settling of finite-size particles in an ambient fluid: a ......8th international conference on...

10
8 th International Conference on Multiphase Flow ICMF 2013, Jeju, Korea, May 26 - 31, 2013 1 Settling of finite-size particles in an ambient fluid: A Numerical Study Todor Doychev 1 , Markus Uhlmann 1 1 Institute for Hydromechanics, Karlsruhe Institute of Technology, 76128 Karlsruhe, Germany Keywords: particulate flow, resolved particles, DNS, immersed boundary method, clusters, Voronoї analysis Abstract We have investigated the gravity induced settling of finite-size particles in an ambient fluid by means of direct numerical simulation (DNS). Such configurations are relevant to a large number of applications such as meteorology, mechanical and environmental engineering. For single particle a variety of motion patterns exist, from straight vertical to fully chaotic paths, for which the fluid motion in the near field around the particles and the particle wake play a dominant role. Here, the interface between the dispersed- and carrier-phase was fully resolved by means of an immersed boundary method. Particular care has been taken to meet the respective resolution requirements. We have performed simulations of particles settling with two different Galileo numbers, = 121 and = 178. The settling regimes for a single particle settling with this Galileo numbers are: (a) steady axisymmetric regime and (b) steady oblique regime. We observed, that in the steady oblique regime ( = 178) the particles exhibit wake induced clustering, while in the steady axisymmetric regime ( = 121) this was not observed. Furthermore, the mean settling velocity of the particles with = 178 was strongly enhanced compared to the velocity of a single settling particle. 1. Introduction Particle-laden flows are found in a large number of environmental natural and technical processes. Examples include pollution dispersion in the atmosphere, raindrop formation in clouds, sediment transport, fluidized bed reactors and combustion devices. Thus, the accurate prediction of such flows is of great importance. This naturally leads to the necessity of reliable theoretical description of the mechanisms that take place in such flows. Despite the progress made in the past, there is still a large scatter of the available data. A recent review of the subject can be found in (Balachandar and Eaton 2010). Here we consider the sedimentation of finite-size particles in an (initially) ambient fluid under the influence of gravity. Under such conditions, the system is characterized by a set of non-dimensional parameters. Given the fluid density , the kinematic fluid viscosity , the vector of gravitational acceleration on the one hand, and the particle diameter , particle density and solid volume fraction on the other hand, dimensional analysis shows that the problem is determined by three non-dimensional parameters. One has already been mentioned, the solid volume fraction . The other two can be taken as the density ratio / and the Galileo number defined as the ratio between the gravity-buoyancy and the viscosity forces = (|| 3 | / − 1| ) 1/2 /. The particles under consideration in the present work have a particle Galileo (Reynolds) number of the order of (100). For single particle, the Galileo number and the particle-to-fluid density ratio characterize the regime of particle settling and in particular the particle wake (Jenny et al. 2004). A variety of motion patterns exist, from straight vertical to fully chaotic paths, for which the fluid motion in the near field around the particles play a dominant role (Ern et al. 2012). Therefore, the proper resolution of the flow field in vicinity of the particles is crucial. In the present work the motion of the fluid and the dispersed phase were fully resolved by means of direct numerical simulation and the interface between the dispersed- and carrier-phase was fully resolved by means of an immersed boundary method (Uhlmann 2005). The interaction between the (turbulent) carrier flow and the solid particles can lead to a number of hydro-dynamical coupling phenomena which are most prominently manifested by the following open questions: (a) Do finite-size particles exhibit clustering or not? (b) How does the flow field and/or particle clustering affect the settling rate of the particles? (c) What are the characteristics of the wake-induced turbulence? Particle clustering has been investigated for long time. Majority of the previous studies have been concentrated on the so-called “preferential concentration” of small particles in turbulent flows (Squires and Eaton 1991; Fessler et al. 1994). Few studies have been devoted to the question, whether finite-size particles exhibit clustering, e.g. (Qureshi et al. 2008; Fiabane et al. 2012). Furthermore, the interaction between the flow filed and the solid particles have been from great interest in the scientific community. Especially the question, whether the particles enhance or attenuate the flow turbulence. Lucci et al. 2010 have investigated the interaction of finite-size heavy particles with a decaying homogeneous-isotropic turbulence in the absence of gravity. In the present study the analysis will focus primary on: (i) the mean apparent velocity lag; (ii) the flow field fluctuations induced by the particles as they settle through the computational domain as well as the velocity fluctuations of the particles; (iii) the spatial structure of the dispersed phase, i.e. do the particles form cluster and what

Upload: others

Post on 21-Oct-2020

0 views

Category:

Documents


0 download

TRANSCRIPT

  • 8th

    International Conference on Multiphase Flow ICMF 2013, Jeju, Korea, May 26 - 31, 2013

    1

    Settling of finite-size particles in an ambient fluid: A Numerical Study

    Todor Doychev

    1, Markus Uhlmann

    1

    1Institute for Hydromechanics, Karlsruhe Institute of Technology, 76128 Karlsruhe, Germany

    Keywords: particulate flow, resolved particles, DNS, immersed boundary method, clusters, Voronoї analysis

    Abstract

    We have investigated the gravity induced settling of finite-size particles in an ambient fluid by means of direct numerical

    simulation (DNS). Such configurations are relevant to a large number of applications such as meteorology, mechanical and

    environmental engineering. For single particle a variety of motion patterns exist, from straight vertical to fully chaotic paths,

    for which the fluid motion in the near field around the particles and the particle wake play a dominant role. Here, the interface

    between the dispersed- and carrier-phase was fully resolved by means of an immersed boundary method. Particular care has

    been taken to meet the respective resolution requirements. We have performed simulations of particles settling with two

    different Galileo numbers, 𝐺𝑎 = 121 and 𝐺𝑎 = 178. The settling regimes for a single particle settling with this Galileo numbers are: (a) steady axisymmetric regime and (b) steady oblique regime. We observed, that in the steady oblique regime

    (𝐺𝑎 = 178) the particles exhibit wake induced clustering, while in the steady axisymmetric regime (𝐺𝑎 = 121) this was not observed. Furthermore, the mean settling velocity of the particles with 𝐺𝑎 = 178 was strongly enhanced compared to the velocity of a single settling particle.

