Open access peer-reviewed chapter

General Drag Correlations for Particle-Fluid System

Written By

Zheng Qi, Shibo Kuang, Liangwan Rong, Kejun Dong and Aibing Yu

Submitted: 13 June 2022 Reviewed: 08 July 2022 Published: 19 August 2022

DOI: 10.5772/intechopen.106427

From the Edited Volume

Boundary Layer Flows - Modelling, Computation, and Applications of Laminar, Turbulent Incompressible and Compressible Flows

Edited by Vallampati Ramachandra Prasad, Valter Silva and João Cardoso

Chapter metrics overview

123 Chapter Downloads

View Full Metrics

Abstract

Particle-fluid flows are commonly encountered in industrial applications. It is of great importance to understand the fundamentals governing the behavior of such a flow system for better process design, control, and optimization. Generally, the particle-fluid flow behavior is strongly influenced by the interaction forces between fluid and particles. Among the various kinds of particle-fluid interaction forces, the drag force is the most essential. This chapter reviews the modeling of drag force for particle-fluid systems: from single particle to multiple particles, monosize to multisize, spherical to nonspherical, and Newtonian fluid to non-Newtonian fluid. Typical drag correlations in the literature are compared and assessed in terms of physical meaning, consistency, and generality.

Keywords

  • drag force
  • particle-fluid flow
  • computational fluid dynamics
  • Lattice-Boltzmann method

1. Introduction

Particle-fluid flows are commonly encountered in industrial applications. It is of great importance to understand the fundamentals governing the behavior of such a flow system for better process design, control, and optimization. The flow behaviors of particles and fluid are strongly influenced by their interaction forces. Thus, it is critical to model the particle-fluid interaction forces accurately when simulating particle-fluid flows. The particle-fluid interaction forces include the pressure gradient force (or buoyancy force), drag force, virtual mass force, Basset force, and lift forces. The drag force, which is usually the dominant force in many particle-fluid flow systems, is undoubtedly the most critical and most studied.

Numerous efforts have been made to quantify the drag force using experimental or numerical methods in the past decades. Early studies in this area were mainly conducted by experiments [1, 2, 3, 4]. Due to the limitations of techniques, the experimental conditions are difficult to control. Therefore, the drag force models proposed in these studies somewhat lack generality and consistency. Nevertheless, these pioneer studies [1, 2, 3, 4] provide a solid foundation for subsequent research. With the rapid development of computational technology, various numerical methods have become attractive for studying the fluid-particle interaction on a subparticle scale, such as direct numerical simulation (DNS) and Lattice-Boltzmann (LB) model. With these numerical approaches, various fluid-particle systems can be studied under well-controlled conditions, considering more important and complicated factors that affect fluid-particle drag force. Nowadays, the studies on the fluid-particle interaction associated with the development of the drag force model have been extended extensively: from single particle to multiple particles, monosize to multisize, spherical particle to nonspherical particle system, and Newtonian fluid to non-Newtonian fluid. This chapter will review the modeling of particle-fluid drag force from these perspectives.

Advertisement

2. Mathematical model

The flow of a fluid is governed by the momentum and mass conservation equations when ignoring the compressible and viscous heat dissipation effect:

ρt+·ρu=0E1
ρut+·ρuu=p+·ρνuE2

where ρ, p, u, and νf are the fluid density, pressure, velocity, and fluid kinematic viscosity, respectively.

In a traditional computational fluid dynamics (CFD) method, Eqs. (1) and (2) with initial and boundary conditions are often solved by the finite volume method (FVM) or finite element method (FEM). However, due to the difficulty of grid generation and boundary treatment in the DNS, the FVM and FEM are not widely used in studying the interaction between fluid and particles on a subparticle scale. Instead of FVM and FEM, the LB model is more widely used to study the drag force in particle-fluid systems. Its detailed description can be found elsewhere [5]. For brevity, the following only briefly introduces the LB model used.

The equations of the LB model are solved in two steps. First is the collision step:

fi+xt=fixtM1Sf·mmeqxtiE3

which is followed by a propagation step:

fix+ciδtt+δt=fi+xtE4

where m=Mf and f is the column vector of fi is the velocity distribution function at the lattice node x and the time t with the discrete velocity vector ci in the ith direction. The number of directions of discrete velocity vectors ci depends on the velocity model used. For example, provided that the fluid particle can move in the 19 directions in a three-dimensional case, as shown in Figure 1, the velocity model is referred to as a D3Q19 scheme. For the D3Q19 scheme, ci is given as

Figure 1.

Schematic illustrations of (a) D3Q19 LB method, and (b) bounce-back boundary treatment.

ci=c011000011111111000000011001111000011110000011000011111111E5

where c=δxδt is the lattice speed, δx is the lattice length, and δt is the time step. Other velocity models are also available, such as D3Q15 and D3Q27, and their details can be found elsewhere [5, 6].

The transfer matrix M in Eq. (3) defines the transformation of the distribution function to the moment space, which can be chosen to be the same as that of D’Humieres et al. [6], as given in Eq. (6). Note that Sf in Eq. (7) is a diagonal matrix. Provided that all the elements of Sf equal the same value as τ, Eq. (3) can be reduced to the well-known LBGK scheme, fi+xt=fixtfixtfieqxtτ. D’Humieres et al. [6] gives an optimized value of Sf, where s0f=s3f=s5f=s7f=0, s1f=1.19, s2f=s10f=s12f=1.4, s16f=s17f=s18f=1.98, s9f=s11f=s13f=s14f=s15f=sν, s4f=s6f=s8f=sq, and sq=82sν8sν. sq are set to sq=82sν8sν to satisfy the nonslip boundary condition. It should be noted that sν has a relationship with the kinematic viscosity ν, as given by Eq. (8).

