We study the evolution equation for magnetic energy density for a non-relativistic magnetized plasma in the (Lagrangian) reference frame comoving with the electron bulk velocity. Analyzing the terms that arise due to the ideal electric field, namely, perpendicular electron compression and magnetic field line bending, we recast them to reveal a quantity with a functional form analogous to the often-studied pressure–strain interaction term that describes one piece of internal energy density evolution of the species in a plasma, except with the species pressure tensor replaced by the magnetic stress tensor. We dub it the “magnetic stress–strain interaction.” We discuss decompositions of the magnetic stress–strain interaction analogous to those used for pressure–strain interaction. These analogies facilitate the interpretation of the evolution of the various forms of energy in magnetized plasmas and should be useful for a wide array of applications, including magnetic reconnection, turbulence, collisionless shocks, and wave–particle interactions. We display and analyze all the terms that can change magnetic energy density in the Lagrangian reference frame of the electrons using a particle-in-cell simulation of magnetic reconnection.

Many modern studies of fundamental processes in magnetized plasmas rely on understanding the evolution of energy and its conversion between different forms.1,2 While this is relatively well-understood for strongly collisional plasmas, it is challenging for collisionless or weakly collisional plasmas, such as many found in space, astrophysical, and fusion plasmas. In this brief communication, we present new insights about the time evolution of the local magnetic energy density in a magnetized plasma, specifically its evolution in the Lagrangian, i.e., comoving, reference frame that locally moves with the bulk velocity of the electrons.

To establish the context for this study, we briefly review results about the evolution of energy density in a fully-ionized net charge neutral collisionless plasma. The evolution equations of the bulk kinetic energy density Ek,σ=(1/2)mσnσuσ2, internal energy density Eint,σ=(3/2)nσkBTσ, and electromagnetic energy density EEM=(E2+B2)/8π are
(1)
(2)
(3)
where σ denotes the species (electron e or ion i), mσ is the constituent mass, nσ=d3vfσ is the number density, fσ is the phase space density, v is the velocity space coordinate, uσ=(1/nσ)d3vvfσ is the bulk flow velocity, kB is Boltzmann's constant, Tσ=(2/3kBnσ)d3v(1/2)mσvσ2fσ is the effective temperature, vσ=vuσ is the peculiar velocity, E is the electric field, B is the magnetic field, Pσ is the pressure tensor with elements Pσ,jk=d3vmσvσ,jvσ,kfσ, Jσ=qσnσuσ is the species current density, qσ is the constituent charge, qσ=d3v(1/2)mσvσ2vσfσ is the vector heat flux density, S=(c/4π)E× B is the Poynting flux, c is the speed of light, and J=σJσ is the current density.

Equations (1)–(3) are in the form of conservation laws, where the local energy density of each kind changes in time due to a locally diverging/converging energy flux (the divergence terms on the left-hand side) or through sources/sinks (the terms on the right-hand side). If the equations are integrated over a volume of interest to describe the change in energy over a whole domain, Gauss' divergence theorem implies the flux terms simply become the flux through the outer surface of the domain. If the domain being considered is closed or isolated, infinite in extent, or periodic (such as in a simulation), then the volume-integrated flux term vanishes and the total energy in each form only changes via the source/sink terms. However, we emphasize that this is only for “global” energy evolution; local energy density evolution is impacted by the divergence of the flux terms.

We briefly discuss some of the important terms in Eqs. (1)–(3). The so-called pressure–strain interaction is defined (with the minus sign) as
(4)
As seen in Eq. (2), it describes the rate of change of internal energy density resulting from a bulk flow velocity strain via compressible and/or incompressible effects.3–6 Significant effort has been expended studying this term in space satellite7–13 and theoretical/numerical14–29 studies. The Jσ·E term in Eq. (1) describes the local rate of conversion of energy density due to the electric field accelerating or decelerating charged particles of species σ. Adding this term up over all the species gives J·E, which shows up with a minus sign in Eq. (3); this enforces conservation of energy, i.e., a gain in local kinetic energy density due to Jσ·E must be offset by an equal loss in local electromagnetic energy density and vice versa. The divergence of the Poynting flux in Eq. (3) describes the local divergence of electromagnetic energy flux. It can be significant for magnetized collisionless plasma systems as shown by observational30–34 and theoretical/numerical35–39 studies.

Equations (1)–(3) are written in the (Eulerian) stationary “laboratory” reference frame, which is at rest and the plasma and electromagnetic field energy move through it. However, at a point of interest in the Eulerian reference frame, the convection of a plasma or field from another point due to the flow of the plasma can change the energy density. Convection does not represent a genuine change in the properties of species σ, so the study of energy evolution is preferably carried out in the reference frame comoving with species σ, i.e., the Lagrangian reference frame. (See, e.g., Refs. 40 and 41.)