    1. Introduction

    Particle-laden flows are found in a large number of

    environmental natural and technical processes. Examples

    include pollution dispersion in the atmosphere, raindrop

    formation in clouds, sediment transport, fluidized bed

    reactors and combustion devices. Thus, the accurate

    prediction of such flows is of great importance. This

    naturally leads to the necessity of reliable theoretical

    description of the mechanisms that take place in such flows.

    Despite the progress made in the past, there is still a large

    scatter of the available data. A recent review of the subject

    can be found in (Balachandar and Eaton 2010).

    Here we consider the sedimentation of finite-size particles

    in an (initially) ambient fluid under the influence of gravity.

    Under such conditions, the system is characterized by a set

    of non-dimensional parameters. Given the fluid density 𝜌𝑓,

    the kinematic fluid viscosity 𝜈, the vector of gravitational acceleration 𝒈 on the one hand, and the particle diameter 𝐷, particle density 𝜌𝑝 and solid volume fraction 𝛷𝑠 on the

    other hand, dimensional analysis shows that the problem is

    determined by three non-dimensional parameters. One has

    already been mentioned, the solid volume fraction 𝛷𝑠. The other two can be taken as the density ratio 𝜌𝑝/𝜌𝑓 and the

    Galileo number defined as the ratio between the

    gravity-buoyancy and the viscosity forces

    𝐺𝑎 = (|𝒈|𝐷3|𝜌𝑝/𝜌𝑓 − 1| )1/2/𝜈.

    The particles under consideration in the present work have a

    particle Galileo (Reynolds) number of the order of 𝑂(100). For single particle, the Galileo number and the

    particle-to-fluid density ratio characterize the regime of

    particle settling and in particular the particle wake (Jenny et

    al. 2004). A variety of motion patterns exist, from straight

    vertical to fully chaotic paths, for which the fluid motion in

    the near field around the particles play a dominant role (Ern

    et al. 2012). Therefore, the proper resolution of the flow

    field in vicinity of the particles is crucial. In the present

    work the motion of the fluid and the dispersed phase were

    fully resolved by means of direct numerical simulation and

    the interface between the dispersed- and carrier-phase was

    fully resolved by means of an immersed boundary method

    (Uhlmann 2005).

    The interaction between the (turbulent) carrier flow and the

    solid particles can lead to a number of hydro-dynamical

    coupling phenomena which are most prominently

    manifested by the following open questions: (a) Do

    finite-size particles exhibit clustering or not? (b) How does

    the flow field and/or particle clustering affect the settling

    rate of the particles? (c) What are the characteristics of the

    wake-induced turbulence?

    Particle clustering has been investigated for long time.

    Majority of the previous studies have been concentrated on

    the so-called “preferential concentration” of small particles

    in turbulent flows (Squires and Eaton 1991; Fessler et al.

    1994). Few studies have been devoted to the question,

    whether finite-size particles exhibit clustering, e.g. (Qureshi

    et al. 2008; Fiabane et al. 2012). Furthermore, the

    interaction between the flow filed and the solid particles

    have been from great interest in the scientific community.

    Especially the question, whether the particles enhance or

    attenuate the flow turbulence. Lucci et al. 2010 have

    investigated the interaction of finite-size heavy particles

    with a decaying homogeneous-isotropic turbulence in the

    absence of gravity.

    In the present study the analysis will focus primary on: (i)

    the mean apparent velocity lag; (ii) the flow field

    fluctuations induced by the particles as they settle through

    the computational domain as well as the velocity

    fluctuations of the particles; (iii) the spatial structure of the

    dispersed phase, i.e. do the particles form cluster and what

  • 8th

    International Conference on Multiphase Flow ICMF 2013, Jeju, Korea, May 26 - 31, 2013

    2

    is the extension and shape of the clusters in case clustering

    takes place.

    The experiments of (Parthasarathy and Faeth 1990a,b)

    provide one of the most relevant datasets for the present

    study (𝑅𝑒𝑝=[65,147,262]). They performed measurements

    of heavy particles settling in an ambient fluid. Their

    experiments provide measurements of the velocity fields for

    both phases. Considering the wake-induced clustering of the

    particles, our parameters are matched most closely by the

    numerical simulations of Kajishima and Takiguchi (2002)

    and the experiments of Nishino and Matsuchita (2004).

    2. Numerical Method

    The numerical code used for the present simulation utilizes

    an efficient and precise formulation of the immersed

    boundary method for the simulation of particulate flows

    (Uhlmann 2005). This method employs direct forcing

    approach by adding a localized volume force term to the

    momentum equations. The basic idea of the immersed

    boundary method is to solve the Navier-Stokes equations in

    the entire computational domain, including the space

    occupied by the particles. The force term is formulated in

    such way, as to impose a rigid body motion upon the fluid at

    the locations of the solid particles and is explicitly computed

    at each time step as a function of the desired particle

    positions and velocities, without recurring to a feed-back

    procedure. Thereby, the stability characteristics of the

    underlying Navier-Stokes solver are maintained in the

    presence of particles, allowing for relatively large time

    steps.

    The necessary interpolation of variable values from Eulerian

    grid positions to particle-related Lagrangian positions and

    vice versa are performed by means of the regularized delta

    function approach of (Peskin 2002). This procedure yields a

    smooth temporal variation of the hydrodynamic forces

    acting on individual particles while these are in arbitrary

    motion with respect to the fixed grid.

    The solution of the Navier-Stokes equations is realized in

    the framework of a standard fractional-step method for

    incompressible flow. The temporal discretization is

    semi-implicit, based on the Crank-Nicholson scheme for the

    viscous terms and a low-storage three-step Runge-Kutta

    procedure for the non-linear part (Verzicco and Orlandi

    1996). The spatial operators are discretized by means of

    central finite-differences on a staggered grid. The temporal

    and spatial accuracy of this scheme are of second order.

