Hostname: page-component-745bb68f8f-s22k5 Total loading time: 0 Render date: 2025-02-05T18:48:41.801Z Has data issue: false hasContentIssue false

Modulation of turbulence intensity by heavy finite-size particles in upward channel flow

Published online by Cambridge University Press:  19 February 2021

Zhaosheng Yu*
Affiliation:
State Key Laboratory of Fluid Power and Mechatronic Systems, Department of Mechanics, Zhejiang University, Hangzhou310027, PR China
Yan Xia
Affiliation:
State Key Laboratory of Fluid Power and Mechatronic Systems, Department of Mechanics, Zhejiang University, Hangzhou310027, PR China
Yu Guo
Affiliation:
State Key Laboratory of Fluid Power and Mechatronic Systems, Department of Mechanics, Zhejiang University, Hangzhou310027, PR China
Jianzhong Lin
Affiliation:
State Key Laboratory of Fluid Power and Mechatronic Systems, Department of Mechanics, Zhejiang University, Hangzhou310027, PR China
*
Email address for correspondence: yuzhaosheng@zju.edu.cn

Abstract

It has been recognized that, generally, large particles enhance the turbulence intensity, while small particles attenuate the turbulence intensity. However, there has been no consensus on the quantitative criterion for particle-induced turbulence enhancement or attenuation. In the present study, interface-resolved direct numerical simulations of particle-laden turbulent flows in an upward vertical channel are performed with a direct forcing/fictitious domain method to establish a criterion for turbulence enhancement or attenuation. The effects of the particle Reynolds number ($Re_p$), the bulk Reynolds number ($Re_b$), the particle size, the density ratio and the particle volume fraction on the turbulence intensity are examined at $Re_b=5746$ (i.e. $Re_\tau =180.8$) and 12 000 ($Re_\tau =345.9$), the ratio of the particle radius to the half channel width $a/H=0.05\text {--}0.15$, the density ratio 2–100, the particle volume fraction $0.3\,\%$$2.36\,\%$ and $Re_p < 227$. Our results indicate that at low $Re_p$ the turbulent intensity across the channel is all diminished; at intermediate $Re_p$ the turbulent intensity is enhanced in the channel centre region and attenuated in the near-wall region; and at sufficiently large $Re_p$ the turbulent intensity is enhanced across the channel. The critical $Re_p$ increases with increasing bulk Reynolds number, particle size and particle–fluid density ratio, while increasing with decreasing particle volume fraction, particularly for the channel centre region. Criteria for enhancement or attenuation are provided for the total turbulence intensity in the channel and the turbulence intensity at the channel centre, respectively, and both are shown to agree well with the experimental data in the literature. The reason for the dependence of the critical particle Reynolds number on the other parameters is discussed.

Type
JFM Papers
Copyright
© The Author(s), 2021. Published by Cambridge University Press

1. Introduction

Particle-laden turbulent flows are commonly encountered in natural and industrial settings, such as fluidized beds, pneumatic transport, fuel combustion and sand storms. Understanding the mechanisms of and developing accurate models for multiphase turbulent flows are of great importance to better designs of relevant industrial devices. Owing to the complicated interactions between particles and turbulence, the mechanisms of turbulence modulation and their parametric dependence remain poorly understood (Balachandar & Eaton Reference Balachandar and Eaton2010). It is still not clear under what conditions the turbulence intensity is enhanced or attenuated by particles, despite extensive investigations over the past several decades (Saber, Lundström & Hellström Reference Saber, Lundström and Hellström2015).

Numerous experiments have been reported on turbulence modulation in gas–solid flows (Lee & Durst Reference Lee and Durst1982; Tsuji & Morikawa Reference Tsuji and Morikawa1982; Tsuji, Morikawa & Shiomi Reference Tsuji, Morikawa and Shiomi1984; Parthasarathy & Faeth Reference Parthasarathy and Faeth1990; Schreck & Kleis Reference Schreck and Kleis1993; Kulick, Fessler & Eaton Reference Kulick, Fessler and Eaton1994; Petersen, Baker & Coletti Reference Petersen, Baker and Coletti2019; Zhu et al. Reference Zhu, Pan, Wang, Liang and Ji2019) and liquid–solid flows (Sato & Hishida Reference Sato and Hishida1996; Suzuki, Ikenoya & Kasagi Reference Suzuki, Ikenoya and Kasagi2000; Kiger & Pan Reference Kiger and Pan2002; Kussin & Sommerfeld Reference Kussin and Sommerfeld2002; Hosokawa & Tomiyama Reference Hosokawa and Tomiyama2004, Reference Hosokawa and Tomiyama2009; Bellani et al. Reference Bellani, Byron, Collignon, Meyer and Variano2012; Shokri et al. Reference Shokri, Ghaemi, Nobes and Sanders2017; Zade, Lundell & Brandt Reference Zade, Lundell and Brandt2019; Mena & Curtis Reference Mena and Curtis2020). It has been recognized that small particles tend to attenuate the turbulence, whereas large particles enhance the turbulence (Tsuji & Morikawa Reference Tsuji and Morikawa1982; Tsuji et al. Reference Tsuji, Morikawa and Shiomi1984; Gore & Crowe Reference Gore and Crowe1989).

Gore & Crowe (Reference Gore and Crowe1989) summarized the previous experimental data on vertical and horizontal pipe and jet flows, and suggested the ratio of the particle diameter $d_p$ to the characteristic length scale of energy-containing eddies $l_e$ as a critical parameter for turbulence modulation: $d_p/l_e < 0.1$ for turbulence attenuation and $d_p/l_e > 0.1$ for turbulence augmentation. Hetsroni (Reference Hetsroni1989) argued that particles suppress the turbulence at low particle Reynolds numbers and enhance the turbulence at high Reynolds numbers due to vortex shedding. He took the critical Reynolds number for the turbulence augmentation as 400, considering that vortex shedding takes place at this Reynolds number. Kulick et al. (Reference Kulick, Fessler and Eaton1994) experimentally observed in a downward channel flow that the turbulence was attenuated by small particles, and that the degree of attenuation increased with increasing particle loading and particle Stokes number. Crowe (Reference Crowe2000) extended the criterion of Gore & Crowe (Reference Gore and Crowe1989) from the balance between the production and dissipation of the turbulent kinetic energy at the pipe centre, and many parameters were involved, including particle volume fraction, density ratio, flow Reynolds number, Froude number, particle-free turbulence intensity, as well as $d_p/l_e$.

Hosokawa & Tomiyama (Reference Hosokawa and Tomiyama2004) demonstrated that the ratio of the eddy viscosity induced by the dispersed phase to the shear-induced eddy viscosity was a good parameter for correlating turbulence modification of their upward vertical pipe flows, which is defined as ${{u_r d_p}}/{{u'l_e}}$, where $u_r$ is the average slip velocity and $u'$ is the single-phase turbulent fluctuating velocity. Righetti & Romano (Reference Righetti and Romano2004) showed that the particle Stokes number $St$ (the ratio of the particle relaxation time to the characteristic time of energy-containing eddies) was a critical parameter for the particle modification of the horizontal turbulent open channel flow: turbulence augmentation at $St>1$ and attenuation at $St<1$. The experimental results of Noguchi & Nezu (Reference Noguchi and Nezu2009) on turbulent open channel flow indicated that the turbulent intensity was enhanced when the particle diameter was larger than the zone-averaged turbulence Kolmogorov length scale. Tanaka & Eaton (Reference Tanaka and Eaton2008) derived a novel dimensionless parameter, called the particle momentum number $Pa = St\,Re_L^2 ({\eta }/{L})^3$ (with $Re_L$ being the Reynolds number based on the large-eddy length scale $L$ and $\eta$ the Kolmogorov length), using dimensional analysis of the particle force in the momentum equation. From the previous experimental data on vertical and horizontal pipe and channel flows, the turbulence was augmented at $Pa<10^3$ or $Pa>10^5$, and attenuated at $10^3<Pa<10^5$.

With a similar dimensional analysis method, Luo, Luo & Fan (Reference Luo, Luo and Fan2016) derived another dimensionless parameter, $Cr$, defined as $Cr=({\rho _{p}}/{\rho _{f}}) ({L}/{d_{p}}) Re_{L}^{-11 /16} Re_{p}$ (with $\rho _p$, $\rho _f$ and $Re_p$ denoting the particle density, fluid density and particle Reynolds number, respectively), and the experimental data on vertical and horizontal pipe and channel flows indicated turbulence augmentation for $Cr>7000$ and attenuation for $Cr<7000$.

Some criteria for turbulence enhancement or attenuation in the literature are presented in table 1. These criteria are based on approximate theories and experimental data, which were generally scattered because of difficulty in measuring the particle-laden turbulent flow accurately.

Table 1. Some criteria for turbulence enhancement or attenuation in the literature.

So far, no criterion has been proposed from the numerical simulations, to the best of our knowledge. However, direct numerical simulations (DNS) with the point-particle-based and interface-resolved methods have provided deep insights into the mechanisms of turbulence modification by particles. The point-particle-based DNS have been applied to isotropic turbulent flows (Squires & Eaton Reference Squires and Eaton1990; Elghobashi & Truesdell Reference Elghobashi and Truesdell1993; Wang & Maxey Reference Wang and Maxey1993; Saito, Watanabe & Gotoh Reference Saito, Watanabe and Gotoh2019) and turbulent channel flows (Li et al. Reference Li, McLaughlin, Kontomaris and Portela2001; Eaton Reference Eaton2009; Wang Reference Wang2010; Zhao, Andersson & Gillissen Reference Zhao, Andersson and Gillissen2010, Reference Zhao, Andersson and Gillissen2013; Vreman Reference Vreman2015; Liu et al. Reference Liu, Tang, Shen and Dong2017; Wang et al. Reference Wang, Fong, Coletti, Capecelatro and Richter2019; Muramulla et al. Reference Muramulla, Tyagi, Goswami and Kumaran2020). In principle, the point-particle method is limited to small particles with size smaller than the Kolmogorov length scale. Therefore, turbulence attenuation was generally observed in the point-particle simulations. Squires & Eaton (Reference Squires and Eaton1990) reported that particles weakened the large-scale structures, whereas they intensified the small-scale structures in the homogeneous isotropic turbulence.

Li et al. (Reference Li, McLaughlin, Kontomaris and Portela2001) studied particle-laden gas flows in a downward channel and found that, at small mass loading, the particles tended to increase the turbulence, while they tended to suppress the turbulence at high mass loadings. Zhao et al. (Reference Zhao, Andersson and Gillissen2010, Reference Zhao, Andersson and Gillissen2013) showed that, at a large Stokes number, the streamwise velocity fluctuation was increased, while the spanwise and wall-normal velocity fluctuations were decreased. The numerical results of Liu et al. (Reference Liu, Tang, Shen and Dong2017) indicated that, in a vertical channel flow, the turbulence was weakened more significantly as the Stokes number increased. Vreman (Reference Vreman2015) demonstrated that the non-uniform part of the mean feedback force contributed significantly to the particle-induced turbulence attenuation in upward channel flow. Recently, Muramulla et al. (Reference Muramulla, Tyagi, Goswami and Kumaran2020) observed that the turbulent kinetic energy (TKE) of upward channel flow decreased suddenly with the increase of particle volume fraction, and argued that this attenuation was caused by the decrease of the TKE generation term, instead of the increase of the turbulent dissipation rate caused by particles.

Interface-resolved direct numerical simulation (IR-DNS) is a better method to study turbulence modulation by finite-size particles with size significantly larger than the Kolmogorov length scale (Balachandar & Eaton Reference Balachandar and Eaton2010; Tenneti & Subramaniam Reference Tenneti and Subramaniam2014; Maxey Reference Maxey2017). The essential features of the interface-resolved methods are that the interfaces between the particles and the fluid are resolved and the hydrodynamic forces on the particles are determined from the solution of the flow fields outside the particle boundaries. Such methods have been applied to simulations of particle-laden isotropic homogeneous flows (Ten Cate et al. Reference Ten Cate, Derksen, Portela and Van Den Akker2004; Lucci, Ferrante & Elghobashi Reference Lucci, Ferrante and Elghobashi2010; Gao, Li & Wang Reference Gao, Li and Wang2013), pipe flows (Wu, Shao & Yu Reference Wu, Shao and Yu2011; Peng & Wang Reference Peng and Wang2019), channel flows (Kajishima et al. Reference Kajishima, Takiguchi, Hamasaki and Miyake2001; Uhlmann Reference Uhlmann2008; Garcia-Villalba, Kidanemariam & Uhlmann Reference Garcia-Villalba, Kidanemariam and Uhlmann2012; Shao, Wu & Yu Reference Shao, Wu and Yu2012; Picano, Breugem & Brandt Reference Picano, Breugem and Brandt2015; Santarelli & Fröhlich Reference Santarelli and Fröhlich2015; Wang et al. Reference Wang, Peng, Guo and Yu2016; Yu, Vinkovic & Buffat Reference Yu, Vinkovic and Buffat2016a; Ardekani et al. Reference Ardekani, Costa, Breugem, Picano and Brandt2017; Yu et al. Reference Yu, Lin, Shao and Wang2017; Costa et al. Reference Costa, Picano, Brandt and Breugem2018; Peng, Ayala & Wang Reference Peng, Ayala and Wang2019; Costa, Brandt & Picano Reference Costa, Brandt and Picano2020; Zhu et al. Reference Zhu, Yu, Shao and Deng2020b), duct flows (Lin et al. Reference Lin, Shao, Yu and Wang2017a,; Fornari et al. Reference Fornari, Kazerooni, Hussong and Brandt2018) and Couette flows (Wang, Abbas & Climent Reference Wang, Abbas and Climent2017). Lucci et al. (Reference Lucci, Ferrante and Elghobashi2010) observed that particles larger than the Kolmogorov scale always reduced the average TKE of the decaying homogeneous isotropic turbulence, and attributed the reason to the particle-induced viscous dissipation near particle surfaces. The finite-size particles reduced the energy spectrum of low wavenumbers and increased the energy spectrum of high wavenumbers (Lucci et al. Reference Lucci, Ferrante and Elghobashi2010), similar to the point-particle case (Squires & Eaton Reference Squires and Eaton1990). The critical wavenumber depended not only on the actual size of the particles, but also on the relative size of the particle with respect to the turbulence scale and the density ratio (Gao et al. Reference Gao, Li and Wang2013).

Regarding the effects of finite-size neutrally buoyant particles on turbulent channel flow, the results of Shao et al. (Reference Shao, Wu and Yu2012) indicated that the presence of particles decreased the streamwise maximum root-mean-square (r.m.s.) velocity near the wall by weakening the intensity of the large-scale streamwise vortices, while increasing the transverse and spanwise r.m.s. velocities in the near-wall region by inducing particle-scale vortices, as later observed by Picano et al. (Reference Picano, Breugem and Brandt2015) and Wang et al. (Reference Wang, Peng, Guo and Yu2016). When gravity was not considered, the particles with large inertia (i.e. large density ratio) significantly suppressed the turbulence (Fornari et al. Reference Fornari, Formenti, Picano and Brandt2016; Yu et al. Reference Yu, Lin, Shao and Wang2017).

Regarding horizontal channel flows, Shao et al. (Reference Shao, Wu and Yu2012) showed that sediments played the role of a rough wall and parts of the vortex structures shedding from the particles ascended into the centre region and substantially increased the turbulence intensity there. Peng et al. (Reference Peng, Ayala and Wang2019) and Vreman & Kuerten (Reference Vreman and Kuerten2018) studied the flow modulation by an array of fixed and moving spherical particles in turbulent channel flows, respectively, and attenuation of TKE was observed. Regarding upward channel flow loaded with heavy spherical particles, Kajishima et al. (Reference Kajishima, Takiguchi, Hamasaki and Miyake2001) and Uhlmann (Reference Uhlmann2008) conducted interface-resolved simulations at particle Reynolds numbers greater than 130 where the particle settling velocity was comparable to the fluid mean velocity. Both results showed that the particles strongly enhanced the turbulence intensity. Santarelli & Fröhlich (Reference Santarelli and Fröhlich2015) simulated upward turbulent channel flow laden with spherical bubble particles at particle Reynolds numbers greater than 200, and also observed turbulence enhancement. Zhu et al. (Reference Zhu, Yu, Pan and Shao2020a) showed that finite-size particles attenuated the turbulence in upward channel flows at relatively low particle Reynolds numbers (smaller than 33) and the increase in the particle Reynolds number (i.e. settling coefficient) gave rise to more pronounced attenuation. Therefore, for upward channel flow, there should exist a critical particle Reynolds number for the transition from turbulence attenuation to augmentation. The main aim of the present study is to determine the dependence of this critical particle Reynolds number on the control parameters, including the particle size, particle volume fraction, density ratio and channel Reynolds number. In other words, we attempt to present a new criterion for turbulence modulation, from the interface-resolved simulations of upward channel flows.

The rest of the paper proceeds as follows. The numerical method and the flow model are presented in the next section (§ 2). The accuracy of our code for the single-phase TKE budget and the collision model are validated in § 3. In § 4, the results of the particle effects on the turbulence statistics and the criteria for turbulent intensity modulation are presented. The reasons for the particle effects on the TKE are discussed in association with the modelling of the interfacial term in the TKE equation. In addition, the results on the wall friction and the total flow drag are reported. Finally, concluding remarks are given in § 5.

2. Methodology and simulation set-up

2.1. Direct forcing/fictitious domain method

In the present study, we employ a direct forcing/fictitious domain (DF/FD) method proposed by Yu & Shao (Reference Yu and Shao2007) to conduct IR-DNS of upward turbulent channel flows. The key idea of this method is that the interior of particles is filled with the fluid and the interior ‘fictitious’ fluid is enforced to satisfy the rigid-body motion constraint through a pseudo body force (i.e. distributed Lagrange multiplier (Glowinski et al. Reference Glowinski, Pan, Hesla and Joseph1999)). For the simplicity of description, we consider only one particle in the following formulae. Let $P(t)$ represent the solid domain and $\varOmega$ represent the entire domain including both the interior and exterior of the solid body. The particle density, volume and moment-of-inertia tensor, translational velocity and angular velocity are denoted by $\rho _s$, $V_p$, $\boldsymbol{\mathsf{J}}$, $\boldsymbol U$ and $\boldsymbol \omega _p$, respectively. The viscosity and density of the fluid are $\mu$ and $\rho _f$, respectively.

