of modelling ballast particle shape in the discrete element method. Granul. Matter 9(1–2), 69–80 (2007)

    Google Scholar 

  36. Man, H.K., van Mier, J.G.M.: Damage distribution and size effect in numerical concrete from lattice analyses. Cement Concr. Compos. 33(9), 867–880 (2011)

    Article  Google Scholar 

  37. Mechtcherine, V., Shyshko, S.: Simulating the behaviour of fresh concrete with the distinct element method-deriving model parameters related to the yield stress. Cement Concr. Compos. 55, 81–90 (2015)

    Article  Google Scholar 

  38. Meguro, K., Hakuno, M.: Fracture analyses of concrete structures by the modified distinct element method. Struct. Eng. Earthq. Eng. 6(2), 283–294 (1989)

    Google Scholar 

  39. Moukarzel, C., Herrmann, H.J.: A vectorizable random lattice. J. Stat. Phys. 68(5), 911–923 (1992)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  40. Murdock, L.J.: The workability of concrete. Mag. Concr. Res. 12(36), 135–144 (1960)

    Article  Google Scholar 

  41. Muth, B., Of, G., Eberhard, P., Steinbach, O.: Collision detection for complicated polyhedra using the fast multipole method or ray crossing. Arch. Appl. Mech. 77(7), 503–521 (2007)

    Article  ADS  MATH  Google Scholar 

  42. PFC 3D Version 5, Computer Software: General Purpose distinct element framework. 2014, Itasca Consulting Group, Minneapolis

  43. Potyondy, D., Cundall, P.A., Lee, C.A.: Modelling rock using bonded assemblies of circular particles. In: Proceedings of the Second North American Rock Mechanics Symposium (1996); Montréal, 1937–1944

  44. Quiroga, P.N., Fowler, D.W.: The effects of aggregates characteristics on the performance of the Portland cement concrete. Report ICAR 104–1F 2003, University of Texas, Austin, Texas

  45. Rao, G.A., Prasad, B.K.R.: Influence of interface properties on fracture behavior of concrete. Sadhana 36(2), 193–208 (2011)

    Article  MathSciNet  Google Scholar 

  46. Reinhardt, H.W., Xu, S.: A practical testing approach to determine mode II fracture energy GIIF for concrete. Int. J. Fract. 105(2), 107–125 (2000)

    Article  Google Scholar 

  47. Rocco, C.J., Elices, M.: Effect of aggregate shape on the mechanical properties of a simple concrete. Eng. Fract. Mech. 76, 286–298 (2009)

    Article  Google Scholar 

  48. Rycroft, C.H.: VORO++: a 3D Voronoi cell library in C++. Chaos 19, 041111 (2009)

    Article  ADS  Google Scholar 

  49. Schlangen, E., Garboczi, E.J.: Fracture simulations of concrete using lattice models: computational aspects. Eng. Fract. Mech. 57(2/3), 319–332 (1997)

    Article  Google Scholar 

  50. Schlangen, E., van Mier, J.G.M.: Experimental and numerical analysis of micro mechanisms of fracture of cement-based composites. Cement Concr. Compos. 14(2), 105–118 (1992)

    Article  Google Scholar 

  51. Sherif, H.A., Kossa, S.S.: Relationship between normal and tangential contact stiffness of nominally flat surface. Wear 151, 49–62 (1991)

    Article  Google Scholar 

  52. Shiu, W.J., Donzé, F., Daudeville, L.: Influence of the reinforcement on penetration and perforation of concrete targets: a discrete element analysis. Eng. Comput. 26(1–2), 29–45 (2009)

    Article  MATH  Google Scholar 

  53. Shyshko, S., Mechtcherine, V.: Developing a discrete element model for simulating fresh concrete: experimental investigation and modelling of interactions between discrete aggregate particles with fine mortar between them. Constr. Build. Mater. 47, 601–615 (2013)

    Article  Google Scholar 

  54. Simo, J.C., Wong, K.K.: Unconditionally stable algorithms for rigid body dynamics that exactly preserve energy and momentum. Int. J. Numer. Methods Eng. 31(1), 19–52 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  55. Smeets, B., Odenthal, T., Vanmaercke, S., Ramon, H.: Polygon-based contact description for modeling arbitrary polyhedra in the discrete element method. Comput. Methods Appl. Mech. Eng. 290, 277–289 (2015)

    Article  ADS  MathSciNet  Google Scholar 

  56. Su, R.K., Bei, C.: The effect of coarse aggregate size on the stress–strain curves of concrete under uniaxial compression. Hong Kong Inst. Eng. Trans. 15(3), 33–39 (2008)

    Google Scholar 

  57. Tavarez, F.A., Plesha, M.E.: Discrete Element method for modelling solid and particulate materials. Int. J. Numer. Methods Eng. 70(4), 379–404 (2007)

