We have developed a local, linear theoretical model for lower hybrid drift waves that can be used for plasmas in the weakly collisional regime. Two cases with typical plasma and field parameters for the current sheet of the magnetic reconnection experiment have been studied. For a case with a low electron beta (βe=0.25, high guide field case), the quasi-electrostatic lower hybrid drift wave is unstable, while the electromagnetic lower hybrid drift wave has a positive growth rate for a high-βe case (βe=8.9, low guide field case). For both cases, including the effects of Coulomb collisions reduces the growth rate but collisional impacts on the dispersion and growth rate are limited (20%).

Magnetic reconnection converts magnetic energy into plasma thermal and flow energy via topological rearrangements of the magnetic field lines. Energy conversion processes during magnetic reconnection result in many free energy sources for waves and instabilities near the diffusion region, such as strong gradients of the magnetic field and plasma parameters. Among them, the lower hybrid drift wave (LHDW) has been widely observed near the diffusion region in both space (e.g., Refs. 1–7) and laboratory plasmas (e.g., Refs. 8–10). The free energy source of LHDWs is the cross field current.11 The large density gradients near the separatrix can particularly be a free energy source by inducing a perpendicular current via a diamagnetic drift.

LHDWs have been a candidate for generating anomalous resistivity because it can interact differently with magnetized electrons and non-magnetized ions, resulting in momentum exchange between the two species (e.g., Refs. 7–9 and 12–16). For reconnection with a negligible guide field, the fast-growing, short-wavelength (kρe1; k is the magnitude of the wave vector k, ρe is the electron gyroradius), quasi-electrostatic LHDW (ES-LHDW) is found to be localized at the edge of the current sheet8 due to the stabilization by the high plasma beta (β).17 On the other hand, the long-wavelength (kρeρi1; ρi is the ion gyroradius), electromagnetic LHDW (EM-LHDW) that propagates obliquely to the magnetic field exists in the electron diffusion region.9 However, extensive efforts via numerical particle-in-cell (PIC) simulations15,16 show that the EM-LHDW does not play an important role in fast reconnection and electron energization near the electron diffusion region during antiparallel reconnection.

Recent observations by the magnetospheric multiscale (MMS) mission show that the ES-LHDW can be generated inside or near the electron diffusion region,5–7 when there is a sizable guide field. The ES-LHDW can drive electron heating and vortical flows6 near the electron diffusion region. Moreover, the ES-LHDW is capable of generating anomalous drag between electrons and ions.7 

Motivated by these observations, Yoo et al.7 have developed a local, linear theoretical model that explains the dynamics of both ES- and EM-LHDWs in the presence of a guide field. This model is based on collisionless closures for the electron heat flux with the assumption of a gyrotropic electron pressure tensor. The results from the model agree with the activities of the ES- and EM-LHDWs inside a current sheet at the magnetopause.7 

In laboratory experiments, such as the magnetic reconnection experiment (MRX), the effects of Coulomb collisions on magnetic reconnection and electron heating are not negligible. The classical Spitzer resistivity,18 for example, can balance the reconnection electric field in the collisional regime and can even account for 10%–20% of that in the collisionless regime.19,20 This indicates that Coulomb collisions may also affect the dynamics of LHDWs in laboratory plasmas.

These collisional effects on LHDWs have not been considered previously, even though LHDWs in the reconnection current sheet have been extensively studied via theoretical analyses and numerical simulations (e.g., Refs. 11, 14, 21–23). This paper provides the first quantitative study of the effects of Coulomb collisions on LHDWs. Through this model, we can address how the dynamics of LHDWs in laboratory plasmas are different from those in collisionless plasmas and when collisional effects become important. To include the effects from collisions, we have advanced the previous models7,24 by using closures of the electron heat flux, heat generated by collisions, and resistivity that can be used for plasmas with arbitrary collisionality.25,26 For a self-consistent modeling of the heat flux and energy conservation, we also have allowed a first-order perturbation of the perpendicular electron temperature (Te1), which was set to be zero in a previous model by Yoo et al.7 Unlike previous models, the zeroth-order electron temperature anisotropy is not allowed in the current model because the available closures were developed under the assumption of isotropic electron pressure at equilibrium. Except these changes, all other assumptions are the same: we used a kinetic equation for unmagnetized ions, fluid equations for electrons, and a gyrotropic pressure tensor for electrons.

This linear model can be used to quantify the effects of LHDWs on electron heating and reconnection dynamics in weakly collisional plasmas; with measured wave amplitudes and quasi-linear arguments, wave-associated anomalous terms and heat generated by collisions with ions can be directly estimated. It should be noted that the wave-associated heating power cannot be estimated by collisionless models.

In Sec. II, we explain the theoretical model for LHDWs in a local geometry. Then, in Sec. III, we numerically calculate dispersion relations of LHDWs for two cases. The biggest difference in the two cases is the value of electron beta, βe. For the low-βe case, which represents conditions near the electron diffusion region during reconnection with a strong guide field, the ES-LHDW is unstable. For the high-βe case, which represents conditions in the same region but with a negligible guide field, the EM-LHDW has positive growth rates. In both cases, collisional effects on LHDWs with typical MRX parameters are not significant (20 %). Finally, in Sec. IV, we discuss the results and propose future research.

Figure 1 shows the geometry of our local theoretical model for a LHDW inside a current sheet. Here, the subscript 0 indicates equilibrium quantities. We chose the ion rest frame, and electrons have velocity (ue0) on the xz plane. The equilibrium magnetic field is along the z direction and the density gradient direction is along the y direction. In this model, there is neither equilibrium temperature gradient nor ion temperature anisotropy. The equilibrium electron temperature is also assumed to be isotropic, but anisotropy is allowed in the perturbed electron temperature. The wave vector (k) lies on the xz plane due to our assumption of negligible ky. Thus, our theoretical model is local and valid only when the wavelength of the LHDW is much smaller than the thickness of the current sheet in the y direction.24 

FIG. 1.

Geometry of the local theory for the LHDW dispersion calculation. We are working in the ion rest frame with the z direction toward the equilibrium magnetic field (B0) and the y direction along the density gradient direction. Due to the force balance, the equilibrium electric field E0 is also along the y direction. The equilibrium electron flow velocity ue0 and wave vector k reside on the xz plane. The angle between k and B0 is given by θ.

FIG. 1.

Geometry of the local theory for the LHDW dispersion calculation. We are working in the ion rest frame with the z direction toward the equilibrium magnetic field (B0) and the y direction along the density gradient direction. Due to the force balance, the equilibrium electric field E0 is also along the y direction. The equilibrium electron flow velocity ue0 and wave vector k reside on the xz plane. The angle between k and B0 is given by θ.

Close modal

To balance the force associated with the pressure (density) gradient, there is an equilibrium electric field along the y direction. By using the ion and electron force balance equations, the equilibrium electric field E0 can be expressed in terms of other plasma parameters. From the ion force balance along the y direction, we have

en0E0=Ti0dn0dy=εn0Ti0,
(1)

where n0 is the equilibrium density, Ti0 is the equilibrium ion temperature, and ε=(dn0/dy)/n0 is the inverse of the density gradient scale. From the y component of the electron momentum equation, we have

en0(E0ue0xB0)=Te0dn0dy,
(2)

where ue0x is the x component of the equilibrium electron flow velocity and Te0 is the equilibrium electron temperature. Then, the equilibrium electric field is

E0=Ti0Te0+Ti0ue0xB0.
(3)

The inverse of the gradient scale is given by

ε=eue0xB0Te0+Ti0.
(4)

Note that Eqs. (3) and (4) are the same as those in the collisionless model in Yoo et al.,7 because the resistivity term is zero along the y direction.

All perturbed quantities have a normal mode decomposition proportional to exp[i(k·xωt)] with the wave vector k=(k,0,k). Here, the subscript 1 indicates perturbed quantities. For the dispersion relation, Maxwell's equations without the displacement current term are used,

k×(k×E1)=iωμ0J1.
(5)

The displacement current term is ignored because the phase velocity of the wave is much smaller than the speed of light.

Assuming the equilibrium ion distribution function to be locally Maxwellian, the perturbed ion current density (Ji1) is given by24 

Ji1=in0e2mikvti[Z(ζ)E1+Z(E1·k̂)2k̂i(ε2k)ZE1yk̂],
(6)

where mi is the ion mass, vti=2Ti0/mi is the ion thermal speed, ζ=ω/kvti, and Z(ζ) is the plasma dispersion function. This is from a perturbed Vlasov equation for unmagnetized ions. This means that any dynamics slower than the ion cyclotron frequency have been ignored, including collisional effects on ion dynamics. In our regime of interest, the ion collision frequency is smaller than the ion cyclotron frequency. The perturbed ion temperature can be also obtained, which is

Ti1=iek[E1·k̂(2Z+Z4)iE1y(εk)(Z+Z4)].
(7)

The perturbed electron current density Je1 is obtained from fluid equations. This is different from the classical formulation of LHDWs, where the kinetic (Vlasov) equation is used for electron dynamics (e.g., Refs. 17, 27, and 28). Since electrons are magnetized, a gyrotropic electron pressure tensor is assumed. In this case, the 3 + 1 fluid model (n, u, p, and p; p and p are the parallel and perpendicular pressure, respectively) is appropriate.25 In this fluid model, off diagonal terms of the electron pressure tensor are ignored.

