SIMULATION OF IRREGULARLY SHAPED PARTICLES USING COUPLING METHOD OF LATTICE BOLTZMANN AND DISCRETE ELEMENT MODELLING by Mohammadhassan Ahmadian B.S., Imam Hossein University, 2013 M.S., Ferdowsi University of Mashhad, 2018 THESIS SUBMITTED IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF MASTER OF SCIENCE IN NATURAL RESOURCES AND ENVIRONMENTAL STUDIES UNIVERSITY OF NORTHERN BRITISH COLUMBIA April 2024 ยฉ Mohammadhassan Ahmadian, 2024 Abstract In this study, the lattice Boltzmann method (LBM) and the discrete element method (DEM) are coupled to simulate the interaction between the fluid phase and irregularly shaped solid particles. For this purpose, the geometry of real particle shapes is represented as clumps of overlapping spheres and then simulated through a multi-sphere model in DEM and coupled with LBM using two open-access codes, LIGGGHTS and Palabos. The accuracy of the coupling method with clumps is demonstrated by simulating several benchmark cases and comparing them with the results from the literature. The coupled LBM-DEM method is then used to simulate the collapse and transport of submerged granular particles with four different irregular particle shapes in addition to sphere particle shapes to highlight the influence of grain morphology in the solidfluid interaction. To the authorsโ€™ best knowledge, it is the first time that LIGGGHTS and Palabos were used to simulate the LBM-DEM coupling of irregular particles. This study demonstrates the potential and accuracy of LBM-DEM method in simulating particles with irregular shapes. The research also provides insight into the importance of considering real shape particles in simulating granular flows in geotechnical engineering. ii Table of Contents Abstract ........................................................................................................................................... ii Table of Contents ........................................................................................................................... iii List of Tables .................................................................................................................................. v List of Figures ................................................................................................................................ vi Glossary ......................................................................................................................................... ix Acknowledgement .......................................................................................................................... x Chapter 1: Introduction ................................................................................................................... 1 1.1 Motivation and background .................................................................................................. 1 1.2 Literature review ................................................................................................................... 2 1.2.1 Numerical approaches for solid-fluid interaction .......................................................... 2 1.2.2 Eulerian-Lagrangian methods ........................................................................................ 5 1.2.3 LBM-DEM coupling method ......................................................................................... 8 1.2.4 Simulation of irregularly shaped particles ................................................................... 10 1.3 Research objective and current study ................................................................................. 11 1.4 Thesis organization ............................................................................................................. 12 Chapter 2: Methodology ............................................................................................................... 14 2.1 Introduction ......................................................................................................................... 14 2.2 Lattice Boltzmann method .................................................................................................. 14 iii 2.3 Discrete element method..................................................................................................... 17 2.4 Immersed moving boundary method .................................................................................. 21 2.5 LBM-DEM coupling scheme .............................................................................................. 24 2.6 Irregular shape simulation................................................................................................... 26 Chapter 3: Validation .................................................................................................................... 31 3.1 Introduction ......................................................................................................................... 31 3.2 Mesh sensitivity, relaxation time, and sub-particle study ................................................... 31 3.3 Benchmark Cases ................................................................................................................ 34 3.3.1 Settling of a single clump ............................................................................................ 34 3.3.2 Settling of two clumps ................................................................................................. 37 3.3.3 Settling of a disc-shaped clump ................................................................................... 40 Chapter 4: Immersed granular collapse ........................................................................................ 43 4.1 Introduction ......................................................................................................................... 43 4.2 Irregularly shaped particles ................................................................................................. 43 4.3 Simulation of column collapse............................................................................................ 47 Chapter 5: Conclusion................................................................................................................... 56 5.1 Contribution ........................................................................................................................ 56 5.2 Limitations .......................................................................................................................... 57 5.3 Future work ......................................................................................................................... 57 References ..................................................................................................................................... 59 iv List of Tables Table 2.1: Values of ๐œ๐’Š , ๐‘ค๐‘– for a D3Q19 lattice structure (G. C. Yang, Jing, Kwok, & Sobral, 2019). ............................................................................................................................................ 16 Table 2.2: Parameters used in CLUMP algorithm for generating clumps .................................... 29 Table 3.1: Non-dimensional number values for two cases of disc-shaped particle benchmark. .. 41 Table 3.2: Characteristics of two cases related to disc-shaped particle benchmark. .................... 41 Table 4.1: 3D morphological descriptors (Zheng et al., 2021) ..................................................... 44 Table 4.2: Morphological descriptors of irregularly shaped particles .......................................... 45 Table 4.3: Characteristics of fluid and solid ................................................................................. 48 v List of Figures Figure 1.1: Solid-fluid interaction simulation methods (Ibrahim & Meguid, 2020) ...................... 3 Figure 2.1: D3Q19 lattice structure .............................................................................................. 15 Figure 2.2: Interaction of two particles in DEM a) when two particles are in contact, b) the springdashpot model ............................................................................................................................... 18 Figure 2.3: The process of calculating solid fraction of cells covered by a clump of overlapping spheres. In this figure, the red box is the bounding box of each sub-particles. ............................ 22 Figure 2.4: The left figure shows how an irregularly shaped particle is modelled by clump. The right figure depicts how a clump is mapped on an LBM grid. The zoomed-in cell is a cell partially saturated by solids and is divided into five sub-cells in each direction. ....................................... 22 Figure 2.5: Flowchart of LBM-DEM coupling method (G. C. Yang et al., 2019) ....................... 25 Figure 2.6: CLUMP algorithm flowchart ..................................................................................... 28 Figure 2.7: Illustration of original STL file (left column), corresponding clump (middle column) and the subsequent shape (right column) of particles after transformation into clump ................ 29 Figure 2.8: Monte Carlo algorithm for calculating the center of mass and moment of inertia in clumps ........................................................................................................................................... 30 Figure 3.1: Domain of sedimentation of a single clump. All boundaries are considered as no-slip walls. ............................................................................................................................................. 32 Figure 3.2: Study the effect of gird resolution on the simulation of sedimentation of a spherical particle in a tank of fluid. .............................................................................................................. 33 Figure 3.3: Study the effect of relaxation time on the simulation of sedimentation of a spherical particle in a tank of fluid. .............................................................................................................. 33 vi Figure 3.4: Sedimentation velocity of an irregularly shaped particle with different numbers of subparticles in the clump. The actual shape of the particle is shown at the side................................ 34 Figure 3.5: Comparison of current simulation and experimental results (ten Cate et al., 2002) of a single clump settling in ambient fluid. .......................................................................................... 35 Figure 3.6: Comparing the simulation results with the experiment (ten Cate et al., 2002) for settling trajectory of the sphere particle at various Reynolds numbers ..................................................... 36 Figure 3.7: Vertical velocity contour of a clump at different times (Re = 11.6) .......................... 37 Figure 3.8: Domain of the benchmark case settling of two clumps. All boundaries are considered as no-slip walls. ............................................................................................................................. 38 Figure 3.9: Settling velocity of two clumped particles in comparison with the results from (Dash & Lee, 2015) ................................................................................................................................. 39 Figure 3.10: Sedimentation of two clumps at different times ....................................................... 39 Figure 3.11: Domain of the benchmark case disc shape clump. All boundaries are considered as no-slip walls. ................................................................................................................................. 40 Figure 3.12: Settling velocity of disc-shaped particle in two different cases. .............................. 42 Figure 3.13: Z-velocity contour of settling disc-shaped clump at different times (Case 2). ...... 42 Figure 4.1: Zingg diagram. (a) Jordan No. 1, (b) Jordan No. 26, (c) Jordan No. 59, (d) Jordan No. 70................................................................................................................................................... 44 Figure 4.2: Shape of Jordan sands used in this study ................................................................... 45 Figure 4.3: Sedimentation velocity of four different Jordan sand particles .................................. 46 Figure 4.4: Sedimentation of four irregularly shaped particles at t=0.12 s. (a) particle No. 1, (b) particle No. 26, (c) particle No. 59, (d) particle No. 70. ............................................................... 47 vii Figure 4.5. Domain dimensions for immersed column collapse. All boundaries are no-slip walls. ....................................................................................................................................................... 48 Figure 4.6: Collapse of column of irregularly shaped particle No. 1 at different times a) t = 0.08 s, b) t = 0.16 s, c) t = 0.39 s, d) t = 0.52 s ......................................................................................... 49 Figure 4.7: Collapse of column of irregularly shaped particle No. 26 at different times a) t = 0.08 s, b) t = 0.16 s, c) t = 0.33 s, d) t = 0.44 s ..................................................................................... 50 Figure 4.8: Collapse of column of irregularly shaped particle No. 59 at different times a) t = 0.08 s, b) t = 0.16 s, c) t = 0.46 s, d) t = 0.62 s ..................................................................................... 