    The particle motion is determined by the Runge-Kutta

    discretized Newton equations for translational and rotational

    rigid-body motion, which are explicitly coupled to the fluid

    equations. The hydrodynamic forces acting upon a particle

    are readily obtained by summing the additional volume

    forcing term over all discrete forcing points. Thereby, the

    exchange of momentum between the two phases cancels out

    identically and no spurious contributions are generated. The

    analogue procedure is applied for the computation of the

    hydrodynamic torque driving the angular particle motion. In

    the case of periodic boundary conditions, the spatial average

    of the force term needs to be subtracted from the momentum

    equation for compatibility reasons (Fogelson and Peskin

    1988; Höfler and Schwarzer 2000).

    Since particles are free to visit any point in the

    computational domain and in order to ensure that the

    regularized delta function verifies important identities (such

    as the conservation of the total force and torque during

    interpolation and spreading), a Cartesian grid with uniform

    isotropic mesh width is used. For reasons of efficiency,

    forcing is only applied to the surface of the spheres, leaving

    the flow field inside the particles to develop freely.

    During the course of the simulation, particles can approach

    each other closely. However, very thin inter particle films

    cannot be resolved by a typical grid and therefore the

    correct build-up of repulsive pressure is not captured which

    in turn lead to possible partial “overlap” of the particle

    position. In order to prevent such non-physical situations,

    we use the artificial repulsion potential of (Glowinski at al.

    1999), relying upon a short-range repulsion force.

    For detailed description of the method and for information

    on the validation tests and grid convergence please refer to

    (Uhlmann 2005, 2008; García-Villalba et al. 2012;

    Kidanemariam 2013) and further references therein.

    3. Setup of the Simulations

    The sedimentation of multiple heavy spherical particles in

    an otherwise ambient fluid under the influence of gravity

    was studied. We carried out numerical experiments with two

    different Galileo numbers, 𝐺𝑎 = 121 and 𝐺𝑎 = 178 . In both cases the particle-to-fluid density ratio and the solid

    volume fraction were kept constant to 𝜌𝑝/𝜌𝑓 = 1.5

    and 𝛷𝑠 = 0.5 % . The corresponding particle Reynolds number based on the balance between drag and immersed

    weight, using the standard drag formula (Clift 1978, p.112)

    𝑅𝑒𝑝∞ = 𝑤𝑝∞𝑐𝑙𝑖𝑓𝑡

    𝐷/𝜈 , was calculated to 𝑅𝑒𝑝∞ = 141 and

    𝑅𝑒𝑝∞ = 245. Under this conditions the flow is considered

    to be dilute, thus dominant effects of inter-particle collisions

    are avoided. Here and in the following, we will refer to the

    particles settling with 𝐺𝑎 = 121 as case M120 and to the particles settling with 𝐺𝑎 = 178 as case M180. Additionally we performed simulations of the settling of a

    single particle with the exact physical parameters as in case

    M120 and case M180. The corresponding simulations are

    denoted by S120 and S180.

    The computational domain Ω for both cases M120 and

    M180 is a box elongated in the vertical direction. The

    domain extends in terms of the particle diameter 𝐷 to: 68𝐷 × 68𝐷 × 341 for case M120 and 85𝐷 × 85𝐷 ×170 for case M180. In all three directions periodic boundary conditions are applied. The selected solid volume

    fraction corresponds then to a total of 15190 particles in case M120 and 11867 particles in case M180. The particles are initially placed randomly in the computational

    domain. The initial particle position and the extensions of

    the computational domains are depicted in figure 1. The

    flow is resolved by 1024 × 1024 × 5120 grid points in case M120 and 2048 × 2048 × 4096 grid points in case M180. Consequently the particle resolution in case M120

    results in 𝐷/𝛥𝑥 = 15 and in case M180 in 𝐷/𝛥𝑥 = 24. The physical and numerical parameters of the performed

    simulations are summarized in table 1 and table 2.

  • 8th

    International Conference on Multiphase Flow ICMF 2013, Jeju, Korea, May 26 - 31, 2013

    3

    Figure 1: Dimensions of the computational domain with the

    initial particle distribution for (a) case M120 and (b) case M180 . D denotes the particle diameter. Gravity acts in the negative z-direction. Periodic boundary conditions are applied in all three directions in both cases.

    The simulations were initialized from a succession of

    coarse-grained runs with randomly distributed fixed

    particles. We start with a coarse simulation with fixed

    randomly distributed particles and evolve the simulation in

    time until stationary steady state is reached. The results of

    the coarse simulations are then linearly interpolated on a

    finer computational grid and the simulations are resumed.

    This is repeated until the targeted resolution is reached.

    During this procedure the particles were kept at fixed

    positions. This ensures that any particle deviations from a

    random distribution due to insufficient resolution are

    avoided. Once the desired resolution is reached, the particles

    are released to move freely and the actual recording of data

    is started. Here and in the following we arbitrary set the

    time at which the particles are released to zero, 𝑡 = 0. During the course of the present document the following

    nomenclature is applied: the domain is discretized by a

    Cartesian grid (x, y, z), where z is the vertical component in

    direction of the gravity g; velocity vectors and their

    components corresponding to the fluid and the particle

    phases are distinguished by subscripts “f” and “p”

    respectively, as in 𝒖𝑓 = (𝑢𝑓 , 𝑣𝑓 , 𝑤𝑓)𝑇 and 𝒖𝑝 =

    (𝑢𝑝, 𝑣𝑝, 𝑤𝑝)𝑇 ; particle position vector is denoted as

    𝐱𝑝 = (𝑥𝑝, 𝑦𝑝, 𝑧𝑝)𝑇 . The fluctuations of particle quantities

    over time are denoted by a single prime, i.e. 𝑢𝑝′ and are

    defined as the difference of the instantaneous values and the

    averaged value at that same time e.g. 𝑢𝑝′ (𝑥𝑝 (𝑡), 𝑡) =

    𝑢𝑝(𝑥𝑝 (𝑡), 𝑡) −< 𝑢𝑝 >𝑝 (𝑥𝑝 (𝑡), 𝑡) . Similarly, the

    fluctuations of the fluid velocity field with respect to the

    average over the volume occupied by the fluid are defined

    as 𝑢𝑓′ (𝑥 (𝑡), 𝑡) = 𝑢𝑓(𝑥 (𝑡), 𝑡) −< 𝑢𝑓 >Ω𝑓 (𝑥 (𝑡), 𝑡) . The

    time in the present work is scaled by the gravitational time

    scale 𝜏𝑔, which is defined as 𝜏𝑔 = (|𝒈|𝐷|𝜌𝑝/𝜌𝑓 − 1| )1/2.