    Article  MATH  Google Scholar 

  58. Tran, V.T., Donzé, F.V., Marin, P.: A discrete element model of concrete under high triaxial loading. Cement Concr. Compos. 33(9), 936–948 (2011)

    Article  Google Scholar 

  59. Tran, V.D., Uji, K., Ueno, U., Ohno, K., Wang, B.: Evaluation of shear bond test methods of concrete repair. In: Dehn et al.(ed.) Concrete Repair, Rehabilitation and Retrofitting IV. Taylor & Francis Group, London (2016)

  60. van Mier, J.G.M. et al.: Strain softening in concrete in uniaxial compression. Materials and Structures (1997); RILEM 148-SSC, 30, 195–209

  61. van Mier, J.G.M., van Vliet, M.R., Wang, T.K.: Fracture mechanisms in particle composites: statistical aspects in lattice type analysis. Mech. Mater. 34(11), 705–724 (2002)

    Article  Google Scholar 

  62. Van Vliet, M.R.A., van Mier, J.G.M.: Softening behavior of concrete under uniaxial compression. In: Wittmann, F.H. (ed.)Proceedings FracMCos-2: Fracture Mechanics of Concrete Structures. AEDIFICATIO Publishers, Freiburg (1995)

  63. Van Vliet, M.R.A., van Mier, J.G.M.: Experimental investigation of size effect in concrete and sandstone under uniaxial tension. Eng. Fract. Mech. 65(2–3), 165–188 (2000)

    Article  Google Scholar 

  64. Walton, O.R.: Simulation of gravity flow and packing of spheres. In: Proceedings of the Nisshin Engineering Powder Technology International Seminar (NEPTIS-1), Osaka, Japan (1993)

  65. Wischers, G., Lusche, M.: Influence of internal stress distribution on the behaviour of normal weight and light weight concrete in compression. Beton-technische Berichte 8(22), 343–347 (1972)

    Google Scholar 

  66. Xu, R., Yang, X.H., Yin, A.Y., Yang, S.F., Ye, Y.: A three-dimensional aggregate generation and packing algorithm for modeling asphalt mixture with graded aggregates. J. Mech. 26(2), 165–171 (2010)

    Article  Google Scholar 

  67. Yang, W., Zhou, Z., Pinson, D., Yu, A.: Periodic boundary conditions for discrete element method simulation of particle flow in cylindrical vessels. Ind. Eng. Chem. Res. 53(19), 8245–8256 (2014)

    Article  Google Scholar 

  68. Zhang, D., Huang, X., Zhao, Y.: Algorithms for generating three-dimensional aggregates and asphalt mixture samples by the discrete-element method. J. Comput. Civil Eng. 27(2), 111–117 (2013)

    Article  Google Scholar 

  69. Zhang, M., Morrison, C.N., Jivkov, A.P.: A meso-scale site bond model for elasticity: theory and calibration. Mater. Res. Innov. 18(S2), 982–986 (2014)

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Arghya Deb.

Ethics declarations

Conflict of Interest

The authors declare that they have no conflict of interest.

Appendices

Appendix A: Contact algorithm: details of modifications

1.1 A.1 Finding the intersection between the valid regions for any two vertices in 3D

The valid region for vertex i is formed by the ‘bounding planes’ of the vertex, each bounding plane being formed by the normals of the adjacent faces of particle A that meet at vertex i. These normals thus comprise the ‘bounding vectors’ of the bounding planes. The valid region for vertex j can be similarly determined from the normals of the faces of particle B that meet at vertex j (Fig. 29a, b). The intersection of the valid regions of vertex i belonging to particle A and vertex j belonging to particle B may just be the valid region of vertex i if the valid region of vertex i is contained entirely within the valid region of vertex j, or vice versa. This is the case illustrated in the top figure in Fig. 29c.

Another possibility is that the two valid regions actually intersect (bottom figure in Fig. 29c). In such a case, the lines of intersection between the bounding planes defining the valid region of vertex i and the bounding planes defining the valid region of vertex j have to be found. For an intersection to be valid, the line of intersection has to lie within the valid region of both the planes (Fig. 30). Once the vector defining the line of intersection has been determined, this vector replaces one of the two bounding vectors for both the bounding planes being considered. Thus when a bounding plane of j finds valid intersections with a bounding plane of i, it has to be determined which of the two bounding vectors defining the bounding plane of j will remain within the intersection region and which of the two bounding vectors has to be discarded. Simple geometric tests were developed to do this.

