13. Lattice-Boltzmann

For an implicit treatment of a solvent, ESPResSo can couple the molecular dynamics simulation to a lattice-Boltzmann fluid. The lattice-Boltzmann method (LBM) is a fast, lattice-based method that, in its “pure” form, allows to calculate fluid flow in different boundary conditions of arbitrarily complex geometries. Coupled to molecular dynamics, it allows for the computationally efficient inclusion of hydrodynamic interactions into the simulation. The focus of the ESPResSo implementation of the LBM is, of course, the coupling to MD and therefore available geometries and boundary conditions are somewhat limited in comparison to “pure” LB codes.

Here we restrict the documentation to the interface. For a more detailed description of the method, please refer to the literature.

Note

Please cite Godenschwager et al. [2013] and Bauer et al. [2021] (BibTeX keys godenschwager13a and bauer21a in doc/bibliography.bib) if you use the LB fluid. When generating your own kernels with pystencils and lbmpy, please also cite Bauer et al. [2019] and Bauer et al. [2021] (BibTeX key bauer19a resp. bauer21b in doc/bibliography.bib).

Note

Requires external feature WALBERLA, enabled with the CMake option -D ESPRESSO_BUILD_WITH_WALBERLA=ON.

13.1. Setting up a LB fluid

The following minimal example illustrates how to use the LBM in ESPResSo:

import espressomd
import espressomd.lb
system = espressomd.System(box_l=[10, 20, 30])
system.time_step = 0.01
system.cell_system.skin = 0.4
lbf = espressomd.lb.LBFluidWalberla(agrid=1.0, density=1.0, kinematic_viscosity=1.0, tau=0.01)
system.lb = lbf
system.integrator.run(100)

To use the GPU-accelerated variant, replace line 6 in the example above by:

lbf = espressomd.lb.LBFluidWalberlaGPU(agrid=1.0, density=1.0, kinematic_viscosity=1.0, tau=0.01)

Note

Feature CUDA required for the GPU-accelerated variant

To use the (much faster) GPU implementation of the LBM, use LBFluidWalberlaGPU in place of LBFluidWalberla. Please note that the GPU implementation uses single precision floating point operations. This decreases the accuracy of calculations compared to the CPU implementation. In particular, due to rounding errors, the fluid density decreases over time, when external forces, coupling to particles, or thermalization is used. The loss of density is on the order of \(10^{-12}\) per time step.

The command initializes the fluid with a given set of parameters. It is also possible to change parameters on the fly, but this will only rarely be done in practice. Before being able to use the LBM, it is necessary to set up a box of a desired size. The parameter is used to set the lattice constant of the fluid, so the size of the box in every direction must be a multiple of agrid.

In the following, we discuss the parameters that can be supplied to the LBM in ESPResSo. The detailed interface definition is available at LBFluidWalberla.

The LB scheme and the MD scheme are not synchronized: In one LB time step typically several MD steps are performed. This allows to speed up the simulations and is adjusted with the parameter tau, the LB time step. The parameters density and kinematic_viscosity set up the density and kinematic viscosity of the LB fluid in (usual) MD units. Internally the LB implementation works with a different set of units: all lengths are expressed in agrid, all times in tau and so on. LB nodes are located at 0.5, 1.5, 2.5, etc. (in terms of agrid). This has important implications for the location of hydrodynamic boundaries which are generally considered to be halfway between two nodes for flat, axis-aligned walls. For more complex boundary geometries, the hydrodynamic boundary location deviates from this midpoint and the deviation decays to first order in agrid. The LBM should not be used as a black box, but only after a careful check of all parameters that were applied.

In the following, we describe a number of optional parameters. Thermalization of the fluid (and particle coupling later on) can be activated by providing a non-zero value for the parameter kT. Then, a seed has to be provided for the fluid thermalization:

lb = espressomd.lb.LBFluidWalberla(kT=1.0, seed=134, ...)