We introduce the following characteristic scales for the non-dimensionalization: $L_c$ for length, $U_c$ for velocity, $L_c/U_c$ for time and ${\rho _f}U_c^2/L_c$ for the body force. Then, the dimensionless FD formulation for the incompressible fluid can be written as

(2.1)\begin{gather} \frac{\partial {\boldsymbol{u}}}{\partial t}+{\boldsymbol{u}} \boldsymbol{\cdot} \boldsymbol{\nabla} {\boldsymbol{u}}= \frac{\nabla^2 {\boldsymbol{u}}}{Re}-\boldsymbol{\nabla} p +{\boldsymbol{\lambda}} \text{ in }{\varOmega}, \end{gather}
(2.2)\begin{gather}{\boldsymbol{u}}={\boldsymbol{U}}+{\boldsymbol{\omega}_p} \times {\boldsymbol{r}} \text{ in }{P(t)}, \end{gather}
(2.3)\begin{gather}\boldsymbol{\nabla} \boldsymbol{\cdot} {\boldsymbol{u}} =0 \text{ in }{\varOmega}, \end{gather}
(2.4)\begin{gather}(\rho_r -1)V_p^*\left(\frac{\textrm{d} {\boldsymbol{U}}}{\textrm{d} t}-Fr \frac{\boldsymbol{g}}{ g}\right)={-}\int_P {\boldsymbol{\lambda}}\, \textrm{d}{\boldsymbol {x}}, \end{gather}
(2.5)\begin{gather}(\rho_r -1)\frac{\textrm{d} ({\boldsymbol{\mathsf{J}}}^* \boldsymbol{\cdot}{\boldsymbol{\omega }}_p)}{\textrm{d} t}={-}\int_P{\boldsymbol r}\times{\boldsymbol{\lambda}}\,\textrm{d}{\boldsymbol {x}}. \end{gather}

Here $\boldsymbol u$, $p$, $\boldsymbol \lambda$ and $\boldsymbol r$ represent the fluid velocity, pressure, pseudo body force and position vector with respect to the centre of mass of the particle, respectively, $\rho _r$ is the particle–fluid density ratio defined by $\rho _r=\rho _s/\rho _f$, $Re$ is the Reynolds number defined by $Re=\rho _{f} U_c L_c/\mu$, $Fr$ is the Froude number defined by $Fr=gL_c/U_c^2$, with $g$ being the gravitational acceleration, and $V_p^*$ and ${\boldsymbol{\mathsf{J}}}^*$ are the dimensionless particle volume and moment-of-inertia tensor defined by $V_p^*=V_p/L_c^3$ and ${\boldsymbol{\mathsf{J}}}^*={\boldsymbol{\mathsf{J}}}/\rho _s L_c^5$.

A fractional-step time scheme is used to decouple the system (2.1)–(2.5) into the following two subproblems.

(1) Fluids subproblem for ${\boldsymbol {u}}^*$ and $p$:

(2.6)\begin{gather} \frac{{\boldsymbol u}^*-{\boldsymbol u}^n}{{\rm \Delta} t}-\frac{1}{2}\frac{\nabla^2{\boldsymbol u}^*}{Re}={-}\boldsymbol{\nabla} p-\frac{1}{2}[3({\boldsymbol u} \boldsymbol{\cdot} \boldsymbol{\nabla} {\boldsymbol u})^n-({\boldsymbol u} \boldsymbol{\cdot} \boldsymbol{\nabla} {\boldsymbol u})^{n+1}]+\frac{1}{2}\frac{\nabla^2{\boldsymbol u}^n}{Re}+{\boldsymbol \lambda}^n, \end{gather}
(2.7)\begin{gather}\boldsymbol{\nabla} \boldsymbol{\cdot} {\boldsymbol u}^*=0. \end{gather}

A finite-difference-based projection method on a homogeneous half-staggered grid is used for the solution of the above fluid subproblem. All spatial derivatives are discretized with a second-order central-difference scheme.

(2) Particle subproblem for ${\boldsymbol U}^{n+1}$, ${\boldsymbol \omega }_p^{n+1}$, ${\boldsymbol \lambda }^{n+1}$ and ${\boldsymbol u}^{n+1}$:

(2.8)\begin{equation} \rho_r V_p^* \frac{{\boldsymbol U}^{n+1}}{{\rm \Delta} t}=(\rho_r-1) V_p^* \left(\frac {{\boldsymbol U}^n}{{\rm \Delta} t}-Fr\frac{\boldsymbol g}{g}\right)+\int_P \left({\frac{{\boldsymbol u}^*}{{\rm \Delta} t}- {\boldsymbol \lambda}^n}\right)\, \textrm{d}{\boldsymbol{x}}, \end{equation}
(2.9)\begin{equation}\rho_r \frac{{\boldsymbol{\mathsf{J}}}^{*}\boldsymbol{\cdot} {\boldsymbol \omega }_p^{n+1}}{{\rm \Delta} t}=(\rho_r-1) \left[\frac{{\boldsymbol{\mathsf{J}}}^{*} \boldsymbol{\cdot} {\boldsymbol \omega }_p^{n}}{{\rm \Delta} t}-{\boldsymbol \omega }_p^{n} \times ({ \boldsymbol{\mathsf{J}}}^{*} \boldsymbol{\cdot} {\boldsymbol \omega }_p^{n})\right]+\int_P {\boldsymbol r} \times \left({\frac{{\boldsymbol u}^*}{{\rm \Delta} t}- {\boldsymbol \lambda}^n}\right)\, \textrm{d}{\boldsymbol {x}}. \end{equation}

The pseudo body force $\boldsymbol \lambda$ is then updated from

(2.10)\begin{equation} {\boldsymbol \lambda}^{n+1}=\frac{{\boldsymbol U}^{n+1}+{\boldsymbol \omega}_p^{n+1}\times {\boldsymbol r}-{\boldsymbol u}^*}{{\rm \Delta} t}+{\boldsymbol \lambda}^n. \end{equation}

Finally, the fluid velocities ${\boldsymbol u}^{n+1}$ at the Eulerian nodes are corrected from

(2.11)\begin{equation} {\boldsymbol u}^{n+1}={\boldsymbol u}^{*}+{\rm \Delta} t ({\boldsymbol \lambda}^{n+1}-{\boldsymbol \lambda}^{n}). \end{equation}

In the above manipulations, the trilinear function is used to transfer the fluid velocity from the Eulerian nodes to the Lagrangian nodes and the pseudo body force from the Lagrangian nodes to the Eulerian nodes.

2.2. Collision model

A state-of-the-art collision model based on the combination of a soft-sphere collision model and the lubrication force correction has been developed for the IR-DNS by Kempe & Fröhlich (Reference Kempe and Fröhlich2012), Brändle de Motta et al. (Reference Brändle de Motta, Breugem, Gazanion, Estivalezes, Vincent and Climent2013), Costa et al. (Reference Costa, Boersma, Westerweel and Breugem2015) and Biegert, Vowinckel & Meiburg (Reference Biegert, Vowinckel and Meiburg2017). A similar collision model is used here. The lubrication force correction has the following form:

(2.12)\begin{equation} \boldsymbol{F}_{i j}^{l}={-}6 {\rm \pi}\mu a \boldsymbol{u}_n[\lambda(\epsilon)-\lambda(\epsilon_{al})], \end{equation}

where $\boldsymbol {u}_n$ is the normal relative velocity between the $i$th and $j$th objects (particle or wall), $\lambda (\epsilon )$ is a function of the normalized gap distance $\epsilon = {{{\zeta _n}} / a}$ and $\epsilon _{al}={\rm \Delta} x/a$ is the threshold gap below which the lubrication correction is activated, where $a$ represents the particle radius and ${\rm \Delta} x$ is the mesh size. The lubrication force is kept constant for $\epsilon < \epsilon _1$, where $\epsilon _1$ is set to 0.001 in our simulations.

The functions $\lambda$ for the particle–particle and particle–wall interactions are (Jeffrey Reference Jeffrey1982):

(2.13)\begin{gather} \text{particle--particle}\quad\lambda ( \epsilon )= \frac{1}{{4\epsilon }} - \frac{9}{{40}}\ln ( \epsilon ) - \frac{3}{{92}}\epsilon \ln ( \epsilon ) + 0.673, \end{gather}
(2.14)\begin{gather}\text{particle--wall}\quad\lambda ( \epsilon )= \frac{1}{\epsilon } - \frac{1}{5}\ln ( \epsilon ) - \frac{1}{{21}}\epsilon \ln ( \epsilon ) + 0.9713. \end{gather}

A discrete element model is employed as a soft-sphere collision model. The components of the collision force on the object $i$ applied by another object $j$ are (Crowe et al. Reference Crowe, Schwarzkopf, Sommerfeld and Tsuji2011)

(2.15)\begin{gather} {{\boldsymbol{F}}_{ij,n}} = ({-{k_n}\delta _n^{{3 / 2}} - {\eta _n}{\boldsymbol{G}} \boldsymbol{\cdot} {\boldsymbol{n}}} ){\boldsymbol{n}}, \end{gather}
(2.16)\begin{gather}{{\boldsymbol{F}}_{ij,t}} ={-} {k_t}{{\boldsymbol{\delta}}_t} - {\eta _t}{{\boldsymbol{G}}_{ct}}, \end{gather}

where ${{\boldsymbol {F}}_{ij,n}}$, $\delta _n$, $k_n$ and $\eta _n$ are the contact force, overlapping distance, stiffness coefficient and damping coefficient in the normal direction, respectively, and ${{\boldsymbol {F}}_{t}}$, $\delta _t$, $k_t$ and $\eta _t$ are the corresponding parameters in the tangential direction. Here $\boldsymbol {n}$ is the unit normal vector pointing from the centre of particle $i$ to that of particle $j$, $\boldsymbol {G}$ is the relative velocity between particles and ${\boldsymbol {G}}_{ct}$ is the tangential relative velocity. The definitions of $k_n$ and $k_t$ are given by Hertzian contact theory (Hertz Reference Hertz1882) and Mindlin's theory (Mindlin Reference Mindlin1953) as

(2.17)\begin{gather} {k_n}= \frac{4}{3}{\left( {\frac{{1 - \sigma _i^2}}{{{E_i}}} + \frac{{1 - \sigma _j^2}}{{{E_j}}}} \right)^{ - 1}}{\left( {\frac{{{a_i} + {a_j}}}{{{a_i}{a_j}}}} \right)^{ - {1 / 2}}}, \end{gather}
(2.18)\begin{gather}k_{t}=8\left(\frac{2-\sigma_{i}}{G_{i}}+\frac{2-\sigma_{j}}{G_{j}}\right)^{{-}1} \left(\frac{a_{i}+a_{j}}{a_{i} a_{j}}\right)^{{-}1/2}\delta_{n}^{1/2}, \end{gather}

where $E$ and $\sigma$ are Young's modulus and Poisson's ratio, respectively, and $G = {E / {2({1 + \sigma } )}}$ is the particle shear modulus. The damping coefficients are given by Tsuji, Tanaka & Ishida (Reference Tsuji, Tanaka and Ishida1992) and Barnocky & Davis (Reference Barnocky and Davis1988) as

(2.19a,b)\begin{equation} \eta _n = {\alpha_n}\sqrt {{m_p}{k_n}} \delta _n^{{1 / 4}},\quad \eta_t = {\alpha_t}\sqrt {{m_p}{k_t}} \delta _t^{{1 / 4}}, \end{equation}

where ${m_p} = {{{m_i}{m_j}} / {( {{m_i} + {m_j}} )}}$ is the effective mass and $\alpha _n$ is a constant related to the dry coefficient of restitution $e_d$ as

(2.20)\begin{equation} \alpha_n ( e_d )=\alpha_t ( e_d ) = 2.22 - 2.26{e_d^{0.4}}. \end{equation}

If $| {{\boldsymbol {F}_t}} | \ge f| {{\boldsymbol {F}_n}} |$, the tangential force is given by the Coulomb-type friction law:

(2.21)\begin{equation} {\boldsymbol{F}}_{t}={-}f|{\boldsymbol{F}}_{n}|{\boldsymbol{t}}, \end{equation}

where $f$ is the friction coefficient and $\boldsymbol {t}$ is the unit tangential vector. In the present study, we set $E/(\rho _f u_b^2)=2 \times 10^3$ (here $u_b$ being the mean velocity of the channel flow), $e_d=0.97$, $\sigma = 0.33$, and $f = 0.3$ for particle–particle collisions and $f = 0.2$ for particle–wall collisions. The time step for the collision is set to be one-tenth of that for the flow solution.

We have recently investigated the effects of the collision model on the turbulent channel flow laden with neutrally buoyant particles (Xia et al. Reference Xia, Xiong, Yu and Zhu2020a). The results showed that the lubrication force correction for the particle pairs had an important effect on the particle pair statistics at the near-contact regime, and could lead to the decrease in the flow friction by around 1.4 % for a particle volume fraction of 2.36 %. For an upward channel flow laden with heavy particles here, the effect of the collision model is expected to be less significant, compared to the neutrally buoyant case, since the heavy particles tend to migrate away from the wall and there are fewer particles in the near-wall region where the particles would collide more frequently and strongly due to high shear rates.

2.3. Simulation set-up

The schematic diagram of the upward channel flow is delineated in figure 1. The computational domain is $[0,8H]\times [-H,H]\times [0,4H]$ in the streamwise ($x$), wall-normal ($y$) and spanwise ($z$) directions, where $H$ is the half channel width. A periodic boundary condition is imposed in the streamwise and spanwise directions, and the no-slip condition is imposed on the sidewalls.

Figure 1. Schematic diagram of the upward channel flow laden with particles; with $x$, $y$ and $z$ representing the streamwise, transverse and spanwise coordinates, respectively.

The flow is driven upwards by a mean pressure gradient $-{\textrm {d}p_e}/{\textrm {d} x}$ in the direction of the $x$ coordinate to maintain a constant flow rate. The pressure gradient is varied as the particle parameters are changed. We choose $H$ and the bulk velocity $u_b$ as the characteristic length and velocity. The channel Reynolds number is defined as $R{e_b}= {{u_b}(2H)/\nu }$, where $\nu$ is the fluid kinematic viscosity; $R{e_b}$ is constant in the simulations, since the bulk velocity $u_b$ is kept constant. Finally, $\boldsymbol g = (-g, 0, 0)$ is the vector of gravitational acceleration. We introduce a dimensionless parameter $u_i/u_b$ as the settling coefficient to control the particle settling effect, where $u_i$ is the velocity of a particle settling in a quiescent fluid when the standard drag coefficient is unity, defined by

(2.22)\begin{equation} {u_i} = \sqrt {\frac{{8a}}{3}( {{\rho_{r}}-1})g}. \end{equation}

Since the terminal velocity of a sphere settling in an unbounded domain is $u_T=\sqrt {({1}/{C_{D}})({8 a}/{3})(\rho _{r}-1) g}$, where $C_D$ represents the standard drag coefficient, $u_i$ is related to $u_T$ via

(2.23)\begin{equation} \frac{u_{T}}{u_{i}}=\sqrt{\frac{1}{C_{D}}}. \end{equation}

For a relatively strong fluid inertial effect, $u_i$ is close to $u_T$, since the drag coefficient is close to unity.

The Galileo number, defined as $G a= \sqrt {8 a^3(\rho _{r}-1) g} / \nu$ (with $\nu$ being the fluid kinematic viscosity), is widely used to measure the particle settling effect. The particle Reynolds number based on $u_i$ is related to the Galileo number via $Re_i = {{u_i}(2a)/\nu }= \sqrt {4/3} Ga$. We note that the Archimedes number is the square of the Galileo number, and $Re_{i}^{2}$ was defined as the ‘best number’ for the particle settling problem (Yu, Phan-Thien & Tanner Reference Yu, Phan-Thien and Tanner2004; Clift, Grace & Weber Reference Clift, Grace and Weber2005). For a given value of $u_i/u_b$, the Galileo number can be computed by

(2.24)\begin{equation} Ga=\frac{\sqrt{3}}{2} \frac{a}{H} \frac{u_{i}}{u_{b}}Re_{b}. \end{equation}

The Froude number $Fr$ can be determined from other control parameters:

(2.25)\begin{equation} Fr =\frac{{gH}}{{u_b^2}} = \frac{{3{{( {{{{u_i}} / {{u_b}}}} )}^2}}}{{8( {{a/ H}})({{\rho _{r}}-1})}}. \end{equation}

The turbulent kinetic energy of the particle-laden channel flow $k$ is controlled by the following eight parameters (seven dimensional and one dimensionless): $k = f(u_b,H,a,\rho _s,\rho _f,\mu ,g,N_p)$, where $N_p$ represents the number of particles in the system. Accordingly, the dimensionless turbulent kinetic energy should be a function of five dimensionless parameters and we choose them as $k/u_b^2 = f(Re_b,a/H,\rho _r,u_i /u_b ,\varPhi _0 )$, where $\varPhi _\textrm {{0}}$ is the mean particle volume fraction in the entire channel. Since the particle wake is more closely related to the particle Reynolds number $Re_p$ than the settling coefficient, the settling coefficient is replaced with $Re_p$ in the criterion for turbulence enhancement. Note that the particle Reynolds number is based on the average slip velocity between the two phases at the channel centre: $Re_p = {{| { \langle u_f \rangle - \langle u_p \rangle } |2a} }/{\nu }$, where $\langle u_f \rangle$ and $\langle u_p \rangle$ are the average velocities of the fluid and particles at the channel centre, respectively.

To examine the effects of the particle size, particle volume fraction, channel Reynolds number and density ratio, nine groups of parameters are chosen. For each group, the settling coefficient $u_i/u_b$ is varied to determine the critical particle Reynolds number for turbulence augmentation. We note that when the other parameters ($Re_b,a/H,\rho _r$) are fixed, the settling coefficient is varied by changing the gravitational acceleration, which we have chosen as the dimensional control parameter. The parameter settings for the nine groups are presented in table 2. There are in total 61 cases, with the consideration of different values of $u_i/u_b$ for each group (see tables 6 and 7).