It is not uncommon to write Eqs. (1) and (2) in the Lagrangian frame of species σ. It is less common to see Poynting's theorem [Eq. (3)] written in a Lagrangian frame (see Ref. 39 for one example). Notably, Zenitani et al.42 discussed the rate of energy density conversion due to the acceleration of charged particles in the reference frame comoving with the bulk velocity of the electrons. The so-called “Zenitani parameter” was introduced as a means to identify where non-ideal electron-scale physics takes place, such as the electron diffusion region of collisionless magnetic reconnection. It was originally thought to describe dissipation,42 but later it was recognized that it may include both reversible and irreversible contributions, particularly in weakly collisional plasmas.43 

In this study, we analyze the local magnetic energy density evolution in the Lagrangian reference frame comoving with the bulk velocity of electrons. By breaking the electric field into ideal and non-ideal contributions, we show the contribution due to the ideal electric field is described by terms analogous to those in ideal-magnetohydrodynamics (ideal-MHD) describing perpendicular compression and magnetic field line bending, but due to bulk electron motion. We recast these two terms to reveal a quantity analogous to the pressure–strain interaction term described in Eq. (4), but with the magnetic stress tensor TM replacing the pressure tensor of the species, where the jkth element is given by TM,jk=(1/4π)[BjBk(1/2)δjkB2] and δjk is the Kronecker delta. We call the magnetic analog to the pressure–strain interaction the “magnetic stress–strain interaction.” This term does not represent new physics concerning the evolution of magnetic energy density; rather it identifies a symmetry in the energy density evolution and conversion between the different forms of energy densities, which, given the broad interest in the pressure-strain interaction, may facilitate insight and understanding of the physical processes during energy conversion.

We first write Poynting's theorem in Eq. (3) in the reference frame comoving with the electrons in the non-relativistic limit. We start by writing the electric field as E=ue×B/c+E, where ue×B/c is the convective electric field and would be the whole electric field if the plasma were ideal with the electrons frozen-in to the magnetic field, and E is the non-ideal electric field. Then, ignoring the electric energy density because we are in the non-relativistic limit, Eq. (3) becomes (see also Ref. 39)
(5)
where the local magnetic energy density is EM=B2/8π. Simplifying Eq. (5) by gathering terms related to E, using Ampère's law J=(c/4π)×B to eliminate J in the ideal term, and using vector identities to gather terms gives the magnetic energy density evolution equation as
(6)
where S=(c/4π)E×B is the non-ideal Poynting flux based solely on the non-ideal electric field E, and ue,=ueb̂(b̂·ue) is the bulk electron velocity perpendicular to the magnetic field and b̂=B/B is the unit vector in the direction of the magnetic field. The component of ue parallel to the magnetic field does not enter the convective electric field and, therefore, cannot contribute to the magnetic energy density equation.

The form of Eq. (6) facilitates the identification of the physical processes at play. The first term on the left-hand side, EM/t, quantifies the local change in magnetic energy density in the stationary (Eulerian) reference frame. The second term on the left-hand side, (ue,·)EM, describes the rate of change of magnetic energy density due to convection of magnetic energy density by electrons perpendicular to the magnetic field. These two terms together describe the rate of magnetic energy density conversion in the Lagrangian reference frame comoving with the perpendicular part of the bulk electron velocity ue,.

Consequently, the four terms on the right-hand side describe mechanisms by which the local magnetic energy density changes in the reference frame comoving with the perpendicular part of the bulk electron velocity. The first term on the right-hand side, (B2/4π)(·ue,), is the rate of change of magnetic energy density due to compression of electron bulk flow perpendicular to the magnetic field. The second term on the right-hand side, (1/4π)B·[(B·)ue,], is the rate of change of magnetic energy density due to magnetic field bending by the electrons. Physically, the bending term is non-zero if the electron bulk flow perpendicular to the magnetic field changes along the direction of the magnetic field. Both of these terms are due to the ideal electric field and are analogous to the terms that arise in ideal-MHD, except the perpendicular electron bulk velocity ue, plays the role of the single-fluid bulk velocity in ideal-MHD. We note this analogy persists here despite the fact that we are not making any assumptions about the validity of ideal-MHD or Hall-MHD in our analysis.

The third term on the right-hand side, ·S, is the rate of change of magnetic energy density due to the divergence of the non-ideal Poynting flux.39 This term describes the transport of magnetic energy density rather than a pure source/sink of it, as integrating it over the entire volume of a system in question results solely in the Poynting flux across the boundary of the system.

Finally, the fourth term on the right-hand side, J·E, is the local rate of change of magnetic energy density due to the acceleration of charged particles by the non-ideal electric field.39 The Zenitani parameter42 is defined as De=γe[J·(E+ue×B/c)ρc(ue·E)], where γe=(1ue2/c2)1/2 is the relativistic gamma factor for electrons and ρc=σqσnσ is the net charge density. In the non-relativistic quasineutral limit, γe1 and ρc0, so the Zenitani parameter is approximately given by J·E. The appearance of the Zenitani parameter as a source term in Eq. (6) is consistent with the expectations in Ref. 42.