M=111111111111111111130111111111111888888888888124444441111111111110110000111111110000044000011111111000000011001111000011110004400111100001111000001100001111111100000440000111111110221111111111112222044222211111111222200011111111111100000002222111111110000000000011110000000000000000000000011110000000000011110000000000011111111000000000001111000011110000000000011111111E6
Sf=diags0fs1fs2fs3fs4fs5fs6fs7fs8fs9fs10fs11fs12fs13fs14fs15fs16fs17fs18fE7
ν=131sν0.5E8

meq in Eq. (3) are the velocity equilibrium moments defined by Eq. (9). The jx, jy, and jz are defined as jx=ρux, jy=ρuy, and jz=ρuz. The constant ρ0 is the mean fluid density of the system, which is set to unity for an incompressible fluid.

mfeq=ρ,11ρ+19jx2+jy2+jz2ρ0,3ρ11jx2+jy2+jz22ρ0,jx,2jx3,jy,2jy3,jz,2jz3,2jx2jy2jz2ρ0,jx2jy2jz22ρ0,jy2jz2ρ0,jy2jz22ρ0,jxjyρ0,jyjzρ0,xjzρ0,0,0,0TE9

When the effect of fluid rheology is considered, ν is dependent on the shear rate for a non-Newtonian fluid other than a constant for a Newtonian fluid. It is described here by the power-law model, which is suitable for a wide range of non-Newtonian fluids [7]:

ν=ν0ėn1=ν02eαβeαβn1/2E10

where ν0 is the flow consistency index, ė denotes the shear rate, and n is the power-law index. The fluid shows shear-thickening behavior when n > 1, shear-thinning behavior when n < 1, and Newtonian behavior when n = 1. In Eq. (10), Einstein’s summation convention is applied and eαβ represents an element of the tensor ė at the position αβ, calculated by [8]

eαβ=32ρτi=0fi1ccE11

where α and β denote spatial indices, c and c are the elements of the vector ci at positions α and β, and fi1=fifieq is the non-equilibrium part of the distribution function.

Through the Chapman-Enskog multiple-scale expansion, Eqs. (3) and (4) can be recovered to the Navier-Stokes equations in a low Mach number limit. The macroscopic variables such as density ρ, momentum density ρu, and stress σ,can be obtained through the moments of the velocity and temperature distributions:

ρ=ifi,ρu=ifici,σ=ificiciE12

The force exerted on a solid particle is calculated by the momentum exchange method proposed by Ladd [9], given by

fixbt+0.5=2fi+xftfi¯+xstciE13

where xf, xs, and xb denote the nodes in the fluid region and solid region and at the boundary for a stationary boundary on a uniform lattice, and the subscript i¯ indicates the opposite direction of i. The curved wall of spherical solid particles is treated by the halfway boundary scheme (Figure 1b), which sets the wall node at the mid-point between fluid and solid nodes. The selection of the halfway boundary treatment has various advantages. First, it is the simplest and most efficient boundary treatment. Secondly, compared with interpolation schemes, the halfway boundary scheme is very space-saving in modeling particles that are placed closely (e.g., in a packed bed) [10]. Thirdly, the high numerical stability of the halfway boundary scheme ensures simulation reliability under various complicated flow conditions. Lastly, the halfway boundary scheme is a second-order scheme; its accuracy in calculating forces can be as good as other interpolation schemes and better than immersed boundary schemes once a calibration step is applied [11].

Advertisement

3. Model validation

It is necessary to verify the validity of the mathematical model before its application for numerical experiments. The LB model has proven to be valid for studying the fluid-particle interactions under different conditions involving a single particle, two interactive particles, and packed beds of spherical and nonspherical particles over a wide range of conditions [12, 13, 14, 15]. For example, the LB prediction of drag coefficient CD0=8fdπρde2U02 for an isolated particle in the Newtonian fluid has been compared with the experimental measurements of Schlichting and Gersten [16], and with those calculated by three correlations given by Stokes [17], Dallavalle [4], and Clift et al. [18], respectively. Figure 2a shows good agreement between the simulations and experiments, including the three correlations. Specifically, in the intermediate flow region of Re = 100–1000, the drag coefficient of the sphere for the velocity inlet and outflow conditions under open BCs falls closer to the experimental data than the one under periodic boundary conditions. This may be due to the nonuniformity in the flow far away from the sphere for high values of Re, the state of which can be different from that of the experiments. Drag coefficient in the flow of power-law fluid has also been compared against the measurements by Chhabra [19] and Peden and Luo [20] over 5 ≤ Re ≤ 100 and 0.6 ≤ n ≤ 1, as shown in Figure 2b. In the comparison, the numerical data obtained by Dhole et al. [21] using the FVM are also considered. Figure 2 shows that both the LB model and FVM can well predict the drag coefficient on particles in the power-law fluid flows at different Re and n. The prediction errors are less than 4.2% for the present LB model with a calibration step and 7.6% for the FVM.

Figure 2.

Comparison between simulated and measured drag coefficients in (a) Newtonian, and (b) non-Newtonian fluids.

Advertisement

4. Drag force for single-particle system

The fluid flow past an isolated particle is one of the most basic flow phenomena. It can be seen as a special case of a particle-fluid system when the particle volume fraction is close to zero. Also, it is very natural to start from a single-particle system to investigate the drag force in a complicated particle-fluid system. In the past decades, extensive efforts have been made to study various physical characteristics (e.g., drag, wake, shear, heat transfer, and vortex shedding) of the flow past an isolated particle. Tiwari et al. [22, 23] have given a comprehensive review on this topic, and interested readers can refer to the review for details.

For an isolated particle in fluid flows, the particle-fluid drag force has been well established, which is expressed as [4]

fd=18CD0πρfdp2ufupufupE14

where fd is the drag force, CD0 is the drag coefficient for a single particle in fluid flows, which is determined by the Reynolds number Re=ρfdpufupμf, ρf is the fluid density, dp is the particle diameter, μf is the fluid viscosity, and uf and up are the fluid and particle velocities, respectively.

Figure 3 shows the relationship between CD0 and Re in different flow regimes established based on experimental and numerical data. It can be seen that the effect of Re on CD0 varies in different flow regimes. Generally, the flow regimes can be classified into nine groups: laminar regime or axisymmetric wake regime (regime I), planar symmetric wake regime (regime II), vortex shedding regime (regime III), separating vortex tubes regime (regime IV), subcritical regime (regime V), critical regime (regime VI), supercritical regime (regime VII), and transcritical regime (regime VIII). It should be noted that the regimes I–IV may also be called the transition regime. In the laminar regime, the drag coefficient CD0 decreases significantly as Re increases. In contrast, in the transition regime (regimes I–-IV), the drop of CD0 decreases as Re increases and CD0 approaches a constant in the subcritical regime. Then, CD0 experiences a rapid drop in the critical regime but increases again and reaches a near constant in the subsequent supercritical and transcritical regimes.

Figure 3.

Drag coefficient for the flows past a stationary sphere in different flow regimes based on experimental and numerical data [23].

Due to the complicated effect of Re on CD0, it is still challenging to formulate a drag coefficient correlation that can accurately describe the complicated variation of CD0 in different flow regimes. However, a general drag coefficient can be formulated to meet the needs of engineering applications if the following requirements are satisfied:

  1. CD0 decreases in the laminar region.

  2. The drop of CD0 decreases in the transition regime.

  3. CD0 approaches a positive constant value when Re is relatively large.

Table 1 summarizes some of the well-known drag coefficient correlations. Notably, all these drag coefficient correlations can meet the above requirements.