51 Figure 4.9: Collapse of column of irregularly shaped particle No. 70 at different times a) t = 0.08 s, b) t = 0.16 s, c) t = 0.33 s, d) t = 0.44 s ..................................................................................... 52 Figure 4.10: Collapse of column of spherical particle at different times a) t = 0.08 s, b) t = 0.16 s, c) t = 0.25 s, d) t = 0.33 s .............................................................................................................. 53 Figure 4.11: Final deposition of column collapse of spherical particles. ๐ฟ๐‘“ = 3.45053 ๐‘๐‘š ....... 54 Figure 4.12: Final deposition of column collapse of a) irregularly shaped particle No. 1 with ๐ฟ๐‘“ = 3.24631 ๐‘๐‘š, b) irregularly shaped particle No. 26 with ๐ฟ๐‘“ = 3.45091 ๐‘๐‘š, c) irregularly shaped particle No. 59 with ๐ฟ๐‘“ = 3.78061 ๐‘๐‘š , d) irregularly shaped particle No. 70 with ๐ฟ๐‘“ = 3.31487 ๐‘๐‘š .................................................................................................................................. 55 viii Glossary BGK Bhatnagar-Gross-Krook CFD Computational Fluid Dynamics CFD-DEM Coupling of Computational Fluid Dynamics and Discrete Element Method CLUMP A Code Library to generate Universal Multi-sphere Particles D3Q19 A 3D lattice structure in LBM DEM Discrete Element Method FVM Finite Volume Method IMB Immersed Moving Boundary LBM Lattice Boltzmann Method LBM-DEM Coupling of Lattice Boltzmann Method and Discrete Element Method LIGGGHTS Open-Source Discrete Element Method Particle Simulation Software MPM Material Point Method MPS Moving Particle Semi-Implicit Palabos Parallel Lattice Boltzmann Solver SPH Smoothed Particle Hydrodynamics TFM Two-Fluid Model ix Acknowledgement I would like to express my deepest gratitude to my supervisor, Dr. Wenbo Zheng, for his invaluable guidance, patience, and support throughout the duration of my research over the past two years. His wisdom and mentorship have been pivotal to my development and success in my studies. My sincere thanks also go to Dr. Lu Jing, who was instrumental in helping me compile the LBM-DEM code. His willingness to answer my questions and provide assistance during my initial phase of engaging with the current code has been greatly appreciated. A special note of thanks is reserved for Dr. Nikhil Aravindakshan, the UNBC HPC Senior Lab Instructor. His expertise and assistance in compiling the code on the UNBC workstation computer were crucial to my research's progress, and for that, I am truly thankful. I am also grateful to Natural Sciences and Engineering Research Council of Canada (NSERC) for providing the funding that supported my master's program. This financial support has been fundamental to my research and academic endeavors. My heartfelt appreciation extends to my committee members, Dr. Jueyi Sui, Dr. Fan Jiang, and Dr. Sumi Siddiqua, in advance, for their time, insightful comments, and constructive criticism on my thesis. Their feedback has been invaluable in enhancing the quality of my work. x Chapter 1: Introduction 1.1 Motivation and background The fluid-solid interaction is a fundamental concept in geotechnical engineering with practical importance in various issues involving water and soils, such as scouring phenomena around offshore mono-pile foundations (Nielsen, Probst, Petersen, & Sumer, 2015) and internal erosion in embankment dams (Foster, Fell, & Spannagle, 2000; Teng, Johansson, & Hellstrรถm, 2023). The interaction between fluids and sediment particles, characterized by a strong coupling effect, plays a crucial role in not only shaping particle movements but also influencing the fluid flow patterns (Jansen & Harting, 2011; Stratford, Adhikari, Pagonabarraga, & Desplat, 2005; T. Wang, Zhang, & Zheng, 2023). The essence of fluid-solid system behaviour lies at the microscopic scale, where the interaction between solid particles and fluid depends on parameters like fluid velocity, solid volume fraction, and particle shape. Such parameters eventually contribute to the interaction in terms of forces and momentum exchange. As conventional experimental measurements are unable to capture the intricate interplay of fluid and particle interactions, employing numerical modelling is more promising in studying the underlying mechanisms involved in a fluid-solid system (Lai, Zhao, Zhao, & Huang, 2023). Most research simulating the interaction of solid particles and fluid simplifies the shape of particles as spheres (Lai et al., 2023). However, the influence of particle irregularity on the overall characteristics of a granular flow, including its packing density, stiffness, and strength, is widely acknowledged as significant (Cho, Dodds, & Santamarina, 2007). In addition, the effect of particleโ€™s shape might get even more apparent as the goal is to eventually simulate real-world problems such as geohazards where the systems are considered at macroscopic scale. Therefore, it is crucial to study more on the feasibility of current computational methods in simulating system 1 of real shape particles submerged in fluid and consequently improving the level of accuracy in granular flows. 1.2 Literature review Finding solutions for the governing equations of granular flows through analytical methods becomes inherently complicated because of nonlinear and complex boundary conditions nature of such equations in real-world situations (Ibrahim & Meguid, 2020). As a result, the conventional recourse lies in employing numerical techniques to address these equations. Numerical analysis presents researchers with two fundamental approaches: the Eulerian method and the Lagrangian method (Ibrahim & Meguid, 2020). Within the Eulerian framework, flow variables are conceptualized as a continuous entity within spatially fixed or moving grids. The evolution of these variables over time is tracked locally within each computational cell. Conversely, the Lagrangian approach involves the tracing of trajectories and various flow parameters concerning fluid or solid particles over time, considering the behaviour of each individual particle. 1.2.1 Numerical approaches for solid-fluid interaction The complexity of granular flows necessitates a nuanced classification, which predominantly depends on the numerical treatment of both the fluid and solid phases. This classification as shown in Figure 1.1 gives rise to three main categories in granular flow modelling: (1) Eulerian-Eulerian, where both phases are treated using the Eulerian framework; (2) Eulerian-Lagrangian, where the fluid phase is handled using the Eulerian method while the solid phase follows the Lagrangian approach; and (3) Lagrangian-Lagrangian, wherein both phases are approached from a Lagrangian perspective (Ibrahim & Meguid, 2020). Each of these categories offers distinct advantages and limitations, making them suitable for specific scenarios and research objectives. Researchers often choose these methods based on the intricacies of the problem at hand, aiming to strike a balance 2 between computational efficiency and accuracy in capturing the complexities of solid-fluid interactions. Figure 1.1: Solid-fluid interaction simulation methods (Ibrahim & Meguid, 2020) The Eulerian-Eulerian approach, also known as purely Eulerian approach relies on averaging of flow variables pertaining to both solid particles and fluids. Among the array of averaging techniques available, volume averaging emerges as a prominent method, where computations occurs within finite volumes, delineated as computational cells (Crowe, Schwarzkopf, Sommerfeld, & Tsuji, 2011). A known model in using Eulerian-Eulerian approach is Two-Fluid Model (TFM) (L. Yang, Padding, & Kuipers, 2016). This model is a direct product of volume averaging methodologies, offering a unique perspective on granular flows. In the TFM, the solid phase within the flowing medium is conceptualized as a fluid-like continuum. A defining characteristic of the TFM lies in its treatment of fluid and solid phases as one continuum medium (L. Yang et al., 2016). This intricate concept postulates that both phases exist as continuous entities within the same averaging volume, albeit partially occupying it. This notion 3 challenges visualization due to the coexistence of the phases within the same space yet occupying only specific portions of that space. However, despite its applicability to specific simulation cases, the TFM grapples with limitations, particularly in its fundamental depiction of particulate dynamics at the microscale. A critical shortcoming arises from the indirect representation of particle-particle and particle-fluid interplay using averaged values derived from closures (Crowe et al., 2011). While the averaging technique could be helpful in gaseous flows involving solid particles of similar size to the averaging of volume, its application cannot be applied to cases such as dispersed solids within a fluid. The reason is that TFM lacks the granularity necessary to capture the intricacies of particulate interactions at a microscale level. In cases where precise microscale descriptions of granular interactions are crucial, such as in studies of sedimentation, particle aggregation, or multiphase flows with significant size disparities between particles, more sophisticated and nuanced modelling approaches are imperative (Ibrahim & Meguid, 2020). In contrast, fully Lagrangian methods take a markedly different route by adopting tracking of fluid and solid phase trajectories in granular flows. This Lagrangian perspective offers a comprehensive understanding of the movement and interaction of individual particles within the flow. One approach to implementing such tracking involves the utilization of a dynamically adaptive grid, where the grid adjusts its structure to follow the movement of fluid particles throughout the simulation process (Ibrahim & Meguid, 2020). Alternatively, the fluid can be represented by a collection of discrete particles devoid of any mesh structure. Methods such as Smoothed Particle Hydrodynamics (SPH) and Moving Particle Semi-Implicit (MPS) methods embrace this meshless or mesh-free paradigm, relying on the interactions between individual particles to model fluid behaviour accurately (Gingold & 4 Monaghan, 1977). In these approaches, the absence of a predefined grid structure allows for a more flexible and adaptable representation of fluid dynamics, especially in cases where traditional grid-based methods may face challenges. Additionally, there are hybrid methods like the Material Point Method (MPM), which amalgamate the utilization of a computational mesh for computational calculations (Koshizuka & Oka, 1996). Despite this grid incorporation, MPM is fundamentally categorized as a particle method due to its emphasis on tracking particle movements and interactions. This hybrid nature highlights the versatility within the realm of particulate flow modelling, where methodologies often blur traditional boundaries, emphasizing the importance of tailoring the approach to the specific intricacies of the problem at hand (Koshizuka & Oka, 1996). Thus, the landscape of granular flow simulation continues to evolve, driven by innovative techniques that blend computational precision with adaptability to capture the rich dynamics of solid-fluid interactions. 1.2.2 Eulerian-Lagrangian methods Employing Lagrangian tracking in granular flow modelling offers distinct advantages, particularly in accurately simulating flows of high convection and flows involving free surfaces. Unlike Eulerian methods, including finite volume and finite difference methods, which are prone to errors related to discretizing the convective terms, Lagrangian methods circumvent these issues (Ibrahim & Meguid, 2020). This freedom from discretization errors enhances the precision of modelling, especially in cases involving rapid fluid movements or the intricate dynamics of free surfaces. However, it is essential to acknowledge that Lagrangian methods, despite their accuracy benefits, come with their own set of challenges. Tracking particles on a large scale and employing semi-implicit solutions, which often necessitate solving Poisson's equation to get the pressure field, can significantly escalate computational costs (Pozorski & Olejnik, 2023). Additionally, the 5 dynamic adaptation of meshes in Lagrangian methods adds another layer of complexity, further contributing to higher computational demands. In contrast, the Finite Volume Method (FVM), as a conventional grid-based method, provides a relatively more streamlined computational approach, making it advantageous in scenarios where computational efficiency is a primary concern (Ibrahim & Meguid, 2020). While Lagrangian methods excel in capturing intricate flow phenomena, their computational demands prompt researchers to weigh the benefits against the costs, considering factors such as the scale of the simulation, available computational resources, and the specific intricacies of the problem at hand. In the Eulerian-Lagrangian approach, solid particle motion is followed by employing Newtonian laws of motion, allowing for discrete particle representation without relying on a computational grid. This method brings the advantage of accurately capturing the discrete characteristics of solid particles (Ibrahim & Meguid, 2020). Unlike the averaging approach of the Two-Fluid Model (TFM) that is indirect, Eulerian-Lagrangian approach enables direct implementation of solid-fluid interaction forces, including drag force and pressure difference forces. Although the estimation of these forces remains complex, surpassing the challenges of volume averaging, the Lagrangian approach provides unparalleled insights into microscale interactions inaccessible through Eulerian-Eulerian methods (Ibrahim & Meguid, 2020). Various models have been developed