Skip to main content
Log in

Identifying spatial transitions in heterogenous granular flow

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

Abstract

It is well known that heterogeneous granular flows exhibit collisional, dense and creep regimes that can coexist in space. How to correctly predict and control such complex phenomena has many applications in both mitigation of natural hazards and optimization of industrial processes. However, it still remains a challenge to establish a predictive granular rheology model due to the lack of understanding of the internal structure variation across different regimes and its interaction with the boundary. In this work, we use DEM simulations to investigate the internal structure of heterogeneous granular flow developed at the center of rotating drum systems. By systematically varying the side wall conditions, we are able to generate various heterogeneous flow fields under different levels of boundary effects. Our extensive simulation results reveal a highly relevant micro-structural quantity \(\delta \theta = |\theta _c - \theta _f|\), where \(\theta _c\) and \(\theta _f\) are the preferred direction of inter-particle contacts and the preferred direction of inter-particle force transmissions, respectively. We show that \(\delta \theta \) can characterize the internal structure of granular flow in collisional, dense and creep regimes, and its variation can identify the transition between them. In particular, in dense and collisional regimes, the classical rheological relation between bulk friction \(\mu \) and inertia number I holds, while in the creep regime, such relation breaks down and \(\mu \) instead depends on \(\delta \theta \). Our findings hold for all investigated flow fields regardless of the level of boundary effect imposed, and regardless of the amount of shear experienced. \(\delta \theta \) thus provides a unified micro-structural characterization for heterogeneous granular flow in different regimes, and lays the foundation of establishing microstructure-informed granular rheology models.

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.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5

Similar content being viewed by others

References

  1. Jaeger, H.M., Nagel, S.R., Behringer, R.P.: Granular solids, liquids, and gases. Rev. Mod. Phys. 68(4), 1259 (1996)

    ADS  Google Scholar 

  2. Mueth, D.M., Debregeas, G.F., Karczmar, G.S., Eng, P.J., Nagel, S.R., Jaeger, H.M.: Signatures of granular microstructure in dense shear flows. Nature 406(6794), 385 (2000)

    ADS  Google Scholar 

  3. Komatsu, T.S., Inagaki, S., Nakagawa, N., Nasuno, S.: Creep motion in a granular pile exhibiting steady surface flow. Phys. Rev. Lett. 86(9), 1757 (2001)

    ADS  Google Scholar 

  4. Bonamy, D., Daviaud, F., Laurent, L.: Experimental study of granular surface flows via a fast camera: a continuous description. Phys. Fluids 14(5), 1666–1673 (2002)

    ADS  MATH  Google Scholar 

  5. Bonamy, D., Daviaud, F., Laurent, L., Bonetti, M., Bouchaud, J.P.: Multiscale clustering in granular surface flows. Phys. Rev. Lett. 89(3), 034301 (2002)

    ADS  Google Scholar 

  6. Choi, J., Kudrolli, A., Rosales, R.R., Bazant, M.Z.: Diffusion and mixing in gravity-driven dense granular flows. Phys. Rev. Lett. 92(17), 174301 (2004)

    ADS  Google Scholar 

  7. Forterre, Y., Pouliquen, O.: Flows of dense granular media. Annu. Rev. Fluid Mech. 40, 1–24 (2008)

    ADS  MathSciNet  MATH  Google Scholar 

  8. MiDi, G.: On dense granular flows. Eur. Phys. J. E 14(4), 341–365 (2004)

    Google Scholar 

  9. da Cruz, F., Emam, S., Prochnow, M., Roux, J.-N., Chevoir, F.: Rheophysics of dense granular materials: discrete simulation of plane shear flows. Phys. Rev. E 72(2), 021309 (2005)

    ADS  Google Scholar 

  10. Jop, P., Forterre, Y., Pouliquen, O.: A constitutive law for dense granular flows. Nature 441, 727–730 (2006)

    ADS  MATH  Google Scholar 

  11. Baran, O., Ertaş, D., Halsey, T.C., Grest, G.S., Lechman, J.B.: Velocity correlations in dense gravity-driven granular chute flow. Phys. Rev. E 74(5), 051302 (2006)

    ADS  Google Scholar 

  12. Staron, L.: Correlated motion in the bulk of dense granular flows. Phys. Rev. E 77(5), 051304 (2008)

    ADS  Google Scholar 

  13. Orpe, A.V., Kudrolli, A.: Velocity correlations in dense granular flows observed with internal imaging. Phys. Rev. Lett. 98(23), 238001 (2007)