Fig. 29
figure 29

Intersection region between the valid regions of vertex i, particle A and vertex j, particle B. a Valid region of A at i, b valid region of B at j, c intersection between valid regions at i at j

Fig. 30
figure 30

Lines in darker outline in b show boundaries of valid region obtained using line of intersection and retained bounding vectors of planes ‘a’ and ‘b

1.2 A.2 Kinetic energy based local tracking algorithm

A kinetic energy based algorithm was developed to determine the optimum interval \(t_{2}\)\(t_{1}\) between global trackings. It involves performing detailed checks, as described in the flow chart in Fig. 31, to determine the optimum size of the interval \([t_{1}, t_{2}]\) independently for each particle involved in a contact interaction. Once this has been done for both particles, the minimum of the two independently determined optimum intervals is taken to be the actual time interval between successive global trackings for that contact.

Since the flow chart is self-explanatory only basic features of the algorithm are discussed here. If the total kinetic energy of the system at time t is substantially less than the peak value of kinetic energy attained over the entire loading history up to time t, then the system as a whole is close to a state of static equilibrium. In such a situation, in order to determine whether the configuration of a particular particle is changing sufficiently slowly, one only needs to examine the history of the kinetic energy of that particle (Conditions I and II in the flow chart). On the other hand when the total kinetic energy of the system is significant, one has to compare the kinetic energy of a particular particle with the kinetic energy of the fastest moving particle in the system in order to determine whether that particle is truly a slow moving particle (Conditions III and IV in the flow chart). Examining the time history of the kinetic energy of individual particles is then not sufficient, since the kinetic energy of a particle may have reduced substantially from its peak kinetic energy, but compared to the other particles in the system, it may still be a relatively fast moving particle. In addition, special provision is made for interactions between particle pairs that are ‘dormant’. Dormant interactions are interactions where two particles are sufficiently close to each other to require that their configurations be tracked to determine whether they are coming in contact. However, their relative velocities are so small that it is unlikely that this would happen at any proximate time. Many such interactions are found to exist during the packing phase of the computations, and monitoring such interactions at every increment to determine if they are near to establishing contact is a significant computational burden. This can be reduced if the dormant interactions are identified. If the kinetic energies of both particles in an interaction are very low compared to the kinetic energy of the fastest moving particle in the system, and the gap between the particles is significant, then such particle pairs are flagged as dormant and contact detection computations are only performed after a significantly large number \((n_{{\textit{dormant}}})\) of time increments. The kinetic energy of each dormant pair is however monitored every increment; if there is a substantial increase in the kinetic energy of either particle in the dormant pair then the pair is reactivated i.e. contact detection computations are performed every increment.

The default values of the parameters used in the algorithm are given in Fig. 31. These values were found to be robust for initial particle impact velocities in the range of 0.25-3 m/s and particle masses in the range of \(1\times 10^{-5}\) to \(1\times 10^{-1}\) kg, which were thought to be typical of the concrete casting (packing) process. However the parameter values are independent of the logical structure of the algorithm, and may be modified if required.

Fig. 31
figure 31

Kinetic energy based algorithm to find intervals between global tracking

Fig. 32
figure 32

Contact point calculation. a Multiple penetrations at nodes ijk and l, b increased penetration at nodes k and l due to penetration reduction at i and j, c single contact point at centroid of all penetrated nodes

1.3 A.3 Determining a single contact point between contacting particles

Once the common plane has been determined, all vertices of the slave particle that are within a distance \(d_{{\textit{gap}}}\) (Sect. 3.3) of the common plane are found, and are treated as vertices that have positive penetration. One possibility is to enforce contact at the vertex of the slave particle A, say vertex i, that has the largest penetration. This works reasonably well when there are no other vertices of particle A that have positive penetration. However in case there are additional vertices of particle A that have also penetrated, but have penetrations that are smaller than that of vertex i, this leads to undesirable behavior. This can be seen from Fig. 32, where four vertices of slave particle A, say vertices ijk and l have positive penetrations. Vertices i and j have penetrated by the same amount and their penetrations are larger than that of vertices k and l (Fig. 32a). If the contact point is taken to be the centroid of vertices i and j, and the contact force on particle A applied at this location, particle A rotates, resulting in increased penetration at vertices k and l (Fig. 32b). Since the objective is to reduce the penetration of all the vertices ijk and l, this is not desirable. To prevent this from happening, the contact point is chosen to be the centroid of all vertices of slave particle A that have positive penetrations i.e. the centroid of the vertices ijk and l (Fig. 32b). This ensures that all these vertices experience a translational force pushing them apart from the common plane.