ReferencesDrag correlation CD0
Schiller and Nauman (1935) [24]CD0=24Re1+0.15Re0.687Re<10000.44Re>1000
Dallavalle (1948) [4]CD0=0.63+4.8Re2
Turton and Levenspiel (1986) [25]CD0=24Re1+0.173Re0.657+0.4131+16300Re1.09

Table 1.

Drag correlations for an isolated particle in fluid flows.

Advertisement

5. From single-particle system to multiparticle system

Extension from an isolated particle system to a multi-particle system is a big challenge due to the complexity of the particulate system. The presence of other particles reduces the space for fluid and generates a sharp fluid velocity gradient, yielding increased shear stress on particle surfaces. The drag force enhancement is closely associated with particle configuration, particle-fluid slip velocity, and particle and fluid properties.

Generally, the methods to formulate drag correlations for multi-particle systems can be grouped into Ergun and Wen-Yu types. The Ergun-type model, also called the capillary model, focuses on the flow resistance of the whole system, exemplified by the Ergun Eq. (1). This kind of model originates from the idea of treating the void space in a porous medium as a bunch of tortuous conduits, as illustrated in Figure 4. Using equivalent tubes instead of the void space of a packed bed, one can model the total pressure drop across the whole system by considering the contributions of viscous force and inertial force. Correspondingly, its friction factor f=dPUs2ε31ε falls into the form of f=ARe+B. The Ergun-type model is an easier way to formulate correlations by fitting the data in the creeping flow regime or the extremely high Reynolds flow regime. However, as pointed out previously [26], because of the original idea of treating a multiparticle system as tortuous conduits, the Ergun-type correlations are more suitable for densely packed beds but usually have a discontinuity at high porosity. For example, when particles are immersed in a highly dilute system, the drag force determined by the Ergun-type model is usually inconsistent with the value given by the drag correlation for a single particle, which is apparently against the real physical world. This defect may cause problems in some situations, such as CFD-DEM simulations, and limit the application range of the Ergun-type model. The Wen-Yu-type model, also called the submerge object model, focuses on the enhanced drag force on a particle due to neighboring particles, exemplified by the Wen-Yu Eq. (2). Such a model is more straightforward as it directly considers the effect of neighboring particles on the drag force, as illustrated in Figure 4. In the model, the enhancement of drag force on particles in a multi-particle system can be expressed as a ratio to the drag force on a single particle in an unhindered environment. Correspondingly, the drag force in the Wen-Yu type can be written in fd=18CDπρfdp2ufupufup, where CD=CD0εf2χ. Unlike Ergun-type models, Wen-Yu-type models naturally ensure the consistency of drag force between the predictions by the correlation for a multi-particle system under dilute conditions and the correlation for a single-particle system, leading to its popularity in CFD-DEM simulations.

Figure 4.

Schematic representation of idealization of Ergun-type model (a), and Wen-Yu-type model (b).

Irrespective of the different origins of Ergun-type and Wen-Yu-type models, the drag correlations proposed by either way can be written in the following equation:

fd=18CDπρfdp2ufupufupE15

where CD is a coefficient relying on Re=ρfdpεfufupμf and εf is the porosity. CD is difficult to derive theoretically for a multiparticle system. Formulating an empirical correlation of CD should meet the following requirements for physical consistency and generality:

  1. CD should be able to reduce to CD0 when εf is approaching 1;

  2. CD should be continuous over the whole application range.

  3. CD should always be larger than CD0 when εf is smaller than 1;

Note that the requirement (2-1) must be satisfied in a CFD-DEM simulation. Otherwise, inconsistency may occur in the dilute regime. The requirement (2-1) can be written in CD0=limε1CD, which should satisfy the requirements (1-1) to (1-3). This is important for evaluating different drag correlations when new drag correlations are considered for a single particle. The requirement (2-3) ensures that CD is physically meaningful because neighboring particles reduce the space for fluid and generate a higher particle-fluid velocity in the surrounding region under the same Re, increasing the drag force.

Table 2 summarizes the correlations of drag coefficient CD reported for multi-particle systems in recent years. For easy comparison, the value of CD when εf=1 is given to assess if limε1CD satisfies the requirements from (1-1) to (1-3). Table 2 also indicates whether the correlations meet the requirements from (2-1) to (2-3). Note that the requirement (2-3) is not easy to be assessed directly. The values of void function χ=2logεfCDCD0 (Figure 5) are calculated from different drag correlations under different conditions to solve this problem. The value of χ in Figure 5 must be positive to satisfy the requirement (2-3). Otherwise, CD would be smaller than CD0.