    ADS  Google Scholar 

  14. Mills, P., Rognon, P., Chevoir, F.: Rheology and structure of granular materials near the jamming transition. EPL (Europhys. Lett.) 81(6), 64005 (2008)

    ADS  Google Scholar 

  15. Pouliquen, O., Forterre, Y.: A non-local rheology for dense granular flows. Philos. Trans. R. Soc. Lond. A Math. Phys. Eng. Sci. 367(1909), 5091–5107 (2009)

    ADS  MATH  Google Scholar 

  16. Kamrin, K., Koval, G.: Nonlocal constitutive relation for steady granular flow. Phys. Rev. Lett. 108(17), 178301 (2012)

    ADS  Google Scholar 

  17. Bouzid, M., Trulsson, M., Claudin, P., Clément, E., Andreotti, B.: Nonlocal rheology of granular flows across yield conditions. Phys. Rev. Lett. 111(23), 238301 (2013)

    ADS  Google Scholar 

  18. Bouzid, M., Izzet, A., Trulsson, M., Clément, E., Claudin, P., Andreotti, B.: Non-local rheology in dense granular flows. Eur. Phys. J. E 38(11), 125 (2015)

    Google Scholar 

  19. de Coulomb, A.F., Bouzid, M., Claudin, P., Clément, E., Andreotti, B.: Rheology of granular flows across the transition from soft to rigid particles. Phys. Rev. Fluids 2(10), 102301 (2017)

    ADS  Google Scholar 

  20. Saitoh, K., Tighe, B.P.: Nonlocal effects in inhomogeneous flows of soft athermal disks. Phys. Rev. Lett. 122(18), 188001 (2019)

    ADS  Google Scholar 

  21. Cortet, P.-P., Bonamy, D., Daviaud, F., Dauchot, O., Dubrulle, B., Renouf, M.: Relevance of visco-plastic theory in a multi-directional inhomogeneous granular flow. EPL (Europhys. Lett.) 88(1), 14001 (2009)

    ADS  Google Scholar 

  22. Renouf, M., Bonamy, D., Dubois, F., Alart, P.: Numerical simulation of two-dimensional steady granular flows in rotating drum: on surface flow rheology. Phys. Fluids 17(10), 103303 (2005)

    ADS  MATH  Google Scholar 

  23. Fan, Y., Umbanhowar, P.B., Ottino, J.M., Lueptow, R.M.: Shear-rate-independent diffusion in granular flows. Phys. Rev. Lett. 115(8), 088001 (2015)

    ADS  Google Scholar 

  24. Miller, T., Rognon, P., Metzger, B., Einav, I.: Eddy viscosity in dense granular flows. Phys. Rev. Lett. 111(5), 058002 (2013)

    ADS  Google Scholar 

  25. Azéma, E., Radjai, F.: Internal structure of inertial granular flows. Phys. Rev. Lett. 112(7), 078001 (2014)

    ADS  Google Scholar 

  26. Jop, P., Forterre, Y., Pouliquen, O.: Crucial role of sidewalls in granular surface flows: consequences for the rheology. J. Fluid Mech. 541, 167–192 (2005)

    ADS  MATH  Google Scholar 

  27. Pohlman, N.A., Ottino, J.M., Lueptow, R.M.: End-wall effects in granular tumblers: from quasi-two-dimensional flow to three-dimensional flow. Phys. Rev. E 74(3), 031305 (2006)

    ADS  Google Scholar 

  28. Du Pont, S.C., Gondret, P., Perrin, B., Rabaud, M.: Wall effects on granular heap stability. EPL (Europhys. Lett.) 61(4), 492 (2003)

    ADS  Google Scholar 

  29. Zhang, Q., Kamrin, K.: Microscopic description of the granular fluidity field in nonlocal flow modeling. Phys. Rev. Lett. 118(5), 058001 (2017)

    ADS  Google Scholar 

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

    Google Scholar 

  31. Kloss, C., Goniva, C., Hager, A., Amberger, S., Pirker, S.: Models, algorithms and validation for opensource DEM and CFD–DEM. Prog. Comput. Fluid Dyn. Int. J. 12(2–3), 140–152 (2012)

    MathSciNet  Google Scholar 

  32. Börzsönyi, T., Ecke, R.E., McElwaine, J.N.: Patterns in flowing sand: understanding the physics of granular flow. Phys. Rev. Lett. 103(17), 178302 (2009)

    ADS  Google Scholar 

  33. Cheng, H., Shuku, T., Thoeni, K., Yamamoto, H.: Probabilistic calibration of discrete element simulations using the sequential quasi-monte carlo filter. Granul. Matter 20(1), 11 (2018)