The first-order electron momentum equation is given by

imen0(ωk·ue0)ue1=ik·Pe1+en0(E1+ue1×B0+ue0×B1)+e(E0+ue0×B0)ne1Re1,
(8)

where Pe1 is the perturbed electron pressure tensor and Re1 is the perturbed resistivity. The perturbed electron density ne1 is given by the electron continuity equation, which is

(ωk·ue0)ne1=(k·ue1iεue1y)n0.
(9)

To close the momentum equation, we need closures for Pe1 and Re1. For Pe1, we only need closures for pe1 and pe1||, since we assume a gyrotropic pressure tensor as mentioned earlier. To obtain pe1 and pe1, we start from the following kinetic equation:

fet+v·feeme(E+v×B)·fev=C(fe),
(10)

where fe is the electron distribution function and C(fe) is the collision operator. First, multiplying the kinetic equation with me(vzuez)2 and integrating over the velocity space yield

pet+·(uepe)+·qe+2uezzpe=Ce||,
(11)

where

pe=me(vzuez)2fedv,
(12)
qe=me(vue)(vzuez)2fedv,
(13)
Ce||=C(fe)me(vue)2dv.
(14)

Similarly, multiplying the kinetic equation with me[(vxuex)2+(vyuey)2]/2 and integrating over the velocity space yield

pet+·(uepe)+·qe+(uexx+ueyy)pe=Ce,
(15)

where

pe=me12[(vxuex)2+(vyuey)2]fedv,
(16)
qe=me12[(vxuex)2+(vyuey)2](vue)fedv,
(17)
Ce=12C(fe)[(vxuex)2+(vyuey)2]dv.
(18)

Linearizing Eq. (11) yields

iωpe1+εue1yn0Te0+i(k·u0)pe1+i(k·ue1)n0Te0+ik·qe1+2ikue1zn0Te0=Ce1||.
(19)

By using pe1||=ne1Te0+n0Te1|| and Eq. (9), Eq. (19) can be written as

i(ωk·u0)n0Te1||=ik·qe1+2ikue1zn0Te0Ce1||.
(20)

Similarly, linearizing Eq. (15) yields

i(ωk·u0)n0Te1=ik·qe1+ikue1xn0Te0Ce1.
(21)

We now need fluid closures for qe1||,qe1,Ce1||, and Ce1. First, the 3 + 1 fluid model gives us7 

qe=ẑmeωce×(peTe+TepeTe2πeTepe)+qezẑ,
(22)

where ωce=eB0/me,πe=2(pepe)/3 and Te||=pe||/ne. After linearization, the x component of qe1 is

qe1x=2Te03(Te0+Ti0)n0ue0x(Te1||Te1)=rten0ue0x(Te1||Te1),
(23)

where rte=2Te0/3(Te0+Ti0). For qe, we derive a closure in  Appendix A, which can be written as

qe=ẑmeωce×[(56pe+176pe)Te(29Te+49Te)pe+(89Te29Te)pe]+qezẑ.
(24)

After linearization, the x component of qe1 is

qe1x=2Te03(Te0+Ti0)n0ue0x(Te1||Te1)=rten0ue0x(Te1||Te1).
(25)

For qe1z and qe1z, we employ a closure for plasmas with arbitrary collisionality, which can be written as25 

qe1z=65he1+σe1,
(26)
qe1z=25he112σe1,
(27)

where

he1=12ik¯||K¯hhn0vteTe1*+ik¯||K¯hσvteπe1||+K¯hRn0Te0(ue1zui1z)+iK¯hSvteπe1||,
(28)
σe1=43ik¯||K¯hσn0vteTe1*ik¯||K¯σσvteπe1||+K¯σRn0Te0(ue1zui1z)+iK¯σSvteπe1||.
(29)

Here, Te1*=Te1+2πe1||/5n0,vte=2Te0/me is the electron thermal speed, and k¯||=k||λc is the normalized parallel wave number. The electron collision length is defined as λcvteτee, and the electron–electron collision time τee is given by

τee=62π3/2ε02meTe03/2n0e4lnΛee,
(30)

where lnΛee is the Coulomb logarithm for electron–electron collisions and ε0 is the permittivity of free space. In Eqs. (28) and (29), K¯AB represents a kernel function that is obtained from a 6400 moment solution.25 The kernel function K¯AB has the following form:

K¯AB=ak¯||α1+d1k¯||δ+d2k¯||2δ+d3k¯||3δ+d4k¯||4δ+d5k¯||5δ+d6k¯||6δ,
(31)

where the values of coefficients, such as a, α, and δ in Eq. (31), are given in Table I in Ji and Joseph.25 For a negative k¯||,K¯AB(k¯||)=K¯AB(k¯||) if α = 0 or α = 2. When α = 1, K¯AB(k¯||)=K¯AB(k¯||). These closures are consistent with those of Hammett and Perkins29 in the collisionless limit, and they become consistent with those of Braginskii30 in the collisional limit.

The heat generated by the collision terms Ce1|| and Ce1 also needs a closure and can be written as

Ce1||=23Qe1+Se1||,
(32)
Ce1=23Qe112Se1||,
(33)

where Qe is the heat generated by collisions and Se|| is related to the temperature anisotropy.25 The closure for Se1|| is given by25 

Se1=43k¯||K¯hSn0τeeTe1*+k¯||τeeK¯σSπe1||+i83K¯RSn0Te0vteτee(ue1zui1z)2.05K¯SSτeeπe1||.
(34)

The heat generated by collisions can be written as26 

Qe=3menemiτei(TiTe)uei·Re,
(35)

where τei is the electron–ion collision time and uei=ueui is the relative flow velocity between electrons and ions. Assuming the ion charge status Zi is unity, τei is

τei=62π3/2ε02meTe03/2n0e4lnΛei,
(36)

where lnΛei is the Coulomb logarithm for electron–ion collisions. Linearizing Qe yields

Qe1=3mene1miτei(Ti0Te0)+3men0miτei(Ti1Te1)ue0·Re1uei1·Re0.
(37)

We also need an expression for the resistivity. Since there is no temperature gradient in the equilibrium quantities, the zeroth-order resistivity Re0 can be written as26 

Re0=α||men0τeiue0zẑαmen0τeiue0xx̂.
(38)

For Zi=1, the two coefficients are26 

α||=0.504,
(39)
α=11.46r+1.06r530.081r43+2.97r+2.13,
(40)

where r=ωceτee. There are additional terms in Re1 since temperature gradients exist in the first order. The parallel (z) component of Re1 is25 

Re1||=ik¯||K¯hRvteτeen0Te1*i34k¯||K¯σRvteτeeπe1||(1K¯RR)n0meτeeuei1z+i2K¯RSvteτeeπe1||.
(41)

Equation (41) can be written as

Re1||=ikn0γez||Te1||ikn0γezTe1(men0/τee)(1K¯RR)uei1z,
(42)

where

γez||=35K¯hR+12K¯σR4K¯RS3k¯||,
(43)
γez=25K¯hR12K¯σR+4K¯RS3k¯||.
(44)

The x component of Re1 is26 

Re1=αmen0τeiuei1xαmeue0xτeine1ikβn0Te1,
(45)

where β for Zi=1 is given by26 

β=6.33r+2.47r83+2.75r73+3.99r2+5.31r53+8.23r+3.52.
(46)

Finally, the y component of Re1 is given by Re1×=α×men0uei1y/τei. Here, the coefficient α× for Zi=1 is26 

α×=r(2.53r+0.81)r83+2.54r73+6.14r2+7.35r53+11.22r+4.09.
(47)

With these closures, the first-order momentum equation [Eq. (8)] can be used to obtain the perturbed electron current density Je1. Then, the Maxwell equation [Eq. (5)] can be written as

(DxxDxyDxzDyxDyyDyzDzxDzyDzz)(E1xE1yE1z)=0.
(48)

The detailed derivation of each component of tensor D can be found in  Appendix B.

Dispersion relations for the lower hybrid drift waves are obtained from |D|=0, where |D| is the determinant of the tensor D; from this equation, the normalized angular frequency Ω is computed numerically for the given k and θ. Required input parameters are B0, n0, Te0,Ti0,ue0z, and ue0x. In addition, the ion mass has to be specified.

Compared to the previous collisionless model in Yoo et al.,7 there are two significant changes in the current model: the inclusion of the first-order perturbation of the perpendicular electron temperature (Te1) and the use of collisional closures. To understand the effects of each change, we obtain dispersion relations from four different models—(i) the collisionless model in Ref. Yoo et al.,7 (ii) a model with collisional closures but without Te1, (iii) the current model in the collisionless limit τee, and (iv) the current model.

First, we obtain dispersion relations with typical plasma and field parameters near the electron diffusion region of the MRX during reconnection with a guide field; B0=180 Gauss, n0=2×1013 cm−3, Te0=Ti0=10 eV, ue0z=130 km/s, and ue0x=50 km/s. Here, the ion species is singly ionized helium. Justified by previous measurements in MRX,19,31 we assume that Zi=1. With these parameters, τeeωce=157,βe is 0.25 and VA is 44 km/s. Note that ue0x exceeds VA, which is a necessary condition for LHDWs to have large growth rates.

Figure 2 shows dispersion relations from the four models. Left (right) panels are contour plots of the real (imaginary) part of the angular frequency as a function of kρe and θ. Here ρe=vte/ωce is the electron gyroradius. From now on, ω represents the real part of the angular frequency and γ represents the imaginary part. Both ω and γ are normalized to the (angular) lower hybrid frequency, ωLH. All four models are qualitatively similar, showing strong growth rates (γ0.6ωLH) for the ES-LHDW. The ES-LHDW propagates almost perpendicular to B0 (θ90°) with ωωLH. The peak growth rate occurs at kρe0.7 and θ91°. Here kρe0.7 corresponds to λ0.6 cm. These similarities among the four models indicate that the effects of Coulomb collisions on the ES-LHDW are limited for typical MRX parameters. Moreover, inclusion of Te1 also has a limited impact on the dispersion.

FIG. 2.

Dispersion relation of the LHDW with typical MRX parameters near the electron diffusion region with a high guide field. Left (right) panels show the real (imaginary) part of the angular frequency as a function of k and θ. (a) Collisionles model without Te1. (b) Collisional model without Te1. (c) Model with Te1 in the collisionless limit (τee). (d) Collisional model with Te1 (the most complete model). The results from the four models qualitatively agree with each other; the quasi-electrostatic LHDW that propagates almost perpendicular to B0 is unstable. The maximum growth rate appears around kρe0.7 and θ91°. The growth rate of the mode decreases with the collisional effects (b) and (d), compared to the corresponding collisionless cases (a) and (c).