for both Lagrangian and Eulerian treatments in granular flows. For the fluid phase (Eulerian), the Navier-Stokes equations that are locally averaged are employed commonly (Mao, Wang, Yang, Huang, & Zheng, 2023). Alternatively, the Lattice Boltzmann method can be employed, modelling fluid flow through streaming and collision processes on a structured lattice (Sayyari, Ahmadian, & Kim, 2022). For the Lagrangian treatment 6 of solid particles, the Discrete Element Model (DEM) is the model that has been used widely (Tavarez & Plesha, 2007). These sophisticated approaches collectively refine the state of the simulation, enabling researchers to explore microscale interactions that were previously beyond reach using Eulerian-Eulerian methods. Traditional computational fluid dynamics (CFD) methods, relying on finite element or finite volume approaches, have been created for multiphase scenarios to replicate intricate fluid-solid interactions (Chu, Wang, Yu, & Vince, 2009; Peng, Doroodchi, Luo, & Moghtaderi, 2014). Alongside these established techniques, the lattice Boltzmann method (LBM) has emerged as an alternative for addressing multiphase fluid dynamics. In contrast to conventional CFD methods that solve macroscopic conservation equations (mass, momentum, energy) using numerical means, LBM takes a unique approach by simulating a system of fictitious particles that adhere to the same conservation laws. LBM offers multiple advantages in terms of intricate boundary scenarios, handling microscopic interactions, and easier parallelization (Jiang, Matsumura, Ohgi, & Chen, 2021). The Discrete Element Method (DEM) stands as a widely utilized numerical method for modelling the flow of discrete materials based on Newtonโ€™s principles of motion. The prevalence of granular flow implementations within industrial processes has spurred the creation of numerous DEM studies (Cortรฉs & Gil, 2007; Tsuji, 2007; Z. Y. Zhou, Kuang, Chu, & Yu, 2010). Through a direct approach, DEM has the capability to dynamically simulate the trajectory of each particle in dense phase flows, accurately tracking both its translational and rotational motions within the defined domain. This comprehensive simulation accounts for a particleโ€™s interactions, encompassing contact forces and couples, both among particles and between particles and boundaries (Almeida, Spogis, & Silva, 2016; El-Emam, Shi, & Zhou, 2019). 7 1.2.3 LBM-DEM coupling method The development of coupling methods in fluid-particle interactions has been marked by several studies, starting through 2D simulations of seepage flow (F. Chen, Drumm, & Guiochon, 2011), piping erosion (Lominรฉ, Scholtรจs, Sibille, & Poullain, 2013), sedimentation of circular particles (H. Zhang et al., 2014), and granular column collapse (Soundararajan, 2015). The LBM-DEM coupling, in particular, stands out for its potential in simulating fluid-particle dynamics. This preference is attributed not only to LBM's efficiency in parallel computing but also to its versatility in adopting various coupling techniques, such as immersed boundary (Feng & Michaelides, 2004), momentum exchange (Ladd, 1994), and immersed moving boundary methods (Noble & Torczynski, 1998). Moreover, LBM-DEM has seen advancements in accuracy, especially with the adoption of a two-relaxation-time collision operator, enhanced solid fraction calculation (D. Wang, Leonardi, & Aminossadati, 2018), and more realistic particle shapes. In exploring the effects of particle shape, Gardner and Sitar (Gardner & Sitar, 2019) studied the interactions of arbitrarily shaped polyhedral particles using LBM-DEM. More recently, Rettinger and Rude (Rettinger & Rรผde, 2022) introduced an innovative four-way coupling scheme that significantly enhances granular flow simulation. This method integrates a combination of the multiple-relaxation-time LBM with an adaptive strategy to control bulk viscosity near particles, thereby augmenting simulation stability. Furthermore, the incorporation of polygonal discrete elements within LBM has enabled the simulation of interactions between convex and concave polygonal particles, broadening the scope of particle dynamics analysis (M. Wang, Feng, Qu, & Zhao, 2021). Coupling LBM and DEM has been enabled by accounting for the interaction along the solid and fluid interface. Noble and Torczynski (Noble & Torczynski, 1998) introduced the concept of Immersed Moving Boundary (IMB), later combined with LBM and DEM by Cook et al. (Cook, 8 Noble, & Williams, 2004) to simulate particle flow. In this approach, the collision of interface lattice distribution functions corresponds to the solid fraction, while momentum exchange at the particle boundary calculates the hydrodynamic force reaction (R. Zhang, Kim, & Dinoy, 2021). IMBโ€™s compatibility with LBMโ€™s parallel nature and its ease of coding make it a practical choice. The accurate hydrodynamic force using IMB is easily attainable due to its physical basis. Solid fraction, crucial for hydrodynamic force calculations, has been the focus of various studies. Owen et al. (Owen, Leonardi, & Feng, 2011) divided lattice cells into subcells, using their positions relative to particle radii to determine solid fraction. Galindo-Torres (Galindo-Torres, 2013) introduced an edge-intersection averaging method to estimate the solid fraction based on interaction lengths. Another method by Jones and Williams (Jones & Williams, 2017) calculates the solid fraction directly from volume integrals, applicable to specific interaction cases. A recent study on nodal solid ratio improvement by Wang et al. (M. Wang, Feng, Owen, & Qu, 2019), reduced the computing cost by half using Gaussian quadrature for nodal solid ratio calculation. In a different paper by Wang et al. (M. Wang, Feng, Qu, Tao, & Zhao, 2020), the situation where a single nodal cell is covered by more than one particle is considered. This situation is particularly important when an irregularly shaped particle is simulated by a clump of overlapping spheres. The study proposed a method to resolve the numerical instability that may arise in calculating the weighting function in such situation. To date, LBM-DEM simulations have been used to simulate a wide range of engineering scenarios, such as sand production (Y. Han & Cundall, 2013), sediment settlement (Jiang, Liu, Chen, & Tsuji, 2022), and internal erosion (W. Zhou, Ma, Ma, Cao, & Cheng, 2020). 9 1.2.4 Simulation of irregularly shaped particles Most of the abovementioned studies simulated the solid phase as spherical particles. In the field of granular materials, especially those in geotechnical engineering, the irregular shapes of granular particles have been recognized as a crucial factor affecting the deformation, strength, and permeability of granular materials (Zheng, Hu, Tannant, & Zhou, 2021). These particles, diverging from the traditional spherical form, encompass a wide array of shapes (Zheng, Hu, Tannant, Zhang, & Xu, 2019), which can be broadly categorized into regular and irregular configurations. Regular non-spherical particles include ellipsoids, super-ellipsoids, and regular polyhedral, each possessing distinct geometric characteristics. An important distinction between non-spherical and spherical particles lies in their particle orientation. Notably, studies by Oschmann (Oschmann, Hold, & Kruggel-Emden, 2014) investigated elongated particle shapes, such as cylinders, plates, and cuboids, within fluidized beds, employing numerical simulations to explore the averaged particle orientation. These investigations show the strong influence of accurate shape representation on simulation results. Early investigations into irregular particle shapes employing coupling methods primarily concentrated on 2D simulations. Han et al. (K. Han, Feng, & Owen, 2007) studied irregular particles with polygonal shapes using the LBM-DEM coupling method within turbulent flows. Their findings showed the robustness of the coupled approach up to a certain Reynolds number threshold. Wachs et al. (Wachs, 2009) conducted relevant research by integrating a DEM solver with Finite Element methods, comparing the behaviour of disks and triangles in a 2D sedimentation scenario involving 300 particles within a closed box. This comparative analysis showed that owing to drag forces, disks settled faster than triangles. 10 As research progressed, a shift towards more realistic simulations of irregular particle shapes became evident. Shen et al. (Shen, Wang, Huang, & Jin, 2022) conducted a study toward this trend by employing a resolved CFD-DEM coupling approach to simulate the impact of a dambreak wave on a rock pile. To represent irregular particle shapes accurately, the study introduced a multisphere clump in the DEM. Moreover, the two-phase interaction involving air and fluid was resolved through the adoption of the volume of fluid method. In a more recent study, LBM-DEM was applied to simulate hydraulic plucking using three-dimensional shape of rock particles and comparing the results with the experiment (Gardner, 2023). The study showed that LBM-DEM was capable of successfully capturing dominant kinetic modes and transient displacement of the rock particle from a fractured rock mass. Moreover, Lai et al. (Lai et al., 2023) presentend a signed distance filed (SDF)-based CFD-DEM for simulating arbitrary-shaped particles in multiphase flow. The authors verified their proposed model through various benchmark cases. This evolution in research reflects that to date, there is a research gap specifically focusing on the LBM-DEM simulation of systems composed of irregularly shaped particles, indicating a novel area in the field. 1.3 Research objective and current study The main objective of this thesis is to develop a three-dimentional LBM-DEM coupling framework of a system of irregularly shaped particles using CLUMP to reconstruct the shape of particles, and then evaluate the accuracy of the method in simulating such particlesโ€™ shape. Consequently, the research tasks for this objective are as follows: โ€ข Regenerate the geometry of irregularly shaped particles through a clump of overlapping spheres using CLUMP โ€ข Investigate the potential and accuracy of LBM-DEM method in simulating irregularly shaped particles by simulating benchmark cases from the literature 11 โ€ข Study the effect of real shape particles compare to regular shape particles by simulating the well-known case of granular column collapse. Therefore, in this study, based on the research objective, an LBM-DEM coupling framework was developed to simulate the interaction between fluid and solid granular particles with irregular shapes. Then, according to CLUMP method a MATLAB code was employed to represent irregularly shaped particles with sub-particles bonded together to form clumps. Two open-source codes named Palabos and LIGGGHTS were configured for the LBM-DEM coupling of solid clumps with fluid. Three benchmark cases were considered to validate our model using clumps. Finally, to study the impact of realistic particle morphology in granular flow simulations, four real shape Jordan sand particles were chosen to simulate the collapse and transport of a granular column of irregularly shaped particles immersed in a fluid. Comparison of the results with that from regular shape particles (spheres) demonstrates the importance of particle shape in granular flow simulations. This study represents a step towards more realistic modelling of granular flows in geotechnical engineering. 1.4 Thesis organization Based on the objective outlined earlier, the organization of the current thesis is as follows. Chapter 2 explains the methodology implemented in this study, including explaining lattice Boltzmann method, discrete element modelling, and immersed moving boundary. Chapter 3 provides the benchmark cases used to validate our adopted computational method. Chapter 4 focuses on one application of current research in simulating a well-known problem in geotechnical engineering, i.e., the collapse of an immersed granular column. The chapter also investigate the mesh sensitivity, relaxation time study, and the minimum number of sub-particles that has to be used to 12 reconstruct an irregularly shaped particle. Finally, Chapter 5 provides the conclusion and limitations of the computational method and some suggestions for future studies. 13 Chapter 2: Methodology 2.1 Introduction This chapter introduces the computational method adopted throughout this study. The chapter starts with explaining the Lattice Boltzmann and Discrete Element methods. Then Immersed Moving Boundary is described as a method for handling solid boundaries. Afterward, the details of LBM-DEM implementation are provided. Eventually, the CLUMP method is explained for generating irregularly shaped particles by a cluster of overlapping spheres. 2.2 Lattice Boltzmann method The lattice Boltzmann method (LBM) is a mesoscopic scale fluid flow simulation method based on the kinetic theory (S. Chen & Doolen, 1998). In this method, the computational domain is divided into structured Cartesian lattice nodes, and then a probability distribution function (PDF) is defined on each lattice node. In other words, ๐‘“๐‘– (๐ฑ, ๐‘ก) is the PDF of the point ๐ฑ at time ๐‘ก at direction ๐‘–. In contrast to conventional CFD methods that solve nonlinear Navier-Stokes partial differential equations, the LBM takes a different approach. In LBM, the Boltzmann equation describes how PDFs evolve throughout time. Based on the BGK approximation, named after its contributors Bhatnagar-Gross-Krook (Bhatnagar, Gross, & Krook, 1954), the well-known Boltzmann equation turns into Equation (2.1). As a result of this approximation, the nonlinear Boltzmann equation becomes an explicit linear equation efficient for parallel computing. 