    Table 1: Physical parameters for particulate flow with a

    single and multiple settling particles in an ambient fluid.

    Solid volume fraction 𝛷𝑠 , density ratio 𝜌𝑝/𝜌𝑓 , Galileo

    number 𝐺𝑎 = (|𝒈|𝐷3|𝜌𝑝/𝜌𝑓 − 1| )1/2/𝜈 , Reynolds

    number 𝑅𝑒𝑝∞ = 𝑤𝑝∞𝑐𝑙𝑖𝑓𝑡

    𝐷/𝜈 based on the terminal

    velocity of a single particle 𝑤𝑝∞𝑐𝑙𝑖𝑓𝑡

    , particle diameter 𝐷

    and fluid viscosity 𝜈 and number of particles 𝑁𝑝 . The

    gravitational constant 𝒈 is applied against the vertical direction 𝒛.

    𝛷𝑠 𝜌𝑝/𝜌𝑓 𝐺𝑎 𝑅𝑒𝑝∞ 𝑁𝑝

    𝑀120 0.005 1.5 121 141 15190 𝑀180 0.005 1.5 178 245 11867 𝑆120 𝑂(10−5) 1.5 121 141 1 𝑆180 𝑂(10−5) 1.5 178 245 1

    Table 2: Numerical parameters for particulate flow of

    single and multiple settling particles in an ambient fluid.

    Particle resolution 𝐷/𝛥𝑥, number of grid nodes 𝑁𝑖 in the 𝑖-th coordinate direction.

    𝐷/𝛥𝑥 𝑁𝑥 × 𝑁𝑦 × 𝑁𝑧

    𝑀120 15 1024 × 1024 × 5120 𝑀180 24 2048 × 2048 × 4096 𝑆120 15 128 × 128 × 2048 𝑆180 24 256 × 256 × 4096

    4. Results and Discussion

    4.1 Settling velocity

    Figure 2 depicts the temporal evolution of the mean

    apparent velocity lag 𝑤𝑝,𝑙𝑎𝑔 for both cases M120 and

    M180. The velocity 𝑤𝑝,𝑙𝑎𝑔 is defined as the difference in

    the mean streamwise velocity components of the two

    phases:

    𝑤𝑝,𝑙𝑎𝑔(𝑡) =< 𝑤𝑝(𝑡) >𝑝− < 𝑤𝑓(𝑡) >Ω𝑓 , (3.1)

    where Ω𝑓 defines a spatial averaging operator over the

    domain occupied by the fluid and 𝑝 denotes an

    averaging operator over the particles. The velocities have

    been scaled as a Reynolds number with the particle diameter

    and the fluid kinematic viscosity, 𝑅𝑒𝑝,𝑙𝑎𝑔 = |𝑤𝑝,𝑙𝑎𝑔 |𝐷/𝜈.

    The terminal settling velocity of the single settling particles,

    case S120 and case S180, is shown as well. As can be seen

  • 8th

    International Conference on Multiphase Flow ICMF 2013, Jeju, Korea, May 26 - 31, 2013

    4

    the mean apparent lag in case M120 remains over the entire

    course of the simulation at approximate value of 𝑅𝑒𝑝,𝑙𝑎𝑔 ≈

    141. Moreover, the velocity 𝑤𝑝,𝑙𝑎𝑔 in case M120 is well

    represented by the settling velocity of the single particle

    from case S120. This implies that, the presence of many

    freely moving particles in case M120 does not have strong

    influence upon the mean apparent velocity lag of the

    particles. In case M180 on the other hand, after the particles

    are released they begin to accelerate for approximately two

    hundred gravitational time units and the velocity 𝑤𝑝,𝑙𝑎𝑔

    reaches a maximum of approximately 𝑅𝑒𝑝,𝑙𝑎𝑔 = 271. After

    reaching maximum, the mean settling velocity decelerates

    and levels up, still exhibiting some fluctuations, at an

    approximate value of 𝑅𝑒𝑝,𝑙𝑎𝑔 = 260. In contrast to case

    M120, we found out that the velocity 𝑤𝑝,𝑙𝑎𝑔 in case M180

    deviates significantly from the one for single particle in case

    S180. As will be discussed below, the increase of the mean

    apparent velocity lag in case M180 in comparison to the

    settling velocity of a single particle is a direct result of the

    clustering of the dispersed phase in case M180.

    Figure 2: Average particle settling velocity for case M120

    (black solid line) and case M180 (red solid line)

    (normalized with the particle diameter and kinematic

    viscosity) as function of time. The time is normalized with

    the gravitational time scale 𝜏𝑔 = (|𝒈|𝐷|𝜌𝑝/𝜌𝑓 − 1| )1/2 .

    Terminal settling velocity of a single particle in ambient

    fluid for case S120 (black dashed line) and case S180 (red

    dashed line).

    4.2 Velocity fluctuations

    As mentioned above, the particles are settling in an

    (initially) ambient fluid. Therefore, any fluctuations of the

    fluid are induced by the settling of the particles. The

    relevant question in this context is aimed at determining the

    intensity of the self-induced fluid motion. This analysis is

    different (but related to) the study of the modulation of

    existing (background) turbulence due to the addition of

    particles. Figure 3a shows the temporal evolution of the root

    mean square (r.m.s.) values of the fluid velocity components

    for both cases, M120 and M180. The r.m.s. values are

    normalized with the terminal settling velocity of the

    particles from case 𝑆120 and case 𝑆180 respectively.