FIG. 2.

Dispersion relation of the LHDW with typical MRX parameters near the electron diffusion region with a high guide field. Left (right) panels show the real (imaginary) part of the angular frequency as a function of k and θ. (a) Collisionles model without Te1. (b) Collisional model without Te1. (c) Model with Te1 in the collisionless limit (τee). (d) Collisional model with Te1 (the most complete model). The results from the four models qualitatively agree with each other; the quasi-electrostatic LHDW that propagates almost perpendicular to B0 is unstable. The maximum growth rate appears around kρe0.7 and θ91°. The growth rate of the mode decreases with the collisional effects (b) and (d), compared to the corresponding collisionless cases (a) and (c).

Close modal

For a better comparison between the four models, the dispersion relation and growth rate of the ES-LHDW are presented in Fig. 3 for θ=91°. It is worth noting that including Coulomb collisions decreases the growth rate γ. This is understandable since collisions decrease the reaction of electrons to the external perturbation, such that they reduce the positive feedback from the plasma. The change in ω is not straightforward but is related to frequency shift due to additional terms of ue1x and ue1z. For example, the parallel force balance equation Eq. (B48) has the resistivity Re1||, which adds additional terms in αez in Eq. (B50). These additional terms can cause a shift in ω (note that αez has a dependency on ω via αe).

FIG. 3.

1D dispersion relation of the ES-LHDW for θ=91°. (a) ω/ωLH as a function of kρe. Including the collisional effects (solid lines) increases the real frequency, while models with Te1 (red lines) have lower ω. (b) γ/ωLH as a function of kρe. Collisional effects (solid lines) decrease γ, compared to the results from the corresponding collisionless cases (dashed lines).

FIG. 3.

1D dispersion relation of the ES-LHDW for θ=91°. (a) ω/ωLH as a function of kρe. Including the collisional effects (solid lines) increases the real frequency, while models with Te1 (red lines) have lower ω. (b) γ/ωLH as a function of kρe. Collisional effects (solid lines) decrease γ, compared to the results from the corresponding collisionless cases (dashed lines).

Close modal

It is interesting to see that including Te1 in the electron dynamics decreases both ω and γ of the ES-LHDW. Interpreting this trend is complicated, because Te1 impacts both the x and z components of the electron momentum equation. For the x component, the first term (ikn0Te1) on the right side of Eq. (B55), which is the perturbed perpendicular electron pressure gradient term, directly contains Te1. For the parallel momentum balance of Eq. (B48), Te1 affects Te1|| via qe1x|| in Eq. (23). The parallel resistivity [Eq. (42)] also has a term with Te1 (ikn0γezTe1).

The dispersion relation is calculated after setting γez=0 to remove contributions from Te1 in the z component of the electron force balance equation. As shown in Fig. 4, this change (green line) decreases ω and increases γ, compared to the reference case with Te1 (red line). Changes in ω and γ are not significant.

FIG. 4.

1D dispersion relation of the ES-LHDW for θ=91°. (a) ω/ωLH as a function of kρe for four cases with collisional effects. The blue (red) line indicates the reference case without (with) Te1. If Te1 is removed from the x component of the electron momentum equation (cyan line), ω becomes significantly larger. Removing the contribution from Te1 in the z component of the electron momentum equation (green line), on the other hand, reduces ω. (b) γ/ωLH as a function of kρe for four cases with collisional effects. Effects of Te1 on γ are not important, as all four cases show similar values.

FIG. 4.

1D dispersion relation of the ES-LHDW for θ=91°. (a) ω/ωLH as a function of kρe for four cases with collisional effects. The blue (red) line indicates the reference case without (with) Te1. If Te1 is removed from the x component of the electron momentum equation (cyan line), ω becomes significantly larger. Removing the contribution from Te1 in the z component of the electron momentum equation (green line), on the other hand, reduces ω. (b) γ/ωLH as a function of kρe for four cases with collisional effects. Effects of Te1 on γ are not important, as all four cases show similar values.

Close modal

The change in ω with Te1 is caused by the ikn0Te1 term in the x component of the electron momentum equation. As shown in Fig. 4(a), without the term (magenta line), ω increases significantly compared to the reference case with Te1 (red line). Removing the ikn0Te1 term also increases γ for most values of k. Again, these changes are caused by the frequency shift due to the additional term with ue1x; from Eqs. (B35) and (B55), the inertial term effectively changes from imen0(ωk·ue0)ue1x to imen0(ωk·ue00.5c¯uxkvte)ue1x.

We have repeated the dispersion calculation for the EM-LHDW that propagates obliquely to B0. The plasma and field parameters used for calculations are B0=30 Gauss, n0=2×1013 cm−3, Te0=Ti0=10 eV, ue0z=50 km/s, and ue0x=130 km/s. Again, the ion species is singly ionized helium and Zi=1. With these parameters, τeeωce=26.2,βe is 8.9 and VA is 7.3 km/s. These parameters represent typical MRX values near the electron diffusion region during reconnection with a negligible guide field.

As shown in Fig. 5, dispersion relations from the four models again qualitatively agree with each other; these models expect positive growth rates for the EM-LHDW. Models without Te1 have the maximum growth rate around kρe0.6 and θ55°, while those with Te1 have the maximum growth rate around kρe0.5 and θ50°. The wavelength with the largest growth rate is about 4 cm. In is interesting to see that all models expect that the mode has frequency significantly less than ωLH in the ion rest frame. This agrees with measurements in MRX and numerical simulations that show that most of the power of the EM-LHDW exists below ωLH.9,16

FIG. 5.

Dispersion relation of the LHDW with typical MRX parameters near the electron diffusion region with a negligible guide field. Left (right) panels show the real (imaginary) part of the angular frequency as a function of k and θ. (a) Collisionless model without Te1. (b) Collisional model without Te1. (c) Model with Te1 in the collisionless limit (τee). (d) Collisional model with Te1 (the most complete model). Again, the results from the four models qualitatively agree with each other; the electromagnetic LHDW that propagates obliquely to B0 is unstable. The maximum growth rate appears around kρe0.5 and θ50°. The growth rate of the mode decreases with collisional effects (b) and (d), compared to the corresponding collisionless cases (a) and (c).

FIG. 5.

Dispersion relation of the LHDW with typical MRX parameters near the electron diffusion region with a negligible guide field. Left (right) panels show the real (imaginary) part of the angular frequency as a function of k and θ. (a) Collisionless model without Te1. (b) Collisional model without Te1. (c) Model with Te1 in the collisionless limit (τee). (d) Collisional model with Te1 (the most complete model). Again, the results from the four models qualitatively agree with each other; the electromagnetic LHDW that propagates obliquely to B0 is unstable. The maximum growth rate appears around kρe0.5 and θ50°. The growth rate of the mode decreases with collisional effects (b) and (d), compared to the corresponding collisionless cases (a) and (c).

Close modal

For comparison between the four models, ω and γ as a function of k for θ=55° are presented in Fig. 6. Similar to the ES-LHDW case, collisional effects decrease γ regardless of the existence of Te1 in the model. This is consistent with the aforementioned explanation; collisions decrease the reaction of electrons to the external perturbation, thereby decreasing the positive feedback. For the EM-LHDW, collisions generally decrease ω especially when Te1 is not included in the model (blue lines). Including Te1 further decreases both ω and γ for this mode (red lines).

FIG. 6.

1D dispersion relation of the EM-LHDW for θ=55°. (a) ω/ωLH as a function of kρe. Models with Te1 (red lines) have lower ω. The impact of Coulomb collisions on ω is negligible. (b)γ/ωLH as a function of kρe. Collisional effects (solid lines) decreases γ, compared to the results from the corresponding collisionless cases (dashed lines).

FIG. 6.

1D dispersion relation of the EM-LHDW for θ=55°. (a) ω/ωLH as a function of kρe. Models with Te1 (red lines) have lower ω. The impact of Coulomb collisions on ω is negligible. (b)γ/ωLH as a function of kρe. Collisional effects (solid lines) decreases γ, compared to the results from the corresponding collisionless cases (dashed lines).

Close modal

In summary, we have developed a local, linear model of LHDWs that includes effects of Coulomb collisions and Te1. This model works best for plasmas with weak collisionality. Without collisions, some assumptions for the 3 + 1 model may not be valid, as the zeroth-order distribution function is not close to a Maxwellian. In addition, in the collisionless plasma, agyrotropy can be developed, while a gyrotropic electron pressure tensor is assumed in this model. For collisional plasmas, we need to consider the zeroth-order electric field along the x and z directions; for the zeroth-order electron force balance, additional components of E0 are needed to balance the zeroth-order resistivity Re0. If there are too many collisions, we need additional first-order terms (eE0xne1 and eE0zne1) in the x and z components of the electron momentum equation [Eq. (8)]. From Eq. (38), required equilibrium electric field components are given by E0z=α||B0ue0z/ωceτei and E0x=αB0ue0x/ωceτei. From Eq. (3), E0x/E0 is given by

E0xE0=αTe0Te0+Ti01ωceτei1ωceτee,
(49)

because αTe0/(Te0+Ti0)1 and τeiτee for Zi=1. This means that E0x is negligible compared to E0, as long as electrons are fully magnetized (ωceτee1), which is one of the basic assumptions of this model. From a similar argument, E0z is also negligible unless |ue0z||ue0x|. For the two cases presented here, the effects of both E0x and E0z are expected to be minimal since |ue0z||ue0x| and ωceτee1.