    Google Scholar 

  34. Turkia, S.B., Wilke, D.N., Pizette, P., Govender, N., Abriak, N.-E.: Benefits of virtual calibration for discrete element parameter estimation from bulk experiments. Granul. Matter 21(4), 110 (2019)

    Google Scholar 

  35. Richard, P., Valance, A., Métayer, J.-F., Sanchez, P., Crassous, J., Louge, M., Delannay, R.: Rheology of confined granular flows: scale invariance, glass transition, and friction weakening. Phys. Rev. Lett. 101(24), 248002 (2008)

    ADS  Google Scholar 

  36. Artoni, R., Richard, P.: Effective wall friction in wall-bounded 3d dense granular flows. Phys. Rev. Lett. 115(15), 158001 (2015)

    ADS  Google Scholar 

  37. Brodu, N., Richard, P., Delannay, R.: Shallow granular flows down flat frictional channels: steady flows and longitudinal vortices. Phys. Rev. E 87(2), 022202 (2013)

    ADS  Google Scholar 

  38. Taberlet, N., Richard, P., Valance, A., Losert, W., Pasini, J.M., Jenkins, J.T., Delannay, R.: Superstable granular heap in a thin channel. Phys. Rev. Lett. 91(26), 264301 (2003)

    ADS  Google Scholar 

  39. Taberlet, N., Richard, P., Hinch, E.J.: S shape of a granular pile in a rotating drum. Phys. Rev. E 73(5), 050301(R) (2006)

    ADS  Google Scholar 

  40. Christoffersen, J., Mehrabadi, M., Nemat-Nasser, S.: A micromechanical description of granular material behavior. J. Appl. Mech. 48(2), 339–344 (1981)

    ADS  MATH  Google Scholar 

  41. Rycroft, C.H., Grest, G.S., Landry, J.W., Bazant, M.Z.: Analysis of granular flow in a pebble-bed nuclear reactor. Phys. Rev. E 74(2), 021306 (2006)

    ADS  Google Scholar 

  42. Camenen, J.-F., Descantes, Y., Richard, P.: Effect of confinement on dense packings of rigid frictionless spheres and polyhedra. Phys. Rev. E 86(6), 061317 (2012)

    ADS  Google Scholar 

  43. Koval, G., Roux, J.-N., Corfdir, A., Chevoir, F.: Annular shear of cohesionless granular materials: from the inertial to quasistatic regime. Phys. Rev. E 79(2), 021306 (2009)

    ADS  Google Scholar 

  44. Gaume, J., Chambon, G., Naaim, M.: Quasistatic to inertial transition in granular materials and the role of fluctuations. Phys. Rev. E 84(5), 051304 (2011)

    ADS  Google Scholar 

  45. Azéma, E., Descantes, Y., Roquet, N., Roux, J.-N., Chevoir, F.: Discrete simulation of dense flows of polyhedral grains down a rough inclined plane. Phys. Rev. E 86(3), 031303 (2012)

    ADS  Google Scholar 

  46. Goldhirsch, I.: Rapid granular flows. Annu. Rev. Fluid Mech. 35(1), 267–293 (2003)

    ADS  MathSciNet  MATH  Google Scholar 

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

    ADS  Google Scholar 

  48. Rothenburg, L., Bathurst, R.: Analytical study of induced anisotropy in idealized granular materials. Geotechnique 39(4), 601–614 (1989)

    Google Scholar 

  49. Börzsönyi, T., Szabó, B., Törös, G., Wegner, S., Török, J., Somfai, E., Bien, T., Stannarius, R.: Orientational order and alignment of elongated particles induced by shear. Phys. Rev. Lett. 108(22), 228302 (2012)

    ADS  Google Scholar 

  50. Hidalgo, R., Szabó, B., Gillemot, K., Börzsönyi, T., Weinhart, T.: Rheological response of nonspherical granular flows down an incline. Phys. Rev. Fluids 3(7), 074301 (2018)

    ADS  Google Scholar 

  51. Nadler, B., Guillard, F., Einav, I.: Kinematic model of transient shape-induced anisotropy in dense granular flow. Phys. Rev. Lett. 120(19), 198003 (2018)

    ADS  Google Scholar 

  52. Gray, J.M.N.T.: Particle segregation in dense granular flows. Annu. Rev. Fluid Mech. 50, 407–433 (2018)

    ADS  MathSciNet  MATH  Google Scholar 

  53. Cantor, D., Azéma, E., Sornay, P., Radjai, F.: Rheology and structure of polydisperse three-dimensional packings of spheres. Phys. Rev. E 98(5), 052910 (2018)