1 ๐‘“๐‘– (๐ฑ + ๐’„๐‘– ๐›ฟ๐‘ก , ๐‘ก + ๐›ฟ๐‘ก ) โˆ’ ๐‘“๐‘– (๐ฑ, ๐‘ก) = โˆ’ [๐‘“๐‘– (๐ฑ, ๐‘ก) โˆ’ ๐‘“๐‘’๐‘ž (๐ฑ, ๐‘ก)] ๐‘– ๐œ (2.1) The governing Equation (2.1) consists of two main processes: streaming and collision. On the left-hand side (LHS) of the equation, the streaming process occurs. It involves the propagation of 14 the PDFs from one lattice node (๐ฑ) (Figure 2.1) to its neighboring node (๐ฑ + ๐œ๐‘– ๐›ฟ๐‘ก ) in the direction ๐‘– at the time ๐‘ก + ๐›ฟ๐‘ก , where ๐œ๐‘– represents the lattice velocity and ๐›ฟ๐‘ก is the increment in time t. Figure 2.1: D3Q19 lattice structure On the other hand, in Equation (2.1), the right-hand side describes the occurrence of the collision process. In this process, the PDFs undergo a linear relaxation to the equilibrium distribution functions (EDFs), ๐‘“๐‘–๐‘’๐‘ž (๐‘ฅ, ๐‘ก), using a single dimensionless relaxation time (๐œ). The EDF employed in this context is the Maxwellian distribution, which uses the Taylor series and can be expanded regarding the macroscopic fluid velocity (๐ฎ๐‘“ ) (Aidun & Clausen, 2010). 2 ๐œ๐‘– . ๐ฎ๐‘“ (๐œ๐‘– . ๐ฎ๐‘“ ) ๐‘ข๐‘“ 2 ๐‘’๐‘ž ๐‘“๐‘– = ๐‘ค๐‘– ๐œŒ๐‘“ [1 + 2 + โˆ’ ] (๐‘– = 0, โ€ฆ ,18) ๐‘๐‘  2๐‘4๐‘  2๐‘2๐‘  (2.2) where ๐‘ค๐‘– is the weight factor related to the corresponding ๐œ๐‘– direction. The speed of sound ๐‘๐‘  in lattice is 1โ„โˆš3. Note that a D3Q19 lattice structure, as shown in Figure 2.1, is adopted in this study. The associated lattice velocity and weight factors are shown in Table 2.1. 15 Table 2.1: Values of ๐œ๐‘– , ๐‘ค๐‘– for a D3Q19 lattice structure (G. C. Yang, Jing, Kwok, & Sobral, 2019). PDFs, ๐‘“๐‘– Lattice velocity, ๐’„๐‘– Weights, ๐‘ค๐‘– ๐‘“0 (0,0,0) 1/3 ๐‘“1 โˆ’ ๐‘“6 (ยฑ1,0,0), (0, ยฑ1,0), (0,0, ยฑ1) 1/18 ๐‘“7 โˆ’ ๐‘“18 (ยฑ1, ยฑ1,0), (ยฑ1,0, ยฑ1), (0, ยฑ1, ยฑ1) 1/36 After calculating ๐‘“๐‘– at each node, the fluid density ๐œŒ๐‘“ and velocity ๐ฎ๐‘“ can be reconstructed using the following equations according to the conservation laws of mass and momentum (He & Luo, 1997): 18 (2.3) ๐œŒ๐‘“ = โˆ‘ ๐‘“๐‘– ๐‘–=0 18 (2.4) ๐œŒ๐‘“ ๐ฎ๐‘“ = โˆ‘ ๐œ๐‘– ๐‘“๐‘– ๐‘–=0 Moreover, the kinematic fluid viscosity, ๐‘ฃ๐‘“ , is a material property constant and has the following relationship with the relaxation time, LBM timestep, and lattice spacing (He & Luo, 1997): 1 ๐›ฟ๐‘ฅ2 ๐‘ฃ๐‘“ = ๐‘๐‘ 2 (๐œ โˆ’ ) 2 ๐›ฟ๐‘ก (2.5) Pressure as other macroscopic variable is related to fluid density based on the equation of state (He & Luo, 1997): ๐‘ = ๐‘๐‘ 2 ๐œŒ๐‘“ (2.6) One of the primary sources of compressibility error in LBM arises from the truncated terms of Taylor expansion of the equilibrium distribution functions. In order to approximate an incompressible flow, the Mach number (Ma) must remain below unity (๐‘€๐‘Ž โ‰ช 1). The necessity 16 for incompressibility in LBM simulations imposes constraints on selecting the LBM relaxation time. In other words, while based on Equation (2.5), the relaxation time has a bottom threshold of 0.5, the higher relaxation time should be set in a way to ensure that the LBM accurately captures the behaviour of incompressible flows while minimizing compressibility-related errors (G. C. Yang et al., 2019). 2.3 Discrete element method In the discrete element method (DEM), granular materials are represented as a collection of individual particles, where interactions between neighbouring particles are explicitly computed using simplified contact laws (Sun, Sakai, & Yamada, 2013). In this study, an open-source DEM software, LIGGGHTS, was employed to simulate the interactions among solid particles. In each iteration of the DEM process, Newtonโ€™s second law determines the amount of particle translation and rotation. Each iteration starts with particle-particle contacts assessment, and accordingly, the corresponding forces are computed. Subsequently, particlesโ€™ linear and rotational velocities are updated. Afterward, new particle positions will be determined based on Newtonโ€™s law of motion (Ding & Xu, 2018). Figure 2.2 illustrates particle โ€œaโ€ (shown as green) and particle โ€œbโ€ (shown as red) are in contact with each other, which ๐œน๐‘› denotes the overlap. The overlap can be calculated using the following equation (G. C. Yang et al., 2019): (2.7) ๐œน๐‘› = (๐‘Ÿ๐‘Ž + ๐‘Ÿ๐‘ โˆ’ ๐‘Ÿ๐‘Ž๐‘ )๐ง where ๐‘Ÿ๐‘Ž and ๐‘Ÿ๐‘ are radii of particles ๐‘Ž and ๐‘, respectively. The distance between two particles is represented by ๐‘Ÿ๐‘Ž๐‘ , and ๐ง is the normal unit vector directed toward particle centers. 17 (a) (b) Figure 2.2: Interaction of two particles in DEM a) when two particles are in contact, b) the spring-dashpot model To calculate the contact force between particles, constitutive contact models can be used, and this study uses the spring-dashpot model. In this case, the following equation gives the normal force (Cundall & Strack, 1979): (2.8) ๐…๐‘› = ๐‘˜๐‘› ๐›ฟ๐‘› + ๐‘๐‘› โˆ†๐ฎ๐‘› where ๐‘˜๐‘› is stiffness and ๐‘๐‘› is damping coefficient, both in the normal direction. โˆ†๐ฎ๐‘› is the relative normal velocity of two interacting particles. In this study, we chose Hertz contact model to calculate the stiffness and damping coefficient as follows (Di Renzo & Di Maio, 2004): 4 โˆ— โˆ— ๐ธ โˆš๐‘… ๐›ฟ๐‘› 3 (2.9) 5 ๐‘๐‘› = โˆ’2โˆš ๐›ฝโˆš๐‘šโˆ— ๐‘†๐‘› 6 (2.10) ๐‘˜๐‘› = where ๐ธ โˆ— , ๐‘… โˆ— , ๐‘šโˆ— are Young modulus, radius and mass equivalent, ๐›ฝ is damping ratio, and ๐‘†๐‘› can be calculated using the following equations (Di Renzo & Di Maio, 2004). 1 (1 โˆ’ ๐œ—12 ) (1 โˆ’ ๐œ—22 ) = + ๐ธโˆ— ๐ธ1 ๐ธ2 (2.11) 18 1 1 1 = + โˆ— ๐‘… ๐‘…1 ๐‘…2 (2.12) 1 1 1 = + โˆ— ๐‘š ๐‘š1 ๐‘š2 (2.13) ๐›ฝ= ln(๐‘’) (2.14) โˆšln2 (๐‘’) + ๐œ‹ 2 (2.15) ๐‘†๐‘› = 2๐ธ โˆ— โˆš๐‘… โˆ— ๐›ฟ๐‘› In the above equations, ๐ธ is the Young modulus, ๐œ— is the Poisson ration, ๐‘’ is the coefficient of restitution, ๐‘š is the mass, and ๐‘… is the radius of each two interacting particles. In the same manner, the tangential force is given as follows (Cundall & Strack, 1979): ๐‘ก๐‘ (2.16) ๐…๐‘ก = โˆ’๐‘˜๐‘ก โˆซ โˆ†๐ฎ๐‘ก ๐‘‘๐‘ก โˆ’ ๐‘๐‘ก โˆ†๐ฎ๐‘ก ๐‘ก๐‘,0 where ๐‘˜๐‘ก and ๐‘๐‘ก are corresponding stiffness and damping coefficient in the tangential direction, and โˆ†๐ฎ๐‘ก denotes the relative tangential velocity of two particles. The integral term in the equation represents the increment of spring, which accumulates energy resulting from the relative motion in the tangential direction between contacting particles. This integral accounts for the deformation of the particle surface (elastic) throughout the time the particles are in contact from ๐‘ก๐‘,0 to ๐‘ก๐‘ . The tangential force exerted between the particles acts in the opposite direction of the displacement in the tangential direction. Additionally, the tangential force magnitude is restricted by the Coulomb friction, denoted as ๐œ‡๐น๐‘› . The value of ๐œ‡ corresponds to the smaller of the two friction coefficients of the contacting particles, and it represents the threshold at which the particles initiate sliding motion against each other. Also based on Hertz model adopted in this study, ๐‘˜๐‘ก and ๐‘๐‘ก can be calculated as follows (Di Renzo & Di Maio, 2004): 19 ๐‘˜๐‘ก = 8๐บ โˆ— โˆš๐‘… โˆ— ๐›ฟ๐‘› (2.17) 5 ๐‘๐‘ก = โˆ’2โˆš ๐›ฝโˆš๐‘šโˆ— ๐‘†๐‘ก 6 (2.18) where the following equations give ๐‘†๐‘ก and ๐บ โˆ— . 1 2(2 โˆ’ ๐œ—1 )(1 + ๐œ—1 ) 2(2 โˆ’ ๐œ—2 )(1 + ๐œ—2 ) = + โˆ— ๐บ ๐ธ1 ๐ธ2 โˆ— โˆ— ๐‘†๐‘ก = 2๐บ โˆš๐‘… ๐›ฟ๐‘› (2.19) (2.20) As mentioned, there are different proposed contact models that calculate contact force (๐…๐‘ = ๐…๐‘› + ๐…๐‘ก ) as a function of relative velocity and distance. In our study, we adopted the Hertz contact model (Jing, Yang, Kwok, & Sobral, 2019). To update the velocities of a particle (both linear and angular), Newtonโ€™s second law of motion takes into account the forces (contact force ๐…๐‘ , gravity ๐†, fluid drag force ๐…๐‘“ ) and torques (contact torque ๐“๐‘ , fluid drag torque ๐“๐‘“ ) exerting on the particle. The linear velocity is updated by considering the net force acting on the particle and dividing it by its mass, while the angular velocity is updated by considering the net torque acting on the particle and dividing it by its moment of inertia. By applying Newtonโ€™s second law, the particleโ€™s linear and angular velocities can be appropriately adjusted to account for the combined effects of these forces and torques. ๐‘š๐š = ๐…๐‘ + ๐† + ๐…๐‘“ (2.21) ๐ผ๐›šฬ‡ = ๐“๐‘ + ๐“๐‘“ (2.22) where ๐‘š is mass and ๐ผ is the moment of inertia of particles. ๐š and ๐›šฬ‡ are translational and rotational acceleration of particles, respectively. Eventually, by using Equations (2.21) and (2.22) via the Verlet method (Verlet, 1967), the particlesโ€™ position and orientation can be updated at each timestep. 20 2.4 Immersed moving boundary method The Immersed Moving Boundary (IMB) method operates based on the principle of introducing a new collision operator ฮฉ that depends on the solid ratio ๐œ€. While the ideal value of this ratio can be obtained through a geometric analysis, it often demands significant computational resources (G. C. Yang et al., 2019). To address this challenge, a cell decomposition method is employed (Owen et al., 2011), as depicted in Figure 2.3. According to this method, fluid cells that are partially saturated by solids are subdivided into equal-sized sub-cells, ๐‘›๐‘ ๐‘ข๐‘ . An algorithm is applied to the centers of these sub-cells to estimate the solid ratio, ๐œ€, by dividing the number of 3 sub-cells located inside the solid boundary over the total sub-cell number ๐‘›๐‘ ๐‘ข๐‘ with one cell (G. C. Yang et al., 2019). For the implementation of IMB, first, each sphere is encased within its own bounding box (Figure 2.3). The core of this method involves calculating the solid fraction for each cell of the bounding box in relation to the sphere's center. Specifically, cells within a distance less than the radius from the center are classified as solid (solid fraction = 1). Conversely, cells beyond the radius have a solid fraction of 0. In the case, the cell is covered by more that two particles (Figure 2.3), the solid fraction of the cell will be calculated based on the particle that has a higher solid fraction in the cell. For the partially saturated cells, the method divides them into smaller, equalsized sub-cells, as shown in Figure 2.3 (denoted as ๐‘›๐‘ ๐‘ข๐‘ ). Then the cell decomposition method is employed (Owen et al., 2011), as depicted in Figure 2.3. An algorithm is then applied to the centers of these sub-cells to estimate the solid ratio, ๐œ€, by dividing the number of sub-cells located inside 3 the solid boundary over the total sub-cell number ๐‘›๐‘ ๐‘ข๐‘ with one cell (G. C. Yang et al., 2019). The solid fraction (๐œ€) in these cells is then determined by comparing the number of sub-cells within the solid boundary to the total number of sub-cells. 21 An important aspect of this process is the assignment of particle IDs. Cells that are fully or partially solid receive the ID of the occupying particle, while purely fluid cells are assigned an ID of zero. This approach offers two significant benefits. Firstly, for cells in the overlapping region of two spheres, the cell is assigned only one particle ID, preventing the duplication of force and momentum calculations. Secondly, as each cell carries a particle ID, calculating the total hydrodynamic force on each particle becomes more efficient โ€“ the code simply sums the forces across cells sharing the same particle ID. Figure 2.3: The process of calculating solid fraction of cells covered by a clump of overlapping spheres. In this figure, the red box is the bounding box of each sub-particles. Figure 2.4: The left figure shows how an irregularly shaped particle is modelled by clump. The right figure depicts how a clump is mapped on an LBM grid. The zoomed-in cell is a cell partially saturated by solids and is divided into five sub-cells in each direction. In fluid cells where ๐œ€ is equal to 0, the aforementioned fluid hydrodynamic collision occurs, which is represented by the BGK collision operator ฮฉ ๐‘“ , as provided in the Equation (2.1). In solid cells where the solid ratio is equal to 1, a collision operator given by Noble and Torczynski (Noble 22 & Torczynski, 1998), according to the concept of non-equilibrium bounce-back (Zou & He, 1997), is employed. This collision operator is denoted as ฮฉ๐‘  , is given by the following expression: ๐‘’๐‘ž โ„ฆ๐‘–๐‘  = ๐‘“โˆ’๐‘– (๐ฑ, ๐‘ก) โˆ’ ๐‘“โˆ’๐‘– (๐œŒ๐‘“ , ๐ฎ๐‘“ ) + ๐‘“๐‘–๐‘’๐‘ž (๐œŒ๐‘“ , ๐ฎ๐‘  ) โˆ’ ๐‘“๐‘– (๐ฑ, ๐‘ก) (2.23) where ๐ฎ๐‘  is the macroscopic velocity of the solid at the location of the lattice node position. The subscript, โˆ’๐‘–, means the opposite direction of ๐‘–. The purpose of the solid collision operator (ฮฉ๐‘  ) is to enforce a no-slip boundary condition between two phases of fluid and solid. This is achieved by setting the PDF ๐‘“๐‘– (๐ฑ + ๐’„๐‘– ๐›ฟ๐‘ก , ๐‘ก + ๐›ฟ๐‘ก ) equal to the equilibrium distribution function (EDF), ๐‘“๐‘’๐‘ž , ๐ฎ , plus the bounce-back of the non-equilibrium part in the opposite direction (G. C. Yang ๐‘– (๐œŒ๐‘“ ๐‘  ) et al., 2019). The non-equilibrium bounce-back term ensures that the fluid particles reflect off the solid surface with a velocity corresponding to the macroscopic velocity of the solid phase at that lattice node. In case the cell is partially saturated, namely ๐œ€ is between 0 and 1; a weighting factor incorporates both collision operators into account. (2.24) โ„ฆ = ๐ตโ„ฆ๐‘  + (1 โˆ’ ๐ต)โ„ฆ ๐‘“ where the weighting factor proposed by Noble and Torczynski (Noble & Torczynski, 1998) can be calculated based on the relaxation time and the solid ratio as follows: ๐ต(๐œ€, ๐œ) = ๐œ€(๐œ โˆ’ 1โ„2) (1 โˆ’ ๐œ€) + (๐œ โˆ’ 1โ„2) (2.25) Afterward, the following equations provide the hydrodynamic force and torque at each node according to the lattice structure that was adopted earlier. ๐‘› 18 ๐…๐‘“ = โˆ‘ ๐ต๐‘— โˆ‘ โ„ฆ๐‘–๐‘  ๐œ๐‘– ๐‘—=1 (2.26) ๐‘–=0 23 ๐‘› 18 ๐“๐‘“ = โˆ‘ [๐ต๐‘— (๐ฑ๐‘— โˆ’ ๐ฑ ๐‘  ) ร— โˆ‘ โ„ฆ๐‘–๐‘  ๐œ๐‘– ] ๐‘—=1 (2.27) ๐‘–=0 where ๐ฑ ๐‘  is the center of mass of solid particles and ๐ฑ๐‘— is the jth lattice cell coordinates. When the hydrodynamic force and torque are calculated from Equations (2.26) and (2.27), then they are incorporated again in Equations (2.21) and (2.22) to update solid particlesโ€™ velocity and location. By utilizing these equations, the particleโ€™s motion and position are adjusted based on the calculated hydrodynamic forces and torques, ensuring the accurate representation of the particleโ€™s dynamics within the simulation. 