To understand the differences with [11] one may consider the situation depicted in Fig. 32a. Chang and Chen [11] would treat this case as an ‘edge’ contact between slave edge ij and the master. Equal contact forces would act at the vertices i and j pushing the two particles apart. In the current algorithm however, a single contact force would act at the centroid (Fig. 32c). If, as in Fig. 32c, the penetration of the centroid is smaller than that of the vertices i and j, the magnitude of the contact force computed using the current algorithm would be somewhat smaller. Hence it may take a larger number of increments to resolve the penetration of edge ij. However, unlike [11], the current algorithm would not result in a reduction in the gap between the common plane and the vertices k and l (Fig. 32b).

While the above comparison is for a particular situation, the current approach seems to work well overall. There is no increase in penetration in the long run as compared to [11] (cf. Fig.13), and the savings in computational time are enormous.

Appendix B: Rotary scaling factor for very flaky/elongated particles

The following procedure ensures that flaky particles in the assembly, with very small values of rotary inertia about their axes of maximum elongation, do not give rise to very low values of \(\Delta t_{{\textit{stable}}}\).

The largest eigen value for the undamped interactions of each two particle system is found by solving the eigen problem:

$$\begin{aligned}&{[}\tilde{{{\varvec{M}}}} -\omega ^{2} {\tilde{{{\varvec{K}}}}}{]} {\varvec{\uppsi }}=\mathbf{0 }\nonumber \\&{\tilde{{{\varvec{M}}}}} =\left[ {{\begin{array}{cccc} \mathbf{M}_\mathrm{A} &{} \mathbf{0 }&{} \mathbf{0 }&{} \mathbf{0 } \\ \mathbf{0 }&{} \mathbf{M}_\mathrm{B}&{} \mathbf{0 }&{} \mathbf{0 } \\ \mathbf{0 }&{} \mathbf{0 }&{} \mathbf{I}_\mathrm{A}&{} \mathbf{0 } \\ \mathbf{0 }&{} \mathbf{0 }&{} \mathbf{0 }&{} \mathbf{I}_\mathrm{B} \\ \end{array}}}\right] {\tilde{{{\varvec{K}}}}} =\left[ {{\begin{array}{cccc} \mathbf{K}&{} {-\mathbf{K}}&{} {\mathbf{K}{\hat{\mathbf{r}}}_\mathrm{A}^\mathrm{T} }&{} -\mathbf{K}{\hat{\mathbf{r}}}_\mathrm{B}^\mathrm{T} \\ {-\mathbf{K}}&{} \mathbf{K}&{} {-\mathbf{K} {\hat{\mathbf{r}}}_\mathrm{A}^\mathrm{T} }&{} \mathbf{K}{\hat{\mathbf{r}}} _\mathrm{B}^\mathrm{T} \\ \mathbf{K}{\hat{\mathbf{r}}}_\mathrm{A} &{} -\mathbf{K}{\hat{\mathbf{r}}} _\mathrm{A} &{} {\hat{\mathbf{r}}}_\mathrm{A} \mathbf{K}{\hat{\mathbf{r}}}_\mathrm{A}^\mathrm{T} &{} -{\hat{\mathbf{r}}} _\mathrm{A} \mathbf{K}{\hat{\mathbf{r}}}_\mathrm{B}^\mathrm{T} \\ -\mathbf{K}{\hat{\mathbf{r}}} _\mathrm{B} &{} \mathbf{K}{\hat{\mathbf{r}}} _\mathrm{B} &{} \hbox {}-{\hat{\mathbf{r}}} _\mathrm{B} \mathbf{K}{\hat{\mathbf{r}}} _\mathrm{A}^\mathrm{T} &{} {\hat{\mathbf{r}}} _\mathrm{B} \mathbf{K}{\hat{\mathbf{r}}}_\mathrm{B}^\mathrm{T} \\ \end{array}}} \right] .\nonumber \\ \end{aligned}$$
(B.1)

\(\tilde{{{\varvec{M}}}}\) and \(\tilde{{{\varvec{K}}}}\) are the generalized mass and stiffness matrices of dimension \(12\times 12\) and \({\varvec{\uppsi }}\) and \(\omega \) are the eigenvectors and eigen values. The generalized mass matrix \(\tilde{{{\varvec{M}}}}\) contains contributions from the mass as well as the rotary inertias of the two particles, A and B, and is evaluated at their respective centers of mass. \(\mathbf{M}_{\mathrm{A}}\) and \(\mathbf{M}_{\mathrm{B}}\) are \(3\times 3\) diagonal matrices. \(\mathbf{M}_{\mathrm{A}}\) is the mass matrix of particle A, with mass \(M_{A}\) and \(\mathbf{M}_{\mathrm{B}}\) is the mass matrix of particle B with mass \(M_{B}\). \(\mathbf{I}_{\mathrm{A}}\) and \(\mathbf{I}_{\mathrm{B}}\) are the \(3\times 3\) rotary inertia matrices for particles A and B. The generalized stiffness matrix \(\tilde{{{\varvec{K}}}}\) involves \(3\times 3\) skew symmetric matrices \({\hat{\mathbf{r}}}_{\mathrm{A}}\) and \({\hat{\mathbf{r}}}_\mathrm{B}\), whose axial vectors are the vectors \(\mathbf{r}_{\mathrm{A}}\) and \(\mathbf{r}_{\mathrm{B}}\) (Fig. 33) as well as the diagonal matrix K whose diagonal term is the normal spring stiffness \(K_{n}\).