To verify this argument, we have calculated dispersion relations of LHDWs after including two additional terms (eE0xne1 and eE0zne1) and have found that impacts from these terms are actually negligible. The basic reason for not including additional components of E0 in the current model is that including E0x may require an additional electron flow component along the y direction, since there will be a corresponding E×B drift of electrons, while ions are unmagnetized. This means that collisions may impact the dynamics of LHDWs by changing the equilibrium itself. A future work will address this effect in a self-consistent manner. As the main purpose of the current study is to study collisional effects on LHDWs, we minimize other changes for simplicity. The parallel component of the equilibrium electric field E0z, on the other hand, can be easily added in the model without creating complexity. Moreover, E0z in the electron diffusion region during reconnection with a strong guide field may significantly exceed the value required to balance the classical resistivity.32 In the future, we will study the possible impacts of E0z on LHDWs with values measured in MRX during guide field reconnection.

With this model, we have calculated two sets of LHDW dispersion relations for typical MRX parameters. The first case uses parameters from the electron diffusion region during reconnection with a significant guide field, while the second one uses those with a negligible guide field. Due to the presence of the guide field, the first case has a low electron beta (βe=0.25), such that the ES-LHDW is unstable in that region. For the second case (βe=8.9), on the other hand, the ES-LHDW is stabilized by the high beta effect17 and the EM-LHDW is unstable instead.

It will be interesting to study the critical value of βe that determines whether the ES- or EM-LHDW is unstable. Initial studies show that the critical value is determined by the value of ue0x/VA; for a relatively low (1) value of ue0x/VA like the first case, βe also has to be low (0.5) to have the ES-LHDW unstable. For a high value (>10) of ue0x/VA, on the other hand, the ES-LHDW exists at the higher βe1. We plan to conduct a statistical study with data from MMS and/or MRX, which will be compared to the results from the current theoretical model.

Based on the two cases we have studied, collisional effects on LHDWs in typical MRX current sheets are limited. In both cases, including Coulomb collisions in the model decreases the growth rate. However, the difference in γ is relatively small (20%). This is because the wavelengths of LHDWs (0.5–5 cm) are smaller than the mean free path of electrons (10 cm) and electrons are fully magnetized (ωceτee1) for these parameters.

To further investigate how collisions may impact on the dispersion relation, we have artificially varied τee and τei. For the ES-LHDW, artificially high collisions significantly affect the dispersion relation and the growth rate, as shown in Figs. 7(a) and 7(b). When the collisions are enhanced by a factor of 5 (red dashed line), the real frequency becomes larger for kρe>0.2 than the reference value (blue solid line). There is also a significant decrease in the growth rate for kρe>0.7. Changes in less collisional cases, on the other hand (green solid and dashed lines), are minimal. With the reduced collision time (τee0.2τee), the mean free path (τeevte) becomes about 2 cm, which corresponds to kρe0.2. This supports the insertion that collisions have large impacts on modes with a wavelength comparable to the mean free path (λ2πτeevte).

FIG. 7.

1D dispersion relations with various collisionalities for the two cases. (a) ω/ωLH as a function of kρe for the ES-LHDW case. When τee is artificially decreased to 0.2τee (red dashed line), which means that collisions are enhanced by a factor of 5, there is a significant increase in ω when kρe>0.4. The same change is also applied to the other collision time, τei. The blue line indicates the reference value without any change in the collision time. (b)γ/ωLH as a function of kρe for the ES-LHDW case. When collisions are enhanced (red solid and dashed lines), there are noticeable changes in γ. (c)ω/ωLH as a function of kρe for the EM-LHDW case. When collisions are enhanced, there are large changes in the dispersion. (d)γ/ωLH as a function of kρe for the EM-LHDW case. When collisions are enhanced (red solid and dashed lines), the growth rate with smaller kρe decreases notably.

FIG. 7.

1D dispersion relations with various collisionalities for the two cases. (a) ω/ωLH as a function of kρe for the ES-LHDW case. When τee is artificially decreased to 0.2τee (red dashed line), which means that collisions are enhanced by a factor of 5, there is a significant increase in ω when kρe>0.4. The same change is also applied to the other collision time, τei. The blue line indicates the reference value without any change in the collision time. (b)γ/ωLH as a function of kρe for the ES-LHDW case. When collisions are enhanced (red solid and dashed lines), there are noticeable changes in γ. (c)ω/ωLH as a function of kρe for the EM-LHDW case. When collisions are enhanced, there are large changes in the dispersion. (d)γ/ωLH as a function of kρe for the EM-LHDW case. When collisions are enhanced (red solid and dashed lines), the growth rate with smaller kρe decreases notably.

Close modal

For the case of the EM-LHDW, the effects from collisions become significant when collisions are enhanced by a factor of 5 or more (τee0.2τee and τei0.2τei). As denoted by the red line in Fig. 7(c), the overall shape of the dispersion relation changes noticeably, when τee is reduced to 0.2τee. The mean free path with 0.2τee is about 2 cm (the same electron temperature and density as the first case), and the change starts around 0.2kρe. When τee reduces even further to 0.1τee (red dashed line), the deviation from the reference line starts around 0.1kρe. For both cases, there are also significant reductions in γ, as shown in Fig. 7(d) especially for kρe<0.7.

This means that parameters for the two cases studied here are actually in the weakly collisional regime and that the dynamics of LHDWs are susceptible to collisional effects only when collisions are strong. For example, if the base electron temperature for both cases is 3 eV, the dispersion relation from this collisional model will be vastly different from that of the collisionless model.

Including Te1 in the model has limited impacts on the dispersion; it generally decreases the frequency and growth rate of LHDWs, but changes in ω and γ are less than 20% for both cases. These changes mostly come from the additional pressure gradient term (ikn0Te1) in the electron momentum equation along the x direction. This limited impact is related to the existence of Lorentz force terms along the perpendicular direction;7 because of these terms, the electron force balance is less sensitive to the pressure gradient term along the perpendicular direction.

It should be noted that the current theoretical model ignores the global structure of the current sheet by assuming that there is no wave propagation along the density gradient direction (y direction in Fig. 1). To address the effects from the global current sheet structure, an eigenmode analysis21,33 or numerical simulations22,23 will have to be carried out, which will be one of our future works. In MRX, where the current sheet is actually broader (10de; de is electron skin depth), this local approximation is generally valid, as the length scale along the y direction is larger than the wavelength of LHDWs.

This model assumes that there is no equilibrium temperature gradient across the current sheet. In MRX, electrons are locally heated in the current sheet.20,34 However, inside the current sheet the temperature gradient is rather small, compared to that of density. Therefore, the effects of the temperature gradient are expected to be negligible.24 

This study will provide a theoretical framework for quantifying anomalous terms and heating associated with LHDWs in MRX. With the solved dispersion relation, we can express every fluctuating quantity in terms of a measurable quantity. For example, the first-order density perturbation [Eq. (B81)] can be expressed in terms of the fluctuation in the reconnection electric field (δErec) that can be measured with a probe.8,35 Then, the wave-associated anomalous drag term D=δneδErec/ne36 can be estimated by measuring δErec. Here, the assumption is that the linear relation holds, such that we can use ne1δne. Furthermore, this model can provide direct estimates of wave-associated heating in Eq. (35) via the same quasi-linear argument. This estimate cannot be done with other collisionless models. In the future, we will establish quasi-linear calculations and conduct measurements of LHDWs in MRX to find out how LHDWs affect the electron and reconnection dynamics.

This work was supported by DOE Contract No. DE-AC0209CH11466, NASA Grant Nos. NNH20ZDA001N and 80HQTR21T0060, NNSFC Contract No. 11975163, the Priority Academic Program Development of Jiangsu Higher Education Institutions (PAPD), and a DOE Grant No. DE-FG02–04ER54746. Digital data used are available in the DataSpace of Princeton University (http://arks.princeton.edu/ark:/88435/dsp01x920g025r).

The authors have no conflicts to disclose.

The data that support the findings of this study are openly available in the DataSpace of Princeton University http://arks.princeton.edu/ark:/88435/dsp01b2773z812, Ref. 38.

From the kinetic equation in the (t,r,wvV) coordinates (V is the fluid velocity),

dfdt(w·V)·wf+·(wf)+w·(Af)+qmw×B·wf=C(f),
(A1)

where

ddt=t+V·,
(A2)
A=1m[F+q(V×B)]dVdt.
(A3)

For the p|| fluid equation, we need to obtain the closure,

q||=d3vmw||2wf=q||||ẑ+q||,
(A4)
q=h=d3v12mw2wf=h||ẑ+h,
(A5)

where

q||||=d3vmw||3f=65h||+σ||,
(A6)
q||=d3v12mw2w||f=25h||12σ||
(A7)

have been obtained in Ji and Joseph,25 and the q|| has been obtained in Yoo et al.7 Now we obtain

q=d3v12mw2wf=q||ẑ+q.
(A8)

Note that q can be obtained from

h=q=d3v12mw2wf=12q||+q.
(A9)

We adopt the closure (transport) ordering d/dt0 and the linear response theory, linear in thermodynamic drives, i.e., T, p|| and p.

We take the moments d3v12mw2w of the kinetic equation:

d3v12mw2wdfdt=ddtq:ignoredbytheclosureordering,d3v12mw2w(w·V)·wf:ignoredbythelinearization,d3v12mw2w·(wf)=·(d3v12mw2wwf).

