- Dassault Systemes, Waltham, MA, United States
A theoretical formulation of lattice Boltzmann models on a general curvilinear coordinate system is presented. It is based on a volumetric representation so that mass and momentum are exactly conserved as in the conventional lattice Boltzmann on a Cartesian lattice. In contrast to some previously existing approaches for arbitrary meshes involving interpolation approximations among multiple neighboring cells, the current formulation preserves the fundamental one-to-one advection feature of a standard lattice Boltzmann method on a uniform Cartesian lattice. The new approach is built on the concept that a particle is moving along a curved path. A discrete space-time inertial force is derived so that the momentum conservation is exactly ensured for the underlying Euclidean space. We theoretically show that the new scheme recovers the Navier-Stokes equation in general curvilinear coordinates in the hydrodynamic limit, along with the correct mass continuity equation.
1 Introduction
Lattice Boltzmann Methods (LBM) have been developed as an advantageous method for computational fluid dynamics during past few decades [1, 2]. The underlying dynamics of a lattice Boltzmann system resides on the kinetic theory, in that it involves motion of many particles according to the Boltzmann equation [3, 4]. There are two fundamental dynamical processes in a basic Boltzmann kinetic system, namely advection and collision. In the advection process, a particle moves from one location to another according to its microscopic velocity, while in the collision process particles interact with each other obeying conservation laws and relax to equilibrium. In a standard LBM model, particle velocity takes on a discrete set of constant values, and the latter form exact links from one lattice site to its neighboring sites on a simple Bravais lattice corresponding to a three dimensional (3D) uniform cubical Cartesian mesh [5–7]. Therefore, having these discrete velocity values, a particle moves from one lattice site to a unique neighboring site per constant time interval (say
However, a uniform and cubical Cartesian mesh poses fundamental limitations. First of all, often in realistic fluid study one deals with a solid geometry with curved surfaces, and obviously a Cartesian mesh does not smoothly conform with such a geometric shape. Secondly, a flow usually has small scale structures concentrated in certain spatial locations and directions. For instance, in the turbulent boundary layer, the flow scale in the direction normal to the wall is much smaller than in the tangential direction or in the fluid region outside of the boundary layer. Consequently, the requirement on spatial resolution is significantly higher in the normal direction inside a boundary layer. A cubic Cartesian mesh does not provide the flexibility with different spatial resolutions in different directions. Therefore, it is of fundamental importance to extend the standard LBM for a general non-cubic Cartesian mesh.
There have been various attempts in the past to extend LBM on arbitrary meshes, including curvilinear mesh (cf [10–12]). The essential idea is to relax the aforementioned exact one-to-one advection between a pair of lattice sites while keeping the advection of particles according to the original constant discrete velocity values. As a consequence, on an arbitrary mesh, a particle after advecting from its original mesh site does not in general land on a single neighboring mesh site. Thus its location needs to be distributed onto a set of neighboring mesh sites via interpolation. As mentioned above, such an effective “one-to-many” advection process destroys the preciseness of a Lagrangian advection characteristics thus resulting in a significantly increased numerical dissipation.
It is of fundamental interest to formulate an LBM on a general mesh that preserves the exact one-to-one feature of the particle advection process. A possible way to accomplish this is to construct the micro-dynamic process on a non-Euclidean space represented by a general curvilinear coordinate system based on Riemann geometry [13]. In such a non-Euclidean space, a constant particle velocity corresponds to a curved and spatially varying path in the underlying Euclidean space. Indeed, the continuum Boltzmann kinetic theory on a Riemannian manifold can be theoretically described [14, 15]. The key difference between a constant particle velocity on a Euclidean and a non-Euclidean space is that the latter is accompanied with an inertial force, due to the 1st law of Newton in Euclidean space. According to this concept, it is entirely conceivable to formulate LBM models on a cubic Cartesian lattice in non-Euclidean space that corresponds to a general curvilinear mesh in the Euclidean space. The first such attempts were made by Mendosa et al. in a series of papers (see [16–18]). The significance of their work is that it established this fundamental approach to LBM on curvilinear meshes via a Riemann geometric framework.