Fig. 33
figure 33

Spring-mass-rotary inertia System for two interacting particles

The equation governing the free vibration of the spring-mass-rotary inertia system with only translational degrees of freedom active is given by:

$$\begin{aligned} \left[ {\begin{array}{cc} \mathbf{M}_\mathrm{A} &{} 0 \\ 0&{} \mathbf{M}_\mathrm{B} \\ \end{array}} \right] \left[ {{\begin{array}{cc} {\ddot{\mathbf{x}}}_\mathrm{A} \\ {\ddot{\mathbf{x}}}_\mathrm{B} \\ \end{array}}}\right] +\left[ {{\begin{array}{cc} \mathbf{K}&{}-\mathbf{K}\\ -\mathbf{K}&{}\mathbf{K}\\ \end{array}}}\right] \left[ {{\begin{array}{cc} \mathbf{X}_\mathrm{A}\\ \mathbf{X}_\mathrm{B}\\ \end{array}}}\right] =\mathbf{0 } \end{aligned}$$
(B.2)

where \(\mathbf{x}_{\mathrm{A}}\) and \(\mathbf{x}_{\mathrm{B}}\) are the spatial degrees of freedom at the center of mass of particles A and B respectively. The eigen value problem corresponding to Eq. (B.2) is solved for all interacting particles to obtain an estimate of the stable time increment of the translation-only system. If this time increment is denoted as \(\Delta t_{stable}^{trans-only}\), the goal is to scale the rotary inertias \(\mathbf{I}_\mathrm{A}\) and \(\mathbf{I}_\mathrm{B}\) such that the scaled rotary inertias \(\mathbf{I}_\mathrm{A}^{scaled}\) and \(\mathbf{I}_\mathrm{B}^{scaled} \) are such that the largest eigen value of the scaled system, \(\omega _{max}^{scaled}\), is less than or equal to the largest eigen value of the translation-only system, \(\omega _{\max }^{{\textit{trans}}-{\textit{only}}}\), i.e.

$$\begin{aligned} \omega _{max}^{scaled} \le \omega _{max}^{trans-only} =2/{\Delta }t_{stable}^{trans-only} \end{aligned}$$
(B.3)

In order to obtain a usable bound on the rotary inertias using the above condition, the following conservative simplifications are made. \({\hat{\mathbf{r}}}_{\mathrm{A}}\) and \( {\hat{\mathbf{r}}}_\mathrm{B}\) in \(\tilde{{{\varvec{K}}}}\hbox { are}\) replaced by the following upper bound matrix r:

$$\begin{aligned} \mathbf{r}=\hbox {}\left[ {{\begin{array}{ccc} 0&{}\quad {-d}&{}\quad d \\ d&{}\quad 0\quad &{}\quad {-d} \\ -d&{}\quad d&{}\quad 0 \\ \end{array}}} \right] ;\Vert \mathbf{r}\Vert _\infty \ge \hbox {max}\left( \Vert {\hat{\mathbf{r}}} _\mathrm{A}\Vert _\infty ,\,\,\Vert {\hat{\mathbf{r}}}_\mathrm{B}\Vert _\infty \right) \nonumber \\ \end{aligned}$$
(B.4)

and \(\mathbf{I}_\mathrm{A}\) and \(\mathbf{I}_\mathrm{B}\) in \(\tilde{{{\varvec{M}}}}\) are replaced by the following lower bound:

$$\begin{aligned} \mathbf{I}=\left[ {{\begin{array}{lll} I&{}\quad 0&{}\quad 0 \\ 0&{}\quad I&{}\quad 0 \\ 0&{}\quad 0&{}\quad I \\ \end{array} }} \right] ;\quad \Vert \mathbf{I}\Vert _\infty \le \mathbf{min} \left( \Vert {{\varvec{I}}}_{{\varvec{A}}}\Vert _\infty ,\Vert {{\varvec{I}}}_{{\varvec{B}}}\Vert _\infty \right) \end{aligned}$$
(B.5)