We revisit the power density due to the ideal electric field. By direct calculation, we find the first two terms on the right-hand side of Eq. (6) are equivalent to
(7)
because using the definition of TM reveals
(8)
The left-hand side of Eq. (8) has the same form as the pressure–strain interaction in Eq. (4), except that the magnetic stress tensor plays the role of the pressure tensor and the strain rate tensor is solely for the bulk flow perpendicular to the magnetic field instead of the total velocity ue. We, therefore, dub (TM·)·ue, the “magnetic stress–strain interaction” (without a minus sign). From Eq. (8), the physical interpretation of the magnetic stress–strain interaction is the combined effects of perpendicular electron compression and magnetic field line bending.
The pressure–strain interaction, (Pσ·)·uσ, is commonly decomposed in ways that elucidate the physical effects underlying it, so we discuss analogous decompositions of (TM·)·ue. First, the compressional part of the pressure–strain interaction is given by Pσ(·uσ), where Pσ=(1/3)tr(Pσ) is the effective (scalar) pressure and tr denotes the trace.5 Its magnetic counterpart is TM(·ue,), where we define the effective (scalar) magnetic stress as TM=(1/3)TM,jj=(1/3)(1/4π)[B2(1/2)δjjB2]=(1/3)EM. Then, the compressible part of the magnetic stress–strain interaction is
(9)
Note, this result elucidates that the second term in Eq. (8) contains a piece describing compression even though it does not come from the compression term. Physically, it describes compression that occurs due to magnetic field line bending. When combined with the first term in Eq. (7), it reveals that the total compression in the magnetic energy evolution equation from the ideal terms is (4/3)EM(·ue,). The factor of 4/3 is crucial; in the ideal-MHD or ideal-double adiabatic (CGL)44 limit, the 4/3 is necessary to achieve conservation of the first and second adiabatic invariants45 (the magnetic moment μ and the parallel action J).
Next, the term that describes the net effect of converging or diverging flow for pressure–strain interaction is called26  PDUσ. We dub the magnetic stress counterpart as TDU, defined in rectangular coordinates as
(10)
Using the definition of the diagonal elements of TM, a brief derivation yields
(11)
The term that isolates the incompressible portion due to normal flow in the pressure–strain interaction was called26, PiDnormal,σ=PDUσPσ(·uσ). We define the magnetic counterpart describing the incompressible portion due to normal flow as
(12)
(13)
since TM=(1/3)EM. The term dubbed PiDshear in Ref. 26 is internal energy density conversion due to purely sheared flow. We refer to the magnetic analog of this as TDshear, which in rectangular coordinates is
(14)
Finally, the analog of the incompressible part of pressure–strain interaction5  PiDσ=PiDnormal,σ+PiDshear,σ for the magnetic stress–strain interaction is
(15)

We now present the terms in the magnetic energy density evolution equation [Eq. (6)] using particle-in-cell simulations of magnetic reconnection. The simulations are identical to those in Ref. 27 and are thoroughly described therein. The simulation data are presented only for the lower current sheet at time t=13Ωci01 with black solid lines denoting in-plane projections of magnetic field lines, where Ωci0 is the ion cyclotron frequency based on the initial asymptotic magnetic field B0. This time is when the reconnection rate is rising most rapidly during the nonlinear growth phase, so the system is not in steady-state. The plots are centered at the X-line (x0,y0), and all distances are given in units of the ion inertial scale di0 based on the difference between the maximum initial number density and the asymptotic value. To minimize particle noise in the simulation, we smooth the raw simulation data over a width of five cells, followed by calculating the desired spatial or temporal derivatives, and finally, smooth the results over five cells. This smoothing approach is chosen after testing various options for the number of cells to smooth over and the number of recursions, ensuring that the signal structure remains largely unaffected by the smoothing. All power densities are given in units of (B02/4π)Ωci0.

Panel (a) of Fig. 1 displays the left-hand side of Eq. (6) that denotes the time rate of change of EM in the Lagrangian frame comoving with ue,. We see that the magnetic energy density decreases (purple) with time in this Lagrangian frame upstream of the X-line and there is a weak increase (green) along the horizontal neutral line (the dashed line) downstream of the X-line. Since the reconnection process converts magnetic energy to energy in the plasma, a decrease in magnetic energy in this Lagrangian frame is expected. The positive signal along the outflow region of the neutral line is because the reconnected magnetic field By increases with distance in the outflow direction from the X-line while the reconnecting magnetic field Bx=0, so EM increases in the reference frame moving with the electrons.

FIG. 1.