2.5 LBM-DEM coupling scheme The coupled LBM-DEM code (Seil & Pirker, 2017) used in this study utilizes Palabos (Latt et al., 2021), an open-source C++ library for LBM, as the master program. This master program then calls an external library, LIGGGHTS (Kloss, Goniva, Hager, Amberger, & Pirker, 2012), for DEM simulations. The flowchart in Figure 2.5 illustrates the coupling scheme employed in this study. The computational cycle begins by generating DEM particles and then initializing the fluid field. Subsequently, interactions of particles are calculated using Equations (2.8) and (2.16). 24 Figure 2.5: Flowchart of LBM-DEM coupling method (G. C. Yang et al., 2019) In order to ensure stability in the DEM simulation, it is crucial for the DEM time step (t) to be smaller than a critical time step (๐‘ก๐‘๐‘Ÿ ) that can be obtained by the stiffness and mass of the particles. Therefore, the calculated critical DEM time step (๐‘ก๐‘๐‘Ÿ ) is generally smaller than the time step (t) used in LBM simulations, particularly for geotechnical engineering problems where soils and rocks exhibit high stiffness (G. C. Yang et al., 2019). To synchronize the DEM simulations with LBM, ๐‘๐‘ ๐‘ข๐‘ DEM sub-cycles are performed for each LBM evolution step, resulting in the following relationship: โˆ†๐‘ก๐ท๐ธ๐‘€ = ๐›ฟ๐‘ก ๐‘๐‘ ๐‘ข๐‘ (2.28) 25 which means that for every LBM timestep ๐›ฟ๐‘ก , there are ๐‘๐‘ ๐‘ข๐‘ DEM sub-cycles during which hydrodynamic force and torque are constant. In an LBM-DEM simulation, once DEM sub-cycles (๐‘๐‘ ๐‘ข๐‘ ) are completed, and the updated positions of the particles are mapped onto the lattice grid. This mapping process identifies the lattice cells that are covered by solid particles. For partially saturated cells, the solid ratio is determined using the cell decomposition method, as depicted in Figure 2.3. Based on the type of the cell (fluid, partially saturated, or solid), the collision process happens. It is worth noting that the modified collision operator is also present in Equations (2.26) and (2.27), enabling the calculation of hydrodynamic forces and torques immediately after the collision process, before the streaming step, thus ensuring high computational efficiency. Subsequently, the hydrodynamic forces are returned to the DEM to determine the motion of the particles. The resulting PDFs are then streamed to the neighbouring lattice nodes. Using the redistributed PDFs, the fluid density (๐œŒ๐‘“ ) and velocity (๐ฎ๐‘“ ) are updated according to Equations (2.3) and (2.4). If necessary, the fluid pressure field can be obtained using Equation (2.6). After all the steps are taken, one cycle of the LBM-DEM simulation concludes, and the simulation continues until it reaches the number of cycles that is preset. 2.6 Irregular shape simulation The practice of modelling real particles as clumps of overlapping spheres or clusters of nonoverlapping spheres has been widely adopted in many studies, particularly when particle morphology is a key consideration. Simulating irregularly shaped particles has various applications in agriculture (Pasha, Hare, Ghadiri, Gunadi, & Piccione, 2016), civil engineering (Suhr & Six, 2020), chemical engineering (Nan, Ghadiri, & Wang, 2017) as well as geotechnical engineering (Kodicherla, Gong, Fan, Wilkinson, & Moy, 2020). In this regard, there are a number 26 of methods proposed in the literature, including Wang et al. (L. Wang, Park, & Fu, 2007), Ferellec and McDowell (Ferellec & McDowell, 2010), Favier et al. (Favier, Abbaspourโ€Fard, Kremmer, & Raji, 1999), and Taghavi (Taghavi, 2011), to name a few. The used open-source code, LIGGGHTS, does not have functions for generating clumps. Therefore, this study adopted CLUMP (Angelidakis, Nadimi, Otsubo, & Utili, 2021), a MATLAB code that has modules to generate and export clumps based on Euclidean distance transform. The code has the capability to even transform the self-generated clump into surface, that is essential if user would like to compare the morphological characterisation of the generated volume to the original particle. In this clump-generating code, the particle morphology undergoes a transformation from a surface mesh into a voxelated representation. The flowchart of the code is shown in Figure 2.6. The user can set a minimum dimension to control the number of voxels representing the particle. Within the voxelated image, all voxels corresponding to the particle are assigned the value zero. The Euclidean distance transform calculates the minimum distance between each zero voxel to its nearest non-zero voxel. Afterward, the voxel that has the maximum value of Euclidean distance transform is set as the first inscribed sphere and the corresponding voxels are set as one. After that, another Euclidean distance transform calculates the new voxelated image. This process repeats until one of the two conditions is met. The first condition happens when the number of spheres passes the maximum user-defined number of spheres. The second condition occurs when the algorithm wants to generate a sphere that is smaller than the user-defined minimum radius. Moreover, this approach allows for generating overlapping spheres by selectively setting a certain percentage of voxels to one, rather than all, forming each new sphere. The extent of overlap is controlled using a variable overlap, which can take values within the range of [0, 1), providing flexibility in managing the degree of overlap among the spheres produced. For this study, all 27 shapes are represented by 100 sub-particles, and the rest of the setting is brought in Table 2.2. Figure 2.7 also provides the illustration of two particle geometries after transformation into a clump. Figure 2.6: CLUMP algorithm flowchart 28 Figure 2.7: Illustration of original STL file (left column), corresponding clump (middle column) and the subsequent shape (right column) of particles after transformation into clump Table 2.2: Parameters used in CLUMP algorithm for generating clumps Minimum dimension Minimum radius Minimum No. of spheres Maximum overlap 400 0 100 95% The generated clumps consist of overlapping spheres, and it is important to correctly calculate the center of mass and moment of inertia for every clump before the simulation starts. This calculation occurs in DEM, and it involves the initial generation of an oriented bounding box for a given multi-sphere clump (Figure 2.8), followed by the implementation of a Monte Carlo simulation process (Y. Zhou, Zhou, Li, & Wang, 2017). Within this process, thousands of points (denoted as N) were randomly distributed within the bounding box. Subsequently, points located outside the clump were eliminated (red dots), while those within the clump were retained (black dots). The number of points remaining inside the clump was denoted as n, and the ratio n/N was 29 considered as an approximate volume ratio representing the proportion of the clump within the bounding box. Figure 2.8: Monte Carlo algorithm for calculating the center of mass and moment of inertia in clumps The next step involved treating the retained points as mass-based voxels. Various kinematic parameters, including mass center and inertia tensor, were then obtained by integrating these voxels in accordance with their mathematical definitions (Y. Zhou et al., 2017). Notably, this approach offers a distinct advantage by excluding the influence of overlapping density. Consequently, it facilitates the extraction of true kinematic parameters for the clump, providing a more accurate representation of its mass distribution and inertial properties (Y. Zhou et al., 2017). This methodology is particularly advantageous in situations where obtaining precise kinematic parameters without the interference of overlapping density is crucial for the analysis of clump behaviour. 30 Chapter 3: Validation 3.1 Introduction This chapter is dedicated to validating LBM-DEM coupling method against known benchmark cases from the literature. Before this, the chapter starts with finding the right grid size, relaxation time, and number of sub-particles that ensure accurate result by conducting a sensitivity study on mesh size, relaxation time, and number of sub-particles needed to reconstruct an irregularly shaped template. Then, the results for the three benchmark cases, which are 1) settling of a single clump, 2) settling of two clumps, and 3) settling of a disc-shaped particle, are provided. In all cases, the particleโ€™s shape is reconstructed through CLUMP as a volume of overlapping sub-particle spheres. 3.2 Mesh sensitivity, relaxation time, and sub-particle study In any computational fluid dynamics simulation, ensuring that the grid size and time step do not affect the results is crucial. In the LBM, this process occurs through investigating different mesh resolutions and relaxation time, which is another representation of time step in LBM. The simulation case chosen for this purpose is the sedimentation of a single sphere in a fluid as shown in Figure 3.1. The sphere shape is reconstructed with 100 sub-particles. The diameter of the clump is 15 mm, and the fluid density and viscosity are 1000 ๐‘˜๐‘”โ„๐‘š3 , and 0.001๐‘ƒ๐‘Ž. ๐‘ , respectively. The density of the particle is 1120 ๐‘˜๐‘”โ„๐‘š3 . 31 Figure 3.1: Domain of sedimentation of a single clump. All boundaries are considered as no-slip walls. For the resolution study, Figure 3.2 demonstrates the effect of four different resolutions, ๐‘ = 10,15, 20, 25. It is worth noting that the number ๐‘ in each case represents the number of cells in one diameter of sphere, which is 15 mm in this case. In order to depict the difference between each grid size, Figure 3.2 only shows the last two seconds of the sedimentation, as the discrepancies in this interval is more apparent. According to the figure, while there is an apparent discrepancy between ๐‘ = 10 and the rest, ๐‘ = 20 is almost identical to ๐‘ = 25 meaning that ๐‘ = 20 would be the right chose for the grid size resolution. Afterward, by setting N to 20, Figure 3.3 displays the effect of relaxation time on the simulation outcome. The figure displays that as the value for the relaxation factor gets closer to 0.5, the velocity trend gets smoother around the time the particle reaches the bottom. Therefore, as a trade-off between accuracy and computational cost, in this study we chose 0.56 or less for most of the simulation cases. 32 Figure 3.2: Study the effect of gird resolution on the simulation of sedimentation of a spherical particle in a tank of fluid. Figure 3.3: Study the effect of relaxation time on the simulation of sedimentation of a spherical particle in a tank of fluid. Moreover, to investigate the impact of the sub-particle number used in generating a clump, Figure 3.4 illustrates the sedimentation velocity of the same irregularly shaped particle using 33 different number of sub-particles. When only 20 spheres are used to create the geometry of the irregularly shaped particle, the terminal velocity is off by 0.005 m/s compared with other results. However, as the number of particles increases, the terminal velocity converges to a certain value. Therefore, in this study, 100 is chosen as the number of sub-particles needed to generate a clump. Figure 3.4: Sedimentation velocity of an irregularly shaped particle with different numbers of sub-particles in the clump. The actual shape of the particle is shown at the side. 3.3 Benchmark Cases 3.3.1 Settling of a single clump For the first benchmark case, the settling of a heavy single particle in an ambient fluid is simulated. The single spherical particle shown in Figure 2.7 is regenerated using 100 sub-particles for a clump. The simulation domain is same as Figure 3.1. The diameter of the clump is 15 mm, and its density is 1120 ๐‘˜๐‘”โ„๐‘š3 . The fluid density and viscosity are according to (ten Cate, Nieuwstad, Derksen, & Van den Akker, 2002). 