Particle Tracing for Fluid Flow
The Particle Tracing for Fluid Flow interface (), found under the Fluid Flow branch () in the Model Wizard, computes the motion of particles in a background fluid. Particle motion can be driven by built-in forces including drag, gravitational, electric, magnetic, and acoustophoretic radiation forces. You can also apply custom user-defined forces, model particle-particle interactions, and solve for particle mass and temperature.
Treatment of Particle Inertia
The Particle Tracing for Fluid Flow interface includes dedicated formulations for tracking inertial particles or neglecting the inertial terms. A full treatment of inertial particles means that the acceleration of the particle in the surrounding fluid is fully resolved in time. For very small particles (usually around a micron or smaller), this acceleration takes place over such a short time that it is more computationally efficient to simply assume that the particles are always in an equilibrium state, with the drag force perfectly counterbalancing all other applied forces such as gravity. This allows you to take significantly larger time steps when tracking small particles in a fluid.
Basic Requirements
In order for the particle tracing approach to be valid, the system should be a dilute or dispersed flow. This means that the particles should occupy a very small fraction of the volume of the surrounding fluid, generally less than 1%. When the volume fraction of the particles is not small, the fluid system is categorized as a dense flow and a different modeling approach is required.
It is important to realize that with the particle tracing approach, particles do not displace the fluid they occupy. Furthermore, the finite radius of the particle is not usually taken into account when detecting and applying particle-wall interactions.
Flow Regimes
Motion of microscopic and macroscopic sized particles is typically dominated by the drag force acting on particles immersed in a fluid. There are two phases in the system: a discrete phase consisting of bubbles, particles, or droplets, and a continuous phase in which the particles are immersed. The velocity of the continuous phase is sometimes entered as an expression but is most often computed using a fluid flow physics interface such as Laminar Flow (), Turbulent Flow, k-ε (), or Creeping Flow ().
Comet tail plot of particle trajectories through a laminar static mixer (colored). In addition, Poincaré maps show the deviation of particle trajectories from their initial position (blue & red).
Sparse Flow
In a sparse flow, the continuous phase affects the motion of the particles via the drag force, but the particles do not have enough inertia to significantly perturb the continuous phase. This is usually true when the particles in the discrete phase are very small or have relatively low number density. This is often referred to as unidirectional coupling. When modeling such a system, it is usually most efficient to solve for the continuous phase first, then compute the trajectories of the discrete particles in a separate study. For example, in the following plot of a static mixer, the stationary velocity field and pressure were first solved using a stationary study, and then the particle trajectories were computed in a time-dependent study.
Dilute Flow
In a dilute flow the continuous phase affects the motion of the particles, and the particle motion in turn disrupts the continuous phase. This is often referred to as a bidirectional coupling. Compared to a sparse flow, the particles in a dilute flow typically have more inertia, due to the particles being bigger, denser, or more numerous.
The Fluid-Particle Interaction () interface automatically sets up a bidirectional coupling between the Particle Tracing for Fluid Flow interface () and the Laminar Flow interface (). The particles exert a volume force on the fluid at their location, equal in magnitude and opposite in direction to the total drag force exerted on the particles. The computational demand is significantly higher when modeling dilute flows than sparse flows. Before setting up a bidirectional coupling, it is often beneficial to first set up the model with a unidirectional coupling and then determine whether the effect of the particles on the surrounding fluid is significant.
Dispersed Flow
In a dispersed flow, the density of particles is greater than in the dilute flow described in the previous section, but their volume fraction is still lower than in a dense flow. The dispersed flow is the upper limit of the applicability of a Lagrangian particle tracking approach. In addition to the bidirectional fluid-particle interaction mentioned in the previous section, particle-particle interactions may also need to be taken into account. This is sometimes referred to as four-way coupling. Particle-particle interactions can be included in models but the following limitations apply:
Particle Diffusion
Particle motion in a fluid is a combination of advective and diffusive transport. Advection is the bulk transport of particles by the mean fluid velocity via the drag force. Advection is typically deterministic; given a particle as a specified location in a laminar flow field, the motion of that particle at future times is completely predictable and reproducible (allowing, of course, some small differences due to the discrete time stepping used by the time dependent study). If 100 identical particles pass through the same point in a stationary flow field, each having the same velocity at that point, then all 100 particles will follow the same path.
Diffusion, on the other hand, is implemented as a stochastic term that can give different results for each particle in the simulation, even if the initial conditions are the same. There are two main mechanisms for diffusive transport in the Particle Tracing for Fluid Flow interface: molecular diffusion and turbulent dispersion.
Molecular Diffusion: The Brownian Force
Molecular diffusion is most significant for extremely small particles, on the order of 0.1 μm or smaller. When such small particles are in a fluid of nonzero absolute temperature, the thermal motion of individual atoms or molecules in that fluid causes them to collide randomly with the particles in directions other than that of the mean flow velocity. These random collisions can cause particles to spread out over time, a phenomenon known as Brownian motion.