Particle-in-cell simulation results of the terms in the magnetic energy density evolution equation [Eq. (6)] during magnetic reconnection. (a) The time rate of change of magnetic energy density EM in the Lagrangian frame comoving with ue,, (b) electron perpendicular compression EM(·ue,), (c) magnetic stress–strain interaction (TM·)·ue,, (d) negative of the divergence of the non-ideal Poynting flux ·S, and (e) the power density by the non-ideal electric field J·E.

FIG. 1.

Particle-in-cell simulation results of the terms in the magnetic energy density evolution equation [Eq. (6)] during magnetic reconnection. (a) The time rate of change of magnetic energy density EM in the Lagrangian frame comoving with ue,, (b) electron perpendicular compression EM(·ue,), (c) magnetic stress–strain interaction (TM·)·ue,, (d) negative of the divergence of the non-ideal Poynting flux ·S, and (e) the power density by the non-ideal electric field J·E.

Close modal

We now examine the terms on the right-hand side of Eq. (6). Figures 1(b)–1(e) contain the perpendicular compression term EM(·ue,), the magnetic stress–strain interaction (TM·)·ue,, the negative divergence of the non-ideal Poynting flux ·S, and the power density by the non-ideal electric field J·E, respectively. We consider the action of these terms in a few key regions for the reconnection process.

Near the X-line, within |xx0|<1 and |yy0|<0.3, perpendicular compression [in (b)] and magnetic stress–strain interaction [in (c)] are small because they are proportional to the magnetic energy and magnetic stress, respectively, which are small because the magnetic field is weak near the X-line. As known from previous work,42 there is a strong J·E [in (e)] in the inner electron diffusion region where the (non-ideal) reconnection electric field E in the z direction accelerates electrons in the z direction. Since the magnetic energy density near the X-line does not change much in time, the non-ideal Poynting flux [in (d)] essentially balances J·E. Thus, S converges near the X-line to provide the energy necessary for the reconnection electric field to accelerate the electrons. Physically, E is in the z direction while Bx>0 for yy0>0 and Bx<0 for yy0<0, so E×B is toward the neutral line both above and below the X-line.

In the outflow region around 1<|xx0|<2, there is a small region of magnetic energy density increase near the neutral line, and otherwise the magnetic energy density is mostly decreasing [panel (a)]. The increase in magnetic energy density is caused by perpendicular compression [panel (b), and it also contributes to the magnetic stress–strain interaction in panel (c) from Eq. (8)]. Near the neutral line, magnetic field lines are strongly kinked, and the y component of ue, points toward the neutral line, thus giving strong compression EM(·ue,). Away from the neutral line, the compression and magnetic stress–strain interaction are negative and significant, consistent with the strongly kinked magnetic field lines losing energy [see Fig. 1(a)] as they straighten; we return to this shortly to identify the physical cause. We note there is also a positive J·E in the outflow region because of a non-zero divergence of the electron pressure tensor.46,47 Since the sign of Ez flips to positive in this region, ·S is positive which nearly balances J·E, so the non-ideal electric field is not the cause of the change in EM at this region.

We next treat the region around the upstream edge of the electron diffusion region, |xx0|<1 and 0.3<|yy0|<0.5. The dominant term is ·S, which arises because the non-ideal reconnection electric field Ez is reasonably uniform, but the strength of the reconnecting magnetic field Bx decreases approaching the X-line. This leads to a converging S associated with the decrease in magnetic energy density in the Lagrangian frame of the electrons. Further upstream, outside the electron diffusion region, the magnetic energy density decreases in the Lagrangian frame of the electrons [panel (a)]; this is dominated by a decrease in EM in the Eulerian frame, i.e., EM/t<0 (not shown). At the time slice plotted here, the system is not in the steady-state, so the magnetic field energy density decreases locally due to the magnetic field getting pulled in toward the X-line as the reconnection process speeds up.

To assess the physical cause of the magnetic stress–strain interaction in Fig. 1(c), we show in Fig. 2 its decomposition, with the compressible TM(·ue,), incompressible normal TDnormal, and incompressible shear TDnormal contributions in panels (a)–(c), respectively. The compressible part in Fig. 2(a) is simply 1/3 of the compression term EM(·ue,) in Fig. 1(b). To see how these plots can help elucidate the physics taking place, consider the outflow region near the neutral line. The increase in EM near the neutral line is caused by the positive compression terms as stated previously, but we also see in Fig. 2(b) that the increase in magnetic energy density is reduced because there is incompressible normal deformation also occurring. Figures 2(a) and 2(c) show that the decrease in EM away from the neutral line in the exhaust is due to negative compression (expansion) and incompressible shear as a result of magnetic field line bending as the strongly curved magnetic field line straightens. Importantly, both Figs. 2(b) and 2(c) reveal changes to the magnetic energy via incompressible effects. This is a reasonable result because the Alfvénic and whistler wave-like effect driving the outflow48–50 are incompressible.