The parameter ext_force_density takes a three dimensional vector as an array_like of float, representing a homogeneous external body force density in MD units to be applied to the fluid.

Before running a simulation at least the following parameters must be set up: agrid, tau, kinematic_viscosity, density.

To detach the LBM solver, use this syntax:

system.lb = None

13.1.1. Performance considerations

The CPU implementation of the LB has an extra flag single_precision to use single-precision floating point values. These are approximately 10% faster than double-precision, at the cost of a small loss in precision.

To enable vectorization, run cmake . -D ESPRESSO_BUILD_WITH_WALBERLA_AVX=ON. The SIMD kernels have better performance over the regular kernels, because they carry out the mathematical operations in batches of 4 values at a time (in double-precision mode) or 8 values at a time (in single-precision mode) along the x-axis. An AVX2-capable microprocessor is required; to check if your hardware supports it, run the following command:

lscpu | grep avx2

13.2. Checkpointing

lb.save_checkpoint(path, binary)
lb.load_checkpoint(path, binary)

The first command saves all of the LB fluid nodes’ populations to an ASCII (binary=False) or binary (binary=True) format respectively. The second command loads the LB fluid nodes’ populations. In both cases path specifies the location of the checkpoint file. This is useful for restarting a simulation either on the same machine or a different machine. Some care should be taken when using the binary format as the format of doubles can depend on both the computer being used as well as the compiler. One thing that one needs to be aware of is that loading the checkpoint also requires the user to reuse the old forces. This is necessary since the coupling force between the particles and the fluid has already been applied to the fluid. Failing to reuse the old forces breaks momentum conservation, which is in general a problem. It is particularly problematic for bulk simulations as the system as a whole acquires a drift of the center of mass, causing errors in the calculation of velocities and diffusion coefficients. The correct way to restart an LB simulation is to first load in the particles with the correct forces, and use:

system.integrator.run(steps=number_of_steps, reuse_forces=True)

upon the first call integrator.run. This causes the old forces to be reused and thus conserves momentum.

13.3. Interpolating velocities

To get interpolated velocity values between lattice nodes, the function:

lb.get_interpolated_velocity(pos=[1.1, 1.2, 1.3])

with a single position pos as an argument can be used.

The interpolation is done linearly between the nearest 8 LB nodes.

A note on boundaries: both interpolation schemes don’t take into account the physical location of the boundaries (e.g. in the middle between two nodes for a planar wall) but will use the boundary node slip velocity at the node position. This means that every interpolation involving at least one boundary node will introduce an error.

13.4. Coupling LB to a MD simulation

MD particles can be coupled to a LB fluid through frictional coupling. The friction force

\[F_{i,\text{frict}} = - \gamma (v_i(t)-u(x_i(t),t))\]

depends on the particle velocity \(v\) and the fluid velocity \(u\). It acts both on the particle and the fluid (in opposite direction). Because the fluid is also affected, multiple particles can interact via hydrodynamic interactions. As friction in molecular systems is accompanied by fluctuations, the particle-fluid coupling has to be activated through the Lattice-Boltzmann thermostat (see more detailed description there). A short example is:

system.thermostat.set_lb(LB_fluid=lbf, seed=123, gamma=1.5)

where lbf is an instance of either LBFluidWalberla or LBFluidWalberlaGPU, gamma the friction coefficient and seed the seed for the random number generator involved in the thermalization.

13.4.1. LB and LEbc

Lees–Edwards boundary conditions (LEbc) are supported by both LB implementations, which follow the derivation in [Wagner and Pagonabarraga, 2002]. Note, that there is no extra python interface for the use of LEbc with the LB algorithm: all the necessary information is internally derived from the currently active MD LEbc protocol in system.lees_edwards.protocol. Therefore, the MD LEbc must be set before the LB actor is instantiated. Use the Off if the system should have no shearing initially; this action will initialize the shear axes, and when the LB actor is instantiated, the Lees-Edwards collision kernels will be used instead of the default ones.