We should decompose wwww into orthogonal polynomials (see Ji and Held37) for the consistent truncation in the expansion of a distribution function.

c=wvT=w2T/m.
(A10)

In terms of orthogonal basis

c2cc=c2(cc13c2I)+13c4I=(c272)(cc13c2I)+72(cc13c2I)+13c4I=p21+72p20+23(12c452c2+158)I+23(52c2158)I=p21+72p20+23p02I+[53(c232)+5254]I=p21+72p20+23p02I+(53p01+54)I,
(A11)
dv12mw2wwf12mvT4[72p20+(53p01+54)I]=7212vT2π+12mvT454nI=72Tmπ+52TmpI=72TmpTmpI.
(A12)

Hereafter → will be used to drop b terms, which will be nullified by the b× operation,

·π=32b||π||12π||12π||.
(A13)

For the w·(Af) term

A=1m[F+q(V×B)]dVdt=1mn(p+·π).
(A14)
dv12mw2ww·(Af)=dvmA·w(12w2w)f=dvm(A·ww+12w2A)f=p·A32pA=A·π52pA.
(A15)

All together ·(wf)+w·(Af)

all=·(72Tmπ+52TmpI)1mn(p+·π)·π52p1mn(p+·π)=72m(T·π+T·π¯1)+52m(pT+Tp¯0)1mn(p+·π)·π52p1mn(p¯0+·π¯1)=72mT·π+1mT·π+52mpT1mn(p+·π)·π,
(A16)
d3v12mw2ww·(w×Bf)=12md3v(w×Bf)·w(w2w)=12md3v(w×Bf)·(2ww+w2I)=12md3vw2w×Bf=h×B,
(A17)
qmd3v12mw2ww·(w×Bf)=Ωh×ẑ.
(A18)

The final equation becomes up to O(Ω0)

(termsdroppedbyclosureordering)+all+(termsb)Ωh×ẑ=(collisiontermsb)
h=1Ωẑ×all
h=1mΩb×[52mT·π+1mT·π+52mpT1mn(·π)·π].
(A19)

Since we are interested in q up to O(Ω1), we consider only the CGL viscosity, which is O(Ω0)

π=32π||(bb13I),
(A20)
·π=32b||π||12π||=12π||+bterms
(A21)

and

(·π)·π=(32b||π||12π||)·32π||(bb13I)=14π||π||+bterms,
(A22)
all=72mT·π+1mT·π+52mpT1mn(p+·π)·π=74mπ||TT2mπ||+52mpT+π||2mnp14mnπ||π||+bterms,
(A23)
q=1mΩb×(74π||TT2π||+52pT+π||2np14nπ||π||),
(A24)
q=q12q||,
(A25)

where7 

q||=1mΩb×(p||T+Tp||T2π||p||np).
(A26)

Finally,

q=q||ẑ+q.
(A27)

One can rewrite equations in terms of p|| and p using

π||=23(p||p),
(A28)
p=13(p||+2p)=nT.
(A29)

In terms of Te1|| and Te1,qe1z and qe1z [Eqs. (26) and (27)] can be expressed as

qe1z=ic¯q||||n0vteTe1||ic¯q||n0vteTe1+c¯qu||n0Te0uei1z,
(B1)
qe1z=ic¯q||n0vteTe1||ic¯qn0vteTe1+c¯qun0Te0uei1z,
(B2)

where six dimensionless parameters are defined as

c¯q||||=925k¯||K¯hh85k¯||K¯hσ+23k¯||K¯σσ45K¯hS23K¯σS,
(B3)
c¯q||=625k¯||K¯hh+415k¯||K¯hσ23k¯||K¯σσ+45K¯hS+23K¯σS,
(B4)
c¯qu||=65K¯hR+K¯σR,
(B5)
c¯q||=325k¯||K¯hh+215k¯||K¯hσ13k¯||K¯σσ415K¯hS+13K¯σS,
(B6)
c¯q=225k¯||K¯hh+815k¯||K¯hσ+13k¯||K¯σσ+415K¯hS13K¯σS,
(B7)
c¯qu=25K¯hR12K¯σR.
(B8)

Here, uei1z=ue1zui1z is the first-order relative flow velocity along the z direction.

With Eqs. (7), (9), (37), (38), (42), and (45), Qe1 can be written as

Qe1=(1K¯RR+α||τeeτei)n0meue0zτeeuei1z+c¯Q||n0τeeTe1||+c¯Qn0τeeTe1+AQ,
(B9)

where

c¯Q||=meτeemiτei+ik¯||ue0zvte(3K¯hR5+K¯σR24K¯RS3k¯||)+ik¯βue0x3vte,
(B10)
c¯Q=2meτeemiτei+ik¯||ue0zvte(2K¯hR5K¯σR2+4K¯RS3k¯||)+2ik¯βue0x3vte,
(B11)
AQ=[3me(Ti0Te0)miτei+αmeue0x2τei]ne1+3men0miτeiTi1+2αmen0.ue0xτeiuei1x.
(B12)

With Eqs. (32)–(34), and (B9), Ce1|| and Ce1 can be written as

Ce1||=c¯C||||n0Te1||τee+c¯C||n0Te1τee+c¯Cu||n0Te0uei1zτeevte+23AQ,
(B13)
Ce1=c¯C||n0Te1||τee+c¯Cn0Te1τee+c¯Cun0Te0uei1zτeevte+23AQ,
(B14)

where six dimensionless parameters are given by

c¯C||||=23c¯Q||+45k¯||K¯hS+23k¯||K¯σS2(2.05K¯SS)3,
(B15)
c¯C||=23c¯Q+815k¯||K¯hS23k¯||K¯σS+2(2.05K¯SS)3,
(B16)
c¯Cu||=43(1K¯RR+α||τeeτei)ue0zvte+8i3K¯RS,
(B17)
c¯C||=23c¯Q||25k¯||K¯hS13k¯||K¯σS+2.05K¯SS3,
(B18)
c¯C=23c¯Q415k¯||K¯hS+13k¯||K¯σS2.05K¯SS3,
(B19)
c¯Cu=43(1K¯RR+α||τeeτei)ue0zvte4i3K¯RS.
(B20)

With these closures, Eqs. (20) and (21) can be written as

irαeTe1||=c¯||||Te1||+c¯||Te1+c¯u||Te0ue1zvtec¯uxTe0ue1xvtec¯nTe0ne1n0+At||,
(B21)
irαeTe1=c¯||Te1||+c¯Te1+c¯uTe0ue1zvte+(ik¯c¯ux)Te0ue1xvtec¯nTe0ne1n0+At,
(B22)

where

αe=(ωk·ue0)/ωce,
(B23)
c¯||||=k¯||c¯q||||c¯C||||+irtekτeeue0x,
(B24)
c¯||=k¯||c¯q||c¯C||irtekτeeue0x,
(B25)
c¯u||=ik¯||c¯qu||c¯Cu||+2ik¯||,
(B26)
c¯ux=8ατeeue0x3τeivte,
(B27)
c¯n=2meτeemiτei(Ti0Te01)+4ατeeue0x23τeivte2,
(B28)
c¯||=k¯||c¯q||c¯C||irtekτeeue0x,
(B29)
c¯=k¯||c¯qc¯C+irtekτeeue0x,
(B30)
c¯u=ik¯||c¯quc¯Cu,
(B31)
At||=2meτeemiτeiTi1(ik¯||c¯qu||c¯Cu||)Te0ui1zvte+8ατeeue0x3τeivteTe0ui1xvte,
(B32)
At=2meτeemiτeiTi1(ik¯||c¯quc¯Cu)Te0ui1zvte+8ατeeue0x3τeivteTe0ui1xvte.
(B33)

With Eqs. (B21) and (B22), Te1|| and Te1 can be written as

Te1||=c¯uz||Te0ue1zvte+c¯ux||Te0ue1xvte+c¯n||Te0ne1n0+Ai||,
(B34)
Te1=c¯uzTe0ue1zvte+c¯uxTe0ue1xvte+c¯nTe0ne1n0+Ai,
(B35)

where

c¯uz||=(irαec¯)c¯u||+c¯||c¯u(irαec¯||||)(irαec¯)c¯c¯||,
(B36)
c¯ux||=(irαec¯)c¯uxc¯||(ik¯c¯ux)(irαec¯||||)(irαec¯)c¯c¯||,
(B37)
c¯n||=(irαec¯+c¯||)c¯n(irαec¯||||)(irαec¯)c¯c¯||,
(B38)
c¯uz=(irαec¯||||)c¯u+c¯||c¯u||(irαec¯||||)(irαec¯)c¯c¯||,
(B39)
c¯ux=(irαec¯||||)(ik¯c¯ux)c¯||c¯ux(irαec¯||||)(irαec¯)c¯c¯||,
(B40)
c¯n=(irαec¯||||+c¯||)c¯n(irαec¯||||)(irαec¯)c¯c¯||.
(B41)

The additional ion terms Ai|| and Ai can be expressed as

Ai||=c¯i||||At||+c¯i||At,
(B42)
Ai=c¯i||At||+c¯iAt,
(B43)

where

c¯i||||=irαec¯(irαec¯||||)(irαec¯)c¯c¯||,
(B44)
c¯i||=c¯||(irαec¯||||)(irαec¯)c¯c¯||,
(B45)
c¯i||=c¯||(irαec¯||||)(irαec¯)c¯c¯||,
(B46)
c¯i=irαec¯||||(irαec¯||||)(irαec¯)c¯c¯||.
(B47)

The z component of Eq. (8) is

imen0(ωk·u0)ue1z=ikpe1+en0(E1z+u0xB1y)Re1||.
(B48)

From the Faraday's Law (ωB1=k×E1), B1y=(kE1xkE1z)/ω. With Eqs. (9), (42), (B34), (B35), and (B48), ue1z is expressed as

iαezue1z=ic¯xzue1x+c¯yzue1y+Aez+Aiz,
(B49)

where

αez=αek||vte2ωce[c¯uz||+γez||c¯uz||+γezc¯uz2i(1K¯RR)k¯||+kvteαeωce(1+c¯n||+γez||c¯n||+γezc¯n)],
(B50)
c¯xz=k||vte2ωce[c¯ux||+γez||c¯ux||+γezc¯ux+kvteαeωce(1+c¯n||+γez||c¯n||+γezc¯n)],
(B51)
c¯yz=εk||vte22αeωce2(1+c¯n||+γez||c¯n||+γezc¯n),
(B52)
Aez=E1zB0+ku0xωE1xcosθE1zsinθB0,
(B53)
Aiz=ikeB0(Ai||+γez||Ai||+γezAi)1K¯RRωceτeeui1z.
(B54)