FIG. 2.

Particle-in-cell simulation results of the decomposition of magnetic stress–strain interaction (TM·)·ue, [Eq. (8)] during magnetic reconnection. (a) The compressible part TM(·ue,) [Eq. (9)], (b) the incompressible normal deformation part TDnormal [Eq. (13)], and (c) the shear deformation part TDshear [Eq. (14)].

FIG. 2.

Particle-in-cell simulation results of the decomposition of magnetic stress–strain interaction (TM·)·ue, [Eq. (8)] during magnetic reconnection. (a) The compressible part TM(·ue,) [Eq. (9)], (b) the incompressible normal deformation part TDnormal [Eq. (13)], and (c) the shear deformation part TDshear [Eq. (14)].

Close modal

Finally, we make two observations. First, the non-ideal terms [in (d) and (e)] are predominantly localized to the electron diffusion region and within an electron inertial length size surrounding the separatrix. This is consistent with the electron frozen-in condition breaking down at sub-electron inertial scales in anti-parallel magnetic reconnection so that E0. Second, although the data presented here are from a time slice when the time evolution is the most rapid, we do not expect significant differences in the structure of these plots near the electron diffusion region in the steady-state because these plots are in the Lagrangian reference frame of the electrons. In the steady-state, /t=0, but the convection term ue,· remains significant.

The reconnection process is used as an example, but we expect the analysis performed here and its interpretation could be useful in other fundamental processes in magnetized plasmas, including turbulence, collisionless shocks, and wave–particle interactions.

This work was partially motivated by P.A.C.'s participation in the Magnetospheric Multiscale (MMS) 10th Community Meeting. We gratefully acknowledge support from NASA Grant Nos. 80NSSC24K0172 and 80NSSC23K0409, NSF Grant No. PHY-2308669, and DOE grant DE-SC0020294. This research used resources of the National Energy Research Scientific Computing Center (NERSC), a U.S. Department of Energy Office of Science User Facility located at Lawrence Berkeley National Laboratory, operated under Contract No. DE-AC02-05CH11231 using NERSC award FES-ERCAP0027083.

The authors have no conflicts to disclose.

M. Hasan Barbhuiya: Data curation (equal); Investigation (equal); Software (equal); Visualization (equal); Writing – review & editing (equal). P. A. Cassak: Conceptualization (equal); Funding acquisition (equal); Investigation (equal); Resources (equal); Supervision (equal); Writing – original draft (equal); Writing – review & editing (equal).

The data that support the findings of this study are openly available in Zenodo at https://doi.org/10.5281/zenodo.14172055.