    ADS  Google Scholar 

  54. Mandal, S., Khakhar, D.: Sidewall-friction-driven ordering transition in granular channel flows: implications for granular rheology. Phys. Rev. E 96(5), 050901(R) (2017)

    ADS  Google Scholar 

  55. Jean, M.: The non-smooth contact dynamics method. Comput. Methods Appl. Mech. Eng. 177(3–4), 235–257 (1999)

    ADS  MathSciNet  MATH  Google Scholar 

  56. Pazouki, A., Kwarta, M., Williams, K., Likos, W., Serban, R., Jayakumar, P., Negrut, D.: Compliant contact versus rigid contact: a comparison in the context of granular dynamics. Phys. Rev. E 96(4), 042905 (2017)

    ADS  Google Scholar 

  57. Silbert, L.E., Ertaş, D., Grest, G.S., Halsey, T.C., Levine, D., Plimpton, S.J.: Granular flow down an inclined plane: Bagnold scaling and rheology. Phys. Rev. E 64(5), 051302 (2001)

    ADS  Google Scholar 

  58. Reagle, C., Delimont, J., Ng, W., Ekkad, S., Rajendran, V.: Measuring the coefficient of restitution of high speed microparticle impacts using a PTV and CFD hybrid technique. Meas. Sci. Technol. 24(10), 105303 (2013)

    ADS  Google Scholar 

  59. Zhou, Y., Wright, B., Yang, R., Xu, B.H., Yu, A.-B.: Rolling friction in the dynamic simulation of sandpile formation. Physica A 269(2–4), 536–553 (1999)

    ADS  Google Scholar 

  60. Frankowski, P. Morgeneyer, M.: Calibration and validation of DEM rolling and sliding friction coefficients in angle of repose and shear measurements, In: AIP Conference Proceedings, vol. 1542, pp. 851–854. AIP (2013)

  61. Phillip Grima, A., Wilhelm Wypych, P.: Discrete element simulations of granular pile formation: method for calibrating discrete element models. Eng. Comput. 28(3), 314–339 (2011)

    MATH  Google Scholar 

  62. Thielicke, W., Stamhuis, E.: PIVlab-towards user-friendly, affordable and accurate digital particle image velocimetry in matlab. J. Open Res. Softw. 2(1), e30

  63. Radjai, F., Roux, S.: Turbulentlike fluctuations in quasistatic flow of granular media. Phys. Rev. Lett. 89(6), 064302 (2002)

    ADS  Google Scholar 

  64. Kharel, P., Rognon, P.: Vortices enhance diffusion in dense granular flows. Phys. Rev. Lett. 119(17), 178001 (2017)

    ADS  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to José E. Andrade.

Ethics declarations

Conflict of interest

The authors declare no potential conflicts of interest.

Additional information

Publisher's Note

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

Appendices

Appendix 1: Discrete element model calibration and validation

In this section we discuss how we determine the model parameters used in our simulations. In general, the interaction between rigid particles can be modeled by either solving a linear-complementarity problem (implicit dynamics, the NSCD) [55] or by penalizing the inter-particle penetration (explicit dynamics, the classical DEM). Despite the different underlying principle, they give consistent results within the scope of rigid particle dynamics [56]. For the classical DEM, there are various inter-particle contact laws with different level of sophistication. Following the discussion in [57], we choose the linear Hookean contact law and pick \(k_n = 2\times 10^5 \bar{m}g/\bar{d}\) (large enough to ensure rigid particle limit for the gravity-driven surface flows considered in our study) with \(k_t = 2/7k_n\), \(\gamma _t = 0\), and \(\gamma _n = -2\text {ln}e\sqrt{\bar{m}k_n/(\pi ^2+\text {ln}^2e)}\). We are left to determine the coefficient of restitution e, the inter-particle friction \(\mu _p\), the particle-wall friction and possibly the addition of rolling friction \(\mu _r\).

1.1 First stage calibration via column collapse tests

We use the column collapse test to perform preliminary model calibration by measuring the angle of repose \(\theta _r\).We first glue glass beads to the center area of an aluminum sheet (\(300\times 300\) mm) and the inner surface of two identical iron angle bars with height 50 mm, length 25 mm and width 16 mm (Fig. 6a). A typical column collapse test can be divided into three steps: (1) filling with glass beads the hollow rectangular tube formed by the two iron angle bars placed over the aluminum sheet center, (2) rapidly removing the bars, and (3) taking picture of the formed pile to measure \(\theta _r\). We repeat the procedure for 50 times and \(\theta _r\) is measured to have a mean of \(13.87^\circ \) and a standard deviation of \(0.576^\circ \).