Table 2. Parameter settings for the particle-laden flows in a vertical channel: $a$ represents the radius of particles, $N_p$ is the number of particles and $\varPhi _\textrm {{0}}$ is the mean particle volume fraction in the entire channel.

Two channel Reynolds numbers $Re_b=5746$ and 12 000 are considered, corresponding to the friction Reynolds numbers $Re_\tau =\rho _f u_\tau H/\mu = 180.8$ and 345.9 for the single-phase flow, respectively, as listed in table 6. Here, $u_\tau$ is the friction velocity, defined by $u_\tau = \sqrt {\tau _w /\rho _f }$, where $\tau _w$ represents the mean wall shear stress. The Kolmogorov length $\eta$ for the turbulent channel flow can be computed by

\[ \frac{\eta}{H} = \frac{1}{H} \left(\frac{\nu^3}{\varepsilon}\right)^{{1}/{4}} = \frac{1}{H} \left( \frac{\nu^3}{\varepsilon^*(u_b^3/H)}\right)^{{1}/{4}} = \left(\frac{1}{\varepsilon^*(Re_b/2)^3}\right)^{{1}/{4}}, \]

where $\nu$ is the kinematic viscosity, $\varepsilon$ is the turbulence dissipation rate and $\varepsilon ^*$ is the dimensionless turbulence dissipation rate normalized with $u_b^3 /H$. The Taylor microscale can be defined as $\lambda = \sqrt {{{15\nu u'^2 } }/{\varepsilon }} = \sqrt {{{10\nu k} }/{\varepsilon }}$, where $u'$ is the turbulence intensity, related to the turbulent kinetic energy via $k = 3u'^2 /2$ (Gao et al. Reference Gao, Li and Wang2013). The Taylor-microscale Reynolds number is defined by $Re_\lambda = {{u'\lambda } }/{\nu }$, and it can be derived that $Re_\lambda = \sqrt {({{10} }/{3}){{(k^* )^2 Re_b } }/{{\varepsilon ^* }}}$, where $k^*$ is the dimensionless turbulent kinetic energy normalized by $u_b^2$. From our simulations, $k^* = 2.86 \times 10^{ - 3}$ and $\varepsilon ^* = 2.52 \times 10^{ - 4}$ for $Re_b=5746$, and $k^* = 2.32 \times 10^{ - 3}$ and $\varepsilon ^* = 1.89 \times 10^{ - 4}$ for $Re_b=12\,000$ at the channel centre. Thus, the Kolmogorov length scale at the channel centre $\eta / H=0.02$ for $Re_b=5746$ and $\eta / H=0.0125$ for $Re_b=12\,000$. The particle size in the present study ranges from $a/H=0.05$ to $a/H=0.15$, corresponding to $a / \eta =2.5\text {--}7.5$ for $Re_b=5746$ and $a / \eta =4\text {--}12$ for $Re_b=12\,000$. The Taylor-microscale Reynolds number $Re_\lambda$ at the channel centre is around 25 for $Re_b=5746$ and 33.8 for $Re_b=12\,000$, respectively.

Two mesh resolutions with respect to the particle size, $d_p/{\rm \Delta} x=12.8$ and 19.2, are used in our simulations, which are comparable to those employed by Uhlmann (Reference Uhlmann2008) and Santarelli & Fröhlich (Reference Santarelli and Fröhlich2015). The accuracy of our DF/FD code for the single-phase turbulent channel and duct flows were validated in Yu et al. (Reference Yu, Lin, Shao and Wang2016b, Reference Yu, Zhu, Wang and Shao2019) and Lin et al. (Reference Lin, Yu, Shao and Wang2017a). Zhu et al. (Reference Zhu, Yu, Shao and Deng2020b) showed that the turbulence statistics for a vertical particle-laden channel flow obtained from $d_p/{\rm \Delta} x=12.8$ were in good agreement with those from $d_p/{\rm \Delta} x=25.6$. Our DF/FD code has been applied to various types of particle-laden turbulent flows, including pipe flows (Wu et al. Reference Wu, Shao and Yu2011), duct flows (Lin et al. Reference Lin, Yu, Shao and Wang2017a,) and channel flows laden with spherical particles (Shao et al. Reference Shao, Wu and Yu2012; Yu et al. Reference Yu, Lin, Shao and Wang2017; Xia, Yu & Guo Reference Xia, Yu and Guo2020b) and spheroidal particles (Zhu, Yu & Shao Reference Zhu, Yu and Shao2018; Zhu et al. Reference Zhu, Yu, Pan and Shao2020a). Thus, we will only validate our codes for the single-phase mean TKE equation and the collision model later. We note that our mesh resolution in table 2 may not ensure that the solution of all turbulent flow details is highly accurate, such as the turbulent dissipation rate in the immediate vicinity of the wall or particle surfaces. However, according to our previous validation tests (Yu et al. Reference Yu, Lin, Shao and Wang2016b; Zhu et al. Reference Zhu, Yu, Shao and Deng2020b), the mesh resolution should be sufficient for the statistical properties of large numbers of particles suspended in a turbulent flow, and their influence on the overall dynamics, which are of interest in the present study. Mesh refinement is limited by our computer resources.

The dimensionless time step is ${\rm \Delta} t=0.001$. The statistics are obtained from the averaging of the data over a period of 500 non-dimensional time units ($H/u_b$) after the statistically steady state is reached. The particle-phase statistics are obtained from the data at the fictitious fluid domain (i.e. the solid domain).

2.4. Definition of turbulence intensity modulation

The turbulence intensity modulation is quantified as the relative change in the turbulence intensity due to the presence of particles (Gore & Crowe Reference Gore and Crowe1989):

(2.26)\begin{equation} {CT}=\frac{{{I}-{I_{\textrm{{sp}}}}}}{{{I_{\textrm{{sp}}}}}}, \end{equation}

where $I$ denotes the mean turbulent intensity of the fluid phase defined as $I = \sqrt {\tfrac {1}{3}( {{{u'_{rms}} ^2} + {{v'_{rms}} ^2} + {{w'_{rms} }^2} } )} /u_b$ and the subscript ‘sp’ stands for the case of ‘single-phase’ flow. We normalize the turbulent intensity with the bulk velocity $u_b$, because $u_b$ is kept constant and is the velocity scale in our computations. The velocity at the pipe centreline was typically used to normalize the turbulent intensity in the experiments (Gore & Crowe Reference Gore and Crowe1989). We consider the change of the turbulent intensity at the channel centre and that of the total turbulent intensity in the channel:

(2.27)\begin{equation} \widetilde {{CT}}= \frac{{\displaystyle\int_{{-H}}^{{H}} {( {{I} - {I_{\textrm{{sp}}}}} )\,\textrm{{d}}y} }}{{\displaystyle\int_{{-H}}^{{H}} {{I_{\textrm{{sp}}}} \,\textrm{{d}}y}}}. \end{equation}

We use the tilde bar $\tilde \cdot$ to indicate the averaging of a quantity over the entire channel in the present study.

3. Validation

The accuracy of our method for the single-phase mean and r.m.s. velocities and the mesh independence of the turbulence statistics for the particle-laden channel flow have been demonstrated in our previous works (Yu et al. Reference Yu, Lin, Shao and Wang2016b, Reference Yu, Zhu, Wang and Shao2019; Zhu et al. Reference Zhu, Yu, Shao and Deng2020b). Here, we provide additional validations for the computation of the single-phase TKE equation and the collision model with the lubrication force correction, since neither has been reported previously.

3.1. Single-phase turbulent channel flow

The first test problem is the particle-free turbulent channel flow at $Re_{\tau }=180$; here $Re_{\tau }$ is the friction Reynolds number defined by $Re_{\tau }=(\rho _{f} u_{\tau } H) / \mu$, where $u_{\tau }=\sqrt {\tau / \rho _{f}}$, with $\tau$ being the mean wall shear stress. The computational domain is $[8 H \times 2 H \times 4 H]$, with the grid number of $1024 \times 256 \times 512$. Our profiles of the production, dissipation and diffusion terms are compared to those of Hoyas & Jiménez (Reference Hoyas and Jiménez2008) in figure 2. The results of Hoyas & Jiménez (Reference Hoyas and Jiménez2008) were obtained using the pseudo-spectral simulations with a computational domain of $12 {\rm \pi}H \times 2 H \times 4 {\rm \pi}H$. From figure 2, our results are in excellent agreement with those of Hoyas & Jiménez (Reference Hoyas and Jiménez2008).

Figure 2. Profiles of the mean TKE production, dissipation and diffusion terms for the particle-free turbulent channel flow, normalized with $(u_{\tau })^{4} / \nu$, as compared to the results of Hoyas & Jiménez (Reference Hoyas and Jiménez2008).

3.2. Collision of a sphere with a wall in a viscous fluid

The bouncing motion of a sphere colliding with a planar wall in a viscous fluid is numerically simulated with our fictitious domain method and the collision model presented earlier. Our results are compared to the experimental results of Gondret, Lance & Petit (Reference Gondret, Lance and Petit2002) for two Stokes numbers $St_c$ = 27 and 152. The Stokes number is defined as $St_c={\rho _{s} u_{in} d_{p}}/{9\mu }$, with $u_{in}$ being the impact velocity before the collision. The parameters are summarized in table 3. The trajectories of the sphere for the two cases are plotted in figure 3. Our results are in excellent agreement with the experiments.

Table 3. Parameter settings for the bouncing motion of a sphere colliding with a planar wall in a viscous fluid.

Figure 3. Trajectories of a sphere colliding with a planar wall in a viscous fluid for (a) $St_c$ = 27 and (b) $St_c$ = 152, compared to the experimental results of Gondret et al. (Reference Gondret, Lance and Petit2002). Here $\zeta _{n}$ is the shortest distance between the sphere surface and the wall, and $t_{ref}=\sqrt {d_{p}/g}$ denotes the reference time scale.

4. Results and discussion

In this section, we first present and discuss the results on the fluid mean velocity, fluctuating velocity, vortex structure and particle concentration distribution at different settling coefficients $u_i/u_b$ for a typical case of $a/H=0.05$, $\varPhi _0=2.36\,\%$, $Re_b=5746$ and $\rho _{r}=2.0$ (i.e. group 5 in table 2). Next, we examine the dependence of the critical Reynolds number on the particle size, particle volume fraction, density ratio and channel Reynolds number, report new criteria for the turbulence modulation, and analyse the mechanisms from the TKE production in the particle-laden turbulent flow. Finally, the friction drag and the total flow drag are reported. Note that, although $y$ is defined as the coordinate in the wall-normal direction, ranging from $-H$ to $H$ (or $-$1 to 1 for its dimensionless value), as used in figure 10, for convenience it is also defined as the distance away from the wall when plotting and discussing the profiles of the statistics.

4.1. Effects of settling coefficient

The effects of the settling coefficient on the fluid and particle statistics are investigated for the case of group 5: $a/H=0.05$, $\varPhi _0=2.36\,\%$, $Re_b=5746$ and $\rho _{r}=2$. The settling coefficient $u_i/u_b$ ranges from 0.092 to 0.7. The cases of $u_i/u_b=0.092$ and 0.159 have been studied by Zhu et al. (Reference Zhu, Yu, Pan and Shao2020a), where the settling coefficient was defined as the ratio of the Stokes settling velocity and the bulk velocity $u_s/u_b$, and these two cases correspond to $u_s/u_b=0.1$ and 0.3.

4.1.1. Mean velocity

Figure 4 shows the mean fluid velocity profiles, with the inset showing the close-up of the velocities in the immediate vicinity of the wall. The mean velocity profile becomes flatter as $u_i/u_b$ increases from zero to 0.25. The mean velocity at $u_i/u_b=0.25$ becomes uniform for $y>0.3H$. As $u_i/u_b$ further increases, the mean velocity profile remains flat in the bulk region, but at $y$ around $0.2H$ the velocity decreases. We suspect that this decrease is related to the local peak of the particle concentration distribution, which could cause a larger local flow drag. Figure 6 shows a local peak of the particle concentration distribution at $y \approx 0.2H$ for $u_i/u_b=0.4$.

Figure 4. Mean fluid velocity profiles of turbulent channel flows for different $u_i/u_b$ at $a/H=0.05$, $\varPhi _0=2.36\,\%$, $Re_b=5746$ and $\rho _{r}=2.0$. The inset shows the close-up of the velocity profiles near the wall.

It was shown that the presence of neutrally buoyant finite-size particles generally leads to drag enhancement (Shao et al. Reference Shao, Wu and Yu2012; Picano et al. Reference Picano, Breugem and Brandt2015; Wang et al. Reference Wang, Peng, Guo and Yu2016). However, heavy particles in upward channel flow could cause a decrease in the wall friction at $u_i/u_b=0.159$ due to the suppression of large-scale vortices (Zhu et al. Reference Zhu, Yu, Pan and Shao2020a). From the inset of figure 4, the velocity gradient on the wall (i.e. flow friction) reaches a minimum at $u_i/u_b=0.2$, and it starts to increase as $u_i/u_b$ increases beyond 0.2. The reason for this increase will be discussed later.

Figure 5 depicts the profiles of the fluid relative mean velocity $\langle u _f\rangle$ with respect to the particle mean velocity $\langle u _p\rangle$ (i.e. mean slip velocity), where the angle brackets denote the phase averaging. It was observed experimentally and numerically that the mean velocity of heavy particles was smaller than the fluid counterpart in the bulk region, but larger than the fluid counterpart near the wall (Tsuji et al. Reference Tsuji, Morikawa and Shiomi1984; Shokri et al. Reference Shokri, Ghaemi, Nobes and Sanders2017; Zhu et al. Reference Zhu, Yu, Pan and Shao2020a). The reason for the larger particle mean velocity near the wall was attributed to the facts that the particles can slip on the wall and are preferentially located in the high-speed streaks (Tsuji et al. Reference Tsuji, Morikawa and Shiomi1984; Zhu et al. Reference Zhu, Yu, Pan and Shao2020a). Figure 5 shows negative slip velocities in the near-wall region at $u_i/u_b=0.159$ and 0.2. For $u_i /u_b \ge 0.25$, there are very few particles in the region of negative slip velocity due to the migration of the particles away from the wall (see figure 6); thus the expected negative slip velocities are not plotted. The position of zero mean slip velocity shifts towards the wall, as the settling coefficient increases. Two facts might be responsible for this observation. Firstly, the preferential distribution of the heavy particles in the high-speed streaks is caused by the entrainment of the large-scale streamwise vortices from the bulk region to the near-wall region under the condition of a higher particle concentration closer to the channel centre, as illustrated by Zhu et al. (Reference Zhu, Yu, Pan and Shao2020a) for $u_i/u_b=0.159$. At $u_i /u_b \ge 0.25$, shear-induced large-scale vortices are significantly suppressed (see figure 10) and the particles are distributed largely uniformly in the bulk region. Thus the preferential distribution of the heavy particles in the high-speed streaks only occurs in the near-wall region for high settling coefficients. Secondly, a larger settling coefficient itself means a larger positive slip velocity, and it also means a larger particle Reynolds number and thereby larger particle inertia. Consequently, at a larger $u_i/u_b$, particles with upward velocity smaller than the mean fluid velocity can approach closer to the wall.

Figure 5. Differences between the fluid and solid mean velocities at $a/H=0.05$, $\varPhi _0=2.36\,\%$, $Re_b=5746$ and $\rho _{r}=2.0$.

Figure 6. Profiles of the particle volume fraction at $a/H=0.05$, $\varPhi _0=2.36\,\%$, $Re_b=5746$ and $\rho _{r}=2.0$. The numerical result of Uhlmann (Reference Uhlmann2008) for $a/H=0.025$, $\varPhi _0= 0.42\,\%$ and $Re_p \approx 136$ is shown for comparison.

The mean particle volume fraction, $\varPhi _s/\varPhi _0$, as a function of the wall-normal distance $y/H$ is depicted in figure 6. The particle migration towards the channel (or pipe) centre was observed in numerical simulations (Zhu et al. Reference Zhu, Yu, Pan and Shao2020a,) and in experiments (Goes Oliveira, van der Geld & Kuerten Reference Goes Oliveira, van der Geld and Kuerten2017; Shokri et al. Reference Shokri, Ghaemi, Nobes and Sanders2017) when the particle settling effect was not strong, as a result of the Saffman effect (Saffman Reference Saffman1965; Auton Reference Auton1987; Auton, Hunt & Prud'Homme Reference Auton, Hunt and Prud'Homme1988). Costa et al. (Reference Costa, Brandt and Picano2020) showed that there exists a strong shear-induced lift force acting on the low-inertia particles in the near-wall region in turbulent channel flow and this lift force was well captured by the Saffman model, even though there was no gravity effect. As mentioned earlier, at $u_i /u_b \ge 0.25$ the particle concentration distribution is almost uniform in the bulk region, and a peak appears in the near-wall region. In fact, the peak position for $0.25 \le u_i /u_b \le 0.4$ is not so close to the wall, and it is around $y \approx 0.3H$ at $u_i/u_b = 0.25$. In the near-wall region, a high shear rate always exists, and a larger slip velocity causes a larger Saffman lift force, which leads to a very small particle volume fraction near the wall and the aforementioned peak. The peak position shifts closer to the wall as $u_i/u_b$ increases. The reason should be related to larger particle inertia at a higher settling coefficient.