ReferencesDrag coefficient CDCD when εf1Req (2-1)Req (2-2)Req (2-3)
Ergun (1952) [1]CD=431501εfRe+1.75CD,εf1=2.33××
Wen and Yu (1966) [2]CD=CD0εf2χ, CD0=24Re1+0.15Re0.687Re<10000.44Re>1000, χ=3.65CD,εf1=24Re1+0.15Re0.687Re<10000.44Re>1000
Gibilaro et al. (1985) [27]CD=CD0εf2χ, CD0=23Re+0.448, χ=3.8CD,εf1=23Re+0.448
Gidaspow (1994) [28]CD=CD0εf2χεf0.8431501εfRe+1.75εf<0.8, CD0=24Re1+0.15Re0.687Re<10000.44Re>1000, χ=3.65CD,εf1=24Re1+0.15Re0.687Re<10000.44Re>1000×
Di Felice (1994) [29]CD=CD0εf2χ, CD0=0.63+4.8Re2, χ=3.70.65exp1.5logRe22CD,εf1=0.63+4.8Re2
Van der Hoef et al. (2005) [30]CD=431801εfRe+18εf41+1.51εfReCD,εf1=24Re××
Benyahia et al. (2006) [31]CD=12εf2FRe, where F=1+38Reεf0.99andReF213/8F3F0+F1Re2εf<0.99andReF3+F324F1F0F22F1F2+F3Reothers,
F0=1w1+31εf/2+135/641εfln1εf+17.41εf1+0.6811εf8.481εf2+8.161εf3+w101εfεf30.6<εf<0.99101εfεf3εf0.6,
F1=2/1εf/400.9εf<0.990.11+0.00051exp11.61εfεf<0.9,
F2=1w1+31εf/2+135/641εfln1εf+17.891εf1+0.6811εf11.031εf2+15.411εf3+w101εfεf3εf>0.6101εfεf3εf0.6,
F3=0.93511εf+0.03667εf>0.90470.0673+0.2121εf+0.0232εf5εf0.9047,
Re=ρfεfdpufup2μf,
w=e10εf0.6/1εf
CCD,εf1=12Re1+0.375ReReF213/8F312Re1+0.03667ReRe>F213/8F3××
Mazzei and Lettieri (2007) [32]CD=CD0εf2χ, CD0=0.63+4.8Re2, χ=124.8+0.42Re3/41+0.175Re3/4 and Re=ρfdpufupμfεfn1CD,εf1=0.63+4.8Re2
Beetstra et al. (2007) [26]CD=431801εfRe+18εf41+1.51εfRe+0.31εf1+31εfεf+8.4Re0.3431+1031εfRe0.521εfCCD,εf1=24Re+0.4131+8.4Re0.3431+Re0.5
Cello et al. (2010) [33]CD=24εf2ReK1+K2εf4+K31εf4, K0=1εf1+3εf, K1=1+128K0+715K02εf1+49.5K0, K2=1+0.13Re+0.000666Re21+0.0342Re+0.00000692Re21,
K3=2Re21+Re410εf+9.2×107K020Re+1900εf20.066Re6600εf+0.000429Re43000εf20.000131Re2+73800εf3
CD,εf1=24Re1+K2××
Tenneti et al. (2011) [34]CD=CD0+43104.581εfRe+8.641εf1/3Reεf+18εf31εf30.95+0.611εf3εf2,
CD0=24Re1+0.15Re0.687Re<10000.44Re>1000
CD,εf1=24Re1+0.15Re0.687Re<10000.44Re>1000×
Rong et al. (2013) [14]CD=CD0εf2χ, CD0=0.63+4.8Re2, χ=2.651+εf5.33.5εfεf2exp1.5logRe22CD,εf1=0.63+4.8Re2
Zaidi et al. (2014) [35]CD=431801εfReεf+18εf41+1.51εfRe+0.612εf0.7Re20043196.21εf0.4Reεf0.3+0.432εf0.86Re200CD,εf1=24Re+0.816Re2000.576Re200×××
Bogner et al. (2015) [36]CD=24Reεf2.7261.751+0.151Re0.6840.4451+Re1.041εf0.161+Re0.00031εfCD,εf1=24Re1.146+0.151Re0.684×
Tang et al. (2015) [37]CD=431801εfRe+18εf41+1.51εfRe+18εf20.111εf2εf0.00456εf4+0.169εf+0.0644εf4Re0.343CD,εf1=24Re10.0046Re+0.233Re0.657××
Zhou and Fan (2015) [38]CD=F024εf2Re+0.256+1.411εf5.611εf2+6.041εf3εf1+3εf1εf+8.4Re0.343εf1+1031εfRe2.5+2εf
where F0=9.91εfεf2+εf31+31εf0.6εf>0.455.87sinπ21εf0.6371.75εf20.363εf0.455.871εf0.637εf2εf<0.363
CD,εf1=24Re+0.2561+8.4Re0.3431+Re0.5×
Sheikh and Qiu (2018) [39]CD=24Reεf0.531.251εf3+Re0.2Re0.5εf10.053εf+0.073CD,εf1=24Re1+Re0.2Re0.5+0.02
Kravets et al. (2019) [40]CD=2401εfRe+24εf41+1.51εfRe+24εf20.16951εf2εf0.004321εf4+0.0719εf+0.02169εf4Re0.2017CD,εf1=24Re10.00432Re+0.0936Re0.7983××

Table 2.

Summary of the drag correlations for multiparticle systems.

Figure 5.

Exponent χ as a function of porosity εf and Reynolds number Re calculated from different drag correlations.

Table 2 shows that most of the drag correlations in Ergun-type cannot satisfy the requirement (2-1), thereby generating nonphysical results in a dilute regime. Combining different correlations was employed in the past to overcome this problem. For example, the Gidaspow correlation [28], adopted in many commercial CFD software such as ANSYS Fluent, combined the Ergun correlation under dense conditions and the Wen-Yu correlation under dilute conditions. A similar treatment is also taken by Benyahia et al. [31] and Zhou and Fan [38]. However, this treatment may cause apparent discontinuity at the switching point, as shown in Figure 5, which is against the requirement (2-2). Some other investigators attempted to formulate an Ergun-type correlation by adding additional items to guarantee continuity based on DNS or LB simulation data for randomly distributed particle systems generated with the Monte Carlo method. Typical work can be exemplified by Van der Hoef et al. [30], Beetstra et al. [26], Cello et al. [33], Tang et al. [37], and Kravets et al. [40]. Among all these Ergun-type correlations, only the correlations proposed by Beetstra et al. [26] and Sheikh and Qiu [39] can meet all the requirements. The correlations of Tenneti et al. [34] and Zhou and Fan [38] can fulfill the requirements (2-1) and (2-3). However, with limε1CD, the correlation of Zhou and Fan [38] cannot be reduced to a single particle. Specifically, the value of the limε1CD is much smaller than the experimental value at a high Re. The void function value χ predicted by the correlation of Tenneti et al. [34] has an opposite trend of other correlations. Also, a discontinuous point exists at a relatively high Re, as shown in Figure 5.

Compared with Ergun-type correlations, Wen-Yu-type drag correlations inherently have the advantage of satisfying both requirements (2-1) and (2-2) due to the introduction of the voidage function χ. In fact, all the Wen-Yu-type correlations in Table 2 can satisfy the requirements from (2-1) to (2-3). One of the most widely used Wen-Yu-type correlations is the one proposed by Di Felice [29], which has good performance in both dilute and dense regimes. However, its voidage function is too simple and ignores the effect of Re and εf. Later, Rong et al. [14] used the LB method to study the drag force in different packed beds with a wide range of porosity generated by the DEM simulations, and on this basis, modified the voidage function of Di Felice [29] to incorporate the effect of porosity. Compared with other formulations, for example, the correlation of Beetstra et al. [26], that of Rong et al. [14] has a simpler form and better performance in a broader range of applications. Note that the existing drag correlations were often established on experimental or numerical data obtained in low-to-intermediate flow regimes (Re < 1000). The same expression has been adopted to deal with the high Re regime in applications. Nevertheless, drag force correlations for mono-size particle systems have been firmly established with the extensive efforts from different investigators in the last decade.

Advertisement

6. From monosize to multisize particle systems

The correlations in Table 2 are all formulated for systems with monosize spheres. However, particle systems composed of multisize spheres are encountered in most applications. The drag correlation formulated for monosize spheres cannot be directly applied to multisize particle systems. Therefore, some investigators have studied the drag forces using LB or DNS simulations for different components in multisize particle systems generated with DEM simulations or Monte Carlo methods, formulating new correlations based on numerical data.