Fig. 6
figure 6

a Setup of the column collapse test, b variation of \(\theta _r\) according to the change of \(\mu _p\) with a fixed \(e = 0.82\), c variation of \(\theta _r\) according to the change of e with a fixed \(\mu _p = 0.4\), and c variation of \(\theta _r\) according to the change of \(\mu _r\) with fixed \(\mu _p = 0.4, e = 0.82\)

Via DEM simulations, we then perform numerical column collapse tests with the same configuration as in the experiments. We carry out two sets of simulations: (1) fixing \(\mu _p = 0.4\) (a common choice for glass beads) and varying e from 0.1 to 0.82, and (2) fixing \(e = 0.82\) [58] and varying \(\mu _p\) from 0.1 to 0.8. From (1) we find that e has negligible effect on \(\theta _r\) (Fig. 6c), and from (2) that \(\theta _r\) first increases but later saturates with the increase of \(\mu _p\) (Fig. 6b). In summary the above results suggest the necessity to incorporate rolling friction \(\mu _r\), a parameter that imposes rotation hinderance [59] to model the interaction between non-spherical particles. Accordingly, we fix \(e = 0.82, \mu _p = 0.4\) and vary \(\mu _r\) from 0.01 to 0.2. Figure 6d shows the variation of \(\theta _r\) against \(\mu _r\). The numerical results indicate \(\mu _r\) to be around 0.07 which is slightly larger than 0.01 in [60] where smooth glass spheres were used and slightly smaller than 0.1 in [61] where plastic spheres were used.

1.2 Second stage calibration and validation via rotating drum experiments

In the rotating drum experiments (Fig. 7a), after the surface flow becomes stationary under rotation speed \(\omega = 2.59^\circ /\hbox {s}, 5.73^\circ /\hbox {s}\) and \(11.23^\circ /\hbox {s}\), we use a high speed camera positioned against the glass plate to take images (288 px \(\times \) 288 px corresponding to a 0.1389 mm/px resolution) for a time period of 10 s with a frame rate of 1000 fps. Accordingly the images cover an area of about 4 cm \(\times 4\) cm at the drum center (Fig. 7b). From the sequence of images, we measure the dynamical angle of repose \(\theta _d\) and the down-stream velocity \(v_{yw}(z)\) near the glass plate. In terms of the former, we first binarize each image, then identify the pixels that represents the slope surface, and lastly use the identified pixels to perform a linear fit whose slope gives \(\theta _d\) (Fig. 7d); as to the latter, we first use the open-source Particle Image Velocimetry (PIV) code [62] to compute the 2D velocity field \((v_1,v_2)\) by correlating boxes with dimension 8 px \(\times 8\) px (corresponding to roughly \(\bar{d}\times \bar{d}\)), we then compute the velocity field under the frame rotating with the drum located at the drum center to get \((v_y ,v_z)\), lastly we compute \(v_{yw}(z)\) by averaging \(v_y\) within bands (\(L_y = 20\bar{d}, L_z = 2\bar{d}\)) positioned in parallel to y (slope surface) over a set of points that are placed every \(\bar{d}\) distance along z (perpendicular to the slope surface) with \(y=0\). Note that \(|v_z| \ll |v_y| \) as the flow is nearly unidirectional. In the simulations, we generate images located at exactly the same location with exactly the same size (and resolution) as the ones taken from experiment (Fig. 7c), from which we follow the same image analysis procedure to find \(\theta _d\). For \(v_{yw}\), under the frame rotating with the drum at the drum center, we first pick particles located within \(2\bar{d}\) away from the front-side flat wall, we then compute \(v_{yw}(z)\) by averaging the particle velocity following the same procedure used in the experiments. Note that the width \(2\bar{d}\) is picked to best represent the glass beads that are captured by the high speed camera.

Fig. 7
figure 7

a The half-filled rotating drum with rear-side wall and inner-cylinder wall being glued with glass beads, b an image taken by the high speed camera at the drum center, c an image generated by numerical simulation with exactly the same location and size (resolution) as (b), d the binarized image of c for \(\theta _d\) estimation, e time-averaged \(\theta _d\) data for three different rotating speed \(\omega \) estimated from experiments (red) and simulations (blue), f the down-stream velocity profile \(y_{yw}(z)\) against the depth at the drum center calculated respectively for \(\omega = 11.23^\circ /\hbox {s}\) from experiment (red triangle) and simulation (blue triangle), for \(\omega = 5.73^\circ /\hbox {s}\) from experiment (red square) and simulation (blue square), and for \(\omega = 2.59^\circ /\hbox {s}\) from experiment (red circle) and simulation (blue circle), and g the corresponding semi-log plot of (f). For eg, the error bars represent the standard deviation associated with each time-averaged quantity