The x component of Eq. (8) is

imen0(ωk·ue0)ue1x=ik(n0Te1+Te0ne1)+en0(E1x+B0ue1yue0zB1y)Re1.
(B55)

With Eqs. (9), (45), (B34), (B35), (B49), and (B55), ue1y can be expressed as

γeyue1y=iαexue1xAexAixc¯zxkvte2αezωce(Aez+Aiz),
(B56)

where γey,αex, and Aex are

γey=1+c¯nxεkvte22αeωce2+c¯zxc¯yzkvte2αezωce,
(B57)
αex=αec¯nxk2vte22αeωce2c¯zxc¯xzkvte2αezωcekvte2ωce[βc¯ux||3+(1+2β3)c¯ux2iατeek¯τei],
(B58)
Aex=E1xB0ku0zωE1xcosθE1zsinθB0,
(B59)
Aix=ikeB0[βAi||3+(1+2β3)Ai]αui1xτeiωce.
(B60)

Here, two dimensionless parameters are given by

c¯nx=1+βc¯n||3+(1+2β3)c¯n2iατeeue0xk¯τeivte,
(B61)
c¯zx=βc¯uz||3+(1+2β3)c¯uz+c¯nxkvteαeωce.
(B62)

Similarly, the y component of Eq. (8) is

imen0(ωk·u0)ue1y=en0(E1yB0ue1xue0xB1z+ue0zB1x)+e(E0ue0xB0)ne1Re1×.
(B63)

With Eqs. (3), (9), and (B49), ue1x can be expressed as

γexue1x=iαeyue1y+3irteku0x2αeαezωce(Aez+Aiz)+Aey+Aiy,
(B64)

where γex,αey,Aey, and Aiy are

γex=1+3rtekue0x2αeωce(1+c¯xzkαezk),
(B65)
αey=αeiα×ωceτei3rteεue0x2αeωce(1+c¯yzkαezε),
(B66)
Aey=E1yB0kω(u0xsinθ+u0zcosθ)E1yB0,
(B67)
Aiy=α×ωceτeiui1y.
(B68)

With Eqs. (B56) and (B64), ue1y is given by

ue1y=i[iCyxe(Aex+Aix)+Cyye(Aey+Aiy)+iCyze(Aez+Aiz)],
(B69)

where

Cyxe=(γeyαexαeyγex)1,
(B70)
Cyye=Cyxeαexγex,
(B71)
Cyze=Cyxe(c¯zxkvte2αezωce+3rteαexkue0x2γexαeαezωce).
(B72)

Similarly, ue1x is given by

ue1x=iCxxe(Aex+Aix)+Cxye(Aey+Aiy)+iCxze(Aez+Aiz),
(B73)

where

Cxye=(γexαexαeyγey)1,
(B74)
Cxxe=Cxyeαeyγey,
(B75)
Cxze=Cxye[3rteku0x2αeαezωce+αeyc¯zxkvte2γeyαezωce].
(B76)

Then, ue1z can be written as

ue1z=iCzxe(Aex+Aix)+Czye(Aey+Aiy)+iCzze(Aez+Aiz),
(B77)

where

Czze=1αez+c¯xzCxzeαez+c¯yzCyzeαez,
(B78)
Czxe=c¯xzCxxeαez+c¯yzCyxeαez,
(B79)
Czye=c¯xzCxyeαez+c¯yzCyyeαez.
(B80)

The final goal is to obtain the perturbed current density of electrons, which is given by J1e=en0ue1eue0ne1. Thus, an expression for ne1 is required. From Eqs. (9), (B69), (B73), and (B77), ne1 is given by

ne1=kn0ωk·ue0[iCxe(Aex+Aix)+Cye(Aey+Aiy)+iCze(Aez+Aiz)],
(B81)

where

Cxe=Cxxesinθ+Cyxeε/k+Czxecosθ,
(B82)
Cye=Cxyesinθ+Cyyeε/k+Czyecosθ,
(B83)
Cze=Cxzesinθ+Cyzeε/k+Czzecosθ.
(B84)

Now, we are ready for computing the dispersion relation. Equation (5) is

k2E1xkkE1ziωμ0J1x=0,
(B85)
k2E1yiωμ0J1y=0,
(B86)
k2E1zkkE1xiωμ0J1z=0.
(B87)

By multiplying by di2 (dic/ωpi is the ion skin depth; ωpi is ion plasma frequency), the above equation can be written as

K2cos2θE1xK2sinθcosθE1ziΩB0en0J1x=0,
(B88)
K2E1yiΩB0en0J1y=0,
(B89)
K2sin2θE1zK2sinθcosθE1xiΩB0en0J1z=0,
(B90)

where Kkdi and Ω=ω/ωci.

From Eq. (6), each component of iΩB0J1i/en0 is

iΩB0en0J1xi=ζZE1x+ζZsinθ2(E1xsinθiεkE1y+E1zcosθ),
(B91)
iΩB0en0J1yi=ζZE1y,
(B92)
iΩB0en0J1zi=ζZE1z+ζZcosθ2(E1xsinθiεkE1y+E1zcosθ).
(B93)

From Eqs. (B73) and (B81), iJ1xe/en0 is given by

iJ1xeen0=Cxxe(Aex+Aix)iCxye(Aey+Aiy)+Cxze(Aez+Aiz),
(B94)

where Cxxe=Cxxe+kue0xCxe/(ωk·ue0),Cxye=Cxye+kue0xCye/(ωk·ue0), and Cxze=Cxze+kue0xCze/(ωk·ue0). Similarly, from Eqs. (B77) and (B81), iJ1ze/en0 is given by

iJ1zeen0=Czxe(Aex+Aix)iCzye(Aey+Aiy)+Czze(Aez+Aiz),
(B95)

where Czxe=Czxe+kue0zCxe/(ωk·ue0),Czye=Czye+kue0zCye/(ωk·ue0), and Czze=Czze+kue0zCze/(ωk·ue0). Since there is no y component in ue0,iJ1ye/en0 is simply

iJ1yeen0=iCyxe(Aex+Aix)+Cyye(Aey+Aiy)+iCyze(Aez+Aiz).
(B96)

In terms of dimensionless parameters, ΩB0Aex,ΩB0Aey, and ΩB0Aez can be written as

ΩB0Aex=(ΩKUe0zcosθ)E1x+(KUe0zsinθ)E1z,
(B97)
ΩB0Aey=[ΩK(Ue0xsinθ+Ue0zcosθ)]E1y,
(B98)
ΩB0Aez=(KUe0xcosθ)E1x+(ΩKUe0xsinθ)E1z.
(B99)

Ue0=ue0/VA and VA=B0/μ0min0=diωci is the Alfvén speed.

With Eq. (7), Aiz in Eq. (B54) is

Aiz=c¯izxiJ1xien0+c¯izziJ1zien0+c¯izTB0[E1·k̂(2Z+Z4)iE1y(εk)(Z+Z4)],
(B100)

where three dimensionless parameters are given by

c¯izx=4α(c¯iz||+c¯iz)τeekue0x3τeiωce,
(B101)
c¯izz=[c¯iz||(ik¯||c¯qu||c¯Cu||)+c¯iz(ik¯||c¯quc¯Cu)]kvte2ωce+i(1K¯RR)ωceτee,
(B102)
c¯izT=2(c¯iz||+c¯iz)meτeecosθmiτei.
(B103)

Here, two additional parameters c¯iz|| and c¯iz are defined as

c¯iz||=(1+γez||)c¯i||||+γezc¯i||,
(B104)
c¯iz=(1+γez||)c¯i||+γezc¯i.
(B105)

Similarly, Aix is

Aix=c¯ixxiJ1xien0+c¯ixziJ1zien0+c¯ixTB0[E1·k̂(2Z+Z4)iE1y(εk)(Z+Z4)],
(B106)

where three dimensionless parameters are given by

c¯ixx=4α(c¯ix||+c¯ix)τeekue0x3τeiωceατeeωce,
(B107)
c¯ixz=[c¯ix||(ik¯||c¯qu||c¯Cu||)+c¯ix(ik¯||c¯quc¯Cu)]kvte2ωce,
(B108)
c¯ixT=2(c¯ix||+c¯ix)meτeesinθmiτei.
(B109)

Two additional parameters c¯ix|| and c¯ix are

c¯ix||=β3c¯i||||+(1+2β3)c¯i||,
(B110)
c¯ix=β3c¯i||+(1+2β3)c¯i.
(B111)

The last ion term is Aiy=(α×/ωceτei)J1yi/en0.