34 For this simulation, the lattice resolution and relaxation times are 20 and 0.56, respectively. By considering the diameter of the clump particle as the length scale and the particleโ€™s terminal velocity as the characteristic velocity, the simulation was performed for four Reynolds numbers (Re), namely 1.48, 4.1, 11.6, and 31.9. The Reynolds number is calculated as follows: ๐‘…๐‘’ = ๐‘ˆ๐‘‘๐‘ ๐œ— 3.1) where ๐‘ˆ is the terminal velocity of the particle, ๐‘‘๐‘ is the diameter of the particle, and ๐œ— is the kinetic viscosity. The results were compared to experimental data, as shown in Figure 3.5. Figure 3.5: Comparison of current simulation and experimental results (ten Cate et al., 2002) of a single clump settling in ambient fluid. In Figure 3.5, for low Reynolds numbers, the velocity of the particle first increases until it reaches its terminal velocity, where the buoyancy force is equal to gravity. Subsequently, as the particle gets closer to the bottom, the velocity gradually decreases. For higher Reynolds numbers, the same process is observed but the particle settles faster. Overall, Figure 3.5 demonstrates that 35 the sedimentation velocity of a spherical particle simulated by a clump is in good agreement with the experimental results at each of the four Reynolds numbers (ten Cate et al., 2002). Figure 3.6 presents the settling trajectory of the sphere at different Reynolds numbers. According to the figure, lower Reynolds number takes much more time for the particle to reach the bottom; and as the Reynolds number increases, the particle falls faster. Figure 3.6 once again verifies that the code can accurately simulate both the velocity and the trajectory of the particle with little discrepancies. Figure 3.6: Comparing the simulation results with the experiment (ten Cate et al., 2002) for settling trajectory of the sphere particle at various Reynolds numbers Figure 3.7 provides an appreciation of the vertical velocity contour of the x-plane that is located in the middle of the tank. The figure depicts that as the particle gains more speed, the wake that occurs behind the particle also gets bigger. The wake region has lower pressure than the surroundings. Therefore, as the wake region grows, the pressure difference between the front and back of the particle increases, leading to a higher drag force. In this regard, the particle loses its velocity as it approaches the bottom of the tank. 36 (a) t = 0.72 s (b) t = 1.48 s Figure 3.7: Vertical velocity contour of a clump at different times (Re = 11.6) 3.3.2 Settling of two clumps In the second benchmark validation, the interaction between two clumps falling from a certain height is investigated. Figure 3.8 shows the dimensions of the simulated domain. All boundaries are no-slip wall. In this case, the density of particles is 1350 kgโ„m3 with the diameter equal to 12.7 mm. The Young modulus and passion ratio of each particle are set 5 MPa, and 0.4 respectively. For this simulation, the lattice resolution and relaxation times are 20 and 0.52, respectively. In addition, fluid density and viscosity are 1195 kgโ„m3 and 0.0305 Pa. s, respectively. Figure 3.9 shows the simulation results in comparison with those by Dash and Lee (Dash & Lee, 2015). The following equations are used to non-dimensionalize the time and velocity in Figure 3.9. ๐‘ก๐‘ = โˆš๐‘‘๐‘ /|๐œŒ๐‘ โ„๐œŒ๐‘“ โˆ’ 1|๐‘” (3.2) ๐‘ˆ๐‘ = โˆš|๐œŒ๐‘ โ„๐œŒ๐‘“ โˆ’ 1|๐‘”๐‘‘๐‘ (3.3) 37 Figure 3.8: Domain of the benchmark case settling of two clumps. All boundaries are considered as no-slip walls. As shown in Figure 3.9, when two particles start to fall (Figure 3.10a), they initially have the same speed, but after some time, the trailing particle gets more speed since it is located at the wake of the leading particle, also called drafting (Figure 3.10b). Then, the particles start to contact each other (kissing, Figure 3.10c) until they begin to tumble and separate at a later stage (Figure 3.10d). While there is a slight discrepancy between the two results, Figure 3.9 shows that our simulation of clump particles can follow all three stages of drafting, kissing, and tumbling, which are associated with this benchmark. Note that in this simulation, in order to capture the aforementioned three stages, a small offset is introduced to the initial position of two clumps, in which the position of leading clump is ( 5.5๐‘‘๐‘ + โˆ†๐‘ฅ, 5.5๐‘‘๐‘ + โˆ†๐‘ฆ, 35๐‘‘๐‘ ) and the position of trailing clump is (5.5๐‘‘๐‘ , 5.5๐‘‘๐‘ , 37๐‘‘๐‘ ). 38 Figure 3.9: Settling velocity of two clumped particles in comparison with the results from (Dash & Lee, 2015) (a) t = 0 s (b) t = 1.2 s (c) t = 1.44 s (d) t = 2.24 s Figure 3.10: Sedimentation of two clumps at different times 39 3.3.3 Settling of a disc-shaped clump The third simulated benchmark case is the settling of a disc-shaped clump. The shape of the disc shown in Figure 2.7 is generated using a clump of 100 sub-particles. The lattice resolution and relaxation times are 20 and 0.55, respectively. Furthermore, the initial orientation of the disc is set to 45ยฐ as shown in Figure 3.11. According to the literature, the settling process of a disc-shaped particle depends on the dimension of the disc. Therefore, we used this point to bring another validation to our code. Figure 3.11: Domain of the benchmark case disc shape clump. All boundaries are considered as no-slip walls. There are two non-dimensional numbers that characterize the settling process of a disc-shaped particle, the Reynolds number and the non-dimensional moment of inertia, with the following definition (Willmarth, Hawk, & Harvey, 1964). ๐ผโˆ— = ๐œ‹๐œŒ๐‘ ๐‘ก 64๐œŒ๐‘“ ๐‘‘ (3.4) In this equation, ๐‘ก, ๐‘‘, and ๐œŒ๐‘ are the thickness, diameter, and density of the disc, respectively. The study presents that the falling of a disc in a fluid depends on the value of the two numbers (๐ผ โˆ— 40 and ๐‘…๐‘’) and could have either stable or unstable falling patterns. Therefore, we specifically set the properties of the fluid and the disc to simulate two scenarios, including stable falling (case 1) and periodic oscillation (case 2). Table 3.1 gives the characteristics of the domain for two cases. In addition, Table 3.2 provides the value of the non-dimensional numbers for each case. It should be mentioned that in both cases, the thickness of the disc is 0.3 times the diameter. Table 3.1: Non-dimensional number values for two cases of disc-shaped particle benchmark. Case 1 Case 2 ๐ผโˆ— 0.0191 0.0398 ๐‘ˆ๐‘ง,๐‘š๐‘’๐‘Ž๐‘› (๐‘š/๐‘ ) 0.021 0.1087 ๐‘…๐‘’๐‘ 21 267 Table 3.2: Characteristics of two cases related to disc-shaped particle benchmark. Case 1 Case 2 ๐œŒ๐‘ (๐‘˜๐‘”โ„๐‘š3 ) 1300 2700 d (mm) 1 mm 2 mm Box size (mm) 6ร—6ร—15 12ร—12ร—48 Figure 3.12 shows the settling velocity versus time for the two cases, which correctly simulated the falling pattern of the disc-shaped particle consistent with those predicted by (Willmarth et al., 1964). Moreover, Figure 3.13 provides further illustration of a disc falling with the oscillating pattern. As the figure shows, based on the non-dimensional number of the particle in Case 1, the particle correctly follows a stable and non-oscillating settling pattern. However, in Case 2, the particle has more moment of inertia compared to Case 1, plus a higher Reynolds number, both of which set the particle unstable and with a certain oscillating frequency. 41 Figure 3.12: Settling velocity of disc-shaped particle in two different cases. (a) t = 0.14 s (b) t = 0.28 s (c) t = 0.38 s Figure 3.13: Z-velocity contour of settling disc-shaped clump at different times (Case 2). 42 Chapter 4: Immersed granular collapse 4.1 Introduction In this chapter, a study is conducted to demonstrate the potential of LBM-DEM in simulating a system of irregularly shaped particles. The case study is a pool of fluid with a column of real shaped particles initially stacked on one side of the domain. As the simulation begins, the column starts to collapse due to gravity force until it reaches its equilibrium state. This case study is performed with four different particle shapes in addition to the one with perfectly sphere particles. The chapter also looks into how morphological descriptors affect a particleโ€™s drag force by studying the sedimentation of each particle shape. It is worth mentioning that the computational resource that is used in this study is a workstation with 16 GB RAM, and 11th Generation Intel CoreTM i7-11700 @ 2.5 GHz ร— 16 of processor. All the simulations are performed using parallel computing and 16 cores were used for each run. 4.2 Irregularly shaped particles Four irregularly shaped particles were chosen to investigate in this study. The particlesโ€™ geometry and their morphological descriptors are taken from a set of Jordan sand collections (Zheng et al., 2019). The particles are chosen in a way that, based on the Zingg diagram (Figure 4.1) each one represents certain shapes i.e. disc, sphere, blade, and rod. The morphological descriptors, including elongation and flatness, are calculated according to Table 4.1. The real shape of the particles is shown in Figure 4.2. In addition, the morphological descriptor values of the four considered irregular particles are given in Table 4.2. 43 Figure 4.1: Zingg diagram. (a) Jordan No. 1, (b) Jordan No. 26, (c) Jordan No. 59, (d) Jordan No. 70 Table 4.1: 3D morphological descriptors (Zheng et al., 2021) Descriptor Symbol Definition Equivalent diameter ๐ท๐‘’๐‘ž Diameter of a sphere with grain volume ๐‘‰ Length ๐ฟ Breadth ๐ต Width ๐‘Š Largest Feret diameter Largest Feret diameter at planes normal to the 3D largest Feret dimeter Smallest Feret diameter Elongation ๐ธ๐‘™ ๐ตโ„๐ฟ Flatness ๐น๐‘™ ๐‘Š โ„๐ต Sphericity ๐‘† 2 โ„ ๐œ‹๐ท๐‘’๐‘ž ๐ด (๐ด is the grain surface area) 44 No. 1 No. 26 No. 59 No. 70 Figure 4.2: Shape of Jordan sands used in this study Table 4.2: Morphological descriptors of irregularly shaped particles Jordan Sand No. 1 26 59 70 Flatness 0.94 0.53 0.53 0.80 Elongation 0.71 0.98 0.53 0.47 Sphericity 0.96 0.92 0.76 0.80 A key effect of particles with irregular shapes is observed in their influence on drag force, where terminal velocity serves as the primary indicator of the drag force. The terminal velocity signifies the point at which the forces of buoyancy, gravity, and drag neutralize each other, resulting in the particle moving at a constant velocity thereafter. While the first two aforementioned forces are only related to the density, the third one is where shape plays its role. To illustrate this point, Figure 4.3 shows the terminal velocity of the four particles while all having the same volume and density as 2460 ๐‘˜๐‘”โ„๐‘š3. The figure points out that when the particle has a more spherical shape, it experiences less drag force, i.e. higher terminal velocity. However, the drag force is stronger for those particles with higher flatness, which eventually leads to less terminal velocity. Figure 4.4 demonstrates that because particles No. 59, and 70 have a more flat 45 shape when they are falling down, they deviate from the central axis of the domain, in addition to less sedimentation velocity compared to particles No. 1, and 26 which have a more round shape. Moreover, the figure displays a stronger wake region behind particles No. 1, and 26 as they are falling with faster velocity compared to the other two particles. Figure 4.3: Sedimentation velocity of four different Jordan sand particles 46 (a) (b) (c) (d) Figure 4.4: Sedimentation of four irregularly shaped particles at t=0.12 s. (a) particle No. 1, (b) particle No. 26, (c) particle No. 59, (d) particle No. 70. 4.3 Simulation of column collapse To demonstrate the application of the current coupling method in geotechnical engineering, a simulation of a column of irregular shape particles collapsing in an ambient fluid is presented. The dimension of the domain is given in Figure 4.5. No slip wall is applied to all domain boundaries. The modelling properties of the fluid and solid particles are given in Table 4.3. The simulations are run for four different particle shapes. The morphological descriptions of each of the four particles used in these simulations are given in Table 4.2 (Zheng et al., 2021). The lattice resolution and relaxation times are set to 15 and 0.508, respectively. To generate the column of 1600 particles (clumps), first, particles stacked at one side of the domain under gravity using only DEM (๐ฟ๐‘– region). Then by removing the wall that holds the particles to one side, the particles start to collapse to the other side of the domain. 47 Table 4.3: Characteristics of fluid and solid Particle Fluid Parameter Equivalent diameter of irregular shape particle Density, ๐œŒ๐‘ Youngโ€™s modulus, E Coefficient of restitution, ๐‘’ Coefficient of friction, ๐œ‡ Poissonโ€™s ratio Density, ๐œŒ๐‘“ Viscosity, ๐œ‡๐‘“ Value 1.1 mm 2468 kgโ„m3 109 Pa 0.65 0.4 0.24 1000 kgโ„m3 0.001 Pa. s Figure 4.5. Domain dimensions for immersed column collapse. All boundaries are no-slip walls. The simulation result for particle No. 1 is shown in Figure 4.6. First, the particles near the front top of the column started to slump down, and larger fluid velocities at t = 0.08 s were near the front face (Figure 4.6a). As time passed, some of the slumping particles reached the floor and created waves with high velocities near the toe (t=0.16 s). Some particles near the top front toppled over and induced relatively large velocities in their vicinity. When t = 0.39 s, the rear and middle portion of the column approximated their final slope grade, while the front toe had particles with 48 great momentum and kept progressing forward. Eventually at t = 0.52 s, the system reaches the final state that is the kinetic energy of the system is less than 10โˆ’10 kJ. (a) (b) (c) (d) Figure 4.6: Collapse of column of irregularly shaped particle No. 1 at different times a) t = 0.08 s, b) t = 0.16 s, c) t = 0.39 s, d) t = 0.52 s Particle No. 26 has the disc shape geometry according to the Zingg diagram. Unlike particle No. 1, at the early stage of collapse (Figure 4.7a), the disc shape particle starts to topple near the front top same as sliding from the very bottom. This early dynamic results in the toppling segment converging with the sliding particles at the lower region, inducing a pronounced and irregular vortex formation around the free surface area (Figure 4.7b). This uneven vortex then leads into a bump at t = 0.33 (Figure 4.7c). Eventually, at the final stage of collapse, some particles roll down from upward and make the final free surface profile (Figure 4.7d). This sequence illustrates the 49 complex interplay between particle geometry, motion dynamics, and fluid interaction, leading to distinctive morphological changes in the collapse structure. (a) (b) (c) (d) Figure 4.7: Collapse of column of irregularly shaped particle No. 26 at different times a) t = 0.08 s, b) t = 0.16 s, c) t = 0.33 s, d) t = 0.44 s Particle No. 59, is the blade-shaped particle. The first point