As Mendosa et al., we formulate LBM on a general curvilinear coordinate system via the concept of Riemannian geometry. There are several critical differences between our approach presented here and that of Mendosa et al. First of all, we adopt a volumetric formulation so that the mass and momentum conservations are exactly ensured. The resulting continuity equation of mass is automatically shown to have the correct form in curvilinear coordinates, and there is no need to introduce any artificial mass source term to correct for any artifacts in the resulting hydrodynamics. Secondly, as in the continuum kinetic theory on a manifold, the only external source term in the extended LBM is associated with an effective inertial force due to curvilinearity. It contributes no mass source. Moreover, such a force term is constructed fully self-consistently within the discrete lattice formulation. This force form in discrete space and time recovers asymptotically the one in the continuum kinetic theory in the hydrodynamic limit. It does not rely on any outside analytical forms borrowed from the continuum kinetic theory [14, 15]. Thirdly, the inertial force enforces the exact momentum conservation for the underlying Euclidean space in the discrete space-time LBM model. Lastly, we demonstrate that the force term must be constructed properly so that it adds momentum in the system at a proper discrete time moment in order to produce the correct resulting Navier-Stokes hydrodynamics to the viscous order. Indeed, the correct Navier-Stokes equation is fully recovered in the hydrodynamic limit with the new formulation without introducing any extra correction terms.
To summarize, our goal is to theoretically demonstrate how discrete kinetic theories can be formulated in curvilinear coordinates in a way that mass and momentum are exactly conserved, and also the correct hydrodynamics is recovered in the microscopic limit. We hope the reader may find such discrete non-equilibrium statistical systems interesting, similar to discrete systems found in other branches of physics, e.g., spin models on a lattice in the equilibrium statistical theory, lattice field theories, etc. In addition to fundamental interest, non-equilibrium discrete systems like LBM may be useful for applications in numerical analysis and for practical use. Computational aspects such as scheme stability, numerical dissipation, performance, and boundary conditions treatment, as well as numerical verification of theoretical results, are outside of the scope of this study. We also restrict ourselves to one of the simplest situations, namely the low speed isothermal flow, which admits a closed self-contained analytical treatment which explicates the main features of volumetric formulation in curvilinear coordinates. Possible extensions and applications of this work are briefly discussed in Section 4 below.
The description of non-equilibrium dynamics and transport at the spatio-temporal scales not treatable by hydrodynamics can often benefit from using kinetic theory methods. Here we show how to construct lattice kinetic theories in a general coordinate system that possess exact conservation laws at both kinetic and continuum scales, as well as correct macroscopic limit behavior. In addition to the fundamental interest, this can be useful for describing interactions at interfaces between different fluids or materials with complex geometry.
The subsequent sections are organized as follows. We describe in Section 2 the formulation of the new LBM on a general curvilinear mesh. In Section 3, we provide a detailed theoretical derivation and show that the new LBM indeed produces the correct Navier-Stokes equation in general curvilinear coordinates. In order for the paper to be more self-contained, we provide, in Supplementary Appendices SA1–SA3 basic theoretical description of the continuum Boltzmann kinetic theory in curvilinear coordinates as in the literature [15], some fundamental properties [13], as well as a derivation of the Navier-Stokes hydrodynamic equation on a general curvilinear coordinate system.
2 Formulation of Lattice Boltzmann in Curvilinear Coordinates
This section is divided into two subsections. First, we construct the geometric quantities that are necessary for defining a general curvilinear coordinate system. Then we present the volumetric lattice Boltzmann formulation on a general curvilinear mesh.