Generally, two main treatments are used to estimate the fluid-particle drag force in a mixture of particles. The first one is to estimate the drag force on each particle directly, calculate the mean drag forces on different components, and finally obtain the total fluid-particle drag force by summing the forces of all particles. Due to the lack of reliable drag correlations for mixtures, different investigators used this treatment in early studies, such as Feng and Yu [41] and Bokkers et al. [42]. It has proven to have a poor performance by Rong et al. [13]. Thus, investigators turned their sight to the second approach, which uses an opposite calculation path. It first estimates the total fluid-particle drag force and then distributes the total force among different components to obtain their mean drag forces according to a specific rule. In this approach, the total drag force is estimated by treating the particle mixture as a monosize particle system with a representative average diameter, that is, the Sauter mean diameter dp=ixi/di1, where xi and di are the volume fraction and diameter of component i. The distributing function can be formulated in different ways. However, it should meet some basic requirements for physical consistency and generality, which have been well-established by Rong et al. [13], given by:

  1. fd,ifd=1 when all di are equal, where fd,i=fd,i/3πμfdiεfufup is the drag force for component i normalized by the Stokes drag and fd is normalized drag force for the equivalent mono-size particle.

  2. fd=i=1Nxifd,iyi2, where yi=di/dp.

  3. fd,i>0.

Note that the requirement (3-1) is required for the drag force model to reproduce the drag law for mono-size particle systems. The requirement (3-2) ensures the sum of the drag forces of all components equals the total fluid-particle drag force. The requirement (3-3) ensures positive mean drag forces.

Table 3 summarizes the distribution functions for multisize particle systems proposed by different investigators. The predictions by different correlations for comparison are given in Figure 6. These correlations were extended from monosize particle systems through distributing functions. Therefore, the performance of the drag correlations of multisize particle systems needs to be evaluated based on the requirements for both monosize and multi-size particle systems. Only the correlations of Rong et al. [13] and Mehrabadi et al. [45] can satisfy all the requirements. Others suffer various discrepancies to different extents. For example, the correlation of Sarkar et al. [43] fails to meet the criteria (3-1) and (3-2), whereas the correlations by Yin and Sundaresan [44] and Cello et al. [33] may generate negative drag forces, which is obviously against the requirement (3-3). It should also be noted that Rong et al. [13] and Mehrabadi et al. [45] used the correlations proposed in their previous studies to estimate the total fluid-particle drag force. The discussion about these two correlations for monosize systems can be found in Section 5.

ReferencesTreatmentsDrag correlationReq. (3-1)Req. (3-2)Req. (3-3)
Sarkar et al. (2009) [43]fd,ifd=yi+0.064yi3Van der Hoef et al. (2005) [30]××
Yin and Sundaresan (2009) [44]fd,i=1εf+fd1εfayi+1ayi2
a=12.661εf+9.0961εf211.3381εf3
Van der Hoef et al. (2005) [30]××
Cello et al. (2010) [33]fd,ifd=yi+1εfεf1εf0.2710.27yi2yij=1NxjyjCello et al. (2010) [33]××
Rong et al. (2014) [13]fd,ifd=0.5εfi=1Nxi/yi2+0.51εfyi2+0.5yiRong et al. (2013) [14]
Mehrabadi et al. (2016) [45]fd,ifd=εfyi+1εfyi2Tenneti et al. (2011) [34]
Duan et al. (2020) [46]fd,i=1εf+fd1εfi=1Nxiyi20.37yi+0.63yi2η
η=2.169ln0.466+1εfεij,max0.2025
εij,max=0.3131χijxi+0.64,xi<0.4140.8701χijxj+0.64,xi>0.414
χij=dj/didj>di
Zhou and Fan (2015) [38]××

Table 3.

Summary of the drag coefficient for multisize particle system.

Figure 6.

fd,i/fd as a function of porosity εf and volume fraction of small particle xs calculated from different drag correlations.

Advertisement

7. From spherical to nonspherical particle system

Particle shape could strongly affect drag forces. Rather than perfectly spherical particles, real particles often show diverse morphology. For example, they can be cubes, cylinders, and ellipsoids, or more generally, particles of irregular shapes. Such morphological diversity adds further complexity to the modeling of particle-fluid interaction forces. In earlier years, nonspherical particles are usually treated as volume-equivalent spheres. However, this simple treatment gives inaccurate results, even for an isolated particle. Several studies have been conducted to overcome this problem, particularly based on ellipsoidal particles. Such efforts are discussed in this section.

First, the definition of the geometry of an ellipsoidal particle is given to understand the factors that should be considered in drag correlations. For a standard axis-aligned ellipsoid particle, the Cartesian coordinates are given by x2a2+y2b2+z2c2=1, where a, b, and c are the principal semi-axes. In most works for ellipsoidal particles, the degenerate cases are considered, where the ellipsoid particles have two equal axes, say a=b, generally referred to as spheroid or ellipsoid of revolution. Variation of a results in different shapes of particles, which can be represented by aspect ratio, defined as Ar=ca=cb. Obviously, for an oblate spheroid, Ar<1; for a sphere, Ar=1; and for a prolate spheroid, Ar>1. The shapes of oblate and prolate spheroids are schematically shown in Figure 7. Note that the angle between the direction of the main flow and the symmetric axis of the ellipsoid body is defined as incident angle θ.

Figure 7.

Characteristics of spheroidal particles prolate spheroid (left) and oblate spheroid (right).

Based on a large number of experimental data, Holzer and Sommerfeld [47] established a drag correlation for an isolated nonspherical particle, given as:

CD=8Re1ψ+16Re1ψ+3Re1ψ3/4+0.42100.4logψ0.21ψE16

where ψ is the sphericity, which is defined as the ratio between the surface area of the equivalent-volume sphere and that of the considered particle. ψ is the mean crosswise sphericity, ψ=ψ,i, when applying to a multi-particle system. Here, the individual crosswise sphericity ψ,i represents the ratio between the cross-sectional area of the equivalent-volume sphere and the projected cross-sectional area of the considered ith particle. For spheroidal particles, these two parameters can be calculated by ψ=Ar2/91+2Ar1.6131/1.61 and ψ,i=Ar2/9Ar2sin2θi+Ar2cos2θi1/2. Later, Zastawny et al. [48] and Ouchene et al. [49] established correlations based on LB and DNS simulation data.

The procedure of extension from single sphere to multi-spheres particle systems is also used in nonspherical particle systems. In this direction, various drag correlations have been established in recent years based on numerical data. Similar to the requirements as discussed for multiparticle systems, the drag correlation CDReεfψ for nonspherical particles should also meet the following requirements for physical consistency and generality:

  1. CDReεfψ should be reduced to the drag correlation for spherical particles when ψ=1.

  2. CDReεfψ should be reduced to the drag correlation CD0Reψ for single nonspherical particle when porosity εf approaches unity.