    (a)

    (b)

    Figure 3: (a) R.m.s. of the fluid velocity fluctuations for

    case M120 (black lines), and case M180 (red lines) as

    function of time. Solid lines show the horizontal

    components, dashed lines show vertical component. (b) As

    in (a), but for the particle velocity. The reference velocity

    𝑤𝑟𝑒𝑓 denotes the terminal settling velocity 𝑤𝑝,𝑙𝑎𝑔 of case

    S120 and case S180 respectively.

    As can be immediately seen, the fluid velocity components

    in case M180 show higher fluctuations than in case M120. It

    can be observed, that in both cases the fluid velocity

    fluctuations of the vertical component are highly dominant.

    This can be attributed to the wake-induced character of the

    fluid motion, since fluid motion is primarily caused by the

    particles moving in streamwise direction. This high level of

    anisotropy also suggests strong effects of the particle wakes,

    where the mean particle wake contributes to the fluctuating

    velocity field. Similar observations were made in the work

    of Parthasarathy and Faeth (1990a,b). In case M180 the

    r.m.s. values increase after releasing the particles for about

    200𝜏𝑔 time units after which they oscillate around a mean

    value of 0.27 for the streamwise component and around

    0.08 for the lateral component. On the other hand the

    fluctuations in case M120 do not experience significant

    increase after particles are released. The r.m.s. values in

  • 8th

    International Conference on Multiphase Flow ICMF 2013, Jeju, Korea, May 26 - 31, 2013

    5

    streamwise direction seem to have a large period oscillating

    behaviour. In the following we assume the flow to be in

    statistically steady state after 200𝜏𝑔. For case M180, after

    reaching statistically stationary state, the fluid velocity

    fluctuations exhibit somehow more prominent peaks in the

    r.m.s. values than in case M120 and they are more

    distinctive for the streamwise component than for the lateral

    velocity component.

    The corresponding fluctuations of the velocity components

    of the dispersed phase are depicted in figure 3b. The

    fluctuations of the dispersed phase experience similar

    evolution over time as for the fluid velocity: the fluctuations

    in case M180 are larger then in case M120 and the

    streamwise component is approximately 2.5 times larger

    then the cross-stream component of the particle velocity.

    The intensity of the fluctuations of the vertical velocity

    component can be attributed entirely to the collective effects,

    since a single particle at both Galileo numbers is in steady

    motion (Jenny et al. 2004).

    By comparison of the velocity fluctuations of both phases,

    we found out that, both phases experience comparable

    values. Similar values for the cross-stream velocity

    component indicate that turbulent dispersion plays an

    important role in the present cases. As mentioned earlier the

    flow is dominated by the wakes generated from the particles

    as they settle through the domain. The particles do not settle

    on straight vertical trajectories, rather they settle on curved

    or oblique paths. As result the particle wakes are not

    oriented exactly in vertical direction, which causes

    momentum along the wake axis to be deposited into the

    lateral direction.

    A measure of the influence of the fluid turbulence on the

    particles is the relative turbulence intensity, defined as the

    ratio between the intensity of the incoming fluid flow

    fluctuations and the apparent slip velocity. In homogeneous

    flows the definition of relative turbulence intensity is often

    based upon the three-component turbulent intensities, viz.

    𝐼𝑟 = (< 𝑢𝑓,𝑖′ >/3)1/2/𝑤𝑝,𝑙𝑎𝑔 . In cases with unidirectional

    mean flow (z-coordinate direction) the following definition

    is commonly employed, 𝐼𝑟𝑉 = (< 𝑤𝑓′ >)1/2/𝑤𝑝,𝑙𝑎𝑔.

    The temporal evolution of the relative turbulence intensity

    𝐼𝑟 (𝐼𝑟𝑉) is depicted in figure 4. Although, the flow in the present simulations is not considered to be turbulent in the

    general sense, the relative turbulence intensities in the

    present simulations are showed to be comparable to the

    turbulence levels often considered in studies of the influence

    of background turbulence upon the particle motion, e.g. (Wu

    and Faeth 1994; Bagchi and Balachandar 2004; Yang and

    Shy 2005; Legendre et al 2006; Poelma et al. 2007; Snyder

    et al 2008; Amoura et al. 2010). For the present cases we

    calculated the relative turbulence intensity to 𝐼𝑟𝑉 =0.2 (0.24) for case M120 (M180). This indicates that the particles in the present cases generated substantial fluid

    “turbulence” as they settle through the domain.

    4.3 Spatial structure of the dispersed phase

    As aforementioned, an interesting feature of particulate

    flows is the ability of the particles to form clusters. As

    already mentioned all the fluctuations of the flow field are

    induced by the settling particles. Moreover, we saw that the

    particles react to the flow field fluctuations. This reaction

    was manifested in the fluctuations of the particle velocity

    field. For the considered Galileo (Reynolds) numbers in this

    study, the wakes are important even at considerably large

    distances behind the particles. Thus, the particles can

    interact with each other over large distances through the

    particle wakes. The most prominent example of the particle

    wake interaction is the “drafting-kissing-tumbling” motion

    of pairs of trailing particles (Fortes et al. 1987; Wu and

    Manasseh 1998). Since the flow field is homogeneous in all

    three directions any deviation of the particles from the

    random distribution can be attributed to the wake character

    of the flow.

    The particle distribution for the present simulations can be

    visually examined in figure 5 (case M120) and figure 6

    (case M180) where the position of the particles is projected

    on the 2 -D horizontal plane. While in case M120 no significant difference in the distribution of the particles at

    the beginning and at later time of the simulation can be

    observed, it is clearly observable that the particle

    distribution in case M180 at later time deviates significantly

    from the random distribution at the beginning of the

    simulation. The distribution of the particles shows regions

    with high number of particles (high particle concentration)

    and regions with small number of particles (low particle

    concentration), or even void regions with complete absence

    of particles. Hereafter, regions with high particle

    concentration are referred as clusters and regions with low

    particle concentration as voids.

