Recently, there has been significant interest in formulations of macro-particle models using variational methods. This is attractive because many of the inherent pathologies of traditional particle-in-cell methods are avoided. Broadly, there are two approaches to constructing these models that on the surface appear incompatible: a Lagrangian method based on the Low Lagrangian and a Hamiltonian method based on the noncanonical Vlasov–Maxwell Poisson bracket. In both cases, a reduction is performed on the distribution function, which is replaced by a finite sum of macro-particles with a fixed spatial structure and definite momentum. The distribution is, thus, replaced as dynamical variable by a collection of macro-particle positions and momenta. Here, we resolve the apparent discrepancy between the two approaches and show (1) that the noncanonical Hamiltonian formulation can be transformed into a canonical Hamiltonian system; (2) that this canonical Hamiltonian system is identical to that obtained from the Lagrangian by Legendre transformation; and (3) that introducing a grid in the Lagrangian setting requires careful treatment and an unusual structure in the variational principle to account for the breaking of gauge invariance.
I. INTRODUCTION
Recently, there has been renewed interest in the variational formulation of macro-particle kinetic plasma models. Early work1,2 involved a reduction of the Low Lagrangian3 using point particles that suffer from significant computational noise. More modern efforts have been based on the phase space action,4 the Low Lagrangian,5–7 or the noncanonical bracket.8,9 Benefits of the variational approach include exact energy conservation, which precludes grid heating,10–12 reduced noise,7 and a rigorous theoretical foundation based on a direct connection to the underlying Maxwell–Vlasov system. While representing the potentials (or fields) with a truncated basis set is compatible with any of these formulations, discretizing space using a grid, which is straightforward in the Lagrangian formulation, seems incompatible with the noncanonical Hamiltonian approach. The use of a grid is essential for computationally performant implementations on large multi-node computer systems. An outstanding question was what are phenomenological differences between macro-particle models obtained from the Lagrangian vs the noncanonical Hamiltonian formulation? Here, we show that once a macro-particle reduction is performed, the noncanonical system can be transformed into a canonical Hamiltonian system and then to a Lagrangian system that is identical to that obtained from the Low Lagrangian by the same macro-particle reduction. Thus, we conclude all formulations are identical. Furthermore, we demonstrate spatial discretization using a grid in both the Lagrangian and canonical Hamiltonian formulations. The latter is of significant interest as it allows for the use of symplectic integrators13 on the combined macro-particle and field phase space.6,14 We show that the grid-based discretization breaks gauge invariance and, consequently, special care is needed in handling the variational principle. Finally, we give a general, minimalist basis formulation that preserves gauge invariance. As a special case, we demonstrate the use of a truncated Fourier representation for the potentials that preserves gauge invariance and conserves momentum.
II. LAGRANGIAN MACRO-PARTICLE REDUCTION
III. REDUCTION OF THE NONCANONICAL SYSTEM
A. Macro-particle reduction
B. Transformation to canonical coordinates
Previously, it has been shown5 that in one-dimension, the macro-particle reduction of the nonrelativistic Vlasov–Ampère system yields dynamical equations equivalent to those obtained from the Lagrangian formulation. Here, we show the equivalence of the two formulations in the general case. The presence of the magnetic field greatly complicates the analysis.
C. Connection to the Lagrangian formulation
D. Constructing the primary Hamiltonian
The primary Hamiltonian (80) and extended bracket (81) is the correct canonical formulation according to the Dirac–Bergmann algorithm, yet the noncanonical system only produced us the constrained Hamiltonian and its corresponding bracket. Since the noncanonical formulation begins with and , it has no need to represent the entire phase space for the potentials and, as can be seen by its Casimirs, is only valid on the constrained phase space. In the noncanonical formulation, we do not need to be a dynamical variable but simply “some scalar function that makes Gauss's Law true.” When starting from the Lagrangian, where the system is derived in terms of potentials, the actual dynamics of are significant and can be carried into the Hamiltonian picture via the primary Hamiltonian.
IV. GAUGE INVARIANCE DISCRETIZATION: TRUNCATED FOURIER BASIS
Previous work has shown that using a Fourier representation for the potential results in exact momentum conservation in the electrostatic case.5 It has also been shown,6 in one-dimension, that if the potentials are expanded in a truncated basis that exactly represents the derivative operator, then a discrete from of the continuity holds and the Lagrangian is gauge invariant and momentum is exactly conserved.