It is interesting that the peak position is well correlated with the trough (local minimum) of the particle r.m.s. velocity component in the wall-normal direction, from the comparison between figure 6 and figure 9(b). Hence, the peak of the particle concentration distribution may also be related to turbophoresis, which drives the particles from the region of high turbulent intensity to that of low turbulent intensity (Caporaloni et al. Reference Caporaloni, Tampieri, Trombetti and Vittori1975; Reeks Reference Reeks1983). The mean momentum equation in the wall-normal (i.e. $y$) direction for the particle phase at the statistically steady state from the spatial averaging theorem (Crowe et al. Reference Crowe, Schwarzkopf, Sommerfeld and Tsuji2011) can be written as

(4.1)\begin{equation} {{\partial (\varPhi _s \langle {\sigma _s } \rangle_{yy} )} }/{{\partial y}} - {{\partial (\varPhi _s \rho _s \langle {v'_s v'_s } \rangle )} }/{{\partial y}} + \bar F_y = 0, \end{equation}

where $\sigma _s$ denotes the particle inner (Cauchy) stress and $\bar F_y$ is the volume-averaged hydrodynamic force on the particles in the $y$ direction. The particle r.m.s. velocity in the wall-normal direction (the second term in the above equation) plays a role in the force balance, similar to the fluid Reynolds stress for the fluid phase. It is also possible that the peak of the particle concentration results in the occurrence of the local minimum of the particle r.m.s. velocity.

The particle concentration distribution obtained by Uhlmann (Reference Uhlmann2008) for $a/H=0.025$ and $Re_p \approx 136$ is plotted in figure 6 for comparison. The two results for high slip velocities are qualitatively similar. The peak position of Uhlmann (Reference Uhlmann2008) is closer to the wall, because his particle size is smaller than ours. As shown in figure 19, the peak position is closer to the wall for smaller particles.

4.1.2. Fluctuating velocity and turbulent kinetic energy

The r.m.s. values of the fluid velocity fluctuations and the Reynolds shear stress for different settling coefficients are depicted in figure 7. The r.m.s. velocities do not change monotonically as $u_i/u_b$ increases from zero to 0.45. The addition of the particles decreases all r.m.s. velocity components across the channel at $u_i/u_b=0.159$. At $u_i/u_b \ge 0.2$, the streamwise r.m.s. velocity at the channel centre is enhanced, whereas the wall-normal and spanwise r.m.s. velocities are attenuated across the channel at $u_i/u_b \le 0.4$, and begin to exceed the single-phase values at the channel centre for $u_i/u_b \ge 0.45$. There is a local minimum in the profile of each r.m.s. velocity component for $0.25 \le u_i /u_b \le 0.4$ in figure 7, and we attribute the reason to the effect of the peak of the particle concentration distribution in figure 6. The Reynolds stress profiles for the particle-free and particle-laden cases are compared in figure 7($d$). The Reynolds stress decreases as $u_i/u_b$ increases up to around 0.25, and then it increases with increasing $u_i/u_b$, although it is still much smaller than the single-phase value at $u_i/u_b=0.45$.

Figure 7. Profiles of the r.m.s. velocity in single-phase and particle-laden turbulent channel flows: (a) streamwise, (b) wall-normal, (c) spanwise and (d) Reynolds shear stress $-\langle u'v'\rangle$ at $a/H=0.05$, $\varPhi _0=2.36\,\%$, $Re_b=5746$ and $\rho _{r}=2.0$.

Figure 8 shows the profiles of TKE and its relative change with respect to the single-phase flow. Turbulence suppression across the entire channel is observed for ${u_i}/{u_b} \le 0.25$. For ${u_i}/{u_b} > 0.25$, TKE augmentation in the bulk region and attenuation in the near-wall region can be observed in figure 8, as also observed in previous experiments (Tsuji et al. Reference Tsuji, Morikawa and Shiomi1984; Hosokawa & Tomiyama Reference Hosokawa and Tomiyama2004; Mena & Curtis Reference Mena and Curtis2020). As $u_i/u_b$ increases, the region for turbulence enhancement is expanded. At $u_i/u_b=0.5$, TKE is enhanced across the entire channel (not shown in figure 8 for convenience of presentation). From table 4, the relative increase in the total TKE at $u_i/u_b=0.5$ reaches $173.7\,\%$.

Figure 8. Profiles of (a) the TKE in single-phase and particle-laden turbulent channel flows and (b) the relative change of the TKE with respect to the single-phase flow, i.e. $(k-k_{{sp}})/k_{{sp}}$, at $a/H=0.05$, $\varPhi _0=2.36\,\%$, $Re_b=5746$ and $\rho _{r}=2.0$.

Table 4. The average fluid-phase r.m.s. velocity components, turbulence intensity $\tilde I$ and TKE $\tilde k$ in the entire channel. The values in parentheses represent the relative differences with respect to the single-phase flow. Here $Re_p$ is the particle Reynolds number based on mean inter-phase slip velocity at the channel centre at $a/H=0.05$, $\varPhi _0=2.36\,\%$, $Re_b=5746$ and $\rho _{r}=2.0$.

In table 4, the average fluid-phase r.m.s. velocity components, turbulence intensity $\tilde I$ and TKE $\tilde k$ in the entire channel for $u_i/u_b\le 0.5$ are presented. We see that the average r.m.s. velocities in all three directions, the mean turbulence intensity and the mean TKE are attenuated at $u_i/u_b \le 0.3$, with the maximum attenuation at $u_i/u_b=0.25$. The mean TKE is enhanced for $u_i/u_b \ge 0.4$, corresponding to $Re_p \ge 96.6$. At $u_i/u_b=0.4$ and 0.5, the increase in the mean TKE is caused by the increase in the streamwise r.m.s. velocity, since the wall-normal and spanwise r.m.s. velocities are reduced, compared to the single-phase flow. As $u_i/u_b$ exceeds 0.5, all components of the average r.m.s. velocities are augmented by the particles.

The solid-phase r.m.s. velocities and kinematic Reynolds shear stresses ($-\langle u'_pv'_p\rangle$ without the density) are plotted in figure 9. The streamwise solid-phase r.m.s. velocities are smaller than those of the fluid in the unladen case for $u_i/u_b < 0.35$, and the wall-normal r.m.s. velocities are larger than those of the fluid in the near-wall region due to the collision between the particles and the wall. The solid-phase r.m.s. velocities generally decrease with increasing $u_i/u_b$ for small $u_i/u_b$, and then increase with increasing $u_i/u_b$, consistent with the results of the fluid r.m.s. velocities in figure 7. As mentioned earlier, the trough (local minimum) in the profile of the wall-normal r.m.s. velocity is correlated with the peak of particle volume fraction in figure 6.

Figure 9. Profiles of the r.m.s. velocity and Reynolds shear stress for the particle phase: (a) streamwise, (b) wall-normal, (c) spanwise and (d) the particle kinematic Reynolds shear stress $-\langle u'_pv'_p\rangle$ at $a/H=0.05$, $\varPhi _0=2.36\,\%$, $Re_b=5746$ and $\rho _{r}=2.0$.

4.1.3. Vortex structure

Figure 10 shows the typical vortex structures in the half channel, identified with the $Q$ criterion ($Q=1.5$), which is defined as

(4.2)\begin{equation} Q=\tfrac{1}{2}(\varOmega_{i j} \varOmega_{i j}-\textit{\textsf{S}}_{i j} \textit{\textsf{S}}_{i j}), \end{equation}

where $\varOmega _{i j}=(u_{i,j}-u_{j,i}) / 2$ and $\textit {\textsf {S}}_{i j}=(u_{i,j} +u_{j,i}) / 2$ are the fluid rotation-rate tensor and the strain-rate tensor, respectively. Zhu et al. (Reference Zhu, Yu, Pan and Shao2020a) showed that the large-scale vortices were weakened more significantly at $u_i/u_b=0.159$ than at $u_i/u_b=0.092$. From figure 10(b), at $u_i/u_b=0.25$ with the particle Reynolds number being approximately 50, the large-scale vortices appear to be completely suppressed. The suppression of the large-scale vortices is clearly responsible for the pronounced attenuation of the turbulence intensity at relatively small settling coefficients, as observed in figures 7 and 8, and table 4.

Figure 10. Vortex structures of the (a) single-phase flow and (b,c) the particle-laden flows for (b) $u_i/u_b=0.25$ and (c) $u_i/u_b=0.45$ at $a/H=0.05$, $\varPhi _0=2.36\,\%$, $Re_b=5746$ and $\rho _{r}=2.0$. The colour of the vortices represents the fluid streamwise velocity, with the same scale for all three cases.

The reason for the suppression of the large-scale vortices is related to the effects of the drag force, the enhanced viscous dissipation, or the particle-induced wake structures. These three factors are interrelated, since a higher settling coefficient implies a larger drag force, a higher viscous dissipation and stronger particle wakes. It is not clear which effect is more important. Significant suppression of the large-scale vortices was observed in the numerical simulations based on the point-particle model (Vreman Reference Vreman2015; Muramulla et al. Reference Muramulla, Tyagi, Goswami and Kumaran2020). Vreman (Reference Vreman2015) demonstrated that the non-uniform part of the mean feedback drag force contributed significantly to particle-induced turbulence attenuation. Xia, Yu & Deng (Reference Xia, Yu and Deng2019) observed that, at the same settling coefficient, light (bubble) particles attenuated the turbulence in a downward channel flow, where the particles migrated towards the channel centre, whereas light particles enhanced the turbulence in an upward channel, where the particles migrated towards the channel wall. From these observations, it seems that the particle feedback force distributed in the bulk region suppresses the self-generation of large-scale vortices in the channel flow. On the other hand, the particle wakes tend to enhance the turbulence intensity when they are sufficiently strong. Figure 10 shows strong particle wakes at $u_i/u_b=0.45$, corresponding to $Re_p=113.5$, in which case the intensity of the particle-induced turbulence becomes stronger than that of large-scale-vortex-induced single-phase turbulence in the bulk region, as shown in figure 8. The stronger particle wakes at a higher $u_i/u_b$ in the near-wall region where the velocity gradient is large are expected to be responsible for the increase in the fluid Reynolds stress with increasing $u_i/u_b$ at $u_i/u_b>0.25$ in figure 7($d$).

4.2. Criterion for turbulence modulation

4.2.1. Effects of other parameters on TKE

We have shown that there is a transition from turbulence attenuation to augmentation as the settling coefficient increases for the typical case of $a/H=0.05$, $\varPhi _0=2.36\,\%$, $Re_b=5746$ and $\rho _{r}=2.0$ (i.e. group 5). The TKE of the particle-laden flow starts to exceed that of the single-phase flow at the channel centre at $u_i/u_b=0.3$($Re_p=64$) (figure 8), while the augmentation in the total TKE starts at $u_i/u_b=0.45$ ($Re_p=96.6$) (table 4). In the following, the effects of the bulk Reynolds number, the particle size, the particle volume fraction and the particle–fluid density ratio on the TKE and the critical particle Reynolds number are examined. We consider two bulk Reynolds numbers $Re_b=5746$ and 12 000, four particle sizes $a/H=0.05$, 0.75, 0.1 and 0.15, two particle volume fractions $\varPhi _0=0.84\,\%$ and $2.36\,\%$, and three density ratios 2, 10 and 100, to inspect the effects of the specific parameter, while keeping the other parameters constant, as shown in table 2. The maximum density ratio is 100, because at this density ratio the particle inertia is already considerably strong for the particle size $a/H=0.1$ (Yu et al. Reference Yu, Lin, Shao and Wang2017). In addition, to investigate the effect of the particle size at the same particle number density, the case of $a/H=0.05$ and $\varPhi _0=0.3\,\%$ (group 7) is added for comparison to the case of $a/H=0.1$ and $\varPhi _0=2.36\,\%$ (Group 1) at the same particle number of 360 in the channel cell.

Figure 11 shows the relative changes of the mean turbulence intensity $\widetilde {{CT}}$ in the entire channel and the turbulence intensity $CT$ at the channel centre as a function of $Re_p$. When the other parameters are kept constant, both critical particle Reynolds numbers increase with increasing particle size $a/H$, bulk Reynolds number $Re_b$ and particle–fluid density ratio $\rho _r$. Although both critical particle Reynolds numbers generally increase with decreasing particle volume fraction $\varPhi _0$, the dependence of the critical $Re_p$ on the particle volume fraction for the turbulence intensity at the channel centre is more pronounced than for the total turbulence intensity in the channel. From figure 11(a), the critical $Re_p$ for the total turbulence intensity is relatively insensitive to the particle volume fraction: its value for $\varPhi _0=0.84\,\%$ is slightly larger than that for $\varPhi _0=2.36\,\%$ at $a/H=0.1$, and almost the same for $\varPhi _0=0.3\,\%$ and $\varPhi _0=2.36\,\%$ at $a/H=0.05$. Nevertheless, the effect of the particle volume fraction on the TKE is significant: a higher particle volume fraction causes more enhancement in the augmentation regime or more reduction in the attenuation regime. In other words, the magnitude of the turbulence enhancement or attenuation increases with increasing particle volume fraction.

Figure 11. The relative change in (a) the mean turbulent intensity $\widetilde {{CT}}$ and (b) the local turbulent intensity at the channel centre $CT$ as a function of $Re_p$.

At a critical particle Reynolds number for the total TKE in the channel, the TKE is enhanced in the bulk region and attenuated in the near-wall region. Figure 12($a$) compares the TKE profiles for different particle volume fractions at comparable $Re_p$ near the critical value. One can see that the increase in the particle volume fraction leads to turbulence enhancement in the bulk region, but it leads to turbulence attenuation in the near-wall region. The amounts of TKE enhancement and attenuation are comparable. Therefore, the total TKE is not sensitive to the variation of the particle volume fraction near the critical state, resulting in the critical $Re_p$ for the total TKE being insensitive to the particle volume fraction. Since the increase in the particle volume fraction leads to the increase in the TKE at the channel centre near the critical state, the critical $Re_p$ for the turbulence augmentation at the channel centre decreases with increasing particle volume fraction.

Figure 12. The TKE profiles of single-phase and particle-laden turbulent channel flows for (a) different particle volume fractions and (b) different particle–fluid density ratios.

It can be observed in figure 12($b$) that, at comparable values of $Re_p$, the flow laden with particles with a larger density (or inertia) has a lower turbulence intensity in the bulk region, which is responsible for the larger critical particle Reynolds number at a larger density ratio.

Figure 11 indicates that the effect of the particle size on the critical particle Reynolds number is most pronounced, compared to those of the other parameters. Some TKE profiles for different particle sizes at the same volume fraction of $\varPhi _0=2.36\,\%$ are shown in figure 13($a$). The TKE in the bulk region decreases significantly with increasing particle size at comparable $Re_p$. Some TKE profiles at the same particle number density (360 particles in the channel domain) for $a/H=0.05$ and $a/H=0.1$ are depicted in figure 13($b$). One can see that the TKEs for $a/H=0.05$ and $\varPhi _0=0.3\,\%$ are still larger than those for $a/H=0.1$ and $\varPhi _0=2.36\,\%$ at comparable $Re_p$. We will explain this particle size effect later.

Figure 13. The TKE profiles of single-phase and particle-laden turbulent channel flows for different particle sizes at (a) the same particle volume fraction of $2.36\,\%$ and (b) the same particle number of $N_p=360$.

The TKE profiles of single-phase and particle-laden turbulent channel flows for $Re_b=5746$ and 12 000 at $a/H=0.05$ and $\varPhi _0=2.36\,\%$ are compared in figure 14. For $Re_b=5746$, the TKE of the particle-laden flow in the near-wall region is considerably attenuated and the large-scale vortices existing in the single-phase flow are completely suppressed at $Re_p=64$ (i.e. $u_i/u_b=0.3$) (see figure 10), and there is appreciable TKE augmentation in the bulk region at $Re_p=96.6$. By contrast, for $Re_b=12\,000$, the TKE of the particle-laden flow in the near-wall region is close to that of the single-phase flow and TKE enhancement takes place only near the channel centre at $Re_p=102.1$. This difference indicates that stronger particle wakes (or drag force) are needed to prevail over the large-scale vortices and result in the turbulence augmentation for a higher channel (or bulk) Reynolds number. This is understandable, considering that the large-scale vortices are stronger (despite being smaller in size) at a higher channel Reynolds number. The direct effect of the particle wakes on the TKE should be related to the particle-induced TKE production rate, and we will show that the particle-induced TKE production rate normalized with the bulk velocity decreases significantly with increasing $Re_b$ at the same particle Reynolds number from (4.14). Our effect of $Re_b$ is consistent with the experimental results of Mena & Curtis (Reference Mena and Curtis2020) on an upward pipe flow, who observed the change from augmentation in the total TKE at $Re_b=2 \times 10^5$ to attenuation at $Re_b=3.5 \times 10^5$.

Figure 14. The TKE profiles of single-phase and particle-laden turbulent channel flows for different bulk Reynolds numbers at $a/H=0.05$ and $\varPhi _0=2.36\,\%$.

4.2.2. Criterion for turbulence modulation

Based on our simulation data in figure 11, we propose a new criterion to distinguish between turbulence augmentation and attenuation in terms of the total TKE and the TKE at the channel centre, respectively. The parameter for the total TKE is

(4.3)\begin{equation} {\chi_{1}} = \frac{{R{e_p}}}{{Re_b^{\textrm{{0}}\textrm{{.33}}}{{( {{d_p / H}} )}^{\textrm{{0.61}}}}\rho _r^{0.05}}}. \end{equation}

We have shown that the particle volume fraction has little effect on the critical Reynolds number for the total TKE, hence it is not included in the new parameter $\chi _1$. The relative change of the mean turbulence intensity $\widetilde {CT}$ as a function of $\chi _1$ is depicted in figure 15($a$). It is observed that the attenuation of the total TKE occurs at $\chi _1<20$, and the augmentation takes place at $\chi _1>20$. We collect experimental data on upward pipe flows in the literature and the numerical data of Uhlmann (Reference Uhlmann2008) on upward channel flow, as summarized in table 5, and compare them with our criterion of $\chi _1=20$ in figure 15($b$). One can see that our criterion agrees well with the experimental data and the numerical data of Uhlmann (Reference Uhlmann2008). There is some discrepancy between our criterion and the recent experiment of Mena & Curtis (Reference Mena and Curtis2020). At the moment, the reason is not clear. The pipe Reynolds number of Mena & Curtis (Reference Mena and Curtis2020) ranges from $Re_b=2 \times 10^5$ to $Re_b=3.5 \times 10^5$, and is much higher than ours. Further studies on the particle–turbulence interactions at high bulk Reynolds numbers are needed.