We use \(\theta _d\) and \(v_{yw}(z)\) measured with \(\omega = 11.23^\circ /\hbox {s}\) for model calibration and the rest two for model validation. Prior to calibration, according to the column collapse test results, we fix \(\mu _{\mathrm{p}} = 0.4\), and \(e = 0.82\) that best represents the property of glass beads, although the latter has negligible effect for simulating steady granular flow [57]. Further, as the front-side plate is also made from glass, we fix the associated wall friction to be 0.4. The only left parameter to calibrate is the inter-particle rolling friction \(\mu _{\mathrm{r}}\). Observing Fig. 6d, we vary \(\mu _{\mathrm{r}}\) to be 0, 0.03, 0.05 and 0.07 and take the particle-wall rolling friction to be zero. As the front-side wall is flat, zero wall rolling friction is a reasonable choice. By solely using \(\theta _d\) we identify that \(\mu _{\mathrm{r}} = 0.03\) gives the best estimation ( \(\mu _{\mathrm{r}} = 0\) underestimates \(\theta _d\) while the rest two lead to overestimation). What’s more, when \(\mu _{\mathrm{r}} = 0.03\), simulation and experiment show excellent agreement (Fig. 7f) in terms of \(v_{yw}(z)\). The choice of \(\mu _{\mathrm{r}}\) = 0.03 is then validated (Fig. 7e–g) by comparing both \(\theta _d\) and \(v_{yw}(z)\) obtained from simulations to those measured from experiments under \(\omega = 2.59^\circ /\hbox {s}\) and \(\omega = 5.73^\circ /\hbox {s}\).

Appendix 2: Additional simulation results

2.1 “S-shape” surface profile

For a direct comparison, Fig. 8 shows the surface shape profile for simulations performed under respectively periodic boundary condition, frictional side walls with \(W/\bar{d} \simeq 21\) and wall friction 0.4 and that with \(W/\bar{d} \simeq 10\) and wall friction 0.4. It can be observed that as the effective drum width is decreased from infinite (periodic boundaries) to \(W/\bar{d} \simeq 21\) and to \(W/\bar{d}\simeq 21\), the “S-shape” profile becomes more obvious under more significant side wall friction effect, especially when the rotation speed is large such as when \(\omega = 67.38^\circ /\hbox {s}\) (Fig. 8j, o) in our case.

Fig. 8
figure 8

Surface shape profiles. Images in first row from ae are those under periodic lateral boundaries, in second row from fj represents those under frictional side walls with \(W/\bar{d} \simeq 21\), and in third row from ko represents those under frictional side walls too but with \(W/\bar{d} \simeq 10\). Wall friction is 0.4

2.2 Effect of lateral boundary condition on the co-directionality between s and \(\dot{\gamma }^d\)

Figure 9 showcases the spatial variation of the misalignment angle \(\alpha \) for all considered drum configurations under \(\omega = 33.69^\circ /\hbox {s}\), where \(\alpha \) is defined as the angle between the principle directions of s and those of \(\dot{\gamma }^d\). It can be observed that the presence of frictional side wall has a great impact on the value of \(\alpha \), and the more frictional the side walls and the narrower the drum, the less co-directional are s and \(\dot{\gamma }^d\). \(\alpha \) is generally large (up to \(60^\circ \)) near the side walls (especially on the bumpy side), and beneath \(z = z_\text {th}\) in the creep flow region. The green solid line indicates locations with local deformation \(|\dot{\gamma }^d|/2\Omega \) equalling one [21] where \(\Omega = \omega /360^\circ \); locations above this line has \(|\dot{\gamma }^d|/2\Omega > 1\) while \(|\dot{\gamma }^d|/2\Omega < 1\) for locations below this line. It thus may be understood that the large \(\alpha \) in the creep flow region is due to the lack of shear. For locations above \(z = z_\text {th}\), shear deformation is sufficient, and the misalignment \(\alpha \) can be attributed to side wall perturbation.