    The clusters and the voids in case M180 (figure 6b) appear

    to extend throughout the entire height of the computational

    domain. This is in line with previous experimental and

    numerical findings (Kajishima and Takiguchi 2002,

    Kajishima 2004a) where similar “columnar particle

    accumulation” (Nishino and Matsushita 2004) was observed.

    Time sequences of such visualization (not shown here)

    shows that these structures are quite robust and they persist

    over long time intervals.

    More quantitative information on the particle clustering can

    be obtained by performing a Voronoї analysis (Monchaux et

    al. 2010). The Voronoї tessellation is a decomposition of the

    space into independent cells, which have the property that

    Figure 4: Temporal evolution of the relative turbulence

    intensity for case M120 (black lines) and case M180 (red

    lines). Solid lines: 𝐼𝑟 = (< 𝑢𝑓,𝑖′ >/3)1/2/𝑤𝑝,𝑙𝑎𝑔 . Dashed

    lines 𝐼𝑟𝑉 = (< 𝑤𝑓′ >)1/2/𝑤𝑝,𝑙𝑎𝑔 .

  • 8th

    International Conference on Multiphase Flow ICMF 2013, Jeju, Korea, May 26 - 31, 2013

    6

    each point in the cell is closer to the cell's site than to any

    other cell's site. As a consequence, the inverse of a Voronoї

    cell's volume is proportional to the local particle

    concentration. In order to quantify preferential concentration

    we compared the probability distribution function (p.d.f.) of

    the normalized Voronoї volumes with the p.d.f. of randomly

    distributed particles (Monchaux et al. 2010, 2012; Fiabane

    et al. 2012). The Voronoї cell volumes are normalized to be

    of unit mean. This normalization allows a qualitative

    comparison of the p.d.f.s for different particle number

    densities, since the so normalized p.d.f.s are independent of

    the particle number density (Ferenc and Neda 2007).

    The distribution of particles experiencing clustering is

    expected to be more intermittent than the distribution of

    randomly distributed particles. This implies that regions

    with higher particle concentration are more probable than

    for random distribution. Respectively, void regions with low

    particle concentration are also more probable.

    Figure 5: (a) Top view of the particle position at the begin

    of the simulation 𝑡/𝜏𝑔 = 0 for case M120. (b) The same,

    but at 𝑡/𝜏𝑔 = 1200.

    Figures 7a and 8a depict the p.d.f. of the normalized

    Voronoї volumes for case M120 and case M180. As can be

    observed, the p.d.f. for both cases at the time when the

    particles were released in the computational domain is well

    represented by the theoretical Gamma distribution for

    random particle fields (Ferenc and Neda 2007), confirming

    that the initial particle distribution was indeed random. At

    later times the p.d.f.s in case M120 (figure 7a) show that the

    extremes for the present data are less probable than in the

    case of randomly distributed particles. This indicates that

    the particles in case M120 become more ordered than

    randomly distributed particles. On the other hand, the p.d.f.s

    in case M180 (figure 8a) deviate significantly from the

    distribution function for the random distributed particles.

    The distribution function of the present data exhibits tails

    with higher probability of finding Voronoї cells with very

    large or very small volumes than in the random case with

    uniform probability, indicating that cluster formation takes

    place in case M180.

    Figure 6: (a) Top view of the particle position at the begin

    of the simulation 𝑡/𝜏𝑔 = 0 for case M180. (b) The

    same, but at 𝑡/𝜏𝑔 = 820.

  • 8th

    International Conference on Multiphase Flow ICMF 2013, Jeju, Korea, May 26 - 31, 2013

    7

    Further, we have analysed the shape of the regions with high

    particle concentration by computing the aspect ratio of each

    Voronoї cell, defined as the ratio of the largest cross-stream

    extension 𝑙𝑥,𝑉𝑖 to the largest streamwise extension 𝑙𝑧,𝑉𝑖 of

    the Voronoї cell, 𝐴𝑉 = 𝑙𝑥,𝑉𝑖/𝑙𝑧,𝑉𝑖 . The aspect ratio 𝐴𝑣

    provides a qualitative measure of the anisotropy of the

    particle clusters and can be seen as measure for the

    stretching of the cells. Figures 7b and 8b show the p.d.f.s of

    the Voronoї cells aspect ratio 𝐴𝑉 for case M120 and case M180. It can be immediately seen that the p.d.f.s in case

    M120 do not show any deviation from the p.d.f. of the

    randomly distributed particles, while in case M180 an

    appreciable difference is observed. This indicates, that the

    majority of the Voronoї cells are squeezed/stretched in the

    vertical/horizontal direction and that the particle structures

    in case M180 are more likely to be aligned in vertical

    direction. This confirms the observations made in figure 6b.

    Figure 7: Case M120: Probability density function of (a)

    the normalized Voronoї cell volumes and (b) the aspect

    ratios 𝐴𝑉. Different lines represent data assembled over different time intervals. The initial random distribution is

    represented by black solid line. 𝑡/𝜏𝑔 = [279,307] (red).

    𝑡/𝜏𝑔 = [559,587] (blue); 𝑡/𝜏𝑔 = [1216,1244] (green);

    𝑡/𝜏𝑔 = [1496,1524] (magenta). The dashed black line

    represents an analytical Gamma function fit (Ferenc and

    Neda 2007).

    An alternative way of characterizing the spatial structure of

    the dispersed phase is by performing nearest-neighbour

    analysis (Kajishima 2004b). Figure 9 depicts the time

    evolution of the average distance to the nearest particle

    neighbour 𝑑𝑚𝑖𝑛. The distance 𝑑𝑚𝑖𝑛 is calculated as:

    𝑑𝑚𝑖𝑛 =1

    𝑁𝑝 ∑ min

    𝑗=1,𝑁𝑝𝑗≠𝑖

    𝑑𝑖,𝑗

    𝑁𝑝

    𝑖=1

    , (4.2)

    where 𝑑𝑖,𝑗 = |𝑥𝑝,𝑖 − 𝑥𝑝,𝑗| is the distance between the centers of particles 𝑖 and 𝑗. The distance 𝑑𝑚𝑖𝑛 has been normalized by its value for a homogeneous distribution with

    the same solid volume fraction, 𝑑𝑚𝑖𝑛ℎ𝑜𝑚 = (|Ω|/𝑁𝑝)1/3. The

    lower limit of 𝑑𝑚𝑖𝑛ℎ𝑜𝑚 corresponds to the minimum value of

    the function, when all particles are in contact with a

    neighbouring particle.