Note

At the moment, LB only supports the case shear_plane_normal="y".

13.5. Reading and setting properties of single lattice nodes

Appending three indices to the lb object returns an object that represents the selected LB grid node and allows one to access all of its properties:

lb[x, y, z].density              # fluid density (scalar)
lb[x, y, z].velocity             # fluid velocity (3-vector)
lb[x, y, z].pressure_tensor      # fluid pressure tensor (symmetric 3x3 matrix)
lb[x, y, z].pressure_tensor_neq  # fluid pressure tensor non-equilibrium part (symmetric 3x3 matrix)
lb[x, y, z].is_boundary          # flag indicating whether the node is fluid or boundary (boolean)
lb[x, y, z].population           # LB populations (19-vector, check order from the stencil definition)

All of these properties can be read and used in further calculations. Only the property population can be modified. The indices x, y, z are integers and enumerate the LB nodes in the three Cartesian directions, starting at 0. To modify is_boundary, refer to Setting up boundary conditions.

Example:

print(lb[0, 0, 0].velocity)
lb[0, 0, 0].density = 1.2

The first line prints the fluid velocity at node (0 0 0) to the screen. The second line sets this fluid node’s density to the value 1.2. Use negative indices to get nodes starting from the end of the lattice.

The nodes can be read and modified using slices. Example:

print(lb[0:4:2, 0:2, 0].velocity)
lb[0:4:2, 0:2, 0].density = [[[1.1], [1.2]], [[1.3], [1.4]]]

The first line prints an array of shape (2, 2, 1, 3) with the velocities of nodes (0 0 0), (0 1 0), (2 0 0), (2 1 0). The second line updates these nodes with densities ranging from 1.1 to 1.4. You can set either a value that matches the length of the slice (which sets each node individually), or a single value that will be copied to every node (e.g. a scalar for density, or an array of length 3 for the velocity).

The LB pressure tensor from property pressure_tensor is calculated as \(\Pi = \rho c_s^2 \mathbb{1} + \rho \mathbf{u} \otimes \mathbf{u}\) with \(\rho\) the fluid density at a particular node, \(\mathbf{u}\) the fluid velocity at a particular node, \(c_s\) the speed of sound and \(\mathbb{1}\) the identity matrix. The non-equilibrium part from property pressure_tensor_neq is defined as \(\Pi^{\text{neq}} = \rho \mathbf{u} \otimes \mathbf{u}\).

13.6. VTK output

The waLBerla library implements a globally-accessible VTK registry. A VTK stream can be attached to a LB actor to periodically write one or multiple fluid field data into a single file using VTKOutput:

vtk_obs = ["density", "velocity_vector"]
# create a VTK callback that automatically writes every 10 LB steps
lb_vtk = espressomd.lb.VTKOutput(
    identifier="lb_vtk_automatic", observables=vtk_obs, delta_N=10)
lb.add_vtk_writer(vtk=lb_vtk)
self.system.integrator.run(100)
# can be deactivated
lb_vtk.disable()
self.system.integrator.run(10)
lb_vtk.enable()
# create a VTK callback that writes only when explicitly called
lb_vtk_on_demand = espressomd.lb.VTKOutput(
    identifier="lb_vtk_now", observables=vtk_obs)
lb.add_vtk_writer(vtk=lb_vtk_on_demand)
lb_vtk_on_demand.write()

Currently supported fluid properties are the density, velocity vector and pressure tensor. By default, the properties of the current state of the fluid are written to disk on demand. To add a stream that writes to disk continuously, use the optional argument delta_N to indicate the level of subsampling. Such a stream can be deactivated.

The VTK format is readable by visualization software such as ParaView 1 or Mayavi2 2, as well as in ESPResSo (see Reading VTK files). If you plan to use ParaView for visualization, note that also the particle positions can be exported using the VTK format (see writevtk()).