1.
G. G.
Howes
, “
A prospectus on kinetic heliophysics
,”
Phys. Plasmas
24
,
055907
(
2017
).
2.
W. H.
Matthaeus
,
Y.
Yang
,
M.
Wan
,
T. N.
Parashar
,
R.
Bandyopadhyay
,
A.
Chasapis
,
O.
Pezzi
, and
F.
Valentini
, “
Pathways to dissipation in weakly collisional plasmas
,”
Astrophys. J.
891
,
101
(
2020
).
3.
D.
Del Sarto
,
F.
Pegoraro
, and
F.
Califano
, “
Pressure anisotropy and small spatial scales induced by velocity shear
,”
Phys. Rev. E
93
,
053203
(
2016
).
4.
Y.
Yang
,
W. H.
Matthaeus
,
T. N.
Parashar
,
P.
Wu
,
M.
Wan
,
Y.
Shi
,
S.
Chen
,
V.
Roytershteyn
, and
W.
Daughton
, “
Energy transfer channels and turbulence cascade in Vlasov-Maxwell turbulence
,”
Phys. Rev. E
95
,
061201
(
2017
).
5.
Y.
Yang
,
W. H.
Matthaeus
,
T. N.
Parashar
,
C. C.
Haggerty
,
V.
Roytershteyn
,
W.
Daughton
,
M.
Wan
,
Y.
Shi
, and
S.
Chen
, “
Energy transfer, pressure tensor, and heating of kinetic plasma
,”
Phys. Plasmas
24
,
072306
(
2017
).
6.
D.
Del Sarto
and
F.
Pegoraro
, “
Shear-induced pressure anisotropization and correlation with fluid vorticity in a low collisionality plasma
,”
Mon. Not. R. Astron. Soc.
475
,
181
(
2018
).
7.
A.
Chasapis
,
Y.
Yang
,
W. H.
Matthaeus
,
T. N.
Parashar
,
C. C.
Haggerty
,
J. L.
Burch
,
T. E.
Moore
,
C. J.
Pollock
,
J.
Dorelli
,
D. J.
Gershman
,
R. B.
Torbert
, and
C. T.
Russell
, “
Energy conversion and collisionless plasma dissipation channels in the turbulent magnetosheath observed by the Magnetospheric Multiscale Mission
,”
Astrophys. J.
862
,
32
(
2018
).
8.
Z. H.
Zhong
,
X. H.
Deng
,
M.
Zhou
,
W. Q.
Ma
,
R. X.
Tang
,
Y. V.
Khotyaintsev
,
B. L.
Giles
,
C. T.
Russell
, and
J. L.
Burch
, “
Energy conversion and dissipation at dipolarization fronts: A statistical overview
,”
Geophys. Res. Lett.
46
,
12693
12701
, https://doi.org/10.1029/2019GL085409 (
2019
).
9.
R.
Bandyopadhyay
,
W. H.
Matthaeus
,
T. N.
Parashar
,
Y.
Yang
,
A.
Chasapis
,
B. L.
Giles
,
D. J.
Gershman
,
C. J.
Pollock
,
C. T.
Russell
,
R. J.
Strangeway
,
R. B.
Torbert
,
T. E.
Moore
, and
J. L.
Burch
, “
Statistics of kinetic dissipation in the Earth's magnetosheath: MMS observations
,”
Phys. Rev. Lett.
124
,
255101
(
2020
).
10.
R.
Bandyopadhyay
,
A.
Chasapis
,
W. H.
Matthaeus
,
T. N.
Parashar
,
C. C.
Haggerty
,
M. A.
Shay
,
D. J.
Gershman
,
B. L.
Giles
, and
J. L.
Burch
, “
Energy dissipation in turbulent reconnection
,”
Phys. Plasmas
28
,
112305
(
2021
).
11.
M.
Zhou
,
H.
Man
,
Y.
Yang
,
Z.
Zhong
, and
X.
Deng
, “
Measurements of energy dissipation in the electron diffusion region
,”
Geophys. Res. Lett.
48
,
e2021GL096372
, https://doi.org/10.1029/2021GL096372 (
2021
).
12.
Y.
Wang
,
R.
Bandyopadhyay
,
R.
Chhiber
,
W. H.
Matthaeus
,
A.
Chasapis
,
Y.
Yang
,
F. D.
Wilder
,
D. J.
Gershman
,
B. L.
Giles
,
C. J.
Pollock
,
J.
Dorelli
,
C. T.
Russell
,
R. J.
Strangeway
,
R. T.
Torbert
,
T. E.
Moore
, and
J. L.
Burch
, “
Statistical survey of collisionless dissipation in the terrestrial magnetosheath
,”
J. Geophys. Res. Space Phys.
126
,
e2020JA029000
, https://doi.org/10.1029/2020JA029000 (
2021
).
13.
J. L.
Burch
,
K. J.
Genestreti
,
S. V.
Heuer
,
A.
Chasapis
,
R. B.
Torbert
,
D. J.
Gershman
,
R.
Bandyopadhyay
,
C. J.
Pollock
,
W. H.
Matthaeus
,
T. K. M.
Nakamura
, and
J.
Egedal
, “
Electron energy dissipation in a magnetotail reconnection region
,”
Phys. Plasmas
30
,
082903
(
2023
).
14.
R. D.
Hazeltine
,
S. M.
Mahajan
, and
P. J.
Morrison
, “
Local thermodynamics of a magnetized, anisotropic plasma
,”
Phys. Plasmas
20
,
022506
(
2013
).
15.
M. I.
Sitnov
,
V. G.
Merkin
,
V.
Roytershteyn
, and
M.
Swisdak
, “
Kinetic dissipation around a dipolarization front
,”
Geophys. Res. Lett.
45
,
4639
4647
, https://doi.org/10.1029/2018GL077874 (
2018
).
16.
T. N.
Parashar
,
