Large-amplitude Alfvén waves are subject to parametric decays, which can have important consequences in space, astrophysical, and fusion plasmas. Although the Alfvén wave parametric decay instability (PDI) was predicted decades ago, observational evidence is limited, stimulating considerable interest in laboratory demonstration of the instability and associated numerical modeling. Here, we report an important step toward direct hybrid simulation of the Alfvén wave PDI in a laboratory plasma, using antenna-like wave injection of a circularly polarized wave and realistic wave-plasma parameters. Considering collisionless damping, we identify the threshold Alfvén wave amplitudes and frequencies required for triggering the instability. These threshold behaviors are corroborated by simple theoretical analysis, considering the balance between PDI growth and ion Landau damping and PDI development in a bounded plasma. Other effects not included in the present model such as finite transverse wave scale and ion–neutral collisions are briefly discussed. These hybrid simulations promise to be a useful tool for investigating laboratory Alfvén wave dynamics and may provide guidance for future laboratory demonstration of the PDI process.

Predicted by Hannes Alfvén1 in 1942, the Alfvén wave (AW) is widely conceived to be the fundamental mode of a magnetized plasma.2–4 Because of little geometrical attenuation, shear AWs are prevalent in space and astrophysical plasmas as a carrier of magnetic and flow energy over a large range of scales.5 In laboratory plasmas, AWs also exist in tokamaks and linear devices, and their interaction with energetic particles is crucial to the performance of fusion reactions.6 Large-amplitude AWs play an important role in several nonlinear processes, such as energy cascade7 and particle acceleration,8 which are essential in plasmas mentioned above. Parametric instabilities are an important class of such nonlinear interactions, arising from parametric excitation of collective plasma modes. These instabilities could potentially contribute to coronal heating,5,9 the observed spectrum and cross-helicity of solar wind turbulence,10–12 and damping of fast magnetosonic waves in fusion plasmas.13,14

Three major types of parametric instabilities driven by a parallel (perpendicular wave vector k = 0 ) AW have been found theoretically:15–22 modulational, beat, and decay instabilities. The modulational instability drives forward upper and lower Alfvénic sidebands as well as a non-resonant acoustic mode at the sideband separation frequency. To allow the forward AWs to interact, the pump wave must be dispersive, i.e., requiring the finite frequency effect through inclusion of the Hall term.22 The beat instability drives a forward upper Alfvénic sideband and a backward lower Alfvénic sideband and is generally negligible in the low-beta plasmas to be studied here. By contrast, the parametric decay instability (PDI), involving the decay of a forward pump AW into a backward daughter AW and a forward ion acoustic wave, is best known for its prominence in low-beta plasmas.

The PDI is of special interest to the space plasma community. First, the backward AW generated during the PDI can interact with the forward AW pump, which is an important ingredient for developing magnetohydrodynamics (MHD) turbulence.23 Second, the PDI-produced ion acoustic wave can strongly heat ions through wave damping.24 Recently, Fu et al.25 showed that the PDI and associated kinetic heating can be robust even in a turbulent environment, which is often the case in space plasmas. The PDI growth rate is largest at low-beta, γ g β 1 / 4 , according to a linear analysis;15 here, β = P / ( B 0 2 / 2 μ 0 ) with P , B 0 , μ 0 being the total plasma pressure, background magnetic field, and vacuum permeability, respectively. A vast region of the solar surface extending from the chromosphere to corona consists of low-beta plasmas.26 Several studies have suggested that the PDI plays an important role in a stratified chromosphere,27 in an expanding solar wind,28,29 and even in the solar wind near 1 AU,30,31 contributing to turbulence development and plasma heating.

The PDI of a shear AW was first predicted by Sagdeev and Galeev15 in 1969 and analyzed in the low-amplitude low-beta limit for a linearly polarized AW. Derby17 and Goldstein18 later independently extended the analyses to finite wave amplitudes for circular polarization in the single-fluid MHD framework. The dispersive effects arising from ion cyclotron resonance were further considered with a two-fluid framework.19,22 Despite the firm theoretical bases, observational evidence of PDI in space has been scarce. The satellite observations in the upstream of the bow shock of Earth's magnetosphere may have found a number of cases with possible AW decay signatures.32–34 A statistical study31 has shown correlations of density or magnetic fluctuations with PDI over a broad parameter space. However, unambiguous observation of individual PDI events in space requires overcoming several challenges,33 such as the turbulent environment, limited sampling locations, broad pump bandwidths, and various kinetic effects.32 Therefore, there is a need for laboratory experiments to validate theoretical predictions. Particularly, the active control of laboratory wave and plasma conditions may help to elucidate the underlying physical processes in a manner not possible with space measurements.