    Figure 8: Case M180: Probability density function of (a)

    the normalized Voronoї cell volumes and (b) the aspect

    ratios 𝐴𝑉. Different lines represent data assembled over different time intervals. Initial random distribution is

    represented by black solid line. 𝑡/𝜏𝑔 = [199,213] (red).

    𝑡/𝜏𝑔 = [345,359] (blue); 𝑡/𝜏𝑔 = [572,576] (green);

    𝑡/𝜏𝑔 = [794,810] (magenta). The dashed black line

    represents an analytical Gamma function fit (Ferenc and

    Neda 2007).

  • 8th

    International Conference on Multiphase Flow ICMF 2013, Jeju, Korea, May 26 - 31, 2013

    8

    The upper limit of the function 𝑑𝑚𝑖𝑛/𝑑𝑚𝑖𝑛ℎ𝑜𝑚 has a value of

    unity and arises for homogeneously distributed particles, i.e.

    particles are positioned on a regular lattice. It can be

    observed, that the average distance to the nearest neighbour

    at initial time is very close to the value for randomly

    distributed particles. As time evolves, the value of

    𝑑𝑚𝑖𝑛/𝑑𝑚𝑖𝑛ℎ𝑜𝑚 in case M120 increases quickly and reaches a

    statistically steady state value of approximately 0.61. This

    finding is in line with the findings of the Voronoї analysis

    that the particles in case M120 are more ordered than

    randomly distributed particles. Contrarily, the value of

    𝑑𝑚𝑖𝑛/𝑑𝑚𝑖𝑛ℎ𝑜𝑚 in case M180 initially decreases for

    approximately 200𝜏𝑔 and undulates around approximate

    value of 0.5. This again corroborates the results obtained by

    the Voronoї analysis that the particles in case M180 tend to

    form agglomerations.

    Figure 9: Temporal evolution of the average distance to

    the nearest neighbor for cases M120 and M180, normalized

    by the value for a homogeneous distribution on a regular

    cubical lattice. Case M120 (black solid line). Case M180

    (red solid line). Random particle distribution with the same

    volume fraction (black dashed line).

    5. Conclusions

    We have simulated the settling of spherical particles at

    moderate Reynolds numbers and low solid volume fractions.

    The solid/fluid interfaces were fully resolved by means of

    an immersed boundary method. The particles were released

    to move freely in an initially ambient fluid. The settling in

    two different single particle regimes was investigated:

    steady axisymmetric regime with Galileo number

    𝐺𝑎 = 121 (𝑅𝑒𝑝,𝑙𝑎𝑔 = 141) and steady oblique regime

    with 𝐺𝑎 = 178 (𝑅𝑒𝑝,𝑙𝑎𝑔 = 250).

    Voronoї analysis of the particle spatial distribution was

    performed. Our analysis revealed that the particles in the

    steady oblique regime exhibit significant clustering, while in

    the steady axisymmetric regime no clustering of the

    dispersed phase was observed. It was observed that, the

    clustering of the particles led to a significant increase of the

    mean apparent velocity lag. Furthermore, the shape of the

    clusters was investigated and it was found that the cluster

    structures have strong anisotropic shape, where particles

    were aggregated in a column like structures which extended

    throughout the entire height of the computational domain.

    It was found that the flow field was considerably anisotropic

    with the vertical direction being the dominant direction. The

    results show that, independent of clustering, considerable

    turbulence levels were induced by the settling particles, with

    relative turbulence intensities reaching values of 0.2 to 0.24.

    Moreover, the velocity fluctuations of both phases were

    comparable indicating that the particles respond strongly to

    the flow field.

    In the future more detailed analysis of the flow will be

    performed, e.g.: (i) The effect of clustering upon the flow

    statistics will be more deeply analysed; (ii) Lagrangian

    analysis of the data; (iii) Statistics of conditionally averaged

    data. The existence of such strong differences in the spatial

    distribution of the particles between the two different

    settling regimes is quite interesting and the role of the

    settling regimes on the spatial distribution of the particles

    will be further analysed. Next step will be the simulation of

    the interaction between finite size particles and forced

    homogeneous turbulence.

    Acknowledgements

    Support through a research grant from DFG (UH 242/1-1) is

    thankfully acknowledged. The authors also want to also

    acknowledge the computer resources, technical expertise

    and assistance provided by the Leibniz Supercomputing

    Center (LRZ), Jülich Supercomputing Center (JSC) as well

    as by the Steinbruch Center for Computing (SCC) at KIT.

    References

    Amoura, Z., Roig, V., Risso, F., Billet, A.M., 2010.

    Attenuation of the wake of a sphere in an intense incident

    turbulence with large length scales. Physics of Fluids 22,

    055105.

    Bagchi, P., Balachandar, S., 2004. Response of the wake of

    an isolated particle to an isotropic turbulent flow. Journal

    of Fluid Mechanics 518, 95–123.

    Balachandar, S., Eaton, J.K., 2010. Turbulent Dispersed

    Multiphase Flow. Annual Review of Fluid Mechanics

    42, 111–133.

    Clift, R., Grace, J., Weber, M., 1978. Bubbles, drops,

    and particles. Academic Press.

    Ern, P., Risso, F., Fabre, D., Magnaudet, J., 2012.

    Wake-Induced Oscillatory Paths of Bodies Freely Rising

    or Falling in Fluids. Annual Review of Fluid Mechanics 44,

    97–121.

    Ferenc, J.S., Néda, Z., 2007. On the size distribution of

    Poisson Voronoi cells. Physica A: Statistical Mechanics

    and its Applications 385, 518–526.

    Fessler, J.R., Kulick, J.D., Eaton, J.K., 1994. Preferential

    concentration of heavy particles in a turbulent channel flow.