In Eq. (B.4) d is the effective radius of the larger of the two particles A and B, and in Eq. (B.5) the isotropic rotary inertia, I, is defined such that its scalar component \(I\hbox { is equal to }{\textit{min}}\left( {{\varvec{\omega }}_{\mathbf{I}_\mathrm{A} } , {\varvec{\omega }}_{\mathbf{I}_\mathrm{B}}}\right) \) where \({\varvec{\omega }}_{\mathbf{I}_\mathrm{A}}\) and \({\varvec{\omega }}_{\mathbf{I}_\mathrm{B}}\) represent the spectrum of \(\mathbf{I}_\mathrm{A}\) and \(\mathbf{I}_\mathrm{B}\) respectively. \(\omega _{max}^{scaled}\) can then be calculated by solving the eigen value problem given by Eq. (B.1) with \(\tilde{{{\varvec{K}}}}\) and \(\tilde{{{\varvec{M}}}}\) modified as above:

$$\begin{aligned} \omega _{max}^{scaled} =\sqrt{\frac{K_n (M_A +M_B )}{M_A M_B }+\frac{6K_n d^{2}}{I}} \end{aligned}$$
(B.6)

The corresponding bound on I is then obtained by enforcing Eq. (B.3):

$$\begin{aligned} I\ge 6K_n d^{2}\big /\left[ 4/\left( {\varDelta t_{stable}^{trans-only}} \right) ^{2}-\hbox {}K_n (M_A +M_B )/M_A M_B \right] \nonumber \\ \end{aligned}$$
(B.7)

The rotary inertias \(\mathbf{I}_\mathrm{A}\) and \(\mathbf{I}_\mathrm{B}\) are scaled to ensure that their smallest eigen values satisfy this lower bound, thereby ensuring that the stability of the system is governed by \(\Delta t_{stable}^{trans-only}\).

Inclusion of viscous damping leads to a reduction in the value of \(\Delta t_{stable}^{trans-only}\) since the largest frequency of the constrained spring mass system described by Eq. (B.2) is affected by the presence of the viscous damping term \(\mathbf{C}\left[ {{\begin{array}{l} {\dot{\mathbf{x}}}_\mathrm{A} \\ {\dot{\mathbf{x}}}_\mathrm{B} \\ \end{array}}}\right] \). The reduced stable time increment for the damped system is then given by:

$$\begin{aligned} \left( \Delta t_{stable}^{trans-only}\right) _{damped} =\Delta t_{stable}^{trans-only} \left( {\sqrt{1+\xi ^{2}}-\xi }\right) \nonumber \\ \end{aligned}$$
(B.8)

where \(\xi \) is the ratio of the damping coefficient to its critical value for the spring-mass-damper system with only translational degrees of freedom active. \((\Delta t_{stable}^{trans-only} )_{damped}\) is then used instead of \(\Delta t_{stable}^{trans-only}\) in Eq. (B.7).

Appendix C: Calibration of meso elastic parameters for hardened concrete

The meso moduli \(K_n^*\hbox { and }K_s^*\) must be calibrated to replicate the elastic response at the macro level, determined by the macro deformability parameters, E (Young’s modulus) and \(\upsilon \) (Poisson’s ratio). The functions \(E\left( {K_n^*,K_s^*}\right) \) and \(\nu \left( {K_n^*,K_s^*}\right) \) are not known for arbitrary polyhedral particles. Trying to obtain them through regression analysis from experimental data requires performing, for every DEM assembly, a large number of experiments to obtain reasonable fits. An alternative calibration procedure was therefore developed. The calibration procedure starts with assuming values of the meso-level material parameters \(K_n^{*} \hbox { and }K_s^*\), say \((K_n^*)^{0}\hbox { and }(K_s^*)^{0}\). Two simple load tests that are independent of each other, e.g. uniaxial compression and triaxial compression, are then performed on the DEM assembly. The elastic strain energy of the DEM assembly \(W_{DEM}^I ,I=1,2\) being the number of independent tests, is then calculated using the assumed values \((K_n^*)^{0}\hbox { and }(K_s^*)^{0}\), and the normal and shear contact forces \(F_k^n\) and \(F_k^s\) at each contact interaction k:

$$\begin{aligned} 2W_{DEM}^I= & {} \left( {1/(K_n^{*} )^{0}} \right) \mathop \sum \nolimits _{k=1}^N \left[ {(F_k^n )^{2}/A_{\mathrm{int}}^k } \right] \nonumber \\&+\left( {1/(K_s^{*} )^{0}} \right) \mathop \sum \nolimits _{k=1}^N \left[ {(F_k^s )^{2}/A_{\mathrm{int}}^k } \right] \end{aligned}$$
(C.1)