Fig. 9
figure 9

Spatial variation of \(\alpha \) under \(\omega = 33.69^\circ /\hbox {s}\) for all drum configuration considered; from left to right: \(W/\bar{d}\simeq 10\) with wall friction of 0.4, periodic boundary, \(W/\bar{d}\simeq 21\) with wall friction of 0.2, 0.4, 0.6 and 0.8, respectively. The grey solid lines indicate \(z = z_\text {th}\) and the green ones represent locations with local deformation \(|\dot{\gamma }^d|/2\Omega = 1\), where \(\Omega = 360^\circ /\omega \)

2.3 Tests of several non-local models

  • The velocity fluctuation model Based on 2D numerical simulation results on granular flow in annular shear cells, the model proposed in [44] show that an improved relation between \(\mu \) and I by adding the effect of velocity fluctuation is able to capture the failure of the one-to-one \(\mu (I)\) relation in the creep flow region. The key prerequisite of this model is that the variation of \(\mu -I\) and that of \(\Delta -I\) are both not one-to-one and are mutually correlated, where \(\Delta = |\delta v|\sqrt{\rho _s/P}\) is called the fluctuation number. However, Fig. 10a shows that for all flows considered, there is a global collapse between \(\Delta \) and I. It appears that this model does not apply to the 3D flows considered in our study.

  • The gradient expansion model In the gradient expansion model [17], the value of \(\mu \) in the \(\mu -I\) relation is modified by considering an additional contribution \(\bar{d}^2\nabla ^2I/I\)which is scaled by a phenomenological constant \(\nu > 0\), assuming short range correlation between particle motions in different locations. Physically, the laplacian term captures on average, how does the I value at a certain point compares to its surrounding area. For instance, if a point is surrounded by a more fluid-like area (higher I), the laplacian term is positive and leads to the decrease of \(\mu \) at that point. Since \(\nu > 0\), essentially as long as \(\nabla ^2I\ne 0\), the model will report a modification on \(\mu \). Figure 10b shows the typical drum-width-averaged I and \(|v|/\sqrt{g\bar{d}}\) against z with the case under \(\omega = 33.69^\circ /\hbox {s}, W/d \simeq 21\) and wall friction of 0.4. The variation of I against z roughly follows the same trend as that of \(|v|/\sqrt{g\bar{d}}\): it linearly decays in the collisional region and exponentially decays in the dense region. Thus in the collisional region \(\nabla ^2 I = 0\) and the model reports no modification on \(\mu \), which is consistent with our observation. However, in the dense region, \(\mu \) will be modified according to the model since \(\nabla ^2I \ne 0\)—this contradicts our computations that confirm the applicability of the \(\mu -I\) relation in the dense region. Again, it appears that this model does not apply to the 3D surface flows considered in our study.

  • The fluidity model The fluidity model [16] implicitly modifies the value of \(\mu \) in the \(\mu -I\) relation by considering the nearby region contribution through the Laplacian of the fluidity parameter \(g = \dot{\gamma }/\mu \) expressed as \(\xi ^2\nabla ^2g\), where \(\xi \) is defined as the cooperative length that diverges as approaching the jamming transition. Although the mathematical expression looks similar to that of the gradient expansion model, the assumed underlying physical mechanism is different. g is found to obey the following microscopic relation [29]: \(\dot{\gamma }^{d}\bar{d}/\mu = \delta v F(\phi )\). Figure 10c shows how the normalized fluidity \(\dot{\gamma }^{d}\bar{d}/(\mu \delta v)\), varies with volume fraction \(\phi \). It can be observed for all data computed from \(W/\bar{d} \simeq 21\) with varying side wall friction, they collapse well onto a single curve who shape resembles the one identified in [29]. However, this curve is clearly drum-width-dependent, as when the effective drum width is respectively infinite (periodic boundary, colored in blue) and \(10\bar{d}\) (colored in yellow), no collapse can be observed, even in the dense and collisional region where the \(\mu (I)\) rheology in its invariant form holds. Memory effect (insufficient shear) observed for \(I < I_\text {th}\) (see Fig. 9) may explain the failure of the model in the creep flow region; while its break down in the fast flow regime (\(I >I_\text {th}\) with sufficient shear deformation) reveals the non-trivial effects of side wall friction that have not been considered in the granular fluidity model [29]—indeed, even though investigated in 3D configuration, the considered flow fields have shear only in z direction as boundaries along both x and y direction are treated as periodic.

Fig. 10
figure 10

a\(\Delta -I\) data for all cases considered collapse onto a single master curve. Data are shown as the drum-with-average values with error bars representing the associated variation. b Variation of drum-width-averaged \(|v|/\sqrt{g\bar{d}}\) and I against z with the case under \(\omega = 33.69^\circ /\hbox {s}, W/d \simeq 21\) and wall friction of 0.4. c Relation between the normalized fluidity \(|\dot{\gamma }^d|\bar{d}/\mu \delta v\) and volume fraction \(\phi \). Data are shown as the drum-with-average values for clarity