    Physics of Fluids 6, 3742.

  • 8th

    International Conference on Multiphase Flow ICMF 2013, Jeju, Korea, May 26 - 31, 2013

    9

    Fiabane, L., Volk, R., Pinton, J.F., Monchaux, R., Cartellier,

    A., Bourgoin, M.M., 2012. Do finite size neutrally buoyant

    particles cluster? arXiv preprint , 1–6.

    Fogelson, A., Peskin, C.S., 1988. A fast numerical method

    for solving the three-dimensional Stokes’ equations in the

    presence of suspended particles. Journal of Computational

    Physics 79, 50–69.

    Fortes, A., Joseph, D., Lundgren, T., 1987. Nonlinear

    mechanics of fluidization of beds of spherical particles.

    Journal of Fluid Mechanics 177, 467–483.

    García-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. International Journal of

    Multiphase Flow 46, 54–74.

    Glowinski, R., Pan, T., Hesla, T., Joseph, D.D., Periaux, J.,

    1999. A distributed Lagrange multiplier/fictitious domain

    method for flows around moving rigid bodies: application

    to particulate flow. Int. J. Numer. Meth. Fluids 1066,

    1043–1066.

    Höfler, K., Schwarzer, S., 2000. Navier-Stokes simulation

    with constraint forces: finite-difference method for

    particle-laden flows and complex geometries. Physical

    review. E, Statistical physics, plasmas, fluids, and related

    interdisciplinary topics 61, 7146–60.

    Jenny, M., Dušek, J., Bouchet, G., 2004. Instabilities and

    transition of a sphere falling or ascending freely in a

    Newtonian fluid. Journal of Fluid Mechanics 508, 201–

    239.

    Kajishima, T., 2004a. Influence of particle rotation on the

    interaction between particle clusters and particle induced

    turbulence. International Journal of Heat and Fluid Flow

    25, 721–728.

    Kajishima, T., 2004b. Numerical investigation of collective

    behavior of gravitationally settling particles in a

    homogeneous field, in: Proc. ICMF, pp. 1–10.

    Kajishima, T., Takiguchi, S., 2002. Interaction between

    particle clusters and particle-induced turbulence.

    International Journal of Heat and Fluid Flow 23, 639–646.

    Kidanemariam, A.G., Chan-Braun, C., Doychev, T.,

    Uhlmann, M., 2013. DNS of horizontal open channel flow

    with finite-size, heavy particles at low solid volume

    fraction. arXiv preprint , 1–44.

    Legendre, D., Merle, A., Magnaudet, J., 2006. Wake of a

    spherical bubble or a solid sphere set fixed in a turbulent

    environment. Physics of Fluids 18, 048102.

    Lucci, F., Ferrante, A., Elghobashi, S., 2010. Modulation of

    isotropic turbulence by particles of Taylor length-scale size.

    Journal of Fluid Mechanics 650, 5.

    Monchaux, R., Bourgoin, M., Cartellier, A., 2010.

    Preferential concentration of heavy particles: A Voronoї

    analysis. Physics of Fluids 22, 103304.

    Monchaux, R., Bourgoin, M., Cartellier, A., 2012.

    Analyzing preferential concentration and clustering of

    inertial particles in turbulence. International Journal of

    Multiphase Flow 40, 1–18.

    Nishino, K., Matsushita, H., 2004. Columnar particle

    accumulation in homogeneous turbulence. Proc. ICMF04

    (5th Int. Conf. Multiphase Flow) , 1–12.

    Parthasarathy, R., Faeth, G.M., 1990a. Turbulence

    modulation in homogeneous dilute particle-laden flows.

    Journal of Fluid Mechanics 220, 485–514.

    Parthasarathy, R., Faeth, G.M., 1990b. Turbulent

    dispersion of particles in self-generated homogeneous

    turbulence. Journal of Fluid Mechanics 220, 515–537.

    Peskin, C.S., 2002. The immersed boundary method. Acta

    Numerica 11, 1–39.

    Poelma, C., Westerweel, J., Ooms, G., 2007. Particlefluid

    interactions in grid-generated turbulence. Journal of Fluid

    Mechanics 589, 315–351.

    Qureshi, M., Arrieta, U., Baudet, C., Cartellier, A., Gagne,

    Y., Bourgoin, M., 2008. Acceleration statistics of inertial

    particles in turbulent flow. The European Physical Journal

    B 66, 531–536.

    Snyder, M.R., Knio, O.M., Katz, J., Le Maitre, O.P., 2008.

    Numerical study on the motion of microscopic oil droplets

    in high intensity isotropic turbulence. Physics of Fluids 20,

    073301.

    Squires, K.., Eaton, J.K., 1991. Preferential concentration

    of particles by turbulence. Physics of Fluids A: Fluid

    Dynamics 3, 1169.

    Uhlmann, M., 2005. An immersed boundary method with

    direct forcing for the simulation of particulate flows.

    Journal of Computational Physics 209, 448–476.

    Uhlmann, M., 2008. Interface-resolved direct numerical

    simulation of vertical particulate channel flow in the

    turbulent regime. Physics of Fluids 20, 053305.

    Verzicco, R., Orlandi, P., 1996. A finite-difference scheme

    for three-dimensional incompressible flows in cylindrical

    coordinates. Journal of Computational Physics 123, 402–

    414.

    Wu, J., Faeth, G.M., 1994. Sphere wakes at moderate

    Reynolds numbers in a turbulent environment. AIAA

    Journal 32, 535–541.

    Wu, J., Manasseh, R., 1998. Dynamics of dual-particles

    settling under gravity. International Journal of Multiphase

    Flow 24, 1343–1358.

  • 8th

    International Conference on Multiphase Flow ICMF 2013, Jeju, Korea, May 26 - 31, 2013

    10

    Yang, T.S., Shy, S.S., 2005. Two-way interaction between

    solid particles and homogeneous air turbulence: particle

    settling rate and turbulence modification measurements.

    Journal of Fluid Mechanics 526, 171–216.