Important: these VTK files are written in multi-piece format, i.e. each MPI rank writes its local domain to a new piece in the VTK uniform grid to avoid a MPI reduction. ParaView can handle the topology reconstruction natively. However, when reading the multi-piece file with the Python vtk package, the topology must be manually reconstructed. In particular, calling the XML reader GetOutput() method directly after the update step will erase all topology information. While this is not an issue for VTK files obtained from simulations that ran with 1 MPI rank, for parallel simulations this will lead to 3D grids with incorrectly ordered data. Automatic topology reconstruction is available through VTKReader:

import pathlib
import tempfile
import numpy as np
import espressomd
import espressomd.lb
import espressomd.io.vtk

system = espressomd.System(box_l=[12., 14., 10.])
system.cell_system.skin = 0.4
system.time_step = 0.1

lbf = espressomd.lb.LBFluidWalberla(
    agrid=1., tau=0.1, density=1., kinematic_viscosity=1.)
system.lb = lbf
system.integrator.run(10)

vtk_reader = espressomd.io.vtk.VTKReader()
label_density = "density"
label_velocity = "velocity_vector"
label_pressure = "pressure_tensor"

with tempfile.TemporaryDirectory() as tmp_directory:
    path_vtk_root = pathlib.Path(tmp_directory)
    label_vtk = "lb_vtk"
    path_vtk = path_vtk_root / label_vtk / "simulation_step_0.vtu"

    # write VTK file
    lb_vtk = espressomd.lb.VTKOutput(
        identifier=label_vtk, delta_N=0,
        observables=["density", "velocity_vector", "pressure_tensor"],
        base_folder=str(path_vtk_root))
    lbf.add_vtk_writer(vtk=lb_vtk)
    lb_vtk.write()

    # read VTK file
    vtk_grids = vtk_reader.parse(path_vtk)
    vtk_density = vtk_grids[label_density]
    vtk_velocity = vtk_grids[label_velocity]
    vtk_pressure = vtk_grids[label_pressure]
    vtk_pressure = vtk_pressure.reshape(vtk_pressure.shape[:-1] + (3, 3))

    # check VTK values match node values
    lb_density = np.copy(lbf[:, :, :].density)
    lb_velocity = np.copy(lbf[:, :, :].velocity)
    lb_pressure = np.copy(lbf[:, :, :].pressure_tensor)
    np.testing.assert_allclose(vtk_density, lb_density, rtol=1e-10, atol=0.)
    np.testing.assert_allclose(vtk_velocity, lb_velocity, rtol=1e-7, atol=0.)
    np.testing.assert_allclose(vtk_pressure, lb_pressure, rtol=1e-7, atol=0.)

13.7. Choosing between the GPU and CPU implementations

ESPResSo contains an implementation of the LBM for NVIDIA GPUs using the CUDA framework. On CUDA-supporting machines this can be activated by compiling with the feature CUDA. Within the Python script, the LBFluidWalberla object can be substituted with the LBFluidWalberlaGPU object to switch from CPU based to GPU based execution. For further information on CUDA support see section CUDA acceleration.

The following minimal example demonstrates how to use the GPU implementation of the LBM in analogy to the example for the CPU given in section Setting up a LB fluid:

import espressomd
system = espressomd.System(box_l=[10, 20, 30])
system.time_step = 0.01
system.cell_system.skin = 0.4
lbf = espressomd.lb.LBFluidWalberlaGPU(agrid=1.0, density=1.0, kinematic_viscosity=1.0, tau=0.01)
system.lb = lbf
system.integrator.run(100)

13.8. Electrohydrodynamics

Note

This needs the feature LB_ELECTROHYDRODYNAMICS.

If the feature is activated, the lattice-Boltzmann code can be used to implicitly model surrounding salt ions in an external electric field by having the charged particles create flow.

For that to work, you need to set the electrophoretic mobility (multiplied by the external \(E\)-field) \(\mu E\) on the particles that should be subject to the field. This effectively acts as a velocity offset between the particle and the LB fluid.