observable about this case is that due to the special shape of the particle, the initial height of the column is higher than other cases (Figure 4.8a). As the simulation starts, the particles located at the front top topple, with none from the bottom of the column sliding forward (Figure 4.8b). Since, in this case, the particles are falling from a higher elevation, by the time the top particles arrive at the bottom, a heavy vortex is all over the free surface (Figure 4.8c). Afterward, the system's kinetic energy dissipates through interactions, and the collapse decelerates, culminating in a new equilibrium where the particles 50 settle into a densely packed arrangement, significantly altering the fluid's flow patterns around them. It is also worth noting that while particle template No. 59 has the least terminal velocity compared to the other three particles, according to Figure 4.3, due to its special shape, the particle tends to oscillate when it gets closer to its terminal velocity. Therefore, oscillation could be the reason this particle template induces more disturbance into the fluid field and takes a longer time to reach its final equilibrium state. (a) (b) (c) (d) Figure 4.8: Collapse of column of irregularly shaped particle No. 59 at different times a) t = 0.08 s, b) t = 0.16 s, c) t = 0.46 s, d) t = 0.62 s The last simulation of irregularly shaped particle is particle No. 70, which is the rod shape. Looking at the first stage of this collapse illustrates how fast this particle shape reacted to gravity force and deformed from its initial static equilibrium state (Figure 4.9a). Comparing Figure 4.9a, and Figure 4.9b shows how the particles at the top are slumping down and forward, causing 51 disturbance in the fluid velocity. Another point in this process is that this particle template reaches its final equilibrium state at a faster pace compared to the disc shape particle. The reason could be the way column collapses, induces the least amount of disturbance in the fluid field. Therefore, the whole process happens smoothly and in less than half a second. It is also worth noticing that according to Figure 4.9, particle No. 70 has one of the lowest terminal velocity magnitudes compared to other particle templates. Thus, it is reasonable to reach the equilibrium states within less time relative to other particle templates. (a) (b) (c) (d) Figure 4.9: Collapse of column of irregularly shaped particle No. 70 at different times a) t = 0.08 s, b) t = 0.16 s, c) t = 0.33 s, d) t = 0.44 s The last simulation belongs to the column collapse of the sphere particles. This simulation is provided mainly as a comparison between a regular sphere particle and other irregularly shaped particles. As shown in Figure 4.10, the sphere shape reaches its final state faster than any other 52 particle templates. The reason is mainly due to the round and symmetric shape of the particle. In other words, the round shape potentially reduces the interparticle friction, leading to particles moving more freely in the domain with less interaction from surroundings. This can also be observed in Figure 4.10c and Figure 4.10d, where those particles that are near the free surface slide down and make their way to the bottom of the fluid tank with no so much friction from the surrounding fluid. (a) (b) (c) (d) Figure 4.10: Collapse of column of spherical particle at different times a) t = 0.08 s, b) t = 0.16 s, c) t = 0.25 s, d) t = 0.33 s 53 Figure 4.11: Final deposition of column collapse of spherical particles. ๐ฟ๐‘“ = 3.45053 ๐‘๐‘š Figure 4.11 and Figure 4.12 show the final deposition state of the spherical and irregularly shaped particles. ๐ฟ๐‘“ is the final length of the granular column. By looking at the figures, it is apparent that all particles, regardless of their shape have eventually almost the same free-surface profile. However, in terms of ๐ฟ๐‘“ , particle No. 59 has the longest deposition length with ๐ฟ๐‘“ = 3.78061 ๐‘๐‘š. This is interesting in particular because the length of the deposition of particle No. 59 is even higher than that of a sphere. We believe that the main reason, as mentioned earlier, is that particle No. 59, due to its shape, builds a stack of higher height; therefore, those particles that are at the top have higher potential energy that later on converts into kinetic energy. In other words, the higher level of the column contributes to the higher kinetic energy of the whole system as well as reaching an equilibrium state at a slower rate. All in all, even though in all cases, the final deposition profile of particles is roughly the same, the amount of time for each column of particle as well as the final length of deposition are not the same among the studied case. This observation once again demonstrates how the particleโ€™s size is important and should be considered in simulating the geohazard processes in geotechnical engineering. 54 (a) (b) (c) (d) Figure 4.12: Final deposition of column collapse of a) irregularly shaped particle No. 1 with ๐ฟ๐‘“ = 3.24631 ๐‘๐‘š, b) irregularly shaped particle No. 26 with ๐ฟ๐‘“ = 3.45091 ๐‘๐‘š, c) irregularly shaped particle No. 59 with ๐ฟ๐‘“ = 3.78061 ๐‘๐‘š, d) irregularly shaped particle No. 70 with ๐ฟ๐‘“ = 3.31487 ๐‘๐‘š 55 Chapter 5: Conclusion 5.1 Contribution In this study, an LBM-DEM coupling framework was established to simulate the fluid-solid interaction considering irregular solid particle shapes. Irregular particles were represented as clumps using an open MATLAB code and were used in two open-source codes, LIGGGHTS and Palabos, for coupled LBM-DEM simulations. Three benchmarks, including the settling of a single clump, the settling of two clumps, and the settling of a disc-shaped clump, were successfully simulated, and the results are consistent with those in the literature. Notably, the simulated settling of two clumps was satisfactory in duplicating the interaction of two clumps in the fluid domain, including drafting, kissing, and tumbling. In addition, the simulation of a disc falling in the fluid domain can adequately produce stable falling and periodic oscillations that were consistent with the theoretical predictions. The number of sub-particles equal to 100 for representing an irregular particle as a clump in the established framework seems to be sufficient in capturing the grain morphology and their influence on the fluid-solid interaction. The established methodology was employed to simulate the collapse of a column of irregularly shaped particles with four different real shape particles. The simulation results can reasonably reproduce the slumping and trending of granular particles and the changes in flow velocity in the fluid domain. The simulation results were compared with the column collapse of spherical particles. It was found that spherical particles possessed larger momentum, travelled faster, and reached a longer run-out distance, which warrants the consideration of realistic particle shape in granular flow when interacting with fluids. To the best of the authorsโ€™ knowledge, this study represents the first time coupling LIGGGHTS and Palabos to simulate the interaction of the fluid with irregularly shaped solid particles. The established methodology is ready to be used for 56 many other geotechnical engineering problems, such as internal erosion in granular packs where particle shape has an important influence on pore structure and the related fine migration. 5.2 Limitations Like every study conducted in the science, this study is also associated with some limitations. While these limitations do not affect the integrity of the current results, addressing them would be a chance to develop the computation model for a better and more accurate simulation outcome in the future. The limitations are as follows: โ€ข In LBM method, this study adopted the single relaxation time BGK collision operator, which is known to have non-physical viscosity dependence on boundary location mapping. Therefore, two-relaxation time collision operator could enhance the accuracy of fluid flow simulation part especially if the model is used to simulate flow in porous media. โ€ข In Chapter 4, for the sake of reducing computational cost, the grid resolution used for simulating the case of immersed granular column collapse is N=15, while based on the grid study, N=20 is the resolution that provides the most accurate results. โ€ข Also in Chapter 4, and again for the sake of reducing the high computational cost, for the disc and sphere shape particles, 50 sub-particles used to reconstruct the geometry, while based on Figure 3.4, the 100 sub-particles is considered as the optimum number of sub-particles. 5.3 Future work Based on the results achieved in this research, the current code is capable of performing the following simulations: 57 โ€ข Simulating the drag force of real irregularly shaped particles at various Reynolds numbers and studying a correlation formula that could relate the drag force to geometric descriptors (such as elongation, flatness, sphericity) and predict the drag force of irregularly shaped particles. โ€ข Simulating the suffusion phenomenon using irregularly shaped particles as fixed, immovable coarse particles along with fine spherical particles that are able to move through the pores according to the applied pressure gradient. 58 References Aidun, C. K., & Clausen, J. R. (2010). Lattice-Boltzmann Method for Complex Flows. Annual Review of Fluid Mechanics, 42(1), 439-472. doi:10.1146/annurev-fluid-121108-145519 Almeida, E., Spogis, N., & Silva, M. (2016). Computational study of the pneumatic separation of sugarcane bagasse. Paper presented at the Proceedings of the 6th International Conference on Engineering for Waste and Biomass Valorisation, Albi, France. Angelidakis, V., Nadimi, S., Otsubo, M., & Utili, S. (2021). CLUMP: A Code Library to generate Universal Multi-sphere Particles. SoftwareX, 15, 100735. doi:https://doi.org/10.1016/j.softx.2021.100735 Bhatnagar, P. L., Gross, E. P., & Krook, M. (1954). A model for collision processes in gases. I. Small amplitude processes in charged and neutral one-component systems. Physical review, 94(3), 511. Chen, F., Drumm, E. C., & Guiochon, G. (2011). Coupled discrete element and finite volume solution of two classical soil mechanics problems. Computers and Geotechnics, 38(5), 638-647. doi:https://doi.org/10.1016/j.compgeo.2011.03.009 Chen, S., & Doolen, G. D. (1998). Lattice Boltzmann method for fluid flows. Annual Review of Fluid Mechanics, 30(1), 329-364. doi:10.1146/annurev.fluid.30.1.329 Cho, G.-C., Dodds, J., & Santamarina, J. C. (2007). Closure to โ€œParticle Shape Effects on Packing Density, Stiffness, and Strength: Natural and Crushed Sandsโ€ by Gye-Chun Cho, Jake Dodds, and J. Carlos Santamarina. Journal of Geotechnical and Geoenvironmental Engineering, 133(11), 1474-1474. doi:10.1061/(ASCE)1090-0241(2007)133:11(1474) Chu, K. W., Wang, B., Yu, A. B., & Vince, A. (2009). CFD-DEM modelling of multiphase flow in dense medium cyclones. Powder Technology, 193(3), 235-247. doi:https://doi.org/10.1016/j.powtec.2009.03.015 Cook, B. K., Noble, D. R., & Williams, J. R. (2004). A direct simulation method for particleโ€ fluid systems. Engineering Computations, 21(2/3/4), 151-168. doi:10.1108/02644400410519721 Cortรฉs, C., & Gil, A. (2007). Modeling the gas and particle flow inside cyclone separators. Progress in Energy and Combustion Science, 33(5), 409-452. doi:https://doi.org/10.1016/j.pecs.2007.02.001 Crowe, C. T., Schwarzkopf, J. D., Sommerfeld, M., & Tsuji, Y. (2011). Multiphase flows with droplets and particles: CRC press. Cundall, P. A., & Strack, O. D. L. (1979). A discrete numerical model for granular assemblies. Gรฉotechnique, 29(1), 47-65. doi:10.1680/geot.1979.29.1.47 59 Dash, S. M., & Lee, T. S. (2015). Two spheres sedimentation dynamics in a viscous liquid column. Computers & Fluids, 123, 218-234. doi:https://doi.org/10.1016/j.compfluid.2015.10.003 Di Renzo, A., & Di Maio, F. P. (2004). Comparison of contact-force models for the simulation of collisions in DEM-based granular flow codes. Chemical Engineering Science, 59(3), 525-541. doi:https://doi.org/10.1016/j.ces.2003.09.037 Ding, W.-T., & Xu, W.-J. (2018). Study on the multiphase fluid-solid interaction in granular materials based on an LBM-DEM coupled method. Powder Technology, 335, 301-314. doi:https://doi.org/10.1016/j.powtec.2018.05.006 El-Emam, M. A., Shi, W., & Zhou, L. (2019). CFD-DEM simulation and optimization of gascyclone performance with realistic macroscopic particulate matter. Advanced Powder Technology, 30(11), 2686-2702. doi:https://doi.org/10.1016/j.apt.2019.08.015 Favier, J. F., Abbaspourโ€Fard, M. H., Kremmer, M., & Raji, A. O. (1999). Shape representation of axi-symmetrical, non-spherical particles in discrete element simulation using multielement model particles. Engineering Computations, 16(4), 467-480. Feng, Z.-G., & Michaelides, E. E. (2004). The immersed boundary-lattice Boltzmann method for solving fluidโ€“particles interaction problems. Journal of Computational Physics, 195(2), 602-628. doi:https://doi.org/10.1016/j.jcp.2003.10.013 Ferellec, J.