W. H.
Matthaeus
, and
M. A.
Shay
, “
Dependence of kinetic plasma turbulence on plasma β
,”
Astrophys. J. Lett.
864
,
L21
(
2018
).
17.
S.
Du
,
F.
Guo
,
G. P.
Zank
,
X.
Li
, and
A.
Stanier
, “
Plasma energization in colliding magnetic flux ropes
,”
Astrophys. J.
867
,
16
(
2018
).
18.
Y.
Yang
,
M.
Wan
,
W. H.
Matthaeus
,
L.
Sorriso-Valvo
,
T. N.
Parashar
,
Q.
Lu
,
Y.
Shi
, and
S.
Chen
, “
Scale dependence of energy transfer in turbulent plasma
,”
Mon. Not. R. Astron. Soc.
482
,
4933
4940
(
2019
).
19.
O.
Pezzi
,
Y.
Yang
,
F.
Valentini
,
S.
Servidio
,
A.
Chasapis
,
W. H.
Matthaeus
, and
P.
Veltri
, “
Energy conversion in turbulent weakly collisional plasmas: Eulerian hybrid Vlasov-Maxwell simulations
,”
Phys. Plasmas
26
,
072301
(
2019
).
20.
L.
Song
,
M.
Zhou
,
Y.
Yi
,
X.
Deng
,
Z.
Zhong
, and
H.
Man
, “
Force and Energy Balance of the Dipolarization Front
,”
J. Geophys. Res. Space Phys.
125
,
e2020JA028278
, https://doi.org/10.1029/2020JA028278 (
2020
).
21.
S.
Du
,
G. P.
Zank
,
X.
Li
, and
F.
Guo
, “
Energy dissipation and entropy in collisionless plasma
,”
Phys. Rev. E
101
,
033208
(
2020
).
22.
O.
Pezzi
,
H.
Liang
,
J. L.
Juno
,
P. A.
Cassak
,
C. L.
Vásconez
,
L.
Sorriso-Valvo
,
D.
Perrone
,
S.
Servidio
,
V.
Roytershteyn
,
J. M.
TenBarge
, and
W. H.
Matthaeus
, “
Dissipation measures in weakly collisional plasmas
,”
Mon. Not. R. Astron. Soc.
505
,
4857
(
2021
).
23.
S.
Fadanelli
,
B.
Lavraud
,
F.
Califano
,
G.
Cozzani
,
F.
Finelli
, and
M.
Sisti
, “
Energy conversions associated with magnetic reconnection
,”
J. Geophys. Res. Space Phys.
126
,
e2020JA028333
, https://doi.org/10.1029/2020JA028333 (
2021
).
24.
G.
Arrò
,
F.
Califano
, and
G.
Lapenta
, “
Spectral properties and energy transfer at kinetic scales in collisionless plasma turbulence
,”
Astron. Astrophys.
668
,
A33
(
2022
).
25.
Y.
Yang
,
W. H.
Matthaeus
,
S.
Roy
,
V.
Roytershteyn
,
T. N.
Parashar
,
R.
Bandyopadhyay
, and
M.
Wan
, “
Pressure–strain interaction as the energy dissipation estimate in collisionless plasma
,”
Astrophys. J.
929
,
142
(
2022
).
26.
P. A.
Cassak
and
M. H.
Barbhuiya
, “
Pressure–strain interaction. I. On compression, deformation, and implications for Pi-D
,”
Phys. Plasmas
29
,
122306
(
2022
).
27.
M. H.
Barbhuiya
and
P. A.
Cassak
, “
Pressure–strain interaction. III. Particle-in-cell simulations of magnetic reconnection
,”
Phys. Plasmas
29
,
122308
(
2022
).
28.
P.
Hellinger
,
V.
Montagud-Camps
,
L.
Franci
,
L.
Matteini
,
E.
Papini
,
A.
Verdini
, and
S.
Landi
, “
Ion-scale transition of plasma turbulence: Pressure–strain effect
,”
Astrophys. J.
930
,
48
(
2022
).
29.
S. A.
Conley
,
J.
Juno
,
J. M.
TenBarge
,
M. H.
Barbhuiya
,
P. A.
Cassak
,
G. G.
Howes
, and
E.
Lichko
, “
The kinetic analogue of the pressure-strain interaction
,”
Phys. Plasmas
(to be published) (2024).
30.
B. T.
Welsch
, “
The photospheric poynting flux and coronal heating
,”
Publ. Astron. Soc. Jpn.
67
,
18
(
2015
).
31.
Y.
Nishimura
,
T.
Kikuchi
,
A.
Shinbori
,
J.
Wygant
,
Y.
Tsuji
,
T.
Hori
,
T.
Ono
,
S.
Fujita
, and
T.
Tanaka
, “
Direct measurements of the poynting flux associated with convection electric fields in the magnetosphere
,”
J. Geophys. Res. Space Phys.
115
,
15491
, https://doi.org/10.1029/2010JA015491 (
2010
).
32.
M.
Volwerk
,
P.
Louarn
,
T.
Chust
,
A.
Roux
,
H.
de Feraudy
, and
B.
Holback
, “
Solitary kinetic Alfvén waves: A study of the Poynting flux
,”
J. Geophys. Res. Space Phys.
101
,
13335
13343
, https://doi.org/10.1029/96JA00166 (
1996
).
33.
D. S.
Payne
,
K. J.
Genestreti
,
K.
Germaschewski
,
M. R.
Argall
,
R. B.
Torbert
,
I.
Dors
, and
A.
Ardakani
, “
Energy balance and time dependence of a magnetotail electron diffusion region
,”
J. Geophys. Res. Space Phys.
125
,