The COMSOL implementation of the Brownian force involves seeding unique random numbers for each particle at each time step taken by the solver. Therefore, a sufficiently large number of time steps should be taken so that the results show statistical convergence. In a similar fashion, the number of model particles should be sufficiently large so that the average behavior of the particles can be clearly seen.
Brownian motion of particles that are released from a single point. They diffuse outward. Upper left t = 0 s, upper right t = 10 s, lower left t = 30 s, and lower right t = 100 s.
Turbulent Dispersion
Turbulence is the dominant mechanism for diffusive transport of macroscopic particles in a high Reynolds number flow. Unlike Brownian motion, turbulent dispersion can remain significant even for particles at the micron scale or larger.
For laminar flow, the drag force only contributes to the advective transport of the particles,
However, if the flow is turbulent, then the fluid velocity u at the particle’s position consists of a mean flow term and an instantaneous fluid velocity perturbation,
This is because most of the turbulence models in COMSOL Multiphysics are variants of the Reynolds Averaged Navier Stokes (RANS) model. In a RANS model, eddies in a turbulent flow are not modeled explicitly; instead, information about the magnitude and lifetime of eddies is only retained in a statistical sense, often through the use of transport equations for additional dependent variables.
Several different expressions for the perturbation term exist, the simplest being
where ζ is a random unit vector and k is the turbulent kinetic energy. Some turbulence models, such as the k-ε model, solve for k as an additional degree of freedom; it represents the amount of energy associated with eddies in the flow.
Unlike molecular diffusion, turbulent diffusion has a finite resolution in time. The turbulent velocity perturbation experienced by a certain particle is assumed to have the same magnitude and direction for a finite time interval, based on the lesser of two time scales: the eddy lifetime (over which eddies or vortices in the flow die out and form again) or the eddy crossing time (over which particles have sufficient inertia to be thrown out of one eddy and into another one). For extremely small particles, the eddy crossing time may grow so large as to become irrelevant.
If the time step taken by the solver is made extremely small, then eventually the turbulent perturbation term between successive iterations no longer consists of uncorrelated random vectors. The built-in discrete random walk and continuous random walk models compensate for this effect by only seeding the random perturbation term at discrete times based on the turbulent dissipation rate ε, if it is available. Therefore, it is important to take a sufficiently small time step so that the changes in this random number seed can be resolved.
Rarefaction Effects
Some forces, such as the drag and thermophoretic forces, include high Knudsen number corrections that improve accuracy when the particles are extremely small or the surrounding fluid is a highly rarefied gas. These rarefaction effects begin to become significant when the Knudsen number, defined as the ratio of the mean free path of surrounding molecules to the particle diameter, is greater than 0.1.
Wall Effects
The drag force and lift force include correction terms to account for the presence of nearby walls, improving their accuracy for channel flows. For turbulent wall-bounded flows, an anisotropic turbulent dispersion model is available, taking different fluctuations in the streamwise, spanwise, and wall normal directions.
Other Built-in Forces
In addition to the drag force and Brownian force, the Particle Tracing for Fluid Flow interface includes a wide variety of other built-in forces that can be used to create high-fidelity multiphysics models, including the following:
Dielectrophoretic separation: of red blood cells (red) and platelets (blue) in a flow channel. The fluid velocity is shown in grayscale. The electric field is indicated by yellow streamlines.
Computing Particle Mass and Temperature
You can solve additional equations for the particle mass and temperature along each particle trajectory. When solving for temperature, built-in features for radiative and convective heating and cooling become available. You can also add user-defined heat sources and sinks.
If the particles are heated or cooled due to radiative or convective heat exchange with their surroundings, then optionally you can compute the dissipated particle heat as a volumetric source or sink term. For example, if hot particles are cooled by being submerged in cold water, you can add the Dissipated Particle Heat node to compute the corresponding heat source term in the fluid, potentially enabling two-way coupled simulation of particle tracing in a nonisothermal flow.
When solving for particle mass, you can make particles grow or shrink as they propagate. When releasing particles, you can enter the initial particle mass directly or sample it from a distribution, such as a lognormal distribution with a specified Sauter mean diameter.
As an alternative to specifying particle growth rate directly, if the particles represent liquid droplets, you can use a dedicated feature to compute the evaporation rate for droplets of a given size and temperature. When solving for both particle mass and temperature, the built-in Stefan-Fuchs evaporation model can optionally include an evaporative cooling term.
Nozzles and Sprays
The Particle Tracing for Fluid Flow interface is not just limited to solid particles in liquids or gases, but can also be used to model sprays of liquid droplets.
A dedicated Nozzle feature is available to release a spray of droplets from points in the geometry. In addition, the Droplet Breakup feature can be used to split the model particles into smaller particles over time. Typically the breakup occurs due to instability on the droplet surface created by external drag forces, or due to sudden acceleration of the droplet.
The Droplet Sprays in Fluid Flow interface (), found under the Fluid Flow branch () in the Model Wizard, is a special case of the Particle Tracing for Fluid Flow interface () that uses default settings and features that are frequently used to model sprays of liquid droplets. This includes options to specify particle viscosity and surface tension as well as an instance of the Droplet Breakup feature.