It is assumed in Eq. (C.1) that there are N contact interactions, each with area of interaction equal to \(A_{\mathrm{int}}^k\). After performing the summations Eq. (C.1) can be written more compactly as:

$$\begin{aligned} W_{DEM}^I =a_I /(K_n^{*})^{0}+b_I /(K_s^{*})^{0} \end{aligned}$$
(C.2)

where:

$$\begin{aligned} a_I =\mathop \sum \nolimits _{k=1}^N (F_k^n )^{2}/A_{\mathrm{int}}^k \hbox { and }b_I =\mathop \sum \nolimits _{k=1}^N (F_k^s )^{2}/A_{\mathrm{int}}^k \end{aligned}$$
(C.3)

The strain energies of the corresponding macro experiments, \(W_{Macro}^1\hbox { and }W_{Macro}^2\), are also found in terms of the components of the macro stress tensor \({\varvec{\sigma }}\) and macro modulus E and Poisson’s ratio \(\nu \). If \(I = 1\) corresponds to a uniaxial test, with the uniaxial stress \(\sigma _{zz}\) applied in the z direction,

$$\begin{aligned} W_{Macro}^1 =\sigma _{zz}^2 /2E \end{aligned}$$
(C.4)

If \(I = 2\) corresponds to a triaxial test, with triaxial stress components \(\sigma _{xx} ,\sigma _{yy} \hbox { and }\sigma _{zz}\) applied in the xy and z directions,

$$\begin{aligned} W_{Macro}^2= & {} (1/2E)\left[ \sigma _{xx}^2 +\sigma _{yy}^2 +\sigma _{zz}^2 \right] \nonumber \\&-\left( {\nu /E} \right) \left[ {\sigma _{xx} \sigma _{yy} +\sigma _{yy} \sigma _{zz} +\sigma _{zz} \sigma _{xx} } \right] \end{aligned}$$
(C.5)

The goal of the calibration procedure is to correct the assumed values of \((K_n^*)^{0}\hbox { and }(K_s^*)^{0}\) to ensure that the strain energy of the DEM model equals that of the macro model, for both \(I = 1, 2\). Denoting the calibration factor for \((K_n^*)^{0}\) as \(c_{K_n^*}\) and the calibration factor for \((K_s^*)^{0}\) as \(c_{K_s^*}\) this requires that the following two equations must be satisfied:

$$\begin{aligned}&a_1 \big /\left( {c_{K_n^{*}}(K_n^{*})^{0}} \right) +b_1 \big /\left( {c_{K_s^{*} } (K_s^{*} )^{0}} \right) =W_{macro}^1 \end{aligned}$$
(C.6)
$$\begin{aligned}&a_2 \big /\left( {c_{K_n^{*}} (K_n^{*} )^{0}} \right) +b_2 \big /\left( {c_{K_s^{*}}(K_s^{*} )^{0}} \right) =W_{macro}^2 \end{aligned}$$
(C.7)

The above two simultaneous equations are solved for \(c_{K_n^{*}}\) and \(c_{K_s^*}\). This yields the calibrated values \(K_n^*=c_{K_n^*} (K_n^*)^{0}\hbox { and }K_s^*=c_{K_s^*} (K_s^*)^{0}\). The calibrated values of \(K_n^*\hbox { and }K_{s}^*\) also depend on the average thickness of the mortar layer between the aggregates; the calibration factors \(c_{K_n^*}\) and \(c_{K_s^{*}}\) incorporate this dependence as well.

While ideally, only two tests are necessary to calibrate \(K_n^*\hbox { and }K_{s}^*\), in practice it was observed that improved results could be obtained by performing three tests (e.g. uniaxial compression, triaxial compression and triaxial tension) instead of two. This resulted in three equations for the two unknowns \(c_{K_n^{{*}}}\) and \(c_{K_s^*}\). The values of \(c_{K_n^\mathrm{*}}\) and \(c_{K_s^*}\) were then found by least squares minimization. The stress states for the three tests are described in Table 7. It is necessary for the stress state corresponding to each test to differ significantly from the other tests, as otherwise the resulting equations may be ill-conditioned or near-singular. In such a case, it may not be possible to obtain accurate corrected values of \(K_n^*\hbox { and }K_{s}^*\) from the equations.

The elastic calibration procedure was verified for assemblies which differed only in the arrangement of packed particles but had the same meso geometry and same aggregate composition and mortar mix. Twelve such assembles were considered. The calibrated values of \(K_{n}^*\hbox { and }K_{s}^*\) obtained from a single (benchmark) assembly were used in all eleven remaining assemblies. All the assemblies were identically loaded in uniaxial compression. The Young’s modulus and Poisson’s ratio were calculated for each assembly by plotting the macro axial stress vs. macro axial strain, and evaluating the ratio of macro lateral strain to macro axial strain. For each assembly the E and \(\upsilon \) values obtained from the DEM simulations were compared with the target values of E and \(\upsilon \), denoted as \(\bar{E}\) and \(\bar{\upsilon }\). The \(E/\bar{E}\) ratios were found to vary between 0.94 to 1.13 while the \(\upsilon /\bar{\upsilon }\) ratios ranged between 0.97 and 1.04. The coefficient of variation for the \(E/\bar{E}\) ratios was 4.936% while that for the \(\upsilon /\bar{\upsilon }\) ratios was 2.458%. The dispersions in \(E/\bar{E}\) and \(\upsilon /\bar{\upsilon }\) ratios compare favorably with values reported in the literature (cf. [24]).