-F., & McDowell, G. R. (2010). A method to model realistic particle shape and inertia in DEM. Granular Matter, 12(5), 459-467. doi:10.1007/s10035-010-0205-8 Foster, M., Fell, R., & Spannagle, M. (2000). The statistics of embankment dam failures and accidents. Canadian Geotechnical Journal, 37(5), 1000-1024. doi:10.1139/t00-030 Galindo-Torres, S. A. (2013). A coupled Discrete Element Lattice Boltzmann Method for the simulation of fluidโ€“solid interaction with particles of general shapes. Computer Methods in Applied Mechanics and Engineering, 265, 107-119. doi:https://doi.org/10.1016/j.cma.2013.06.004 Gardner, M. (2023). Toward a Complete Kinematic Description of Hydraulic Plucking of Fractured Rock. Journal of Hydraulic Engineering, 149(7), 04023015. doi:doi:10.1061/JHEND8.HYENG-13193 Gardner, M., & Sitar, N. (2019). Coupled three-dimensional discrete element-lattice Boltzmann methods for fluid-solid interaction with polyhedral particles. International Journal for Numerical and Analytical Methods in Geomechanics, 43(14), 2270-2287. doi:https://doi.org/10.1002/nag.2972 Gingold, R. A., & Monaghan, J. J. (1977). Smoothed particle hydrodynamics: theory and application to non-spherical stars. Monthly Notices of the Royal Astronomical Society, 181(3), 375-389. doi:10.1093/mnras/181.3.375 60 Han, K., Feng, Y., & Owen, D. (2007). Numerical simulations of irregular particle transport in turbulent flows using coupled LBM-DEM. Computer Modeling in Engineering Sciences, 18(2), 87. Han, Y., & Cundall, P. A. (2013). LBMโ€“DEM modeling of fluidโ€“solid interaction in porous media. International Journal for Numerical and Analytical Methods in Geomechanics, 37(10), 1391-1407. doi:https://doi.org/10.1002/nag.2096 He, X., & Luo, L.-S. (1997). Lattice Boltzmann Model for the Incompressible Navierโ€“Stokes Equation. Journal of Statistical Physics, 88(3), 927-944. doi:10.1023/B:JOSS.0000015179.12689.e4 Ibrahim, A., & Meguid, M. A. (2020). Coupled Flow Modelling in Geotechnical and Ground Engineering: An Overview. International Journal of Geosynthetics and Ground Engineering, 6(3), 39. doi:10.1007/s40891-020-00223-0 Jansen, F., & Harting, J. (2011). From bijels to Pickering emulsions: A lattice Boltzmann study. Physical Review E, 83(4), 046707. doi:10.1103/PhysRevE.83.046707 Jiang, F., Liu, H., Chen, X., & Tsuji, T. (2022). A coupled LBM-DEM method for simulating the multiphase fluid-solid interaction problem. Journal of Computational Physics, 454, 110963. doi:https://doi.org/10.1016/j.jcp.2022.110963 Jiang, F., Matsumura, K., Ohgi, J., & Chen, X. (2021). A GPU-accelerated fluidโ€“structureinteraction solver developed by coupling finite element and lattice Boltzmann methods. Computer Physics Communications, 259, 107661. doi:https://doi.org/10.1016/j.cpc.2020.107661 Jing, L., Yang, G. C., Kwok, C. Y., & Sobral, Y. D. (2019). Flow regimes and dynamic similarity of immersed granular collapse: A CFD-DEM investigation. Powder Technology, 345, 532-543. doi:https://doi.org/10.1016/j.powtec.2019.01.029 Jones, B. D., & Williams, J. R. (2017). Fast computation of accurate sphere-cube intersection volume. Engineering Computations, 34(4), 1204-1216. doi:10.1108/EC-02-2016-0052 Kloss, C., Goniva, C., Hager, A., Amberger, S., & Pirker, S. (2012). Models, algorithms and validation for opensource DEM and CFDโ€“DEM. Progress in Computational Fluid Dynamics, an International Journal, 12(2-3), 140-152. doi:10.1504/PCFD.2012.047457 Kodicherla, S. P. K., Gong, G., Fan, L., Wilkinson, S., & Moy, C. K. S. (2020). Investigations of the effects of particle morphology on granular material behaviors using a multi-sphere approach. Journal of Rock Mechanics and Geotechnical Engineering, 12(6), 1301-1312. doi:https://doi.org/10.1016/j.jrmge.2020.04.005 Koshizuka, S., & Oka, Y. (1996). Moving-Particle Semi-Implicit Method for Fragmentation of Incompressible Fluid. Nuclear Science and Engineering, 123(3), 421-434. doi:10.13182/NSE96-A24205 61 Ladd, A. J. C. (1994). Numerical simulations of particulate suspensions via a discretized Boltzmann equation. Part 1. Theoretical foundation. Journal of Fluid Mechanics, 271, 285-309. doi:10.1017/S0022112094001771 Lai, Z., Zhao, J., Zhao, S., & Huang, L. (2023). Signed distance field enhanced fully resolved CFD-DEM for simulation of granular flows involving multiphase fluids and irregularly shaped particles. Computer Methods in Applied Mechanics and Engineering, 414, 116195. doi:https://doi.org/10.1016/j.cma.2023.116195 Latt, J., Malaspinas, O., Kontaxakis, D., Parmigiani, A., Lagrava, D., Brogi, F., Chopard, B. (2021). Palabos: Parallel Lattice Boltzmann Solver. Computers & Mathematics with Applications, 81, 334-350. doi:https://doi.org/10.1016/j.camwa.2020.03.022 Lominรฉ, F., Scholtรจs, L., Sibille, L., & Poullain, P. (2013). Modeling of fluidโ€“solid interaction in granular media with coupled lattice Boltzmann/discrete element methods: application to piping erosion. Numerical and Analytical Methods in Geomechanics, 37(6), 577-596. doi:https://doi.org/10.1002/nag.1109 Mao, W., Wang, Y., Yang, P., Huang, Y., & Zheng, H. (2023). Dynamics of granular debris flows against slit dams based on the CFDโ€“DEM method: effect of grain size distribution and ambient environments. Acta Geotechnica, 18(11), 5811-5838. doi:10.1007/s11440023-01944-y Nan, W., Ghadiri, M., & Wang, Y. (2017). Analysis of powder rheometry of FT4: Effect of particle shape. Chemical Engineering Science, 173, 374-383. doi:https://doi.org/10.1016/j.ces.2017.08.004 Nielsen, A. W., Probst, T., Petersen, T. U., & Sumer, B. M. (2015). Sinking of armour layer around a vertical cylinder exposed to waves and current. Coastal Engineering, 100, 5866. doi:https://doi.org/10.1016/j.coastaleng.2015.03.010 Noble, D. R., & Torczynski, J. R. (1998). A Lattice-Boltzmann Method for Partially Saturated Computational Cells. International Journal of Modern Physics C, 09(08), 1189-1201. doi:10.1142/S0129183198001084 Oschmann, T., Hold, J., & Kruggel-Emden, H. (2014). Numerical investigation of mixing and orientation of non-spherical particles in a model type fluidized bed. Powder Technology, 258, 304-323. doi:https://doi.org/10.1016/j.powtec.2014.03.046 Owen, D. R. J., Leonardi, C. R., & Feng, Y. T. (2011). An efficient framework for fluidโ€“ structure interaction using the lattice Boltzmann method and immersed moving boundaries. International Journal of Numerical Methods in Engineering(1-5), 66-95. doi:https://doi.org/10.1002/nme.2985 Pasha, M., Hare, C., Ghadiri, M., Gunadi, A., & Piccione, P. M. (2016). Effect of particle shape on flow in discrete element method simulation of a rotary batch seed coater. Powder Technology, 296, 29-36. doi:https://doi.org/10.1016/j.powtec.2015.10.055 62 Peng, Z., Doroodchi, E., Luo, C., & Moghtaderi, B. (2014). Influence of void fraction calculation on fidelity of CFD-DEM simulation of gas-solid bubbling fluidized beds. 60(6), 20002018. doi:https://doi.org/10.1002/aic.14421 Pozorski, J., & Olejnik, M. (2023). Smoothed particle hydrodynamics modelling of multiphase flows: an overview. Acta Mechanica. doi:10.1007/s00707-023-03763-4 Rettinger, C., & Rรผde, U. (2022). An efficient four-way coupled lattice Boltzmann โ€“ discrete element method for fully resolved simulations of particle-laden flows. Journal of Computational Physics, 453, 110942. doi:https://doi.org/10.1016/j.jcp.2022.110942 Sayyari, M. J., Ahmadian, M. H., & Kim, K. C. (2022). Three-dimensional condensation in a vertical channel filled with metal foam using a pseudo-potential lattice Boltzmann model. International Journal of Thermal Sciences, 172, 107352. doi:https://doi.org/10.1016/j.ijthermalsci.2021.107352 Seil, P., & Pirker, S. (2017). LBDEMcoupling: Open-Source Power for Fluid-Particle Systems. Paper presented at the Proceedings of the 7th International Conference on Discrete Element Methods, Singapore. Shen, Z., Wang, G., Huang, D., & Jin, F. (2022). A resolved CFD-DEM coupling model for modeling two-phase fluids interaction with irregularly shaped particles. Journal of Computational Physics, 448, 110695. doi:https://doi.org/10.1016/j.jcp.2021.110695 Soundararajan, K. K. (2015). Multi-scale multiphase modelling of granular flows. University of Cambridge Stratford, K., Adhikari, R., Pagonabarraga, I., & Desplat, J. C. (2005). Lattice Boltzmann for Binary Fluids with Suspended Colloids. Journal of Statistical Physics, 121(1), 163-178. doi:10.1007/s10955-005-8411-1 Suhr, B., & Six, K. (2020). Simple particle shapes for DEM simulations of railway ballast: influence of shape descriptors on packing behaviour. Granular Matter, 22, 1-17. Sun, X., Sakai, M., & Yamada, Y. (2013). Three-dimensional simulation of a solidโ€“liquid flow by the DEMโ€“SPH method. Journal of Computational Physics, 248, 147-176. doi:https://doi.org/10.1016/j.jcp.2013.04.019 Taghavi, R. (2011). Automatic clump generation based on mid-surface. Paper presented at the Proceedings, 2nd international FLAC/DEM symposium, Melbourne. Tavarez, F. A., & Plesha, M. E. (2007). Discrete element method for modelling solid and particulate materials. International Journal for Numerical Methods in Engineering, 70(4), 379-404. doi:https://doi.org/10.1002/nme.1881 ten Cate, A., Nieuwstad, C. H., Derksen, J. J., & Van den Akker, H. E. A. (2002). Particle imaging velocimetry experiments and lattice-Boltzmann simulations on a single sphere settling under gravity. Physics of Fluids, 14(11), 4012-4025. doi:10.1063/1.1512918 63 Teng, P., Johansson, F., & Hellstrรถm, J. G. I. (2023). Modelling erosion of a single rock block using a coupled CFD-DEM approach. Journal of Rock Mechanics and Geotechnical Engineering, 15(9), 2375-2387. doi:https://doi.org/10.1016/j.jrmge.2023.06.001 Tsuji, Y. (2007). Multi-scale modeling of dense phase gasโ€“particle flow. Chemical Engineering Science, 62(13), 3410-3418. doi:https://doi.org/10.1016/j.ces.2006.12.090 Verlet, L. (1967). Computer "Experiments" on Classical Fluids. I. Thermodynamical Properties of Lennard-Jones Molecules. Physical review, 159(1), 98-103. doi:10.1103/PhysRev.159.98 Wachs, A. (2009). A DEM-DLM/FD method for direct numerical simulation of particulate flows: Sedimentation of polygonal isometric particles in a Newtonian fluid with collisions. Computers & Fluids, 38(8), 1608-1628. doi:https://doi.org/10.1016/j.compfluid.2009.01.005 Wang, D., Leonardi, C. R., & Aminossadati, S. M. (2018). Improved coupling of time integration and hydrodynamic interaction in particle suspensions using the lattice Boltzmann and discrete element methods. Computers & Mathematics with Applications, 75(7), 2593-2606. doi:https://doi.org/10.1016/j.camwa.2018.01.002 Wang, L., Park, J.-Y., & Fu, Y. (2007). Representation of real particles for DEM simulation using X-ray tomography. Construction and Building Materials, 21(2), 338-346. doi:https://doi.org/10.1016/j.conbuildmat.2005.08.013 Wang, M., Feng, Y. T., Owen, D. R. J., & Qu, T. M. (2019). A novel algorithm of immersed moving boundary scheme for fluidโ€“particle interactions in DEMโ€“LBM. Computer Methods in Applied Mechanics and Engineering, 346, 109-125. doi:https://doi.org/10.1016/j.cma.2018.12.001 Wang, M., Feng, Y. T., Qu, T. M., Tao, S., & Zhao, T. T. (2020). Instability and treatments of the coupled discrete element and lattice Boltzmann method by the immersed moving boundary scheme. International Journal for Numerical Methods in Engineering, 121(21), 4901-4919. doi:https://doi.org/10.1002/nme.6499 Wang, M., Feng, Y. T., Qu, T. M., & Zhao, T. T. (2021). A coupled polygonal DEM-LBM technique based on an immersed boundary method and energy-conserving contact algorithm. Powder Technology, 381, 101-109. doi:https://doi.org/10.1016/j.powtec.2020.11.081 Wang, T., Zhang, F., & Zheng, W. (2023). Suffusion of Gap-Graded Soil with Realistically Shaped Coarse Grains: A DEMโ€“DFM Numerical Study. International Journal of Geomechanics, 23(1), 04022247. doi:10.1061/(ASCE)GM.1943-5622.0002616 Willmarth, W. W., Hawk, N. E., & Harvey, R. L. (1964). Steady and Unsteady Motions and Wakes of Freely Falling Disks. The Physics of Fluids, 7(2), 197-208. doi:10.1063/1.1711133 64 Yang, G. C., Jing, L., Kwok, C. Y., & Sobral, Y. D. (2019). A comprehensive parametric study of LBM-DEM for immersed granular flows. Computers and Geotechnics, 114, 103100. doi:https://doi.org/10.1016/j.compgeo.2019.103100 Yang, L., Padding, J. T., & Kuipers, J. A. M. (2016). Modification of kinetic theory of granular flow for frictional spheres, Part I: Two-fluid model derivation and numerical implementation. Chemical Engineering Science, 152, 767-782. doi:https://doi.org/10.1016/j.ces.2016.05.031 Zhang, H., Tan, Y., Shu, S., Niu, X., Trias, F. X., Yang, D., Sheng, Y. (2014). Numerical investigation on the role of discrete element method in combined LBMโ€“IBMโ€“DEM modeling. Computers & Fluids, 94, 37-48. doi:https://doi.org/10.1016/j.compfluid.2014.01.032 Zhang, R., Kim, H.-J., & Dinoy, P. R. (2021). Particle Flow Simulation Based on Hybrid IMBDEM-LBM Approach with New Solid Fraction Calculation Scheme. Applied Sciences, 11(8). doi:10.3390/app11083436 Zheng, W., Hu, X., Tannant, D. D., Zhang, K., & Xu, C. (2019). Characterization of two- and three-dimensional morphological properties of fragmented sand grains. Engineering Geology, 263, 105358. doi:https://doi.org/10.1016/j.enggeo.2019.105358 Zheng, W., Hu, X., Tannant, D. D., & Zhou, B. (2021). Quantifying the influence of grain morphology on sand hydraulic conductivity: A detailed pore-scale study. Computers and Geotechnics, 135, 104147. doi:https://doi.org/10.1016/j.compgeo.2021.104147 Zhou, W., Ma, Q., Ma, G., Cao, X., & Cheng, Y. (2020). Microscopic investigation of internal erosion in binary mixtures via the coupled LBM-DEM method. Powder Technology, 376, 31-41. doi:https://doi.org/10.1016/j.powtec.2020.07.099 Zhou, Y., Zhou, B., Li, J., & Wang, H. (2017). Study on the Multi-sphere Method Modeling the 3D Particle Morphology in DEM. Paper presented at the Proceedings of the 7th International Conference on Discrete Element Methods, Singapore. Zhou, Z. Y., Kuang, S. B., Chu, K. W., & Yu, A. B. (2010). Discrete particle simulation of particleโ€“fluid flow: model formulations and their applicability. Journal of Fluid Mechanics, 661, 482-510. doi:10.1017/S002211201000306X Zou, Q., & He, X. (1997). On pressure and velocity boundary conditions for the lattice Boltzmann BGK model. Physics of Fluids, 9(6), 1591-1598. doi:10.1063/1.869307 65