Equations (B88)–(B90) can be written as

(DxxDxyDxzDyxDyyDyzDzxDzyDzz)(E1xE1yE1z)=0.
(B112)

Each component of the tensor D is

Dxx=K2cos2θCxxe(ΩKUe0zcosθ)CxzeKUe0xcosθCxxi(ζZ+ζZsin2θ2)CxziζZcosθsinθ2CxTiΩsinθ(2Z+Z4),
(B113)
Dxy=Cxyeα×ωceτeiζZ+iCxye[ΩK(Ue0xsinθ+Ue0zcosθ)]+i(εk)CxxiζZsinθ2+i(εk)CxziζZcosθ2+i(εk)CxTi(Z+Z4),
(B114)
Dxz=K2sinθcosθCxxeKUe0zsinθCxze(ΩKUe0xsinθ)CxxiζZ2sinθcosθCxzi(ζZ+ζZcos2θ2)CxTiΩcosθ(2Z+Z4),
(B115)
Dyx=i[Cyxe(ΩKUe0zcosθ)+CyzeKUe0xcosθ]iCyxi(ζZ+ζZsin2θ2)iCyziζZcosθsinθ2iCyTiΩsinθ(2Z+Z4),
(B116)
Dyy=K2(1iCyyeα×ωceτei)ζZCyye[ΩK(Ue0xsin   θ+Ue0zcos   θ)](εk)CyxiζZsin   θ2(εk)CyziζZcos   θ2(εk)CyTiΩ(Z+Z4),
(B117)
Dyz=i[CyxeKUe0zsinθ+Cyze(ΩKUe0xsinθ)]iCyxiζZsinθcosθ2iCyzi(ζZ+ζZcos2θ2)iCyTiΩcosθ(2Z+Z4),
(B118)
Dzx=K2sinθcosθCzxe(ΩKUe0zcosθ)CzzeKUe0xcosθCzxi(ζZ+ζZsin2θ2)CzziζZcosθsinθ2CzTiΩsinθ(2Z+Z4),
(B119)
Dzy=Czyeα×ωceτeiζZ+iCzye[ΩK(Ue0xsinθ+Ue0zcosθ)]+i(εk)CzxiζZsinθ2+i(εk)CzziζZcosθ2+i(εk)CzTiΩ(Z+Z4),
(B120)
Dzz=K2sin2θCzxeKUe0zsinθCzze(ΩKUe0xsinθ)CzxiζZ2sinθcosθCzzi(ζZ+ζZcos2θ2)CzTiΩcosθ(2Z+Z4),
(B121)

where

Cxxi=1+Cxxec¯ixx+Cxzec¯izx,
(B122)
Cxzi=Cxxec¯ixz+Cxzec¯izz,
(B123)
CxTi=Cxxec¯ixT+Cxzec¯izT,
(B124)
Cyxi=Cyxec¯ixx+Cyzec¯izx,
(B125)
Cyzi=Cyxec¯ixz+Cyzec¯izz,
(B126)
CyTi=Cyxec¯ixT+Cyzec¯izT,
(B127)
Czxi=Czxec¯ixx+Czzec¯izx,
(B128)
Czzi=1+Czxec¯ixz+Czzec¯izz,
(B129)
CzTi=Czxec¯ixT+Czzec¯izT.
(B130)

Since the current model has been established independently, benchmarking with the classical model is desirable. Here, we used the well-known model by Davidson et al.17 For this benchmarking, we set both k|| and ue0z to be zero as in the classical model.

As shown in Fig. 8, the results from both collisional (blue line) and collisionless (red line) models do not agree with results from the classical model (black line). In particular, our models expect an almost linear dispersion relation, but ω increases slowly for small kρe in the classical model. Another interesting difference is that the peak growth rate occurs around kρe0.6 in our models, while it is around kρe1 in the classical model. This discrepancy is not due to the choice of our heat flux closures; there is not much difference between our two models, which shows the insensitivity of the dispersion to pe1. Moreover, the dispersion relation is independent of pe1|| when k||=0. We also have confirmed that this discrepancy is not due to the inclusion of the perturbed ion current density, which is ignored in the classical model.

FIG. 8.

Dispersion relation for the case of the ES-LHDW (Te=Ti=10 eV, ne=2×1013 cm−3, B0=180 Gauss, ue0x=50 km/s, singly ionized helium). (a) Dispersion relation for four cases. The blue and red lines indicate results from collisional and collisionless models, respectively. The green line denotes the case derived here with Poisson's equation and perturbed quantities in the collisionless model. The black lines indicate the results from the classical models.17 (b) Growth rate of the ES-LHDW for all cases.

FIG. 8.

Dispersion relation for the case of the ES-LHDW (Te=Ti=10 eV, ne=2×1013 cm−3, B0=180 Gauss, ue0x=50 km/s, singly ionized helium). (a) Dispersion relation for four cases. The blue and red lines indicate results from collisional and collisionless models, respectively. The green line denotes the case derived here with Poisson's equation and perturbed quantities in the collisionless model. The black lines indicate the results from the classical models.17 (b) Growth rate of the ES-LHDW for all cases.

Close modal

We note that the basic set of equations used in the classical model by Davidson et al.17 is different. The biggest difference is that Poisson's equation is used in the classical model, while we used Faraday's induction law. To understand the cause of this discrepancy, we have developed another model to calculate the dispersion relation. In this model, we follow the basic equations of the classical model, while using our results for the perturbed density and current density.

In our geometry, the first-order equations in Davidson et al.17 can be written as

E1yiμ0ωk2(1Δ2)J1y=0,
(C1)
E1x+ieε0k(ni1ne1)=0,
(C2)

where Δ=ω/(ck), which is from the displacement current. This contribution is ignored, since the phase velocity of LHDWs is much smaller than the speed of light (|Δ2|1). We have confirmed that the dispersion relation is insensitive to the inclusion of Δ2.

For J1y,ni1, and ne1, we use the results from our models. The perturbed ion density is given by24 

ni1=in0emik2vti2Z(kE1xiεE1y).
(C3)

For the perturbed electron density, we will use one from the collisionless model for simplicity, as there is not much difference between two models. We also assume that Te0=Ti0. With k||=0 and ue0z=0,ne1 can be expressed as7 

ne1=kn0(ωkue0x)B0[iCxnE1x+Cyn(1kue0xω)E1y],
(C4)

where

Cxn=(αe+εk)[1αe2+12αe(εkvte2ωce2+kue0xωce)+12(k2vte2ωce2+εue0xωce)]1,
(C5)
Cyn=(1+εkαe)[1αe2+12αe(εkvte2ωce2+kue0xωce)+12(k2vte2ωce2+εue0xωce)]1.
(C6)

The y component of the perturbed ion current is24 

J1yi=ie2n0miωζZE1y.
(C7)

The y component of the perturbed electron current is7 

J1ye=ien0B0[iCxuE1x+Cyu(1kue0xω)E1y],
(C8)

where

Cxu=(1+kue0x2αeωce)[1αe2+12αe(εkvte2ωce2+kue0xωce)+12(k2vte2ωce2+εue0xωce)]1,
(C9)
Cyu=(αek2vte22αeωce2)[1αe2+12αe(εkvte2ωce2+kue0xωce)+12(k2vte2ωce2+εue0xωce)]1.
(C10)

With Eqs. (C3), (C4), (C7), and (C8), Eqs. (C1) and (C2) can be written as

DyyE1y+DyxE1x=0,
(C11)
DxyE1y+DxxE1x=0,
(C12)

where

Dyy=1ζZK2(1Δ2)ΩKUe0xK2(1Δ2)Cyu,
(C13)
Dyx=iΩCxuK2(1Δ2),
(C14)
Dxy=idi22K2λDi2(εk)Ziωpi2Cynωci2Ω,
(C15)
Dxx=1di22K2λDi2Z+ωpi2Cxnωci2(ΩKUe0x),
(C16)

where λDi=ε0Ti0/e2n0 is the ion Debye Length. The dispersion relation can be obtained by setting DxxDyyDxyDyx=0.

The dispersion relation from this simplified model (green line) agrees with the classical model, as shown in Fig. 8(a). This means that the discrepancy is due to the use of Poisson's equation, where the Faraday induction term is ignored. With the parameters for the ES-LHDW, βe is about 0.25, which means that perturbed magnetic field due to the perturbed plasma current may not be negligible. This argument is supported by observations in laboratory and space,7,10 where magnetic field fluctuations exist when there are strong electric field fluctuations associated with ES-LHDW.

It is interesting to see that the growth rate from the simplified model is considerably lower than that from the classical model, as shown in Fig. 8(b). This difference is likely related to the lack of a rigorous modeling of the heat flux in this simplified model. Although the magnitude is different, both models show that the peak growth rate is around kρe1.

This comparison shows that the use of electron fluid equations is acceptable for dynamics of LHDWs. It should be also noted that only our models include full electromagnetic effects, since the induction term is included. These effects are important when β is not negligible.