V. CHARGE CONSERVATION AND GAUGE SELECTION IN NON-GAUGE INVARIANT SYSTEMS
For the Lagrangian (3), the continuity equation is structural in the reduction (1) and not dynamic as it is, say, in the general Vlasov system. With the noncanonical formulation seen in Sec. III B, Gauss's Law does not appear as an equation of motion but rather as a constraint on the system and instead follows from the equations of motion as a consequence of and satisfying the continuity equation. While discretizations such as that in Sec. IV can maintain this property, representing the fields on a grid, which is desirable for computational performance, generically results in the discrete version of the continuity equation not being exactly satisfied.6 As we will see below, the failure of the continuity equation to hold implies the system is not gauge-invariant.
A. Varying a non-gauge invariant Lagrangian
Gauss's Law and charge conservation are effectively tied to one another in the continuous system. As was seen in the Lagrangian (74), the multiplier that enters the system to produce Gauss's Law in the Hamiltonian picture transforms into the structure needed to allow gauge transformations in the Lagrangian picture. Thus, when the discretization violates local charge conservation, the dynamical equations no longer imply Gauss's Law.
B. Enforcing Gauss's Law via Lagrange multipliers
VI. DISCRETIZING ON A GRID
A. Lagrangian formulation
B. Canonical formulation
VII. NUMERICAL EXAMPLES
Figure 1 shows (top row), (second row), (third row), and the error in Gauss's Law (bottom row), where . We compare , , and quadratic (first and third column) with , , and quartic (second and fourth column). While is insufficient resolution to capture the dynamics, the qualitative features are plausibly correct. The differences in the two cases is largely due to spatial resolution; taking , , quartic and the same time step as the high-resolution gives essentially the same results as the case in Fig. 1.
Electric fields, density, and the error in Gauss's Law for , , and quadratic (126), (a)–(d) (early time) and (i)–(l) (late time) and for , and quartic (128), (e)–(h) (early time) and (m)–(p) (late time): (a), (e), (i), and (m); (b), (f), (j), and (n); (c), (g), (k), and (o); and (d), (h), (l), and (p).
Electric fields, density, and the error in Gauss's Law for , , and quadratic (126), (a)–(d) (early time) and (i)–(l) (late time) and for , and quartic (128), (e)–(h) (early time) and (m)–(p) (late time): (a), (e), (i), and (m); (b), (f), (j), and (n); (c), (g), (k), and (o); and (d), (h), (l), and (p).
Figure 2 shows the RMS error (in z) in Gauss's Law for quadratic (top row) and quartic (bottom row) with (left column), (center column), and (right column) for various values of . For quadratic , the error in Gauss's Law exhibits significant dependence on as the resolution increases; for the range of , the error plateaus for (see Fig. 4). Quartic yields errors that are essentially independent of for all but the highest resolution considered. The error in Gauss's Law arises from the difference between and , thus, we expect the error to be smaller for that span more cells and to scale with due to second-order approximations of spatial derivatives. Both of these expectations are borne out in Fig. 2.
The RMS error in Gauss's Law for quadratic (126) (top row) and quartic (128) (bottom row) and a range of : (a) and (d); (b) and (e); and (c) and (f).
Figure 3 shows the absolute value for the same parameters as in Fig. 2. The departure of from its initial value of 0 is the consequence of the lack of translation invariance of the discretization. The error in is seen to be largely insensitive to both and to the nature of . For quadratic , the behavior for and is inconsistent with the other cases. We do not have an explanation for this observation.
The absolute value of for quadratic (126) (top row) and quartic, (128), (bottom row) and a range of : (a) and (d); (b) and (e); and (c) and (f).
Figures 4 and 5 show the RMS (in z and t) error in Gauss's Law and the RMS value of , respectively, for quadratic (top), cubic (center), and quartic (bottom) . The dashed gray line is proportional to from which we see that both quantities scale consistent with second-order spatial differencing. At fixed , the error in Gauss's Law will become independent of for small . As the macro-particle becomes broader and smoother, the quadratic scaling persists to smaller . We have verified that does not influence this behavior: we repeated the quadratic cases with the smallest value of used in our computations and we obtain virtually identical results to those of Fig. 4(a).