Table 7 Loading states used for elastic calibration

If the same cement and aggregate type as well as concrete mix proportions are used for assemblies with different meso geometries, the volume of mortar remains the same but the mortar gets distributed over very different total aggregate surface areas, since the total surface area of the aggregates varies with meso geometry. As a consequence, the average thickness of the mortar layer between the aggregates changes with meso geometry. Therefore if the calibrated values of \(K_n^*\hbox { and }K_s^*\), obtained for an assembly with meso geometry ‘A’, are used in an assembly with a different meso geometry ‘B’, the macro Young’s modulus and Poisson ratio values may not match. However, according to experiments reported by previous researchers (e.g. [25, 56]), changes in meso geometry should not lead to any significant differences in the initial elastic properties of concrete. To account for this additional scalar correction factors \(\gamma _{n}\) and \(\gamma _{s}\), representing corrections to \(K_{n}^*\hbox { and }K_s^*\) due to differences in average mortar thickness between the two assemblies, have to be determined. These factors are found on a trial and error basis. For assemblies with new meso geometries, a couple of trials were found sufficient to determine modified meso moduli \(\gamma _n K_{n}^*\hbox { and }\gamma _s K_s^*\) that resulted in the \(E/\bar{E}\) and \(\upsilon /\bar{\upsilon }\) ratios becoming sufficiently close to 1.0 (Table 8).

Table 8 Calibrated meso elastic parameters for meso geometries with varying aggregate angularity but identical concrete mix proportions, aggregate size gradation, cement and aggregate type

Appendix D: Calibration of meso damage parameters for hardened concrete

(a) Choice of damage initiation stress: Experimental data on interfacial fracture properties of concrete is limited. Recently, Rao and Prasad [45] performed experiments to determine damage parameters for mortar-rock interfaces. The damage initiation stress values reported in this paper are used as the mean values of the respective Weibull distributions. These damage initiation stress values are not significantly different from those proposed in [61].

(b) Determination of the interface fracture energy and damage evolution parameters: No experimental data is available for the meso level interface fracture energy and hence these values are determined by performing several trials. However some guidance can be obtained from the results of macro experiments [31, 46] which suggest that Mode II fracture energy can be up to 25 to 30 times higher than the Mode I fracture energy. It is assumed that this holds true at the meso scale as well. Hence if a DEM sample, with a large Mode II to Mode I meso fracture energy ratio, is loaded under uniaxial tension, it is expected that interface failure due to tensile cracking in Mode I would be the dominant mode of failure. Several DEM simulations of cubic specimens under uniaxial tension were therefore performed, in which the Mode II fracture energy was kept fixed at a high value and the Mode I fracture energy and Mode I damage evolution parameters varied until the macro stress–strain curve obtained from the DEM simulation showed a close match with the experimental stress strain curve. The Mode I fracture energy and Mode I damage evolution parameters were then held fixed at these values. Subsequently multiple DEM simulations of uniaxial compression tests were performed, in which the Mode II fracture energy and Mode II damage evolution parameters were varied. The Mode II fracture energy and Mode II damage evolution parameters which resulted in best match between the DEM macro stress–strain curve and the experimental curve for uniaxial compression were then selected.

(c) Final check: Finally a further DEM uniaxial tension test was simulated with Mode I and Mode II fracture parameters as obtained in (b) above. If the DEM macro stress–strain curve matched the experimental uniaxial tension test curve, and the crack patterns were along expected lines, the calibration of the meso fracture parameters was regarded as complete.

In case of significant differences, the Mode II fracture parameters were held fixed at their values obtained at the end of (b), and the Mode I fracture parameters varied to obtain a good match with the expereimental uniaxial tension stress strain curve. The entire cycle described in (b) was then repeated. A consistent set of fracture parameters satisfying the final check was usually obtained after a few iterations.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Dandapat, R., Ghosh, S. & Deb, A. A discrete element implementation for concrete: particle generation, contact calculations, packing under gravity and modeling material response. Granular Matter 20, 30 (2018). https://doi.org/10.1007/s10035-018-0803-4

Download citation

  • Received:

  • Published:

  • DOI: https://doi.org/10.1007/s10035-018-0803-4

Keywords

Navigation