2.1 Construction of Geometric Quantities on a Curvilinear Mesh
We describe how to construct a coordinate system for formulating kinetic theory on a curvilinear mesh. Let
It is easy to see that, under such a construction, the coordinate values
When a curvilinear mesh is provided, spatial locations of all the vertices
For instance,
and the inequality only becomes an equality everywhere if the curvilinear mesh is a uniform lattice (so that
Following the basic differential geometry concept, we now construct the basis tangent vectors at
where the scalar factor
and
Notice, in Eq. 3
as well as the volume J of the cell centered at
and we can always choose a proper handedness so that
with
Indeed, the basis tangent vectors and the co-tangent vectors are orthonormal to each other,
where
Clearly the inverse metric tensor is the inverse of the metric tensor,
and
Having these basic geometric quantities defined above, we can now introduce the lattice Boltzmann velocity vectors on a general curvilinear mesh, similar to the ones on a standard Cartesian lattice,
Here and thereafter, the summation convention is used for repeated Roman indices. For convenience, in subsequent derivations we adopt the ‘lattice units’ convention so that
As shown in the subsequent sections, a set of necessary moment isotropy and normalization conditions must be satisfied in order to recover the correct full Navier-Stokes hydrodynamics [6, 7, 19, 20]. These are, when there exists a proper set of constant weights
where the constant
Lastly, we define a set of specific geometric quantities below that are essential for the extended LBM model,
It is easily seen that
Once a curvilinear mesh is specified, all the geometric quantities above are fully determined and can thus be pre-computed before starting a dynamic LBM simulation.
2.2 Volumetric Lattice Boltzmann Model on a Curvilinear Mesh
Although the basic theoretical framework of the work is more general, for simplicity of describing the basic concept, we present the formulation for the so called isothermal LBM in this section. Similar to the standard lattice Boltzmann equation (LBE), we write the evolution of particle distribution below [1, 5, 21–23]
where
The particle density distribution function
where
where
and the velocity component value in the curvilinear coordinate system is given by,
Or for simplicity of notation, we define a three-dimensional fluid velocity array
It is immediately seen that Eq. 18 has the same form for the fluid velocity as that in the standard Cartesian lattice based LBM. Similarly, the momentum conservation of the collision term in Eq. 13 can also be written as,
Often in LBM the collision term takes on a linearized form [1, 20], namely
where
with τ being the collision relaxation time [5, 22–24]. In order to recover the correct Navier-Stokes hydrodynamics, besides Eqs 13,19 the collision matrix needs to satisfy an additional condition [8, 9, 20, 25–27],
Obviously the BGK form trivially satisfies such an additional property.
The extra term
We define the advection process as an exact one-to-one hop from one site to another as in the standard LBM, namely
where
In the above, the left side of the inequality sign is the momentum value at the end of advection process while the right side is the value at the beginning of the process. The inequality is there because the path of particles is curved (as well as stretched or compressed), so that its velocity at the end of the advection is changed from its original value. This is fundamentally different from that on a uniform Cartesian lattice, in that the particles have a constant velocity throughout the advection process. Consequently, we have the following inequalities in the overall momentum values,
where the right side of the unequal sign in Eq. 23 represents the total amount of momentum advected out of all the neighboring cells, while the left side is the total momentum arrived at cell q after the advection along the curved paths. Thus from Eqs 22,23, we see that the net momentum change via advection from all the neighboring cells into cell q is given by,
Similarly, we can realize that the net momentum change via advection out of cell q to all its neighboring cells is given by
Subsequently, if we impose the constraints on
then we preserve an exact mass conservation as well as recover the exact momentum conservation in discrete space-time for the underlying Euclidean space. Here
Using the geometric quantities defined in the previous subsection, it can be immediately shown that
where the geometric function
As to be demonstrated in the next section, in order to recover the full viscous Navier-Stokes equation, an additional constraint on the momentum flux also needs to be imposed below,
with
A specific form of