For more information on this method and how it works, read the publication Hickey et al. [2010].

13.9. Setting up boundary conditions

Currently, only the so-called “link-bounce-back” algorithm for boundary nodes is available. This creates a boundary that is located approximately midway between lattice nodes. With no-slip boundary conditions, populations are reflected back. With slip velocities, the reflection is followed by a velocity interpolation. This allows to create shear flow and boundaries “moving” relative to each other.

Under the hood, a boundary field is added to the blockforest, which contains pre-calculated information for the reflection and interpolation operations.

13.9.1. Per-node boundary conditions

One can set (or update) the slip velocity of individual nodes:

import espressomd.lb
system = espressomd.System(box_l=[10.0, 10.0, 10.0])
system.cell_system.skin = 0.1
system.time_step = 0.01
lbf = espressomd.lb.LBFluidWalberla(agrid=0.5, density=1.0, kinematic_viscosity=1.0, tau=0.01)
system.lb = lbf
# make one node a boundary node with a slip velocity
lbf[0, 0, 0].boundary = espressomd.lb.VelocityBounceBack([0, 0, 1])
# update node for no-slip boundary conditions
lbf[0, 0, 0].boundary = espressomd.lb.VelocityBounceBack([0, 0, 0])
# remove boundary conditions
lbf[0, 0, 0].boundary = None

13.9.2. Shape-based boundary conditions

Adding a shape-based boundary is straightforward:

import espressomd.lb
import espressomd.shapes
system = espressomd.System(box_l=[10.0, 10.0, 10.0])
system.cell_system.skin = 0.1
system.time_step = 0.01
lbf = espressomd.lb.LBFluidWalberla(agrid=0.5, density=1.0, kinematic_viscosity=1.0, tau=0.01)
system.lb = lbf
# set up shear flow between two sliding walls
wall1 = espressomd.shapes.Wall(normal=[+1., 0., 0.], dist=2.5)
lbf.add_boundary_from_shape(shape=wall1, velocity=[0., +0.05, 0.])
wall2 = espressomd.shapes.Wall(normal=[-1., 0., 0.], dist=-(system.box_l[0] - 2.5))
lbf.add_boundary_from_shape(shape=wall2, velocity=[0., -0.05, 0.])

The velocity argument is optional, in which case the no-slip boundary conditions are used. For a position-dependent slip velocity, the argument to velocity must be a 4D grid (the first three dimensions must match the LB grid shape, the fourth dimension has size 3 for the velocity).

The LB boundaries use the same shapes objects to specify their geometry as constraints do for particles. This allows the user to quickly set up a system with boundary conditions that simultaneously act on the fluid and particles. For a complete description of all available shapes, refer to espressomd.shapes.

When using shapes, keep in mind the lattice origin is offset by half a grid size from the box origin. For illustration purposes, assuming agrid=1, setting a wall constraint with dist=1 and a normal vector pointing along the x-axis will set all LB nodes in the left side of the box as boundary nodes with thickness 1. The same outcome is obtained with dist=1.49, but with dist=1.51 the thickness will be 2.

13.10. Prototyping new LB methods

Start by installing the code generator dependencies:

python3 -m pip install --user -c requirements.txt numpy sympy lbmpy pystencils islpy

Next, edit the code generator script to configure new kernels, then execute it:

python3 maintainer/walberla_kernels/generate_lb_kernels.py

The script takes optional arguments to control the CPU or GPU architecture, as well as the floating-point precision. The generated source code files need to be written to src/walberla_bridge/src/lattice_boltzmann/generated_kernels/. These steps can be automated with the convenience shell functions documented in maintainer/walberla_kernels/Readme.md. Edit the CMakeLists.txt file in the destination folder to include the new kernels in the build system. Then, adapt src/walberla_bridge/src/lattice_boltzmann/LBWalberlaImpl.hpp to use the new LB kernels.

1

https://www.paraview.org/

2

http://code.enthought.com/projects/mayavi/