Figure 15. Relative changes of the mean turbulence intensity in the channel $\widetilde {CT}$ as a function of ${\chi _1}$ for (a) the present numerical simulations and (b) previous experimental and numerical data.

Table 5. Summary of the previous experiments on upward pipe flow and the simulation of Uhlmann (Reference Uhlmann2008) on upward channel flow. Here $D$ is the diameter of the pipe and $D$ equals $2H$ for the channel flow.

The parameter of the criterion for the turbulence intensity at the channel centre is

(4.4)\begin{equation} {\chi _\textrm{{2}}} = \frac{{\varPhi_0^{0.1}}}{{Re_b^{0.53} {{( {{d_p / H}} )}^{\textrm{{0}}\textrm{{.61}}}}\rho _r^{0.065}}}R{e_p}. \end{equation}

Figure 16($a$) shows that a value of $\chi _2$ of approximately 1.55 can classify the turbulence augmentation and attenuation in our simulations. From figure 16($b$), our criterion of $\chi _2 = 1.55$ for the turbulence intensity at the channel centre also agrees well with the previous experiments.

Figure 16. Relative changes of the turbulence intensity at the channel centre $CT$ as a function of ${\chi _2}$ for (a) the present numerical simulations and (b) previous experimental and numerical data.

4.3. TKE budget and modelling of interfacial term

In this subsection, we perform a TKE budget analysis for further investigation of the effects of particles on the turbulence modulation. We also provide a preliminary modification on the model for the interfacial term in the TKE equation.

4.3.1. Turbulent kinetic energy budget

The volume-averaged fluid TKE equation for particle-laden flows can be derived as (Crowe et al. Reference Crowe, Schwarzkopf, Sommerfeld and Tsuji2011; Ma et al. Reference Ma, Santarelli, Ziegenhein, Lucas and Fröhlich2017; Vreman & Kuerten Reference Vreman and Kuerten2018; Peng et al. Reference Peng, Ayala and Wang2019)

(4.5)\begin{align} &\frac{\partial(\varPhi \rho k)}{\partial t}+ \frac{\partial(\varPhi \rho\langle u_{j}\rangle k)}{\partial x_{j}} \nonumber\\ &\quad=\underbrace{-\varPhi \rho\langle u_{i}^{\prime} u_{j}^{\prime}\rangle \frac{\partial\langle u_{i}\rangle}{\partial x_{j}}}_{\mathcal{P}} \underbrace{-\varPhi\left\langle\tau_{i j}^{\prime} \frac{\partial u_{i}^{\prime}}{\partial x_{j}}\right\rangle}_\epsilon \nonumber\\ & \qquad \underbrace{-\frac{\partial\left(\varPhi \rho\left\langle u_{j}^{\prime} \dfrac{u_{i}^{\prime} u_{i}^{\prime}}{2}\right\rangle\right.}{\partial x_{j}}- \frac{\partial(\varPhi\langle u_{j}^{\prime} p^{\prime}\rangle)}{\partial x_{j}}+\frac{\partial(\varPhi\langle u_{i}^{\prime} \tau_{i j}^{\prime} \rangle)}{\partial x_{j}}}_\mathcal{D} \nonumber\\ & \qquad +\underbrace{\varPhi(-\langle p\rangle \delta_{i j}+\langle\tau_{i j}\rangle)\left(\frac{\partial\langle u_{i}\rangle}{\partial x_{i}}-\left\langle\frac{\partial u_{i}}{\partial x_{j}}\right\rangle\right)}_{\mathcal{I} {\rm 1}} \nonumber\\ & \qquad \underbrace{-\frac{1}{V} \int_{S_{I}} n_{j} u_{i}({-}p \delta_{i j}+\tau_{i j}) \, \textrm{d} s+\frac{\langle u_{i}\rangle}{V} \int_{S_{I}}({-}p \delta_{i j}+\tau_{i j}) n_{j}\, \textrm{d} s}_{\mathcal{I} {\rm 2}}. \end{align}

Here, $\varPhi$, $\rho$ and $\tau _{i j}$ represent the fluid volume fraction, density and viscous stress, respectively. For convenience, the subscript ‘$f$’ denoting the fluid phase is omitted. The subscripts $i$ and $j$ denote the coordinate component, $V$ represents the volume for averaging, $S_I$ is the interface between the fluid and solid inside the volume $V$, and $n$ is the unit vector on the particle surface pointing to the fluid. The prime in (4.5) and other places below indicates the fluctuating part of the quantity.

For the channel flow at the statistically steady state, the statistics are homogeneous in the streamwise and spanwise directions and do not change with time. Thus the derivative is only non-zero in the wall-normal (i.e. $y$) direction, and the averaging volume $V$ is a thin band stretching over the streamwise and spanwise directions. In (4.5), ${\mathcal {P}}$ and $\epsilon$ represent the production due to the shear flow and the viscous dissipation of TKE; $\mathcal {D}$ denotes the diffusion terms including the turbulent and viscous diffusion terms. The terms ${\mathcal {P}}$, $\epsilon$ and $\mathcal {D}$ exist in the single-phase flow. The terms ${\mathcal {I}}1$ and ${\mathcal {I}}2$ exist only for the multiphase flows due to the presence of the interface between the two phases, and thus are called the interfacial terms or the particle source terms in the TKE equation (Crowe et al. Reference Crowe, Schwarzkopf, Sommerfeld and Tsuji2011; Ma et al. Reference Ma, Santarelli, Ziegenhein, Lucas and Fröhlich2017). We will show that the interfacial term can be approximated with two terms: the product of the mean inter-phase drag force and the mean slip velocity, and the correlation between the particle fluctuating velocity and the drag force on the particle.

Let $\sigma _{ij}$ denote the total stress $\sigma _{ij} = - p\delta _{ij} + \tau _{ij}$. From the spatial averaging theorem (Crowe et al. Reference Crowe, Schwarzkopf, Sommerfeld and Tsuji2011), one can derive

(4.6)\begin{align} \varPhi \left\langle {\frac{{\partial u_i } }{{\partial x_j }}} \right\rangle & = \frac{{\partial (\varPhi \langle {u_i } \rangle )} }{{\partial x_j }} - \frac{1 }{V}\int_{S_I } {n_j } u_i\, \textrm{d} s \nonumber\\ & = \varPhi \frac{{\partial \langle {u_i } \rangle } }{{\partial x_j }} + \frac{1 }{V}\int_{S_I } {n_j } \langle {u_i } \rangle \, \textrm{d} s - \frac{1 }{V}\int_{S_I } {n_j } u_i \, \textrm{d} s \nonumber\\ & = \varPhi \frac{{\partial \langle {u_i } \rangle } }{{\partial x_j }} - \frac{1 }{V}\int_{S_I } {n_j } u'_i \, \textrm{d} s. \end{align}

Then, the interfacial term ${\mathcal {I}}1$ can be written as

(4.7)\begin{equation} {\mathcal{I}}1 = \varPhi ( { - \langle \,p\rangle \delta _{ij} + \langle {\tau _{ij} } \rangle } )\left( {\frac{{\partial \langle {u_i } \rangle } }{{\partial x_i }} - \left\langle {\frac{{\partial u_i } }{{\partial x_j }}} \right\rangle } \right) = \frac{1 }{V}\int_{S_I } {\langle {\sigma _{ij} } \rangle n_j } u'_i \, \textrm{d} s. \end{equation}

The interfacial term ${\mathcal {I}}2$ can be written as

(4.8)\begin{equation} {\mathcal{I}}2 ={-} \frac{1 }{V}\int_{S_I } {\sigma _{ij} n_j } u'_i \, \textrm{d} s. \end{equation}

Thus, the total interfacial term has the form (Ma et al. Reference Ma, Santarelli, Ziegenhein, Lucas and Fröhlich2017; Vreman & Kuerten Reference Vreman and Kuerten2018)

(4.9)\begin{equation} {\mathcal{I}} = {\mathcal{I}}\textrm{{1}} + {\mathcal{I}}\textrm{{2}} ={-} \frac{1}{V}\int_{S_I } {\sigma '_{ij} n_j } u'_i \, \textrm{d} s. \end{equation}

Accurate computation of the interfacial term via the integration of the stress on the particle surface is difficult for the fictitious domain method, and here we obtain it from the TKE equation at the statistically steady state: ${\mathcal {I}}=-({\mathcal {P}}+\epsilon + \mathcal {D})$. Since the terms $\mathcal {P}$, $\epsilon$ and $\mathcal {D}$ are computed by averaging over time, and over the streamwise and spanwise directions (not over the wall-normal direction), the averaging volume can be regarded as a very thin band containing the plane $y=y_i$, where $y=y_i$ is the $y$ coordinate of the computational grid.

The profiles of the production, dissipation, diffusion and interfacial terms in the TKE equation for the case of $a/H=0.05$, $\varPhi _0=2.36\,\%$, $\rho _r=2.0$ and $Re_b=5746$ are depicted in figure 17. Similar to the behaviour of the Reynolds stress and the r.m.s. velocities in figure 7, the production rate of TKE ${\mathcal {P}}$ first decreases and then increases, as $u_i/u_b$ increases; ${\mathcal {P}}$ is attenuated by the particles for $u_i/u_b \le 0.45$, similar to the Reynolds stress in figure 7(b). In the bulk region, the particle wakes do not contribute significantly to the Reynolds stress and ${\mathcal {P}}$, since the mean fluid velocity there is nearly uniform at a high $u_i/u_b$. The reduction in ${\mathcal {P}}$ is clearly caused by the suppression of the large-scale vortices. For the present particle-laden flow where there is a mean interphase slip velocity, the interfacial term is always larger than zero and increases with increasing $u_i/u_b$, as shown in figure 17(c). Hence, the interfacial term can be regarded as the particle-induced TKE production. At a high $u_i/u_b$, the strong particle wakes cause a large TKE and accordingly a large viscous dissipation $\epsilon$. The large drag force at a high $u_i/u_b$ also leads to a large interfacial term ${\mathcal {I}}$. In the bulk region, the shear-induced production ${\mathcal {P}}$ and the diffusion $\mathcal {D}$ are small; therefore, the dissipation $\epsilon$ is nearly equal to the particle-induced production ${\mathcal {I}}$ in magnitude.

Figure 17. Profiles of the terms in the TKE equation for the typical case of $a/H=0.05$, $\varPhi _\textrm {{0}}=2.36\,\%$ and $Re_b=5746$: (a) the production term ${\mathcal {P}}$, (b) the dissipation term $\epsilon$, (c) the interfacial term ${\mathcal {I}}$ and (d) the diffusion term ${\mathcal {D}}$. All terms are normalized with ${{\rho _f}u_b^3}/H$.

The production, dissipation and interfacial terms for different particle sizes at the same particle number density and comparable particle Reynolds numbers are compared in figure 18($a$). It has been shown in figure 13($b$) that smaller particles can cause more augmentation in the TKE at the same particle number density and comparable particle Reynolds numbers. Figure 18($a$) shows the same behaviour for the interfacial and dissipation terms: they are larger for smaller particles at the same particle number density and Reynolds number. It should be noted that the interfacial term is larger for smaller particles ($a/H=0.05$) in the bulk region even if the local particle number density there is significantly smaller for smaller particles (because the particle concentration is higher in the near-wall region for smaller particles), as shown in figure 19. Figure 18($b$) compares the production, dissipation and interfacial terms for $\rho _r=2$ and 100 at $a/H=0.1$. Overall, the interfacial term is greater for the less dense particles with lower inertia. The reason should be related to the particle concentration distribution. For $\rho _r=2$ ($Re_p=163.8$, $a/H=0.1$), there is a peak at around $y=0.4H$ mainly as a result of the Saffman effect, which drives the particles away from the wall, whereas for $\rho _r=100$, the particle concentration distribution is more homogeneous across the channel. The higher particle concentration at around $y=0.4H$ for $\rho _r=2$ ($a/H=0.1$) in figure 19 should be responsible for its larger interfacial term at around $y=0.4H$ in figure 18($b$).

Figure 18. Production, dissipation and interfacial terms in the TKE equation, normalized with ${{\rho _f}u_b^3/H}$, for (a) different particle sizes $a/H$ at the same particle number density and (b) different particle–fluid density ratios $\rho _r$.

Figure 19. Profiles of the particle volume fraction for different particle sizes at the same particle number density and different particle–fluid density ratios.

The reason for the general decrease in the critical Reynolds number (particularly the one for the TKE at the channel centre) with decreasing particle size and density ratio and with increasing particle volume fraction should be related to the larger interfacial term (i.e. particle-induced TKE production) for a smaller particle size or density ratio, or a larger particle volume fraction at the same particle Reynolds number. However, it should be noted that, although the interfacial term ${\mathcal {I}}$ or the dissipation rate $\epsilon$ is generally correlated with the TKE, a higher ${\mathcal {I}}$ or $\epsilon$ does not always mean a higher TKE. For example, the TKE at the channel centre for $\rho _r=2$ and $Re_p=163.8$ is higher than that for $\rho _r=100$ and $Re_p=169.9$, as shown in figure 12($b$), but ${\mathcal {I}}$ and $\epsilon$ at the channel centre are almost the same for the two cases, as shown in 18($b$). Hence, the dependence of the interfacial term on the parameters can be regarded as the main reason for the dependence of the critical particle Reynolds number on the corresponding parameters, but it is not the only reason. As for the effect of the density ratio, another reason may be that less dense particles have stronger hydrodynamic interactions and consequently may have larger variations in the instantaneous slip velocity at the channel centre where the flow is dominated by the particle wakes, when the mean interphase slip velocity is the same.

4.3.2. Analysis of the interfacial term in the dilute limit

We here attempt to provide an explanation for the effects of the particle size and the bulk Reynolds number on the TKE at the same particle Reynolds number by simplifying the interfacial term in the dilute limit. The interfacial term ${\mathcal {I}}1$ is not important, since the average of the velocity gradient should be close to the gradient of the average velocity, as shown in the calculation of Peng et al. (Reference Peng, Ayala and Wang2019). Following the analysis of Crowe et al. (Reference Crowe, Schwarzkopf, Sommerfeld and Tsuji2011), we neglect the effects of the particle rotation. We further assume that all particle surfaces are inside the volume $V$, which is reasonable for a homogeneous system such as particle sedimentation in the bulk region. The fluid velocity on the surface for a non-rotating particle $k$ is equal to the particle translational velocity, which is a constant for the integration on the particle surface. Then, the interfacial term ${\mathcal {I}}2$ can be simplified as

(4.10)\begin{equation} - \frac{1}{V}\int_{{S_I}} {{u_i}( - p{\delta _{ij}} + {\tau _{ij}})} \, \textrm{d} s + \frac{{\langle {{u_i}} \rangle }}{V}\int_{{S_I}} {( - p{\delta _{ij}} + {\tau _{ij}}){n_j}} \, \textrm{d} s = \frac{1}{V}\sum_k {(\langle {{u_i}} \rangle - {u_{pki}}){F_{ki}}}, \end{equation}

where ${F_{ki}}$ is the drag force on the particle $k$ and $u_{pki}$ is the translational velocity of the particle $k$.

For vertical channel flow, the drag force is mainly directed in the streamwise direction and thus we neglect the contributions from the other two directions:

(4.11)\begin{equation} \frac{1}{V}\sum_k {(\langle {{u_i}} \rangle - {u_{pki}}){F_{ki}}} \approx \frac{1}{V}\sum_k {(\langle u_f \rangle - {u_{pk}})F_k}, \end{equation}

where $u_f$, $u_{pk}$ and $F_k$ represent the fluid velocity, the particle translational velocity and the drag force on the particle $k$ in the streamwise direction, respectively. Considering that ${\langle u_f \rangle - {u_{pk}}}={\langle {{u_f}} \rangle - \langle {{u_p}} \rangle - u{'_{pk}}}$ (with $u{'_{pk}}$ being the particle velocity fluctuation), we have

(4.12)\begin{equation} {\mathcal{I}} \approx \frac{1}{V}\sum_k {(\langle {{u_f}} \rangle - \langle {{u_p}} \rangle - u{'_{pk}}){F_k}} = (\langle {{u_f}} \rangle - \langle {{u_p}} \rangle )\frac{1}{V}\sum_k {{F_k} - \frac{1}{V}\sum_k {u{'_{pk}}{F_k}}}.\end{equation}

The first term on the right-hand side is the product of the mean inter-phase drag force and the mean slip velocity, and was regarded as the main contribution to the interfacial term (Troshko & Hassan Reference Troshko and Hassan2001; Crowe et al. Reference Crowe, Schwarzkopf, Sommerfeld and Tsuji2011).

For particle-laden flows in the dilute limit, one may assume that each particle settles with the same slip velocity $u_r=\langle {{u_f}} \rangle - \langle {{u_p}} \rangle$, and then

(4.13)\begin{equation} (\langle {{u_f}} \rangle - \langle {{u_p}} \rangle ) \frac{1}{V}\sum_k {{F_k}}= n\frac{{u_r^3{C_D}{\rm \pi} {a^2}{\rho _f}}}{2} = n\frac{Re_p^3{\nu ^3}{C_D}{\rm \pi} {\rho _f}}{{16a}},\end{equation}

in which $n = \varPhi _s /(4{\rm \pi} {a^3}/3)$ is the particle number density and $C_D$ is the standard drag coefficient. The interfacial term in our simulations is normalized with ${{\rho _f}u_b^3/H}$. From (4.13), one has

(4.14)\begin{equation} \frac{(\langle u_f \rangle - \langle u_p \rangle)\displaystyle\frac{1}{V} \sum\nolimits_k F_k} { \rho _f u_b^3/H} \sim n^*\frac{{Re_p^3{C_D}}}{Re_b^3(a/H)} \sim \varPhi_s\frac{{Re_p^3{C_D}}}{Re_b^3(a/H)^4},\end{equation}

where $n^*$ denotes the dimensionless particle number density. The drag coefficient $C_D$ is a function of the particle Reynolds number alone.