With a simple algebra, one can verify that it satisfies the moment constraints of Eqs 26, 28, 29. Notice, due to the appearance of
The final part for complete specifying the extended LBM on a curvilinear mesh is the form of the equilibrium distribution function, which needs to be defined appropriately in order to recover the correct Euler equation as well as the Navier-Stokes equation in curvilinear coordinates in the hydrodynamic limit. In particular, the following fundamental conditions on hydrodynamic moments must be realized,
In the above,
with
The equilibrium distribution form above is analogous to that of a low Mach number expansion of the Maxwell-Boltzmann distribution expressed in curvilinear coordinates [14]. It reduces to the standard LBM equilibrium distribution if the curvilinear mesh is a uniform Cartesian lattice so that
With all the quantities and dynamic properties defined above, we can show that the lattice Boltzmann Eq. 12 computed on the (non-Euclidean) uniform Cartesian lattice
3 Derivation of the Navier-Stokes Hydrodynamics in Curvilinear Coordinates
In this section, we provide a detailed derivation and show that the extended LBM presented above indeed produces the correct Navier-Stokes equations in general curvilinear coordinates. We rewrite the lattice Boltzmann Eq. 12) below,
Expanding it in both time and space up to the second order leads to
Now we introduce multiple scales in time and space based on the conventional Chapman-Enskog expansion procedure, [21, 28]
and
Here ϵ (
Likewise,
for they introduce no source of mass. Equating the same powers of ϵ, Eq. 36 leads to the following two equations
and
where the collision term has taken the linearized form in Eq. 20. Eq. 38 can be directly inverted,
Therefore, Eq. 39 can also be written as,
Taking the mass moment of Eqs 38, 41, and using mass conservation properties of
Based on the equilibrium moment definitions Eqs 37, 42 is equivalent to
For the next order, we have
Combining Eq. 43 and Eq. 44, we get
Substitute the definitions in Eqs 37, 45 becomes
where
is to be further discussed later in this section. Define the fluid velocity as
and since
Recognizing that
Taking the momentum moment of Eq. 38, we get,
Define (see also in Eq. 32)
together with Eq. 47, Eq. 51 becomes
Similarly, taking the momentum moment of Eq. 41,
Let
and use relation Eq. 40, we can derive
Using the collision matrix property Eq. 21 and its inverse, namely
then Eq. 56 becomes
Substitute
After some simple cancellations and rearrangements, it becomes
Using definitions in Eq. 32,
where in the above we have defined
Now let us examine the forcing terms
with
where
Similarly,
Consequently, when substitute Eqs 63, 64 in Eq. 62, we obtain, to the leading order
where
is the Christoffel symbol as defined in differential geometry [13]. Also in the above, the relationship Eq. 22 is used, namely
Furthermore, since
we have from Eq. 65 the following
Comparing terms of the same order in ϵ, we recognize that
and
Replacing
From the definition Eqs 52, 69, we see that
Similarly, let us also define
where
Hence, from Eq. 72, we have
Taking Eq. 73 to Eq. 53, we have for the leading order the equation below
On the other hand, taking Eqs 74, 76 to Eq. 61, we get in the viscous order
Combining Eqs. 77, 78 together and using the definition in Eq. 48, one obtains formally the Cauchy’s transport equation, namely
where
Eq. 79 can be further expressed in terms of hydrodynamic quantities. From the moment properties of the equilibrium distribution function in Eq. 32, we have
which has exactly the same form as that from the continuum kinetic theory. On the other hand, from Eq. 74 we have
where, from Eq. 32,
which has exactly the same form as that from the continuum kinetic theory. However, the term
with
Taking the long wave length limit, Eq. 84 in the continuum limit becomes
Therefore,
So we finally obtain from Eq. 81 that
Substituting Eqs 80, 87 into Eq. 79, comparing with eqn (B.9) in Supplementary Appendix SA2, together with the definitions of Eqs 80, 82 (having the same forms as that from the continuum kinetic theory), we finally arrive at exactly the same form as derived out of a continuum kinetic theory in curvilinear coordinates. The only difference between the two is that the collision time τ is replaced by
where the pressure
with the standard covariant derivative of the velocity component defined as
Using the standard definitions of differential operators in differential geometry [13], we can recognize that Eq. 88 is indeed the familiar Navier-Stokes equation in a generic coordinate-free operator representation,
4 Discussion