  3. CDReεfψ should be continuous over the entire application range of Re, εf, ψ.

  4. CDReεfψ should always be larger than CD0Reψ when εf is smaller than 1;

Table 4 summarizes the drag correlations proposed by different investigators. Figure 8 shows the values of χ calculated from different drag correlations. All the correlations can meet the requirements (4-1) & (4-2) and can be reduced to the drag correlation for a single spherical particle. These correlations are established based on those for a nonspherical particle and are extensions of earlier studies from different investigators. Specifically, Rong et al. [12] used the drag correlation of Holzer and Sommerfeld [47], and their correlation is the extension of their work for monosize and multisize particle systems [13, 14]. Li et al. [50, 51] extended the work of Zhou and Fan [38]. Cao et al. [52] extended the work of Tenneti et al. [34] and adopted the drag correlation for ellipsoids proposed by Ouchene et al. [49]. Because Li et al. [50, 51] conducted their works for oblate and prolate particles separately, the drag correlation discontinues at the switching point as Ar approaches 1. Cao et al. [52] only examined the drag correlation for prolate particles and their drag correlation may generate the negative value of the void function at relatively high Ar or Re. So far, only the drag correlation proposed by Rong et al. [12] can satisfy all the requirements (4-1) to (4-4). However, it should be noted that all these studies only focused on particular particle shapes and a unified drag correlation with better generality for particles of irregular shapes is still lacking.

ReferencesDrag coefficient CDReq (4-1)Req (4-2)Req (4-3)Req (4-4)
Rong et al. (2015) [13]CD=CD0εf2χ, χ=β+λ, β=2.651+εf5.33.5εfεf2exp1.5logRe22
λ=1ψ101.8ψ0.812+2.439ψ20.6exp3.5logRe22, ψ=Ar2/91+2Ar1.6131/1.61
Li et al. (2019, 2020) [50, 51]CD=CD,0°cos2θ+CD,90°sin2θ
CD,0°=CD,sphereFD0,0°ψ0°,
FD0,0°=832ArAr21+2Ar21Ar213/2lnAr+Ar21ArAr211,Ar>11,Ar=183Ar1/32Ar1Ar2+212Ar21Ar23/2tan11Ar2Ar1,Ar<1
ψ0°=1εf0.103S+0.167Ar10.3260.374Ar0.34S+0.51.05+1,Ar>11,Ar=11εf0.2S+0.498Ar1.651Ar10.7060.550+0.302Ar2.3S+0.51.08+1,Ar<1
CD,90°=CD,sphereFD0,90°ψ90°
FD0,90°=83ArAr21+2Ar23Ar213/2lnAr+Ar211,Ar>11,Ar=183Ar1/3Ar1Ar22Ar231Ar23/2sin11Ar21,Ar<1
ψ90°=1εf0.344S+0.716Ar10.2810.232S+Ar0.8510.401+1,Ar>11,Ar=11εf0.160S+0.293Ar2.181Ar12.651.541.30Ar0.266S+0.51.13+1,Ar<1
S=123cos2ϕ¯1
×
Cao et al. (2020) [52]CD=CD0εf2χ, χ=β+λ, β=4.9880.5139+εf3.1751.493εfεf2exp1.5logRe220.5884logRe
λ=0.5201Ar2+0.6094Ar11.74εf3+1.713Ar23.467Ar+12.65εf+1.056Ar2+0.4316Ar4.168,
×

Table 4.

Summary of the drag correlations for nonspherical particle system.

Figure 8.

Exponent χ as a function of porosity εf, Reynolds numbers Re and Ar calculated from different drag correlations.

Advertisement

8. From Newtonian fluid to non-Newtonian fluid system

Another important factor affecting the fluid-particle interaction force is fluid rheology. In some applications, the fluid does not necessarily follow Newton’s law of viscosity. Fluid may present shear-thinning/shear-thickening behaviors. That is, the viscosity of fluid μ increases/decreases as the shear stress τ or shear rate γ̇ increases. Table 5 lists some widely used viscosity models for non-Newtonian fluid. Among these models, the power-law model is the most popular one in engineering applications because it provides a unified and simple way of describing the rheological characteristics of shear-thinning fluid (n < 1), Newtonian fluid (n = 1), and shear-thickening fluid (n > 1). However, even with the simplest power-law fluid model, predicting the behaviors of particles and fluid or quantifying the fluid-particle interaction force in non-Newtonian fluid is extremely challenging due to the variable viscosity.

Viscosity modelCorrelation
Power-law fluid modelμ=Kγ̇n1, where K is the flow consistency index and n is the power index.
Spriggs fluid modelμ=μ0,γ̇γ0̇μ0γ̇γ0̇n1,γ̇>γ0̇, where γ0̇ is the reference shear rate.
Carreau fluid modelμ=μ+μ0μ1+λγ̇2n12
Bingham fluid modelμ=μ+τγ̇,τmaxτ,τmax<τ, where τmax is the maximum shear stress.
Casson fluid modelμ=μ+τγ̇2,τmaxτ,τmax<τ

Table 5.

Summary of the viscosity models for non-Newtonian fluids.

To date, attention paid to the effect of fluid rheology on drag force is much less than other factors. Table 6 summarizes the drag correlations proposed for a non-Newtonian fluid-particle system. Note that the rheology of all the considered non-Newtonian fluids obeys the power law. As Newtonian fluid is a special case when n=1 for the power-law viscosity model, the drag correlation for a non-Newtonian fluid-particle system should not only satisfy requirements (2-1) to (2-3) but also the following requirements:

ReferencesDrag coefficient CDReq (2-1)Req (2-2)Req (2-3)Req (5-1)
Srinivas and Chhabra (1995) [53]CD=43150Rep+1.75, Rep=1521nρfεfnufup2ndpnK1εfn4n3n+1n××
Sabiri and Comiti (2000) [54]CD=4τ316Rep+0.194, Rep=ρfεfnufup2ndpnτ2n22n3K33n+1/4nn1εfn, τ=10.41lnεf S×××
Dhole et al. (2004) [55]CD=43230.4Rek+0.33, Rek=ρfεf2nufup2ndpφ1εf,
φ=K129+3nn150Sεf1n/2, S=dp2εf31501εf2
××
Qi et al. [57]CD=CD0εf2χ, CD0=2.654n1.9+3.213+4.8n0.08Re2, Re=ρfεf2nufup2ndpnK
χ=2.65nε+1n5.33.5εnε2explogRe1.522+2.65ε+12n2.65nε+1n1+exp2logRe1.5

Table 6.

Summary of the drag correlations for non-Newtonian fluid-particle system.