2.4 Additional results from micro-scale analysis

The micro-structure can be investigated by three kinds of quantities that range from lower-order (L) to higher-order (H) within each kind: (1) geometry-associated ones range from volume fraction \(\phi \) (L), coordination number Z (L) to the angular distribution of contact orientation (H) [47, 48]; (2) inter-particle-force-associated ones range from normal (and tangential) force p.d.f. distribution (L) [47, 48] to their angular distributions (H) [47, 48]; and (3) kinematics-associated ones where the lower order quantity can be the velocity fluctuation \(\delta v\) [29, 44]. Inspired by [48, 63], we may regard \(a_\chi ^v = (\lambda _3 - \lambda _1)/(\lambda _3 + \lambda _1) \) as the higher order quantity where \(\lambda _3(\lambda _1)\) is the maximum (minimum) eigenvalue of the tensor \(\chi = \langle |\delta v| n_i^v n_j^v \rangle /\langle |\delta v|\rangle \). Here “\(\langle \cdot \rangle \)” denotes the average over all particles and \(n^v\) is the direction of velocity fluctuation associated with each particle. Accordingly \(a_\chi ^v\) has range from 0 to 1 and reflects how “cooperative” the particle motions are: a value close to 1 in the quasi-static flow region implies the formation of “granular eddy” [24, 64].

Fig. 11
figure 11

Spatial variation of the coordination number Z and \(a_\chi ^v\) against the depth z. The symbol shape represents different rotation speed, while the symbol color represents different drum configurations in terms of drum width W and side wall friction. Error bars represent the variation of each shown quantity across the drum width

In principle, we are trying to find certain micro-structural quantities whose variations against I, (R1) exhibit a clear transition as passing through \(z = z_{\text {th}}\), and more importantly, (R2) show consistence with that of \(\mu \) against I for rheological considerations: when \(I > z_{\text {th}}\) they are free of drum configuration effect and rotation speed effect while they become lateral boundary dependent and rotation speed dependent as soon as \(I \le I_{\text {th}}\). Many of the aforementioned quantities, as we discover, only satisfy R1, such as Z and \(a_\chi ^v\). Figure 11a shows the variation of Z against the depth z. Z varies case by cases when \(z \le z_\text {th}\) and remains almost constant when \(z>z_0\). However, on the contrary, the stress ratio \(\mu \) shows a global collapse when \(z \le z_\text {th}\) but varies case by case when \(z>z_\text {th}\). \(a_\chi ^v\) shows a slightly different spatial variation (Fig. 11b): it remains constant when \(z\le z_\text {th}\) in a similar way to that of \(\delta \theta \). However, \(a_\chi ^v\) seems to be boundary condition dependent instead. When \(z >z_\text {th}\), it rapidly increases independently with respect to the boundary condition and rotation speed, which is inconsistent with the variation of \(\mu \) either. Following the momentum transferring argument, we have also investigated the portion of formed granular clusters based on either the velocity fluctuation following [5, 64] or the local volume fraction fluctuation (achievable via the computed Voronoi diagram) following [5, 14]. The variation of such portion along the depth show similar trend as that of \(a_\chi ^v\) that does not satisfy (R2).

2.5 Determination of \(z_\text {th}\) based on the drum-width-averaged velocity profile

Figure 12 shows the velocity magnitude profile against depth at the drum center |v|(z) and its variation across the drum width, for the same set of simulations considered here. When the lateral boundaries are periodic, |v|(z) shows negligible differences across the drum width. For the cases with frictional side walls (\(W/\bar{d} \simeq \) 21 and 10), |v|(z) varies majorly as vertical translation without shape alteration. Therefore without sacrificing much the accuracy we determine \(z_\text {th}\) based on the drum-width-averaged |v|(z) profile.

Fig. 12
figure 12

|v|(z) profile variation across the drum width for different rotation speeds with each color representing a certain location x between \(-8\bar{d}\) and \(9\bar{d}\) (\(W/\bar{d}\simeq 21\)) and between \(-3\bar{d}\) and \(4\bar{d}\) (\(W/\bar{d}\simeq 10\)). First row from ae: data extracted from simulations with periodic lateral boundaries; second row from fj: data extracted from simulations with frictional side walls (\(W/\bar{d}\simeq 21\)); third row from ko: data extracted from simulations with frictional side walls (\(W/\bar{d}\simeq 10\)); wall friction is 0.4

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Li, L., Andrade, J.E. Identifying spatial transitions in heterogenous granular flow. Granular Matter 22, 52 (2020). https://doi.org/10.1007/s10035-020-01013-1

Download citation

  • Received:

  • Published:

  • DOI: https://doi.org/10.1007/s10035-020-01013-1

Keywords

Navigation