1.
S. D.
Bale
,
F. S.
Mozer
, and
T.
Phan
, “
Observation of lower hybrid drift instability in the diffusion region at a reconnecting magnetopause
,”
Geophys. Res. Lett.
29
,
33-1
33-4
, (
2002
).
2.
C.
Norgren
,
A.
Vaivads
,
Y. V.
Khotyaintsev
, and
M.
André
, “
Lower hybrid drift waves: Space observations
,”
Phys. Rev. Lett.
109
,
055001
(
2012
).
3.
D. B.
Graham
,
Y. V.
Khotyaintsev
,
C.
Norgren
,
A.
Vaivads
,
M.
André
,
S.
Toledo-Redondo
,
P.-A.
Lindqvist
,
G. T.
Marklund
,
R. E.
Ergun
,
W. R.
Paterson
,
D. J.
Gershman
,
B. L.
Giles
,
C. J.
Pollock
,
J. C.
Dorelli
,
L. A.
Avanov
,
B.
Lavraud
,
Y.
Saito
,
W.
Magnes
,
C. T.
Russell
,
R. J.
Strangeway
,
R. B.
Torbert
, and
J. L.
Burch
, “
Lower hybrid waves in the ion diffusion and magnetospheric inflow regions
,”
J. Geophys. Res.
122
,
517
533
, (
2017
).
4.
D. B.
Graham
,
Y. V.
Khotyaintsev
,
C.
Norgren
,
A.
Vaivads
,
M.
André
,
J. F.
Drake
,
J.
Egedal
,
M.
Zhou
,
O. L.
Contel
,
J. M.
Webster
,
B.
Lavraud
,
I.
Kacem
,
V.
Génot
,
C.
Jacquey
,
A. C.
Rager
,
D. J.
Gershman
,
J. L.
Burch
, and
R. E.
Ergun
, “
Universality of lower hybrid waves at Earth's magnetopause
,”
J. Geophys. Res.: Space Phys.
124
,
8727
8760
, (
2019
).
5.
L.-J.
Chen
,
S.
Wang
,
M.
Hesse
,
R. E.
Ergun
,
T.
Moore
,
B.
Giles
,
N.
Bessho
,
C.
Russell
,
J.
Burch
,
R. B.
Torbert
,
K. J.
Genestreti
,
W.
Paterson
,
C.
Pollock
,
B.
Lavraud
,
O. L.
Contel
,
R.
Strangeway
,
Y. V.
Khotyaintsev
, and
P.-A.
Lindqvist
, “
Electron diffusion regions in magnetotail reconnection under varying guide fields
,”
Geophys. Res. Lett.
46
,
6230
6238
, (
2019
).
6.
L.-J.
Chen
,
S.
Wang
,
O. L.
Contel
,
A.
Rager
,
M.
Hesse
,
J.
Drake
,
J.
Dorelli
,
J.
Ng
,
N.
Bessho
,
D.
Graham
,
L. B.
Wilson
,
T.
Moore
,
B.
Giles
,
W.
Paterson
,
B.
Lavraud
,
K.
Genestreti
,
R.
Nakamura
,
Y. V.
Khotyaintsev
,
R. E.
Ergun
,
R. B.
Torbert
,
J.
Burch
,
C.
Pollock
,
C. T.
Russell
,
P.-A.
Lindqvist
, and
L.
Avanov
, “
Lower-hybrid drift waves driving electron nongyrotropic heating and vortical flows in a magnetic reconnection layer
,”
Phys. Rev. Lett.
125
,
025103
(
2020
).
7.
J.
Yoo
,
J.-Y.
Ji
,
M. V.
Ambat
,
S.
Wang
,
H.
Ji
,
J.
Lo
,
B.
Li
,
Y.
Ren
,
J.
Jara-Almonte
,
L.-J.
Chen
,
W.
Fox
,
M.
Yamada
,
A.
Alt
, and
A.
Goodman
, “
Lower hybrid drift waves during guide field reconnection
,”
Geophys. Res. Lett.
47
,
e2020GL087192
, (
2020
).
8.
T. A.
Carter
,
H.
Ji
,
F.
Trintchouk
,
M.
Yamada
, and
R. M.
Kulsrud
, “
Measurement of lower-hybrid drift turbulence in a reconnecting current sheet
,”
Phys. Rev. Lett.
88
,
015001
(
2001
).
9.
H.
Ji
,
S.
Terry
,
M.
Yamada
,
R.
Kulsrud
,
A.
Kuritsyn
, and
Y.
Ren
, “
Electromagnetic fluctuations during fast reconnection in a laboratory plasma
,”
Phys. Rev. Lett.
92
,
115001
(
2004
).
10.
J.
Yoo
,
M.
Yamada
,
H.
Ji
,
J.
Jara-Almonte
,
C. E.
Myers
, and
L.-J.
Chen
, “
Laboratory study of magnetic reconnection with a density asymmetry across the current sheet
,”
Phys. Rev. Lett.
113
,
095002
(
2014
).
11.
P. H.
Yoon
and
A. T. Y.
Lui
, “
Lower-hybrid-drift and modified-two-stream instabilities in current sheet equilibrium
,”
J. Geophys. Res.
109
,
A02210
, (
2004
).
12.
R.
Kulsrud
,
Plasma Physics for Astrophysics
(
Princeton University Press
,
Princeton
,
2005
).
13.
I.
Silin
,
J.
Büchner
, and
A.
Vaivads
, “
Anomalous resistivity due to nonlinear lower-hybrid drift waves
,”
Phys. Plasmas
12
,
062902
(
2005
).
14.
P. H.
Yoon
and
A. T. Y.
Lui
, “
Anomalous resistivity by fluctuation in the lower-hybrid frequency range
,”
J. Geophys. Res.: Space Phys.
112
,
A06207
, (
2007
).
15.
V.
Roytershteyn
,
W.
Daughton
,
H.
Karimabadi
, and
F. S.
Mozer
, “
Influence of the lower-hybrid drift instability on magnetic reconnection in asymmetric configurations
,”
Phys. Rev. Lett.
108
,
185001
(
2012
).
16.
V.
Roytershteyn
,
S.
Dorfman
,
W.
Daughton
,
H.
Ji
,
M.
Yamada
, and
H.
Karimabadi
, “
Electromagnetic instability of thin reconnection layers: Comparison of three-dimensional simulations with MRX observations
,”
Phys. Plasmas
20
,
061212
(
2013
).
17.
R.
Davidson
,
N.
Gladd
,
C.
Wu
, and
J.
Huba
, “
Effects of finite plasma beta on the lower-hybrid drift instability
,”
Phys. Fluids
20
,
301
(
1977
).
18.
L.
Spitzer
,
Physics of Fully Ionized Gases
, 2nd ed. (
Interscience Publishers
,
New York
,
1962
).
19.
A.
Kuritsyn
,
M.
Yamada
,
S.
Gerhardt
,
H.
Ji
,
R.
Kulsrud
, and
Y.
Ren
, “
Measurements of the parallel and transverse Spitzer resistivities during collisional magnetic reconnection
,”
Phys. Plasmas
13
,
055703
(
2006
).
20.
J.
Yoo
,
M.
Yamada
,
H.
Ji
,
J.
Jara-Almonte
, and
C. E.
Myers
, “
Bulk ion acceleration and particle heating during magnetic reconnection in a laboratory plasma
,”
Phys. Plasmas
21
,
055706
(
2014
).
21.
W.
Daughton
, “
Electromagnetic properties of the lower-hybrid drift instability in a thin current sheet
,”
Phys. Plasmas
10
,
3103
3119
(
2003
).
22.
W.
Daughton
,
G.
Lapenta
, and
P.
Ricci
, “
Nonlinear evolution of the lower-hybrid drift instability in a current sheet
,”
Phys. Rev. Lett.
93
,
105004
(
2004
).
23.
X. Y.
Wang
,
Y.
Lin
,
L.
Chen
, and
Z.
Lin
, “
A particle simulation of current sheet instabilities under finite guide field
,”
Phys. Plasmas
15
,
072103
(
2008
).
24.
H.
Ji
,
R.
Kulsrud
,
W.
Fox
, and
M.
Yamada
, “
An obliquely propagating electromagnetic drift instability in the lower hybrid frequency range
,”
J. Geophys. Res.
110
,
A08212
, (
2005
).
25.
J.-Y.
Ji
and
I.
Joseph
, “
Electron parallel closures for the 3 + 1 fluid model
,”
Phys. Plasmas
25
,
032117
(
2018
).
26.
J.-Y.
Ji
and
E. D.
Held
, “
Closure and transport theory for high-collisionality electron–ion plasmas
,”
Phys. Plasmas
20
,
042114
(
2013
).
27.
N.
Krall
and
P.
Liewer
, “
Low-frequency instabilities in magnetic pulses
,”
Phys. Rev. A
4
,
2094
(
1971
).
28.
N.
Gladd
, “
The lower hybrid drift instability and the modified two stream instability in high density theta pinch environments
,”
Plasma Phys.
18
,
27
(
1976
).
29.
G. W.
Hammett
and
F. W.
Perkins
, “
Fluid moment models for Landau damping with application to the ion-temperature-gradient instability
,”
Phys. Rev. Lett.
64
,
3019
3022
(
1990
).
30.
S. I.
Braginskii
, “
Transport processes in a plasma
,” in
Reviews of Plasma Physics
, edited by
M. A.
Leontovich
(
Consultants Bureau
,
New York
,
1965
), Vol.
1
, pp.
205
311
.
31.
F.
Trintchouk
,
M.
Yamada
,
H.
Ji
,
R. M.
Kulsrud
, and
T. A.
Carter
, “
Measurement of the transverse Spitzer resistivity during collisional magnetic reconnection
,”
Phys. Plasmas
10
,
319
322
(
2003
).
32.
W.
Fox
,
F.
Sciortino
,
A. V.
Stechow
,
J.
Jara-Almonte
,
J.
Yoo
,
H.
Ji
, and
M.
Yamada
, “
Experimental verification of the role of electron pressure in fast magnetic reconnection with a guide field
,”
Phys. Rev. Lett.
118
,
125002
(
2017
).
33.
K.
Tummel
,
L.
Chen
,
Z.
Wang
,
X. Y.
Wang
, and
Y.
Lin
, “
Gyrokinetic theory of electrostatic lower-hybrid drift instabilities in a current sheet with guide field
,”
Phys. Plasmas
21
,
052104
(
2014
).
34.
J.
Yoo
,
B.
Na
,
J.
Jara-Almonte
,
M.
Yamada
,
H.
Ji
,
V.
Roytershteyn
,
M. R.
Argall
,
W.
Fox
, and
L.-J.
Chen
, “
Electron heating and energy inventory during asymmetric reconnection in a laboratory plasma
,”
J. Geophys. Res.
122
,
9264
9281
, (
2017
).
35.
Y.
Hu
,
J.
Yoo
,
H.
Ji
,
A.
Goodman
, and
X.
Wu
, “
Probe measurements of electric field and electron density fluctuations at megahertz frequencies using in-shaft miniature circuits
,”
Rev. Sci. Instrum.
92
,
033534
(
2021
).
36.
F. S.
Mozer
,
M.
Wilber
, and
J. F.
Drake
, “
Wave associated anomalous drag during magnetic field reconnection
,”
Phys. Plasmas
18
,
102902
(
2011
).
37.
J.-Y.
Ji
and
E. D.
Held
, “
Landau collision operators and general moment equations for an electron–ion plasma
,”
Phys. Plasmas
15
,
102101
(
2008
).
38.
J.
Yoo
,
Y.
Hu
,
J. Y.
Ji
,
H.
Ji
,
M.
Yamada
,
A.
Goodman
,
K.
Bergstedt
, and
A.
Alt
(
2021
). “Effects of coulomb collisions on lower hybrid drift waves inside a laboratory reconnection current sheet,”
DataSpace of Princeton University
. https://dataspace.princeton.edu/handle/88435/dsp01x920g025r