e2020JA028290
, https://doi.org/10.1029/2020JA028290 (
2020
).
34.
K. J.
Genestreti
,
P. A.
Cassak
,
A.
Varsani
,
J. L.
Burch
,
R.
Nakamura
, and
S.
Wang
, “
Assessing the time dependence of reconnection with Poynting's theorem: MMS observations
,”
Geophys. Res. Lett.
45
,
2886
2892
, https://doi.org/10.1002/2017GL076808 (
2018
).
35.
Y.-H.
Liu
,
P.
Cassak
,
X.
Li
,
M.
Hesse
,
S.-C.
Lin
, and
K.
Genestreti
, “
First-principles theory of the rate of magnetic reconnection in magnetospheric and solar plasmas
,”
Commun. Phys.
5
,
97
(
2022
).
36.
D. S.
Payne
,
C. J.
Farrugia
,
R. B.
Torbert
,
K.
Germaschewski
,
A. R.
Rogers
, and
M. R.
Argall
, “
Origin and structure of electromagnetic generator regions at the edge of the electron diffusion region
,”
Phys. Plasmas
28
,
112901
(
2021
).
37.
D. S.
Payne
,
R. B.
Torbert
,
K.
Germaschewski
,
T. G.
Forbes
, and
J. R.
Shuster
, “
Relating the Phases of Magnetic Reconnection Growth to Energy Transport Mechanisms in the Exhaust
,”
J. Geophys. Res. Space Phys.
129
,
e2023JA032015
, https://doi.org/10.1029/2023JA032015 (
2024
).
38.
R. A.
Treumann
and
W.
Baumjohann
, “
The usefulness of poynting's theorem in magnetic turbulence
,” in
Annales Geophysicae
(
Copernicus Publications Göttingen
,
Germany
,
2017
), Vol.
35
, pp.
1353
1360
.
39.
R. A.
Treumann
and
W.
Baumjohann
, “
Poynting's theorem in magnetic turbulence
,” arXiv:1701.01266 (
2017
).
40.
P. A.
Cassak
,
M. H.
Barbhuiya
,
H.
Liang
, and
M. R.
Argall
, “
Quantifying energy conversion in higher-order phase space density moments in plasmas
,”
Phys. Rev. Lett.
130
,
085201
(
2023
).
41.
J. M.
TenBarge
,
J.
Juno
, and
G. G.
Howes
, “
Electron energization in reconnection: Eulerian vs Lagrangian perspectives
,”
Phys. Plasmas
31
,
022901
(
2024
).
42.
S.
Zenitani
,
M.
Hesse
,
A.
Klimas
, and
M.
Kuznetsova
, “
New measure of the dissipation region in collisionless magnetic reconnection
,”
Phys. Rev. Lett.
106
,
195003
(
2011
).
43.
P. A.
Cassak
,
K. J.
Genestreti
,
J. L.
Burch
,
T.-D.
Phan
,
M. A.
Shay
,
M.
Swisdak
,
J. F.
Drake
,
L.
Price
,
S.
Eriksson
,
R. E.
Ergun
,
B. J.
Anderson
,
V. G.
Merkin
, and
C. M.
Komar
, “
The effect of a guide field on local energy conversion during asymmetric magnetic reconnection: Particle-in-cell simulations
,”
J. Geophys. Res. Space Phys.
122
,
11,523
11,542
, https://doi.org/10.1002/2017JA024555 (
2017
).
44.
G. F.
Chew
,
M. L.
Goldberger
, and
F. E.
Low
, “
The Boltzmann equation and the one-fluid hydromagnetic equations in the absence of particle collisions
,”
Proc. R. Soc. London, Ser. A: Math. Phys. Sci.
236
,
112
118
(
1956
).
45.
F. F.
Chen
,
Introduction to Plasma Physics and Controlled Fusion
(
Springer International Publishing
,
2016
).
46.
M.
Hesse
,
S.
Zenitani
, and
A.
Klimas
, “
The structure of the electron outflow jet in collisionless magnetic reconnection
,”
Phys. Plasmas
15
,
112102
(
2008
).
47.
J.
Egedal
,
B.
Wetherton
,
W.
Daughton
, and
A.
Le
, “
Processes setting the structure of the electron distribution function within the exhausts of anti-parallel reconnection
,”
Phys. Plasmas
23
,
122904
(
2016
).
48.
M. E.
Mandt
,
R. E.
Denton
, and
J. F.
Drake
, “
Transition to whistler mediated magnetic reconnection
,”
Geophys. Res. Lett.
21
,
73
, https://doi.org/10.1029/93GL03382 (
1994
).
49.
B. N.
Rogers
,
R. E.
Denton
,
J. F.
Drake
, and
M. A.
Shay
, “
The role of dispersive waves in collisionless magnetic reconnection
,”
Phys. Rev. Lett.
87
,
195004
(
2001
).
50.
J. F.
Drake
and
M. A.
Shay
, “
The fundamentals of collisionless reconnection
,” in
Reconnection of Magnetic Fields: Magnetohydrodynamics and Collisionless Theory and Observations
, edited by
J.
Birn
and
E.
Priest
(
Cambridge University Press
,
2007
).