In this paper, we present a theoretical formulation of lattice Boltzmann models in a general curvilinear coordinate system. The formulation is an application of the Riemannian geometry for kinetic theory [14] to discrete space and time. Unlike some previous works [16–18], here we use a volumetric representation so that conservation laws are exactly ensured [11]. Furthermore, in the current formulation, we find that the main and the only additional source term in the extended LBM model is corresponding to the inertial force to ensure the exact momentum conservation in the underlying Euclidean space. This is the same as in the continuum kinetic theory. On the other hand, this forcing term needs to be applied at an appropriate discrete time in order to realize the correct viscous fluid effect associated with non-equilibrium physics. The equilibrium distribution function also needs to be properly modified that is directly analogous to the Maxwell-Boltzmann distribution on a curved space. Unlike the previous formulations [16–18], there are no other terms or treatments added to cancel any discrete artifacts. Through a detailed analysis, we have shown that the current LBM formulation recovers the correct Navier-Stokes behavior in the hydrodynamic limit, as long as a discrete lattice velocity set satisfying a sufficient order of isotropy is used. Extensive numerical validations of this extended LBM for various flows on various curvilinear meshes are to be presented in future publications. The main benefit of this kind of theoretical formulation in a general curvilinear coordinate system is its preservation of the key LBM one-to-one Lagrangian nature of particle advection. Not only this is desirable in algorithmic simplicity, as a standard LBM on a Cartesian lattice, it also has non-trivial implications for flows at finite Knudsen number [7–9]. Although the specific LBM model constructed here is for an isothermal fluid, the fundamental framework is directly extendable to more general fluid flow situations. Possible extensions in the future may include transport of scalars, complex fluids, finite Knudsen flows, as well as higher speed flows with energy and temperature dynamics in curved space. The latter is essential for study of highly compressible flows and flows with substantial temperature variations. Another interesting possible extension of the current theoretical formulation in the future is for a time varying coordinate system. This is useful particularly for studying of fluid flows around a dynamically deforming solid object.
Data Availability Statement
The original contributions presented in the study are included in the article/Supplementary Material, further inquiries can be directed to the corresponding author.
Author Contributions
The author confirms being the sole contributor of this work and has approved it for publication.
Conflict of Interest
Author HC was employed by the company Dassault Systemes.
Acknowledgments
I am grateful to Raoyang Zhang, Pradeep Gopalakrishnan, Ilya Staroselsky and Alexei Chekhlov for insightful discussions.
Supplementary Material
The Supplementary Material for this article can be found online at: https://www.frontiersin.org/articles/10.3389/fams.2021.691582/full#supplementary-material
References
1. Benzi, R, Succi, S, and Vergassola, M. The Lattice Boltzmann Equation: Theory and Applications. Phys Rep (1992). 222:145–97. doi:10.1016/0370-1573(92)90090-m
2. Chen, S, and Doolen, GD. Lattice Boltzmann Method for Fluid Flows. Annu Rev Fluid Mech (1998). 30:329–64. doi:10.1146/annurev.fluid.30.1.329
5. Qian, YH, D'Humières, D, and Lallemand, P. Lattice BGK Models for Navier-Stokes Equation. Europhys Lett (1992). 17:479–84. doi:10.1209/0295-5075/17/6/001
6. Chen, H, Goldhirsch, I, and Orszag, SA. Discrete Rotational Symmetry, Moment Isotropy, and Higher Order Lattice Boltzmann Models. J Sci Comput (2008). 34:87–112. doi:10.1007/s10915-007-9159-3
7. Shan, X, Yuan, X-F, and Chen, H. Kinetic Theory Representation of Hydrodynamics: a Way beyond the Navier-Stokes Equation. J Fluid Mech (2006). 550:413. doi:10.1017/s0022112005008153
8. Chen, H, Zhang, R, Staroselsky, I, and Jhon, M. Recovery of Full Rotational Invariance in Lattice Boltzmann Formulations for High Knudsen Number Flows. Physica A: Stat Mech its Appl (2006). 362:125–31. doi:10.1016/j.physa.2005.09.008
9. Zhang, R, Chen, H, and Shan, X. Efficient Kinetic Method for Fluid Simulation beyond the Navier-Stokes Equation. Phys Rew. E (2006). 74:046703. doi:10.1103/physreve.74.046703
10. He, X, Luo, L-S, and Dembo, M. Some Progress in Lattice Boltzmann Method Part I Nonuniform Mesh Grids. J Comput Phys (1996). 129:357–63. doi:10.1006/jcph.1996.0255
11. Chen, H. Volumetric Formulation of the Lattice Boltzmann Method for Fluid Dynamics: Basic Concept. Phys Rev E (1998). 58:3955–63. doi:10.1103/physreve.58.3955
12. Barraza, JAR, and Dieterding, R. Towards a Generalised Lattice-Boltzmann Method for Aerodynamic Simulations. J Comput Phys (2019). 45:101182. doi:10.1016/j.jocs.2020.101182