Hence, the above dimensionless interfacial term is inversely proportional to the particle size at the same particle number density $n$, $Re_p$ and $Re_b$, which provides an explanation for the early observations that the interface term is larger for smaller particles at the same particle Reynolds number and particle number density in figure 18($a$) and the critical particle Reynolds number increases significantly with increasing particle size in figure 11. Equation (4.14) also explains why the dimensionless TKE in the bulk region is significantly smaller for a higher bulk Reynolds number at comparable particle Reynolds numbers in figure 14 and why the critical particle Reynolds number is higher at a higher bulk Reynolds number in figure 11.

4.3.3. Modelling of the interfacial term

Modelling of the interfacial term is crucial for the Reynolds-averaged Navier–Stokes (RANS) model and the Reynolds stress model (i.e. second-order moment model) for the simulations of industrial-scale multiphase turbulent flows (Ansys 2013; Ma et al. Reference Ma, Santarelli, Ziegenhein, Lucas and Fröhlich2017, Reference Ma, Lucas, Jakirlić and Fröhlich2020). The Troshko–Hassan model (Troshko & Hassan Reference Troshko and Hassan2001) is one of the widely used models for the interfacial term:

(4.15)\begin{equation} {\mathcal{I}} = C_n \bar F{u_r}, \end{equation}

where $\bar F$ is the volume-averaged drag force, $\bar F = ({1 }/{V})\sum _k {F_k }$, and $C_n$ is a coefficient. In the Ansys FLUENT software, a value of 0.75 is recommended for $C_n$ (Ansys 2013). Ma et al. (Reference Ma, Santarelli, Ziegenhein, Lucas and Fröhlich2017) determined $C_n$ as a function of $Re_p$, from the interface-resolved DNS data of Santarelli, Roussel & Fröhlich (Reference Santarelli, Roussel and Fröhlich2016) on the bubbly flow in a vertical channel:

(4.16)\begin{equation} C_n = 0.18Re_p^{0.23}.\end{equation}

In the following, we attempt to evaluate $C_n$ from our DNS data. Following Ma et al. (Reference Ma, Santarelli, Ziegenhein, Lucas and Fröhlich2017), the drag force is determined directly from the simulations, rather than from the drag model. For the channel flow at the statistically stationary state, the fluid mean momentum equation in the streamwise direction can be written as follows (Yu et al. Reference Yu, Lin, Shao and Wang2017):

(4.17)\begin{equation} \frac{\textrm{{d}}}{{\textrm{{d}}y}}\left( {{\varPhi _f}\mu \frac{{\textrm{{d}}\langle {{u_f}} \rangle }}{{\textrm{d} y}}} \right) + {\varPhi _f}\left( { - \frac{{\textrm{{d}}{p_e}}}{{\textrm{{d}}x}}} \right) + \frac{\textrm{{d}}}{{\textrm{{d}}y}}( {{\varPhi _f}{\rho _f}\langle { - {{u'}_f}{{v'}_f}} \rangle } ) - \bar F = 0. \end{equation}

For upward flow, one can compute the pressure gradient from the force balance,

(4.18)\begin{equation} -\textrm{d} p_{e}/\textrm{d} x=\tau_{w}/H+\varPhi_{0} (\rho_{s}-\rho_{f})g, \end{equation}

where $\tau _{w}$ is the mean shear stress on the sidewalls. Then, the volume-averaged drag force can be calculated from (4.17). The calculation of the drag force at the channel centre can be simplified, if the flow is homogeneous there (i.e. $\textrm {d}(\,\cdot \,)/{\textrm {d} y}=0$). The solid-phase mean momentum equation in the streamwise direction is (Yu et al. Reference Yu, Lin, Shao and Wang2017)

(4.19)\begin{equation} \frac{\textrm{d}}{{\textrm{d} y}}(\varPhi_s \langle {\sigma _s } \rangle _{xy}) + \varPhi _s \left( - \frac{{\textrm{d}p_e } }{{\textrm{d} x}}\right) - \varPhi _s (\rho _s - \rho _f )g + \frac{\textrm{d}}{{\textrm{d} y}}(\varPhi _s \rho_s \langle { - {{u'}_s}{{v'}_s}} \rangle ) + \bar F = 0. \end{equation}

For a homogeneous flow, the fluid momentum equation (4.17) reduces to $\bar F = \varPhi _f ( - {{\textrm {d}p_e } }/{{\textrm {d} x}})$, and the solid-phase momentum equation (4.19) reduces to $\bar F = \varPhi _s (\rho _s - \rho _f )g - \varPhi _s ( - {{\textrm {d}p_e } }/{{\textrm {d} x}})$. From these two equations, one can obtain the pressure gradient and the drag force as

(4.20)\begin{equation} \bar F = \varPhi_f \left( - \frac{{\textrm{d}p_e } }{{\textrm{d} x}}\right) = \varPhi _f \varPhi _s (\rho _s - \rho _f )g, \end{equation}

which was used by Ma et al. (Reference Ma, Santarelli, Ziegenhein, Lucas and Fröhlich2017) for the calculation of the drag force. Equation (4.17) is used in the present study, because, for $\rho _r=100$, the gradient of the particle Reynolds stress at the channel centre in (4.19) is found to be not negligibly small. For $\rho _r=2$, the results obtained with the two methods are close to each other for relatively high particle Reynolds numbers.

The mean $C_n$ values for the channel centre region $- 0.1 \le y/H \le 0.1$ (here $y=0$ corresponding to the channel centre) as a function of $Re_p$ for different particle sizes, different volume fractions and different density ratios are plotted in figure 20($a$). When we ran the cases for the study of the criterion on the TKE augmentation and attenuation, the data for the TKE budget were not computed. Owing to our limited computer resources, only some cases were selected for the generation of the data for the modelling of the interfacial term. Figure 20($a$) shows that the coefficient $C_n$ decreases with increasing particle Reynolds number for the same group of control parameters, and is higher for a lower particle volume fraction or a larger particle size at the same particle Reynolds number. The effect of the density ratio is small. The fitting of our data yields

(4.21)\begin{equation} {C_n}=2.1 Re_p^{ - 0.09}{\left( {\frac{a}{H}} \right)^{0.12}}\rho _r^{0.007}{({1 - {\varPhi_s}} )^{8.0}}. \end{equation}

Figure 20($b$) shows that the above model of $C_n$ fits our data well.

Figure 20. Plots of (a) $C_n$ versus $Re_p$ for different particle sizes, particle volume fractions and density ratios, and (b) ${C_n}=2.1 Re_p^{-0.09}{({{a}/{H}})^{0.12}}\rho _r^{0.007}{({1 - {\varPhi _s}})^{8.0}}$ as a function of $Re_p$.

The dependence of our $C_n$ on $Re_p$ appears to be inconsistent with that of Ma et al. (Reference Ma, Santarelli, Ziegenhein, Lucas and Fröhlich2017). Our $C_n$ (4.21) decreases with increasing $Re_p$, whereas theirs (4.16) increases with increasing $Re_p$. However, our dependence of $C_n$ on $Re_p$ is examined at the same particle size and particle volume fraction, whereas theirs is examined at different particle sizes and particle volume fractions. Their results are consistent with ours in that $C_n$ is larger for larger particles and a lower particle volume fraction. Thus, we suspect that the dependence of their $C_n$ on $Re_p$ is caused by the effects of the particle size and the particle volume fraction, rather than $Re_p$. In addition, we note that they used the data on bubbly flow at $Re_p>230$, whereas we consider solid particles at $Re_p<230$.

The reasons for the dependence of $C_n$ on the particle Reynolds number, the particle size, the particle volume fraction and the density ratio are not clear, but should be related to the effects of the second term on the right-hand side in (4.12), i.e. the particle velocity fluctuation term: $({1}/{V})\sum _k {u{'_{pk}}{F_k}}$. If the effect of the particle rotation is not important and (4.12) is a good approximation to the interfacial term, then $C_n \approx 1 - ({1 }/{V})\sum _k {u'_{pk} F_k /} (u_r \bar F)$. Particles may be accelerated by the particle wakes, resulting in the positive correlation between the particle fluctuating velocity and the force on the particle. An increase in the particle Reynolds number or the particle volume fraction, or a decrease in the particle size or the density ratio, may cause a larger particle velocity fluctuation and the increase in $({1 }/{V})\sum _k {u'_{pk} F_k /} (u_r \bar F)$, i.e. the decrease in $C_n$. A more significant effect of the density ratio on $C_n$ than that shown in figure 20($a$) was expected from this reasoning. In fact, if the drag force is calculated with (4.20), instead of (4.17), the difference in $C_n$ for $\rho _r=100$ and $\rho _r=2$ would be more pronounced. Using (4.17), $C_n$ can be determined from:

(4.22)\begin{equation} \frac{{\mathcal{I}} }{{\rho _f u_b^3 /H}} = \frac{{C_n \varPhi _f \varPhi _s (\rho _s - \rho _f )gu_r H} }{{\rho _f u_b^3 }} = C_n \varPhi _f \varPhi _s \frac{3}{{8(a/H)}}\left(\frac{{u_i } }{{u_b }}\right)^2 \frac{{u_r}}{{u_b}}. \end{equation}

Now we compare two cases: $(a/H,\rho _r,u_i/u_b)=(0.1,2,0.3)$ in group 1 and $(a/H, \rho _r,u_i/u_b)=(0.1,100,0.25)$ in group 3. From table 6, the particle Reynolds numbers for these two cases are $Re_p=163.8$ and $Re_p=169.9$, respectively. From figures 18($b$) and 19, the interfacial term and the particle volume fraction at the channel centre for these two cases are nearly the same. The mean slip velocities are also close to each other, since the particle Reynolds numbers are close. The relative difference in $C_n$ is around $16\,\%$, since the settling coefficients $u_i/u_b$ are 0.3 and 0.275, respectively. As mentioned earlier, the effect of the particle Reynolds stress or the shear flow on the drag force even at the channel centre cannot be neglected for $\rho _r=100$ in the range of $Re_p$ studied. When gravity is not considered, i.e. $u_i/u_b=0$, there is still a pronounced mean drag force on the particles at the channel centre for $\rho _r=100$ (Yu et al. Reference Yu, Lin, Shao and Wang2017), which is mainly balanced with the gradient of the particle Reynolds stress.

Table 6. Results for the simulation cases: groups 1–5. Pressure gradients $(-\mathrm {d} p_{e} / \mathrm {d} x)$ are normalized by $\rho _{f} u_{b}^{2} / H$.

Only a preliminary work on the modelling of the interfacial term in the TKE equation is conducted here. We intend to do further studies on the models for the near-wall region and lower particle Reynolds numbers in the future. The drag and lift models, the RANS model and the second-order model for particle-laden turbulent flows based on the interface-resolved DNS data are good subjects for extensive future work.

4.4. Flow friction and total drag

In this subsection, we will explore the effects of the particle sedimentation on the wall friction and total flow drag. The summation of the fluid momentum equation (4.17) and the solid-phase momentum equation (4.19) yields (Picano et al. Reference Picano, Breugem and Brandt2015; Yu et al. Reference Yu, Lin, Shao and Wang2017)

(4.23)\begin{align} &\frac{\textrm{d}}{\textrm{d} y}(\varPhi_{f}\langle\sigma_{f}\rangle_{x y}+\varPhi_{s}\langle\sigma_{s}\rangle_{x y})+\left(-\frac{\textrm{d} p_{e}}{\textrm{d} x}\right) \nonumber\\ &\quad +\frac{\textrm{d}}{\textrm{d} y}(\varPhi_{f} \rho_{f}\langle-u_{f}^{\prime} v_{f}^{\prime}\rangle+\varPhi_{s} \rho_{s}\langle-u_{s}^{\prime} v_{s}^{\prime}\rangle)-\varPhi_{s}(\rho_{s}-\rho_{f}) g=0. \end{align}

Substituting (4.18) into (4.23) yields

(4.24)\begin{align} -\frac{{{\tau_w}}}{H} &= \frac{{\textrm{d}}}{{{\textrm{d}} y}}( {{\varPhi _f}{{\langle {{\sigma _f}} \rangle }_{xy}} + {\varPhi _s}{{\langle {{\sigma _s}} \rangle }_{xy}}} ) + ({\varPhi _0} - {\varPhi _s})( {{\rho _s} - {\rho _f}} )g \nonumber\\ &\quad + \frac{{\textrm{d}} }{{{\textrm{d}} y}}({{\varPhi _f}{\rho _f}\langle { - u{'_f}v{'_f}} \rangle+{\varPhi_s}{\rho_s}\langle{-u{'_s}v{'_s}}\rangle}). \end{align}

The friction coefficient is defined as $C_f=2\tau _{w}/\rho _f u_{b}^{2}$. Applying triple integrations (i.e. $\int _{0}^{1}\textrm {d}y \int _{0}^{y}\textrm {d}y \int _{0}^{y}\textrm {d}y$) to (4.24), one obtains (Fukagata, Iwamoto & Kasagi Reference Fukagata, Iwamoto and Kasagi2002; Zhu et al. Reference Zhu, Yu, Pan and Shao2020a):

(4.25)\begin{align} C_f&= \underbrace {\frac{6}{Re_b} \int_0^1 (1-y) \varPhi _f \left(\frac{\textrm{d}u^*_f}{\textrm{d} y}\right)\, \textrm{d} y} _{C_{fv}} +\underbrace{{6} \int_0^1 (1-y) (\varPhi _f \langle -{u_f^*}'{u_f^*}'\rangle )\, \textrm{d} y} _{C_{fR}} \nonumber\\ &\quad +\underbrace{{6} \int_0^1 (1-y) \frac{\varPhi_s}{\rho_f u_b^2} \langle \sigma_s \rangle _{xy}\,\textrm{d} y}_{C_{pI}} +\underbrace{{6} \int_0^1 (1-y) (\varPhi _s \rho _r\langle -{u_s^*}'{u_s^*}'\rangle )\, \textrm{d} y}_{C_{pR}}\nonumber\\ &\quad + \underbrace{\left({-3} \int_0^1 (1-y)^2 (\varPhi_s-\varPhi_0)(\rho_r-1)Fr\,\textrm{d} y\right) }_{C_\textrm{g}}, \end{align}

where $C_{fV}$, $C_{fR}$, $C_{pI}$, $C_{pR}$ and $C_g$ are the contributions of the fluid viscous stress, the fluid Reynolds stress, the particle inner stress, the particle Reynolds stress and gravity, respectively. Here, $y$ is the dimensionless coordinate normalized with $H$, with the wall position shifted to $y = 0$, and $u^*$ and $v^*$ denote the dimensionless velocity components normalized by $u_b$, and $Fr$ represents the Froude number defined by $Fr=gH /u_{b}^{2}$. Note that the particle Reynolds stress is calculated by using the velocities at the Eulerian grids inside the particle regions rather than the particle translational velocities. To derive (4.25), the following relations are used:

(4.26)\begin{gather} {( {\varPhi _f \langle {\sigma _f } \rangle _{xy} + \varPhi _s \langle {\sigma _s } \rangle _{xy} } )} |_{y = 0} = \tau _w, \end{gather}
(4.27)\begin{gather}\int_0^1 {\int_0^y {f\, \textrm{d} y\, \textrm{d} y = } } \int_0^1 {(1 - y)} \,f\, \textrm{d} y, \quad \int_0^1 {\int_0^y {\int_0^y {f\, \textrm{d} y\, \textrm{d} y\, \textrm{d} y} = } } \frac{1}{2}\int_0^1 {(1 - y)} ^2 f\, \textrm{d} y. \end{gather}

Each contribution, normalized with the friction coefficient of the single-phase case, is plotted in figure 21 for the different settling coefficients. The single-phase results (denoted by SP) are also shown for comparison. Because the particle volume fraction is small ($2.36\,\%$), the contribution from the particle stress is small for $\rho _r=2$. However, it is significant for $\rho _r=100$. Another notable difference between the two density ratios is that, as $u_i/u_b$ increases, the contribution from the fluid Reynolds stress for $\rho _r=2$ first decreases and then increases, whereas it is almost independent of $u_i/u_b$ for $\rho _r=100$. For the particle-laden flows, generally the wall friction first decreases and then increases as $u_i/u_b$ increases. The decrease in $C_f$ for $\rho _r=2$ is caused by the decrease in the fluid Reynolds stress, and the subsequent increase in $C_f$ is mainly due to the increase in the gravity term $C_g$ for $(a/H,\rho _r,Re_b)=(0.05,2,5746)$ and $(a/H,\rho _r,Re_b)=(0.1,100,5746)$, while it is mainly due to the increase in the fluid Reynolds stress for $(a/H,\rho _r,Re_b)=(0.1,2,5746)$ and $(a/H,\rho _r,Re_b)=(0.05,2,12\,000)$. The attenuation in the fluid Reynolds stress for $(a/H,\rho _r,Re_b)=(0.05,2,5746)$ is most pronounced, compared to the other cases, because the particle-induced TKE production (or the effect of the particle wakes) is more significant for smaller particles or a lower bulk Reynolds number, as discussed earlier in association with (4.14). The contributions from the gravity term at relatively large settling coefficients for $(a/H,\rho _r,Re_b)=(0.05,2,5746)$ are also most pronounced, because the gravity term is larger for smaller particles, as implied by $(\rho _r - 1)Fr = 3(u_i/u_b)^2/(8a/H)$ from (2.25), and the particle concentration distribution is more homogeneous at a higher bulk Reynolds number, which means that $\varPhi _s$ is closer to $\varPhi _0$.

Figure 21. Contributions of the fluid viscous stress $C_{fV}$, the fluid Reynolds stress $C_{fR}$, the particle total stress $C_{pR}+C_{pI}$ and the gravity term $C_g$ to the wall friction $C_f$ at different $u_i/u_b$ for: (a) group 1, $(a/H,\rho _r,Re_b)=(0.1,2,5746)$; (b) group 3, $(a/H,\rho _r,Re_b)=(0.1,100,5746)$; (c) group 5, $(a/H,\rho _r,Re_b)=(0.05,2,5746)$; and (d) group 6, $(a/H,\rho _r,Re_b)=(0.05,2,12\,000)$. The results are normalized by $C_f$ for the single-phase case. SP denotes the single-phase case, rather than the particle-laden flow at $u_i/u_b=0$.