  1. CDReεfn should be reduced to the drag correlation CD for a Newtonian fluid-particle system in Table 2 when n=1.

The correlation proposed by Srinivas and Chhabra [53] is established based on the Ergun-drag correlation with a different definition of Reynolds number for non-Newtonian fluids and can satisfy requirements (2-2) and (5-2). However, it also has the same problem faced in the Ergun correlation. Sabiri and Comiti [54] and Dhole et al. [55] further revised the definition of the Reynolds number and introduced the tortuosity of packed beds to consider the effect of fluid rheology. However, their correlations cannot overcome the discrepancies of the Ergun drag correlation. Besides, the correlations of Srinivas and Chhabra [53] and Sabiri and Comiti [54] may generate a negative value of void function χ, as shown in Figure 9, which is against the requirement (2-3). To overcome this problem, Qi et al. [15, 56, 57] conducted a series of studies by extending the studies of Rong et al. [13, 14] from Newtonian fluid to non-Newtonian fluid. Qi et al. [56, 57] also considered both monosize and multisize particle systems. Overall, the resulting drag correlation is consistent with the drag correlation proposed for Newtonian fluid-particle systems by Rong et al. [13, 14] and satisfies all the requirements (Table 6).

Figure 9.

Exponent χ as a function of porosity εf, Reynolds number Re and power-law index n calculated from different drag correlations.

Advertisement

9. Conclusions

Drag force correlations for particle-fluid systems are reviewed, covering from simple to complicated systems, including from single particle to multiple particles, monosize to multisize, spherical to nonspherical, and Newtonian fluid to non-Newtonian fluid. The drag correlations for mono-size and multi-size spherical particle systems are more mature. However, the physical consistency and generality of different drag correlations could be more carefully considered, which are discussed in this review. Several drag correlations show superiority in generality. The understanding of fluid-particle interactions in complicated systems involving factors such as particle shape or fluid rheology is still lacking. For example, the studies on the effect of fluid rheology are largely limited to non-Newtonian fluids obeying the power-law fluid model. Similarly, the studies on nonspherical particles are mostly limited to ellipsoids. Further studies should be conducted to generally consider these important factors to meet various engineering needs.