Again we see that the behavior of is nearly independent of the details of and ; exhibits nearly perfect scaling with . As with Fig. 3, for quadratic , the behavior for and is inconsistent with the other cases. There is an indication of this behavior in the cubic case as the curve appears to be slightly above the other lines for large . Understanding the interplay of spatial resolution and macro-particle shape will be the subject of future work.
The errors in Gauss's Law and momentum generally do not grow in time and scale as expected with spatial resolution. The error in Gauss's Law can be reduced to near machine precision by using the Lagrange multiplier however doing so involves a significant increase in computational cost as an elliptic equation, (115), would have to be solved at each stage of the time stepper. We do not expect imposing the Gauss's Law constraint to effect the error in momentum. Exploring the advantages and trade-off of imposing Gauss's Law as a constraint will be explored in a future publication.
VIII. CONCLUSIONS
We have shown that a macro-particle reduction applied to the Low Lagrangian or to the noncanonical Vlasov–Maxwell system results in equivalent dynamical systems with continuous space. In the case of the noncanonical system, a complicated coordinate transformation yielded a canonical Hamiltonian and bracket. Using the Dirac–Bergmann procedure, we constructed a canonical Hamiltonian system on the extended phase-space from the macro-particle reduced Lagrangian, which differs from the transform noncanonical system as the latter yields the constrained Hamiltonian. We considered two approaches to reducing these systems to a finite degree-of-freedom approximations suitable for computational implementation. We illustrated a general truncated basis function expansion approach that preserves gauge invariance (and, thus, Gauss's Law) and showed that the special case of a truncated Fourier representation of the potentials also conserves total momentum.
The use of a Fourier basis seems unavoidable if exact momentum conservation is desired. The basis must be an eigenfunction of the translation operator which seems to suggest that a Fourier representation is required.5,6 Provided the domain is large enough, the influence of periodic boundary conditions can be minimized. Of more concern is the all-to-all communications necessary a parallel implementation since all of the macro-particle couple to all field modes (in contrast to the localized coupling in a gridded mode). As a result, there currently appears to be no path to obtaining sufficient computational performance to allow use of a Fourier representation for production work. Nonetheless, the Fourier representation is valuable for assessing the quantitative errors introduced by the lack of exact momentum conservation.
We presented a grid-based discretization of both the Lagrangian and canonical Hamiltonian formulations in the Weyl gauge. The spatial grid results in the continuity equation only holding on average and, thus, Gauss's Law must be enforced with a constraint. Of course, all of these reductions have exact energy conservation with continuous time. We have presented numerical examples in one spatial dimension of a combined electromagnetic mode and Langmuir wave. The error in Gauss's Law does not grow secularly and scales as expected with spatial resolution. The equation of motion in canonical Hamiltonian formulation can be numerically integrated using standard symplectic methods which should result in excellent (but not exact) energy conservation.
It has been shown that variational methods can have low level of noise.6,7 Here we have shown that the Weyl gauge, without imposing Gauss's Law with a Lagrange multiplier results in reasonable conservation properties. Thus, it is possible to use the variational formulation without having to solve an elliptic equation, thereby having performance approximately equivalent (in terms of arithmetic operations per time step) to traditional PIC algorithms but with much less noise so fewer particles should be necessary. As our examples show, the departure from exact conservation is rather benign: errors do not grow secularly and scales quadratically with grid resolution. It seems likely that with higher-order spatial approximations, these errors could become negligible.
ACKNOWLEDGMENTS
The authors gratefully acknowledge helpful discussions with P. J. Morrison, John Finn, and Evstati Evstatiev. This work was supported by NSF (Grant No. PHY-2108788). A.J.H. was supported by the University of Nebraska Program of Excellence in Atomic, Molecular, Optical, and Plasma Physics.
AUTHOR DECLARATIONS
Conflict of Interest
The authors have no conflicts to disclose.
Author Contributions
Adam J. Higuet: Conceptualization (supporting); Formal analysis (supporting); Writing – original draft (lead); Writing – review & editing (lead). B. A. Shadwick: Conceptualization (lead); Formal analysis (lead); Funding acquisition (lead); Supervision (lead); Writing – original draft (supporting); Writing – review & editing (supporting).
DATA AVAILABILITY
The data that support the findings of this study are available from the corresponding author upon reasonable request.