Figure 21 shows that there is a reduction in the wall friction at intermediate settling coefficients, compared to the single-phase flow. However, this does not mean drag reduction, since the total flow drag is normally measured with the pressure drop at the same flow rate. The pressure gradients for all cases studied are presented in tables 6 and 7. The total flow drag increases monotonically with increasing $u_i/u_b$ without any drag reduction.

Table 7. Results for the simulation cases: groups 6–9.

5. Conclusion

Interface-resolved direct numerical simulations (IR-DNS) of upward turbulent channel flows laden with finite-size spherical particles have been performed, and the modulation of the turbulence intensity by heavy particles has been investigated at $Re_b=5746$ and 12 000, $a/H=0.05$, 0.075, 0.1 and 0.15, $\rho _r=2$, 10 and 100, $\varPhi _0=0.3\,\%$, $0.84\,\%$ and $2.36\,\%$, and $Re_p < 227$. From our results, the following conclusions can be drawn.

  1. (i) At low $Re_p$, the turbulent intensity across the channel is all diminished; at intermediate $Re_p$, the turbulent intensity is enhanced in the channel centre region and attenuated in the near-wall region; and at sufficiently large $Re_p$, the turbulent intensity is enhanced across the channel. The particles attenuate turbulence by suppressing large-scale vortices, which is related to the effect of the drag force distributed in the bulk region; while they enhance turbulence by generating wake vortices, which is related to the interfacial term (i.e. the particle-induced production rate) in the TKE equation. The critical $Re_p$ increases with increasing bulk Reynolds number, particle size and particle–fluid density ratio, while increasing with decreasing particle volume fraction, particularly for the channel centre region. The effects of the bulk Reynolds number, particle size and particle volume fraction on the TKE and critical $Re_p$ can be explained by the dependence of the interfacial term on these parameters (4.14).

  2. (ii) The augmentation of the total TKE in the channel occurs at a value of ${\chi _{1}} = R{e_p}/[Re_b^{\textrm {{0}}\textrm {{.33}}}{{({{d_p / H}} )}^{\textrm {{0.61}}}} \rho _r^{0.05}] > 20$, while attenuation occurs at ${\chi _{1}} < 20$. The augmentation of the TKE at the channel centre occurs at ${\chi _\textrm {{2}}} = {{\varPhi _0^{0.1}R{e_p}}}/[Re_b^{0.53}$ ${{( {{d_p / H}} )}^{0.61}}\rho _r^{0.065}] > 1.55$, while attenuation occurs at ${\chi _{2}} < 1.55$. These two criteria are in good agreement with the previous experimental data.

  3. (iii) The coefficient in the Troshko–Hassan model for the interfacial term is fitted from our DNS data as ${C_n}=2.1 Re_p^{ - 0.09}{( {{a}/{H}} )^{0.12}}\rho _r^{0.007}{({1 - {\varPhi _s}} )^{8.0}}$. The effects of the parameters on $C_n$ may be explained by the effect of the parameters on the particle velocity fluctuation.

  4. (iv) For the particle-laden flows, the wall friction first decreases and then increases as $u_i/u_b$ increases. The decrease in the wall friction for $\rho _r=2$ is caused by the decrease in the fluid Reynolds stress, and the subsequent increase in $C_f$ is due to the increase in the gravity term $C_g$ or in the fluid Reynolds stress. The attenuation in the fluid Reynolds stress is more pronounced for smaller particles or a lower bulk Reynolds number, since the particle-induced TKE production is more significant for smaller particles or a lower bulk Reynolds number at comparable $u_i/u_b$. The effect of the gravity term is also more significant for smaller particles or a lower bulk Reynolds number at comparable $u_i/u_b$. The contribution to the wall friction from the fluid Reynolds stress for $\rho _r=2$ first decreases and then increases as $u_i/u_b$ increases, whereas it is almost independent of $u_i/u_b$ for $\rho _r=100$.

  5. (v) At relatively low particle Reynolds numbers, the particles migrate towards the channel centre due to the Saffman effect, whereas at relatively high particle Reynolds numbers, the fluid mean velocity and the particle concentration distribution are almost uniform in the bulk region, and a peak of the particle concentration distribution appears in the near-wall region. The peak position shifts closer to the wall as the particle Reynolds number increases. The peak position is well correlated with the trough (local minimum) of the particle r.m.s. velocity component in the wall-normal direction, as well as the fluid r.m.s. velocity components.

We believe that the present work enhances significantly the understanding of the modulation of the TKE by the spherical particles. However, it is not clear whether our criteria are valid for parameters far beyond the range studied, such as the situation at high bulk Reynolds numbers. The particle modulation of the turbulence intensity in a wider parameter range for upward channel flow and for other flows such as downward channel flow and jet flows is worth studying. The drag and lift models, the RANS model and the second-order model for particle-laden turbulent flows based on the IR-DNS data are good subjects for extensive future work.

Funding

This work was supported by the National Natural Science Foundation of China (Grant Nos 91752117 and 12072319 for Z.Y., Grant No. 11632016 for J.L., and Grant Nos 11872333 and 91852205 for Y.G.).

Declaration of interest

The authors report no conflict of interest.

References

REFERENCES

Ansys, Inc. 2013 ANSYS FLUENT Theory Guide. Release 15.0. Canonsburg.Google Scholar
Ardekani, M.N., Costa, P., Breugem, W.-P., Picano, F. & Brandt, L. 2017 Drag reduction in turbulent channel flow laden with finite-size oblate spheroids. J. Fluid Mech. 816, 4370.CrossRefGoogle Scholar
Auton, T.R. 1987 The lift force on a spherical body in a rotational flow. J. Fluid Mech. 183, 199218.CrossRefGoogle Scholar
Auton, T.R., Hunt, J.C.R. & Prud'Homme, M. 1988 The force exerted on a body in inviscid unsteady non-uniform rotational flow. J. Fluid Mech. 197, 241257.CrossRefGoogle Scholar
Balachandar, S. & Eaton, J.K. 2010 Turbulent dispersed multiphase flow. Annu. Rev. Fluid Mech. 42, 111133.CrossRefGoogle Scholar
Barnocky, G. & Davis, R.H. 1988 Elastohydrodynamic collision and rebound of spheres: experimental verification. Phys. Fluids 31 (6), 13241329.CrossRefGoogle Scholar
Bellani, G., Byron, M.L., Collignon, A.G., Meyer, C.R. & Variano, E.A. 2012 Shape effects on turbulent modulation by large nearly neutrally buoyant particles. J. Fluid Mech. 712, 4160.CrossRefGoogle Scholar
Biegert, E., Vowinckel, B. & Meiburg, E. 2017 A collision model for grain-resolving simulations of flows over dense, mobile, polydisperse granular sediment beds. J. Comput. Phys. 340, 105127.CrossRefGoogle Scholar
Brändle de Motta, J.C., Breugem, W.-P., Gazanion, B., Estivalezes, J.-L., Vincent, S. & Climent, E. 2013 Numerical modelling of finite-size particle collisions in a viscous fluid. Phys. Fluids 25 (8), 083302.CrossRefGoogle Scholar
Caporaloni, M., Tampieri, F., Trombetti, F. & Vittori, O. 1975 Transfer of particles in nonisotropic air turbulence. J. Atmos. Sci. 32 (3), 565568.2.0.CO;2>CrossRefGoogle Scholar
Clift, R., Grace, J.R. & Weber, M.E. 2005 Bubbles, Drops, and Particles. Courier Corporation.Google Scholar
Costa, P., Boersma, B.J., Westerweel, J. & Breugem, W.-P. 2015 Collision model for fully resolved simulations of flows laden with finite-size particles. Phys. Rev. E 92 (5), 053012.CrossRefGoogle ScholarPubMed
Costa, P., Brandt, L. & Picano, F. 2020 Interface-resolved simulations of small inertial particles in turbulent channel flow. J. Fluid Mech. 883, A54.CrossRefGoogle Scholar
Costa, P., Picano, F., Brandt, L. & Breugem, W.-P. 2018 Effects of the finite particle size in turbulent wall-bounded flows of dense suspensions. J. Fluid Mech. 843, 450478.CrossRefGoogle Scholar
Crowe, C.T. 2000 On models for turbulence modulation in fluid–particle flows. Intl J. Multiphase Flow 26 (5), 719727.CrossRefGoogle Scholar
Crowe, C.T., Schwarzkopf, J.D., Sommerfeld, M. & Tsuji, Y. 2011 Multiphase Flows with Droplets and Particles. CRC Press.CrossRefGoogle Scholar
Eaton, J.K. 2009 Two-way coupled turbulence simulations of gas-particle flows using point-particle tracking. Intl J. Multiphase Flow 35 (9), 792800.CrossRefGoogle Scholar
Elghobashi, S. & Truesdell, G.C. 1993 On the two-way interaction between homogeneous turbulence and dispersed solid particles. I: turbulence modification. Phys. Fluids A: Fluid Dyn. 5 (7), 17901801.CrossRefGoogle Scholar
Fornari, W., Formenti, A., Picano, F. & Brandt, L. 2016 The effect of particle density in turbulent channel flow laden with finite size particles in semi-dilute conditions. Phys. Fluids 28 (3), 033301.CrossRefGoogle Scholar
Fornari, W., Kazerooni, H.T., Hussong, J. & Brandt, L. 2018 Suspensions of finite-size neutrally buoyant spheres in turbulent duct flow. J. Fluid Mech. 851, 148186.CrossRefGoogle Scholar
Fukagata, K., Iwamoto, K. & Kasagi, N. 2002 Contribution of Reynolds stress distribution to the skin friction in wall-bounded flows. Phys. Fluids 14 (11), L73L76.CrossRefGoogle Scholar
Gao, H., Li, H. & Wang, L.-P. 2013 Lattice Boltzmann simulation of turbulent flow laden with finite-size particles. Comput. Maths Applics. 65 (2), 194210.CrossRefGoogle Scholar
Garcia-Villalba, M., Kidanemariam, A.G. & Uhlmann, M. 2012 DNS of vertical plane channel flow with finite-size particles: Voronoi analysis, acceleration statistics and particle-conditioned averaging. Intl J. Multiphase Flow 46, 5474.CrossRefGoogle Scholar
Glowinski, R., Pan, T.-W., Hesla, T.I. & Joseph, D.D. 1999 A distributed lagrange multiplier/fictitious domain method for particulate flows. Intl J. Multiphase Flow 25 (5), 755794.CrossRefGoogle Scholar
Goes Oliveira, J.L., van der Geld, C.W.M. & Kuerten, J.G.M. 2017 Concentration and velocity statistics of inertial particles in upward and downward pipe flow. J. Fluid Mech. 822, 640663.CrossRefGoogle Scholar
Gondret, P., Lance, M. & Petit, L. 2002 Bouncing motion of spherical particles in fluids. Phys. Fluids 14 (2), 643652.CrossRefGoogle Scholar
Gore, R.A. & Crowe, C.T. 1989 Effect of particle size on modulating turbulent intensity. Intl J. Multiphase Flow 15 (2), 279285.CrossRefGoogle Scholar
Hertz, H.R. 1882 Uber die beruhrung fester elastischer korper und uber die harte. In Verhandlung des Vereins zur Beforderung des GewerbefleiBes, Berlin p. 449. Duncker.CrossRefGoogle Scholar
Hetsroni, G. 1989 Particles-turbulence interaction. Intl J. Multiphase Flow 15 (5), 735746.CrossRefGoogle Scholar
Hosokawa, S. & Tomiyama, A. 2004 Turbulence modification in gas–liquid and solid–liquid dispersed two-phase pipe flows. Intl J. Heat Fluid Flow 25 (3), 489498.CrossRefGoogle Scholar
Hosokawa, S. & Tomiyama, A. 2009 Multi-fluid simulation of turbulent bubbly pipe flows. Chem. Engng Sci. 64 (24), 53085318.CrossRefGoogle Scholar
Hoyas, S. & Jiménez, J. 2008 Reynolds number effects on the Reynolds-stress budgets in turbulent channels. Phys. Fluids 20 (10), 101511.CrossRefGoogle Scholar
Jeffrey, D.J. 1982 Low-Reynolds-number flow between converging spheres. Mathematika 29 (1), 5866.CrossRefGoogle Scholar
Kajishima, T., Takiguchi, S., Hamasaki, H. & Miyake, Y. 2001 Turbulence structure of particle-laden flow in a vertical plane channel due to vortex shedding. JSME Intl J. Ser. B Fluids Therm. Engng 44 (4), 526535.CrossRefGoogle Scholar
Kempe, T. & Fröhlich, J. 2012 Collision modelling for the interface-resolved simulation of spherical particles in viscous fluids. J. Fluid Mech. 709, 445489.CrossRefGoogle Scholar
Kiger, K.T. & Pan, C. 2002 Suspension and turbulence modification effects of solid particulates on a horizontal turbulent channel flow. J. Turbul. 3 (19), 117.CrossRefGoogle Scholar
Kulick, J.D., Fessler, J.R. & Eaton, J.K. 1994 Particle response and turbulence modification in fully developed channel flow. J. Fluid Mech. 277, 109134.CrossRefGoogle Scholar
Kussin, J. & Sommerfeld, M. 2002 Experimental studies on particle behaviour and turbulence modification in horizontal channel flow with different wall roughness. Exp. Fluids 33 (1), 143159.CrossRefGoogle Scholar
Lee, S.L. & Durst, F. 1982 On the motion of particles in turbulent duct flows. Intl J. Multiphase Flow 8 (2), 125146.CrossRefGoogle Scholar
Li, Y., McLaughlin, J.B., Kontomaris, K. & Portela, L. 2001 Numerical simulation of particle-laden turbulent channel flow. Phys. Fluids 13 (10), 29572967.CrossRefGoogle Scholar
Lin, Z.-W., Shao, X.-M., Yu, Z.-S. & Wang, L.-P. 2017 b Effects of finite-size heavy particles on the turbulent flows in a square duct. J. Hydrodyn. 29 (2), 272282.CrossRefGoogle Scholar
Lin, Z., Yu, Z., Shao, X. & Wang, L.-P. 2017 a Effects of finite-size neutrally buoyant particles on the turbulent flows in a square duct. Phys. Fluids 29 (10), 103304.CrossRefGoogle Scholar
Liu, C., Tang, S., Shen, L. & Dong, Y. 2017 Characteristics of turbulence transport for momentum and heat in particle-laden turbulent vertical channel flows. Acta Mechanica Sin. 33 (5), 833845.CrossRefGoogle Scholar
Lucci, F., Ferrante, A. & Elghobashi, S. 2010 Modulation of isotropic turbulence by particles of Taylor length-scale size. J. Fluid Mech. 650, 555.CrossRefGoogle Scholar
Luo, K., Luo, M. & Fan, J. 2016 On turbulence modulation by finite-size particles in dilute gas-solid internal flows. Powder Technol. 301, 12591263.CrossRefGoogle Scholar
Ma, T., Lucas, D., Jakirlić, S. & Fröhlich, J. 2020 Progress in the second-moment closure for bubbly flow based on direct numerical simulation data. J. Fluid Mech. 883, A9.CrossRefGoogle Scholar
Ma, T., Santarelli, C., Ziegenhein, T., Lucas, D. & Fröhlich, J. 2017 Direct numerical simulation–based Reynolds-averaged closure for bubble-induced turbulence. Phys. Rev. Fluids 2 (3), 034301.CrossRefGoogle Scholar
Maeda, M., Hishida, K. & Furutani, T. 1980 Velocity distributions of air-solids suspension in upward pipe flow: effect of particles on air velocity distribution. In Transactions of the Japan Society of Mechanical Engineers, Series B, vol. 46 (412), pp. 2313–2320. Japan Society of Mechanical Engineers.CrossRefGoogle Scholar
Maxey, M. 2017 Simulation methods for particulate flows and concentrated suspensions. Annu. Rev. Fluid Mech. 49, 171193.CrossRefGoogle Scholar
Mena, S.E. & Curtis, J.S. 2020 Experimental data for solid–liquid flows at intermediate and high stokes numbers. J. Fluid Mech. 883, A24.CrossRefGoogle Scholar
Mindlin, R.D. 1953 Elastic spheres in contact under varying oblique forces. J. Appl. Mech. 20, 327344.Google Scholar
Muramulla, P., Tyagi, A., Goswami, P.S. & Kumaran, V. 2020 Disruption of turbulence due to particle loading in a dilute gas–particle suspension. J. Fluid Mech. 889, A28.CrossRefGoogle Scholar
Noguchi, K. & Nezu, I. 2009 Particle–turbulence interaction and local particle concentration in sediment-laden open-channel flows. J. Hydro-Environ. Res. 3 (2), 5468.CrossRefGoogle Scholar
Parthasarathy, R.N. & Faeth, G.M. 1990 Turbulence modulation in homogeneous dilute particle-laden flows. J. Fluid Mech. 220, 485514.CrossRefGoogle Scholar
Peng, C., Ayala, O.M. & Wang, L.-P. 2019 A direct numerical investigation of two-way interactions in a particle-laden turbulent channel flow. J. Fluid Mech. 875, 10961144.CrossRefGoogle Scholar
Peng, C. & Wang, L.-P. 2019 Direct numerical simulations of turbulent pipe flow laden with finite-size neutrally buoyant particles at low flow Reynolds number. Acta Mechanica 230 (2), 517539.CrossRefGoogle Scholar
Petersen, A.J., Baker, L. & Coletti, F. 2019 Experimental study of inertial particles clustering and settling in homogeneous turbulence. J. Fluid Mech. 864, 925970.CrossRefGoogle Scholar
Picano, F., Breugem, W.-P. & Brandt, L. 2015 Turbulent channel flow of dense suspensions of neutrally buoyant spheres. J. Fluid Mech. 764, 463487.CrossRefGoogle Scholar
Reeks, M.W. 1983 The transport of discrete particles in inhomogeneous turbulence. J. Aerosol Sci. 14 (6), 729739.CrossRefGoogle Scholar
Righetti, M. & Romano, G.P. 2004 Particle–fluid interactions in a plane near-wall turbulent flow. J. Fluid Mech. 505, 93121.CrossRefGoogle Scholar
Saber, A., Lundström, T.S. & Hellström, J.G.I. 2015 Turbulent modulation in particulate flow: a review of critical variables. Engineering 7 (10), 597.CrossRefGoogle Scholar
Saffman, P.G.T. 1965 The lift on a small sphere in a slow shear flow. J. Fluid Mech. 22 (2), 385400.CrossRefGoogle Scholar
Saito, I., Watanabe, T. & Gotoh, T. 2019 A new time scale for turbulence modulation by particles. J. Fluid Mech. 880, R6.CrossRefGoogle Scholar
Santarelli, C. & Fröhlich, J. 2015 Direct numerical simulations of spherical bubbles in vertical turbulent channel flow. Intl J. Multiphase Flow 75, 174193.CrossRefGoogle Scholar
Santarelli, C., Roussel, J. & Fröhlich, J. 2016 Budget analysis of the turbulent kinetic energy for bubbly flow in a vertical channel. Chem. Engng Sci. 141, 4662.CrossRefGoogle Scholar
Sato, Y. & Hishida, K. 1996 Transport process of turbulence energy in particle-laden turbulent flow. Intl J. Heat Fluid Flow 17 (3), 202210.CrossRefGoogle Scholar
Schreck, S. & Kleis, S.J. 1993 Modification of grid-generated turbulence by solid particles. J. Fluid Mech. 249, 665688.CrossRefGoogle Scholar
Serizawa, A., Kataoka, I. & Michiyoshi, I. 1975 Turbulence structure of air-water bubbly flow-II. Local properties. Intl J. Multiphase Flow 2 (3), 235246.CrossRefGoogle Scholar
Shao, X., Wu, T. & Yu, Z. 2012 Fully resolved numerical simulation of particle-laden turbulent flow in a horizontal channel at a low Reynolds number. J. Fluid Mech. 693, 319344.CrossRefGoogle Scholar
Shokri, R., Ghaemi, S., Nobes, D.S. & Sanders, R.S. 2017 Investigation of particle-laden turbulent pipe flow at high-Reynolds-number using particle image/tracking velocimetry (piv/ptv). Intl J. Multiphase Flow 89, 136149.CrossRefGoogle Scholar
Squires, K.D. & Eaton, J.K. 1990 Particle response and turbulence modification in isotropic turbulence. Phys. Fluids A: Fluid Dyn. 2 (7), 11911203.CrossRefGoogle Scholar
Suzuki, Y., Ikenoya, M. & Kasagi, N. 2000 Simultaneous measurement of fluid and dispersed phases in a particle-laden turbulent channel flow with the aid of 3-D PTV. Exp. Fluids 29 (1), S185S193.CrossRefGoogle Scholar
Tanaka, T. & Eaton, J.K. 2008 Classification of turbulence modification by dispersed spheres using a novel dimensionless number. Phys. Rev. Lett. 101 (11), 114502.CrossRefGoogle ScholarPubMed
Ten Cate, A., Derksen, J.J., Portela, L.M. & Van Den Akker, H.E.A. 2004 Fully resolved simulations of colliding monodisperse spheres in forced isotropic turbulence. J. Fluid Mech. 519, 233271.CrossRefGoogle Scholar
Tenneti, S. & Subramaniam, S. 2014 Particle-resolved direct numerical simulation for gas-solid flow model development. Annu. Rev. Fluid Mech. 46, 199230.CrossRefGoogle Scholar
Theofanous, T.G. & Sullivan, J. 1982 Turbulence in two-phase dispersed flows. J. Fluid Mech. 116, 343362.CrossRefGoogle Scholar
Troshko, A.A. & Hassan, Y.A. 2001 A two-equation turbulence model of turbulent bubbly flows. Intl J. Multiphase Flow 27 (11), 19652000.CrossRefGoogle Scholar
Tsuji, Y. & Morikawa, Y. 1982 LDV measurements of an air—solid two-phase flow in a horizontal pipe. J. Fluid Mech. 120, 385409.CrossRefGoogle Scholar
Tsuji, Y., Morikawa, Y. & Shiomi, H. 1984 LDV measurements of an air-solid two-phase flow in a vertical pipe. J. Fluid Mech. 139, 417434.CrossRefGoogle Scholar
Tsuji, Y., Tanaka, T. & Ishida, T. 1992 Lagrangian numerical simulation of plug flow of cohesionless particles in a horizontal pipe. Powder Technol. 71 (3), 239250.CrossRefGoogle Scholar
Uhlmann, M. 2008 Interface-resolved direct numerical simulation of vertical particulate channel flow in the turbulent regime. Phys. Fluids 20 (5), 053305.CrossRefGoogle Scholar
Vreman, A.W. 2015 Turbulence attenuation in particle-laden flow in smooth and rough channels. J. Fluid Mech. 773, 103136.CrossRefGoogle Scholar
Vreman, A.W. & Kuerten, J.G.M. 2018 Turbulent channel flow past a moving array of spheres. J. Fluid Mech. 856, 580632.CrossRefGoogle Scholar
Wang, B. 2010 Inter-phase interaction in a turbulent, vertical channel flow laden with heavy particles. Part II: two-phase velocity statistical properties. Intl J. Heat Mass Transfer 53 (11-12), 25222529.CrossRefGoogle Scholar
Wang, G., Abbas, M. & Climent, É. 2017 Modulation of large-scale structures by neutrally buoyant and inertial finite-size particles in turbulent Couette flow. Phys. Rev. Fluids 2 (8), 084302.CrossRefGoogle Scholar
Wang, G., Fong, K.O., Coletti, F., Capecelatro, J. & Richter, D.H. 2019 Inertial particle velocity and distribution in vertical turbulent channel flow: a numerical and experimental comparison. Intl J. Multiphase Flow 120, 103105.CrossRefGoogle Scholar
Wang, L.-P. & Maxey, M.R. 1993 Settling velocity and concentration distribution of heavy particles in homogeneous isotropic turbulence. J. Fluid Mech. 256, 2768.CrossRefGoogle Scholar
Wang, L.-P., Peng, C., Guo, Z. & Yu, Z. 2016 Flow modulation by finite-size neutrally buoyant particles in a turbulent channel flow. J. Fluids Engng 138 (4), 041306.CrossRefGoogle Scholar
Wu, T.-H., Shao, X.-M. & Yu, Z.-S. 2011 Fully resolved numerical simulation of turbulent pipe flows laden with large neutrally-buoyant particles. J. Hydrodyn. 23 (1), 2125.CrossRefGoogle Scholar
Xia, Y., Xiong, H., Yu, Z. & Zhu, C. 2020 a Effects of the collision model in interface-resolved simulations of particle-laden turbulent channel flows. Phys. Fluids 32 (10), 103303.CrossRefGoogle Scholar
Xia, Y., Yu, Z. & Deng, J. 2019 A fictitious domain method for particulate flows of arbitrary density ratio. Comput. Fluids 193, 104293.CrossRefGoogle Scholar
Xia, Y., Yu, Z. & Guo, Y. 2020 b Interface-resolved numerical simulations of particle-laden turbulent channel flows with spanwise rotation. Phys. Fluids 32 (1), 013303.CrossRefGoogle Scholar
Yu, Z., Lin, Z., Shao, X. & Wang, L.-P. 2016 b A parallel fictitious domain method for the interface-resolved simulation of particle-laden flows and its application to the turbulent channel flow. Engng Appl. Comput. Fluid Mech. 10 (1), 160170.Google Scholar
Yu, Z., Lin, Z., Shao, X. & Wang, L.-P. 2017 Effects of particle-fluid density ratio on the interactions between the turbulent channel flow and finite-size particles. Phys. Rev. E 96 (3), 033102.CrossRefGoogle ScholarPubMed
Yu, Z., Phan-Thien, N. & Tanner, R.I. 2004 Dynamic simulation of sphere motion in a vertical tube. J. Fluid Mech. 518, 6193.CrossRefGoogle Scholar
Yu, Z. & Shao, X. 2007 A direct-forcing fictitious domain method for particulate flows. J. Comput. Phys. 227 (1), 292314.CrossRefGoogle Scholar
Yu, W., Vinkovic, I. & Buffat, M. 2016 a Finite-size particles in turbulent channel flow: quadrant analysis and acceleration statistics. J. Turbul. 17 (11), 10481071.CrossRefGoogle Scholar
Yu, Z., Zhu, C., Wang, Y. & Shao, X. 2019 Effects of finite-size neutrally buoyant particles on the turbulent channel flow at a Reynolds number of 395. Appl. Maths Mech. 40 (2), 293304.CrossRefGoogle Scholar
Zade, S., Lundell, F. & Brandt, L. 2019 Turbulence modulation by finite-size spherical particles in newtonian and viscoelastic fluids. Intl J. Multiphase Flow 112, 116129.CrossRefGoogle Scholar
Zhao, L.H., Andersson, H.I. & Gillissen, J.J.J. 2010 Turbulence modulation and drag reduction by spherical particles. Phys. Fluids 22 (8), 081702.CrossRefGoogle Scholar
Zhao, L., Andersson, H.I. & Gillissen, J.J.J. 2013 Interphasial energy transfer and particle dissipation in particle-laden wall turbulence. J. Fluid Mech. 715, 3259.CrossRefGoogle Scholar
Zhu, H.-Y., Pan, C., Wang, J.-J., Liang, Y.-R. & Ji, X.-C. 2019 Sand-turbulence interaction in a high-Reynolds-number turbulent boundary layer under net sedimentation conditions. Intl J. Multiphase Flow 119, 5671.CrossRefGoogle Scholar
Zhu, C., Yu, Z., Pan, D. & Shao, X. 2020 a Interface-resolved direct numerical simulations of the interactions between spheroidal particles and upward vertical turbulent channel flows. J. Fluid Mech. 891, A6.CrossRefGoogle Scholar
Zhu, C., Yu, Z. & Shao, X. 2018 Interface-resolved direct numerical simulations of the interactions between neutrally buoyant spheroidal particles and turbulent channel flows. Phys. Fluids 30 (11), 115103.CrossRefGoogle Scholar
Zhu, C., Yu, Z., Shao, X. & Deng, J. 2020 b Interface-resolved numerical simulations of particle-laden turbulent flows in a vertical channel filled with Bingham fluids. J. Fluid Mech. 883, A43.CrossRefGoogle Scholar
Figure 0