References

  1. 1. Ergun S. Fluid flow through packed columns. Chemical Engineering Progress. 1952;48:89-94
  2. 2. Wen CY, Yu YH. Mechanics of fluidisation. Chemical Engineering Progress Symposium Series. 1966;62:100-111
  3. 3. Richardson JF, Jerónimo S. Velocity-voidage relations for sedimentation and fluidisation. Chemical Engineering Science. 1979;34:1419-1422
  4. 4. Dallavalle JM. Micromeritics: The Technology of Fine Particles. New York: Pitman Publishing corporation; 1948
  5. 5. Chen S, Doolen GD. Lattice Boltzmann method for fluid flows. Annual Review of Fluid Mechanics. 1998;30:329-364
  6. 6. Dhumieres D, Ginzburg I, Krafczyk M, Lallemand P, Luo LS. Multiple-relaxation-time lattice Boltzmann models in three dimensions. Philosophical Transaction A Mathematical Physics, Engineering and Science. 2002;360:437-451
  7. 7. Denier JP, Dabrowski PP. On the boundary-layer equations for power-law fluids. Proceedings of the Royal Society Mathematical Physical and Engineering Sciences. 2004;460:3143-3158
  8. 8. Artoli AM, Hoekstra AG, Sloot PMA. Optimizing lattice Boltzmann simulations for unsteady flows. Computers & Fluids. 2006;35:227-240
  9. 9. Ladd AJC. Numerical simulations of particulate suspensions via a discretized Boltzmann equation. Part 1. Theoretical foundation. Journal of Fluid Mechanics. 1994;271:285-309
  10. 10. Stobiac V, Tanguy PA, Bertrand F. Boundary conditions for the lattice Boltzmann method in the case of viscous mixing flows. Computers & Fluids. 2013;73:145-161
  11. 11. Chen L, Yu Y, Lu JH, Hou GX. A comparative study of lattice Boltzmann methods using bounce-back schemes and immersed boundary ones for flow acoustic problems. International Journal for Numerical Methods in Fluids. 2014;74:439-467
  12. 12. Rong LW, Zhou ZY, Yu AB. Lattice-Boltzmann simulation of fluid flow through packed beds of uniform ellipsoids. Powder Technology. 2015;285:146-156
  13. 13. Rong LW, Dong KJ, Yu AB. Lattice-Boltzmann simulation of fluid flow through packed beds of spheres: Effect of particle size distribution. Chemical Engineering Science. 2014;116:508-523
  14. 14. Rong LW, Dong KJ, Yu AB. Lattice-Boltzmann simulation of fluid flow through packed beds of uniform spheres: Effect of porosity. Chemical Engineering Science. 2013;99:44-58
  15. 15. Qi Z, Kuang S, Rong L, Yu A. Lattice Boltzmann investigation of the wake effect on the interaction between particle and power-law fluid flow. Powder Technology. 2018;326:208-221
  16. 16. Schlichting H, Gersten K, Gersten K. Boundary-Layer Theory. Berlin: Spring; 2000
  17. 17. Stokes GG. On the Effect of the Internal Friction of Fluids on the Motion of Pendulums. Cambridge: Pitt Press; 1851
  18. 18. Clift R, Grace JR, Weber ME. Bubbles, Drops, and Particles. New York: Dover Publications; 2005
  19. 19. Chhabra RP. Non-Newtonian Fluid Particle Systems: Sphere Drag. Monash University; 1980
  20. 20. Peden JM, Luo Y. Settling velocity of variously shaped particles in drilling and fracturing fluids. SPE Drilling Engineering. 1987;2:337-343
  21. 21. Dhole SD, Chhabra RP, Eswaran V. Flow of power-law fluids past a sphere at intermediate Reynolds numbers. Industrial & Engineering Chemistry Research. 2006;45:4773-4781
  22. 22. Tiwari SS, Pal E, Bale S, Minocha N, Patwardhan AW, Nandakumar K, et al. Flow past a single stationary sphere, 1. Experimental and numerical techniques, Powder Technology. 2020;365:115-148
  23. 23. Tiwari SS, Pal E, Bale S, Minocha N, Patwardhan AW, Nandakumar K, et al. Flow past a single stationary sphere, 2. Regime mapping and effect of external disturbances, Powder Technology. 2020;365:215-243
  24. 24. Schiller L, Naumann Z. A drag coefficienty correlation. Z. Ver. Deutsch. Ing. 1933;77:318-320
  25. 25. Turton R, Levenspiel O. A short note on the drag correlation for spheres. Powder Technology. 1986;47:83-86
  26. 26. Beetstra R, Van Der Hoef MA, Kuipers JM. Drag force of intermediate Reynolds number flow past mono- and bidisperse arrays of spheres. AICHE Journal. 2007;53:489-501
  27. 27. Gibilaro LG, Di Felice R, Waldram SP, Foscolo PU. Generalized friction factor and drag coefficient correlations for fluid-particle interactions. Chemical Engineering Science. 1985;40:1817-1823
  28. 28. Gidaspow D. Multiphase Flow and Fluidization: Continuum and Kinetic Theory Descriptions. New York: Academic Press; 1994
  29. 29. Di Felice R. The voidage function for fluid-particle interaction systems. International Journal of Multiphase Flow. 1994;20:153-159
  30. 30. Van Der Hoef MA, Beetstra R, Kuipers JM. Lattice-Boltzmann simulations of low-Reynolds-number flow past mono- and bidisperse arrays of spheres: Results for the permeability and drag force. Journal of Fluid Mechanics. 2005;528:233-254
  31. 31. Benyahia S, Syamlal M, O’brien TJ. Extension of Hill–Koch–Ladd drag correlation over all ranges of Reynolds number and solids volume fraction. Powder Technology. 2006;162:166-174
  32. 32. Mazzei L, Lettieri P. A drag force closure for uniformly dispersed fluidized suspensions. Chemical Engineering Science. 2007;62:6129-6142
  33. 33. Cello F, Di Renzo A, Di Maio FP. A semi-empirical model for the drag force and fluid–particle interaction in polydisperse suspensions. Chemical Engineering Science. 2010;65:3128-3139
  34. 34. Tenneti S, Garg R, Subramaniam S. Drag law for monodisperse gas–solid systems using particle-resolved direct numerical simulation of flow past fixed assemblies of spheres. International Journal of Multiphase Flow. 2011;37:1072-1092
  35. 35. Zaidi AA, Tsuji T, Tanaka T. A new relation of drag force for high Stokes number monodisperse spheres by direct numerical simulation. Advanced Powder Technology. 2014;25:1860-1871
  36. 36. Bogner S, Mohanty S, Rüde U. Drag correlation for dilute and moderately dense fluid-particle systems using the lattice Boltzmann method. International Journal of Multiphase Flow. 2015;68:71-79
  37. 37. Tang Y, Peters EJF, Kuipers JAM, Kriebitzsch SHL, Van Der Hoef MA. A new drag correlation from fully resolved simulations of flow past monodisperse static arrays of spheres. AICHE Journal. 2015;61:688-698
  38. 38. Zhou Q, Fan L-S. Direct numerical simulation of low-Reynolds-number flow past arrays of rotating spheres. Journal of Fluid Mechanics. 2015;765:396-423
  39. 39. Sheikh B, Qiu T. Pore-scale simulation and statistical investigation of velocity and drag force distribution of flow through randomly-packed porous media under low and intermediate Reynolds numbers. Computers & Fluids. 2018;171:15-28
  40. 40. Kravets B, Rosemann T, Reinecke SR, Kruggel-Emden H. A new drag force and heat transfer correlation derived from direct numerical LBM-simulations of flown through particle packings. Powder Technology. 2019;345:438-456
  41. 41. Feng YQ, Yu AB. Assessment of model formulations in the discrete particle simulation of gas−solid flow. Industrial and Engineering Chemistry Research. 2004;43:8378-8390
  42. 42. Bokkers GA, Annaland MVS, Kuipers JM. Mixing and segregation in a bidisperse gas-solid fluidised bed: A numerical and experimental study. Powder Technology. 2004;140:176-186
  43. 43. Sarkar S, Van Der Hoef MA, Kuipers JAM. Fluid–particle interaction from lattice Boltzmann simulations for flow through polydisperse random arrays of spheres. Chemical Engineering Science. 2009;64:2683-2691
  44. 44. Yin X, Sundaresan S. Fluid-particle drag in low-Reynolds-number polydisperse gas-solid suspensions. AICHE Journal. 2009;55:1352-1368
  45. 45. Mehrabadi M, Tenneti S, Subramaniam S. Importance of the fluid-particle drag model in predicting segregation in bidisperse gas-solid flow. International Journal of Multiphase Flow. 2016;86:99-114
  46. 46. Duan F, Zhao L, Chen X, Zhou Q. Fluid–particle drag and particle–particle drag in low-Reynolds-number bidisperse gas–solid suspensions. Physics of Fluids. 2020;32:113311
  47. 47. Hölzer A, Sommerfeld M. New simple correlation formula for the drag coefficient of non-spherical particles. Powder Technology. 2008;184:361-365
  48. 48. Zastawny M, Mallouppas G, Zhao F, Van Wachem B. Derivation of drag and lift force and torque coefficients for non-spherical particles in flows. International Journal of Multiphase Flow. 2012;39:227-239
  49. 49. Ouchene R, Khalij M, Arcen B, Tanière A. A new set of correlations of drag, lift and torque coefficients for non-spherical particles and large Reynolds numbers. Powder Technology. 2016;303:33-43
  50. 50. Li X, Jiang M, Huang Z, Zhou Q. Effect of particle orientation on the drag force in random arrays of prolate ellipsoids in low-Reynolds-number flows. AICHE Journal. 2019;65:e16621
  51. 51. Li X, Jiang M, Huang Z, Zhou Q. Effect of particle orientation on the drag force in random arrays of oblate ellipsoids in low-Reynolds-number flows. AICHE Journal. 2020;67:e17040
  52. 52. Cao Z, Tafti DK, Shahnam M. Development of drag correlation for suspensions of ellipsoidal particles. Powder Technology. 2020;369:298-310
  53. 53. Srinivas BK, Chhabra RP. An experimental-study of non-Newtonian fluid-flow in fluidized-beds - minimum fluidization velocity and bed expansion characteristics. Chemical Engineering and Processing. 1991;29:121-131
  54. 54. Sabiri NE, Comiti J. Experimental validation of a model allowing pressure gradient determination for non-Newtonian purely viscous fluid-flow through packed beds. Chemical Engineering Science. 1997;52:3589-3592
  55. 55. Dhole SD, Chhabra RP, Eswaran V. Power law fluid flow through beds of spheres at intermediate Reynolds numbers. Chemical Engineering Research and Design. 2004;82:642-652
  56. 56. Qi Z, Kuang S, Yu A. Lattice Boltzmann investigation of non-Newtonian fluid flow through a packed bed of uniform spheres. Powder Technology. 2019;343:225-236
  57. 57. Qi Z, Kuang SB, Qiu TS, Yu AB. Lattice Boltzmann investigation on fluid flows through packed beds: Interaction between fluid rheology and bed properties. Powder Technology. 2020;369:248-260

Written By

Zheng Qi, Shibo Kuang, Liangwan Rong, Kejun Dong and Aibing Yu

Submitted: 13 June 2022 Reviewed: 08 July 2022 Published: 19 August 2022