14. Vedenyapin, V, Sinitsyn, A, and Dulov, E. Kinetic Boltzmann, Vlasov and Related Equations. London: Elsevier (2011).
15. Love, PJ, and Cianci, D. From the Boltzmann Equation to Fluid Mechanics on a Manifold. Phil Trans R Soc A (2011). 369:2362–70. doi:10.1098/rsta.2011.0097
16. Mendoza, M, Debus, J-D, Succi, S, and Herrmann, HJ. Lattice Kinetic Scheme for Generalized Coordinates and Curved Spaces. Int J Mod Phys C (2014). 25:1441001. doi:10.1142/s0129183114410010
17. Debus, J-D, Mendoza, M, and Herrmann, HJ. Dean Instability in Double-Curved Channels. Phys Rev E (2014). 90:053308. doi:10.1103/physreve.90.053308
18. Debus, J-D, Mendoza, M, Succi, S, and Herrmann, HJ. Poiseuille Flow in Curved Spaces. Phys Rev E (2016). 93:043316. doi:10.1103/physreve.93.043316
19. Chen, H, and Shan, X, Fundamental Conditions for N-Th-Order Accurate Lattice Boltzmann Models, Physica D 237, issues: 14-17, 2003–8. (2007). doi:10.1016/j.physd.2007.11.010
20. Chen, H, Teixeira, C, and Molvig, K. Digital Physics Approach to Computational Fluid Dynamics: Some Basic Theoretical Features. Intl J Mod Phys C (1997). 9(8):675. doi:10.1142/s0129183197000576
21. Frisch, U, D’Humieres, D, Hasslacher, B, Lallemand, P, Pomeau, Y, and Rivet, JP. Lattice Gas Hydrodynamics in Two and Three Dimensions. Complex Syst (1987). 1:649–707.
22. Chen, S, Chen, H, Martnez, D, and Matthaeus, W. Lattice Boltzmann Model for Simulation of Magnetohydrodynamics. Phys Rev Lett (1991). 67:3776–9. doi:10.1103/physrevlett.67.3776
23. Chen, H, Chen, S, and Matthaeus, W. Recovery of the Navier-Stokes Equations Using a Lattice-Gas Boltzmann Method. Phys.Rev A (1992). 45:5339–42. doi:10.1103/physreva.45.r5339
24. Bhatnagar, PL, Gross, EP, and Krook, M. A Model for Collision Processes in Gases fsI Small Amplitude Processes in Charged and Neutral One-Component Systems. Phys Rev (1954). 94(3):511–25. doi:10.1103/physrev.94.511
25. Latt, J, and Chopard, B. Lattice Boltzmann Method with Regularized Pre-collision Distribution Functions. Math Comput Simulat (2006). 72(2-6):165. doi:10.1016/j.matcom.2006.05.017
26. Chen, H, Gopalakrishnan, P, and Zhang, R. Recovery of Galilean Invariance in thermal Lattice Boltzmann Models for Arbitrary Prandtl Number. Int J Mod Phys C (2014). 25(10):1450046. doi:10.1142/s0129183114500466
Keywords: lattice Boltzmann methods, curvilinear coordinates, Riemann geometry, hydrodynamics, non-equilibrium phenomena
Citation: Chen H (2021) Volumetric Lattice Boltzmann Models in General Curvilinear Coordinates: Theoretical Formulation. Front. Appl. Math. Stat. 7:691582. doi: 10.3389/fams.2021.691582
Received: 06 April 2021; Accepted: 10 May 2021;
Published: 16 June 2021.
Edited by:
Snezhana I. Abarzhi, University of Western Australia, AustraliaReviewed by:
Semion Sukoriansky, Ben-Gurion University of the Negev, IsraelXiaobo Nie, Halliburton, United States
Copyright © 2021 Chen. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.
*Correspondence: Hudong Chen, hudong.chen@3ds.com