Table 1. Some criteria for turbulence enhancement or attenuation in the literature.

Figure 1

Figure 1. Schematic diagram of the upward channel flow laden with particles; with $x$, $y$ and $z$ representing the streamwise, transverse and spanwise coordinates, respectively.

Figure 2

Table 2. Parameter settings for the particle-laden flows in a vertical channel: $a$ represents the radius of particles, $N_p$ is the number of particles and $\varPhi _\textrm {{0}}$ is the mean particle volume fraction in the entire channel.

Figure 3

Figure 2. Profiles of the mean TKE production, dissipation and diffusion terms for the particle-free turbulent channel flow, normalized with $(u_{\tau })^{4} / \nu$, as compared to the results of Hoyas & Jiménez (2008).

Figure 4

Table 3. Parameter settings for the bouncing motion of a sphere colliding with a planar wall in a viscous fluid.

Figure 5

Figure 3. Trajectories of a sphere colliding with a planar wall in a viscous fluid for (a) $St_c$ = 27 and (b) $St_c$ = 152, compared to the experimental results of Gondret et al. (2002). Here $\zeta _{n}$ is the shortest distance between the sphere surface and the wall, and $t_{ref}=\sqrt {d_{p}/g}$ denotes the reference time scale.

Figure 6

Figure 4. Mean fluid velocity profiles of turbulent channel flows for different $u_i/u_b$ at $a/H=0.05$, $\varPhi _0=2.36\,\%$, $Re_b=5746$ and $\rho _{r}=2.0$. The inset shows the close-up of the velocity profiles near the wall.

Figure 7

Figure 5. Differences between the fluid and solid mean velocities at $a/H=0.05$, $\varPhi _0=2.36\,\%$, $Re_b=5746$ and $\rho _{r}=2.0$.

Figure 8

Figure 6. Profiles of the particle volume fraction at $a/H=0.05$, $\varPhi _0=2.36\,\%$, $Re_b=5746$ and $\rho _{r}=2.0$. The numerical result of Uhlmann (2008) for $a/H=0.025$, $\varPhi _0= 0.42\,\%$ and $Re_p \approx 136$ is shown for comparison.

Figure 9

Figure 7. Profiles of the r.m.s. velocity in single-phase and particle-laden turbulent channel flows: (a) streamwise, (b) wall-normal, (c) spanwise and (d) Reynolds shear stress $-\langle u'v'\rangle$ at $a/H=0.05$, $\varPhi _0=2.36\,\%$, $Re_b=5746$ and $\rho _{r}=2.0$.

Figure 10

Figure 8. Profiles of (a) the TKE in single-phase and particle-laden turbulent channel flows and (b) the relative change of the TKE with respect to the single-phase flow, i.e. $(k-k_{{sp}})/k_{{sp}}$, at $a/H=0.05$, $\varPhi _0=2.36\,\%$, $Re_b=5746$ and $\rho _{r}=2.0$.

Figure 11

Table 4. The average fluid-phase r.m.s. velocity components, turbulence intensity $\tilde I$ and TKE $\tilde k$ in the entire channel. The values in parentheses represent the relative differences with respect to the single-phase flow. Here $Re_p$ is the particle Reynolds number based on mean inter-phase slip velocity at the channel centre at $a/H=0.05$, $\varPhi _0=2.36\,\%$, $Re_b=5746$ and $\rho _{r}=2.0$.

Figure 12

Figure 9. Profiles of the r.m.s. velocity and Reynolds shear stress for the particle phase: (a) streamwise, (b) wall-normal, (c) spanwise and (d) the particle kinematic Reynolds shear stress $-\langle u'_pv'_p\rangle$ at $a/H=0.05$, $\varPhi _0=2.36\,\%$, $Re_b=5746$ and $\rho _{r}=2.0$.

Figure 13

Figure 10. Vortex structures of the (a) single-phase flow and (b,c) the particle-laden flows for (b) $u_i/u_b=0.25$ and (c) $u_i/u_b=0.45$ at $a/H=0.05$, $\varPhi _0=2.36\,\%$, $Re_b=5746$ and $\rho _{r}=2.0$. The colour of the vortices represents the fluid streamwise velocity, with the same scale for all three cases.

Figure 14

Figure 11. The relative change in (a) the mean turbulent intensity $\widetilde {{CT}}$ and (b) the local turbulent intensity at the channel centre $CT$ as a function of $Re_p$.

Figure 15

Figure 12. The TKE profiles of single-phase and particle-laden turbulent channel flows for (a) different particle volume fractions and (b) different particle–fluid density ratios.

Figure 16

Figure 13. The TKE profiles of single-phase and particle-laden turbulent channel flows for different particle sizes at (a) the same particle volume fraction of $2.36\,\%$ and (b) the same particle number of $N_p=360$.

Figure 17

Figure 14. The TKE profiles of single-phase and particle-laden turbulent channel flows for different bulk Reynolds numbers at $a/H=0.05$ and $\varPhi _0=2.36\,\%$.

Figure 18

Figure 15. Relative changes of the mean turbulence intensity in the channel $\widetilde {CT}$ as a function of ${\chi _1}$ for (a) the present numerical simulations and (b) previous experimental and numerical data.

Figure 19

Table 5. Summary of the previous experiments on upward pipe flow and the simulation of Uhlmann (2008) on upward channel flow. Here $D$ is the diameter of the pipe and $D$ equals $2H$ for the channel flow.

Figure 20

Figure 16. Relative changes of the turbulence intensity at the channel centre $CT$ as a function of ${\chi _2}$ for (a) the present numerical simulations and (b) previous experimental and numerical data.

Figure 21

Figure 17. Profiles of the terms in the TKE equation for the typical case of $a/H=0.05$, $\varPhi _\textrm {{0}}=2.36\,\%$ and $Re_b=5746$: (a) the production term ${\mathcal {P}}$, (b) the dissipation term $\epsilon$, (c) the interfacial term ${\mathcal {I}}$ and (d) the diffusion term ${\mathcal {D}}$. All terms are normalized with ${{\rho _f}u_b^3}/H$.

Figure 22

Figure 18. Production, dissipation and interfacial terms in the TKE equation, normalized with ${{\rho _f}u_b^3/H}$, for (a) different particle sizes $a/H$ at the same particle number density and (b) different particle–fluid density ratios $\rho _r$.

Figure 23

Figure 19. Profiles of the particle volume fraction for different particle sizes at the same particle number density and different particle–fluid density ratios.

Figure 24

Figure 20. Plots of (a) $C_n$ versus $Re_p$ for different particle sizes, particle volume fractions and density ratios, and (b) ${C_n}=2.1 Re_p^{-0.09}{({{a}/{H}})^{0.12}}\rho _r^{0.007}{({1 - {\varPhi _s}})^{8.0}}$ as a function of $Re_p$.

Figure 25

Table 6. Results for the simulation cases: groups 1–5. Pressure gradients $(-\mathrm {d} p_{e} / \mathrm {d} x)$ are normalized by $\rho _{f} u_{b}^{2} / H$.

Figure 26

Figure 21. Contributions of the fluid viscous stress $C_{fV}$, the fluid Reynolds stress $C_{fR}$, the particle total stress $C_{pR}+C_{pI}$ and the gravity term $C_g$ to the wall friction $C_f$ at different $u_i/u_b$ for: (a) group 1, $(a/H,\rho _r,Re_b)=(0.1,2,5746)$; (b) group 3, $(a/H,\rho _r,Re_b)=(0.1,100,5746)$; (c) group 5, $(a/H,\rho _r,Re_b)=(0.05,2,5746)$; and (d) group 6, $(a/H,\rho _r,Re_b)=(0.05,2,12\,000)$. The results are normalized by $C_f$ for the single-phase case. SP denotes the single-phase case, rather than the particle-laden flow at $u_i/u_b=0$.

Figure 27

Table 7. Results for the simulation cases: groups 6–9.