The long wavelengths of AWs make it difficult to fit the waves in a laboratory device while maintaining a low enough plasma density to allow for the use of physical probes throughout the plasma region. Recently, with the Large Plasma Device (LAPD) at University of California, Los Angeles (UCLA), several controlled AW experiments have been conducted.4,35–42 Particularly, Dorfman and Carter41 excited an ion acoustic mode through beating of two counter-propagating AWs. While this setup is similar to seeding the PDI with an artificial daughter AW, the instability growth was not clearly identified. More recently, Dorfman and Carter42 also recorded a perpendicular modulational instability of kinetic AWs; when a single finite-frequency, finite- k , AW was launched above a certain wave amplitude threshold, three daughter waves were detected: two sideband AWs co-propagating with the pump wave and a low frequency non-resonant mode at the sideband separation frequency. However, the theoretical growth rate for zero- k modulational instability was too small to explain the observation. More critically, the PDI was missing despite having a much larger theoretical growth rate than the zero- k modulational instability under the conditions investigated. So far, no quantitative evidence of PDI has been found in laboratory experiments. In planning future experiments aimed at demonstration of this important instability, it should be of considerable interest to find optimal parameters with first-principle numerical simulations.

In this paper, we report a crucial step in the development of a hybrid simulation framework to directly model laboratory AW dynamics. In particular, we simulate the PDI of nonlinear AWs with a physical setup and wave-plasma parameters that closely resemble LAPD conditions. We consider LAPD-like injection of a planar circularly polarized AW in a bounded (along the background magnetic field direction) plasma with ion kinetics retained (e.g., collisionless Landau damping). These simulations allow us to gain insight into the threshold AW amplitudes and frequencies needed for triggering PDI in the collisionless bounded plasma. The simulation results can be interpreted with simple theoretical estimates, where the key physics are the balance between the PDI growth and ion Landau damping and the development of PDI in a bounded plasma; it is the first time these principles have been applied in the Alfvén wave PDI context. We also briefly discuss complicating effects such as ion–neutral collisions and finite AW source sizes not included in the current simulations. It is hoped that the present work will lay the basis for further simulation developments that would enable quantitative predictions and provide guidance for future demonstration of PDI in LAPD.

We start with introducing our hybrid simulation model. The PDI has been extensively studied via MHD simulations,11,30 hybrid simulations,25,43–47 and even full particle-in-cell simulations.48 In the context of Earth's foreshock and solar wind, the PDI has also been invoked in simulations to explain, for example, the generation of density fluctuations,49 the origin of low-frequency Alfvénic spectrum,29 the effects of wind acceleration or expansion,28 and the evolution of magnetic-field-line switchbacks.50 However, the above-mentioned simulations have mostly considered a periodic system,30,47,48,50 limiting their applicability to laboratory plasmas where the interactions are finite in space-time and nonperiodic. The periodic boundary constraint was lifted in some MHD simulations,11 but they did not capture kinetic effects which are important in laboratory PDI. To address that, we have recently developed novel hybrid simulation capabilities aimed at directly modeling laboratory AW dynamics.

The new developments are based on the three-dimensional hybrid code H3D,51 which treats electrons as a massless fluid and ions as individual macroparticles. Therefore, the simulation captures ion kinetic effects and also saves computing power by ignoring electron dynamics. The code was modified to simulate turbulent plasmas with periodic boundary conditions.24,25 Recently, we have relaxed the periodic boundary constraint and studied the PDI in a large system ( 100 λ 0 with λ0 being the reference AW wavelength) with absorption boundaries for the AWs.52 Compared to usual periodic interaction, several new PDI dynamics were found with the nonperiodic interaction, including reduced energy transfer, and localized density fluctuations and ion heating, as well as the formation of a stable residual wave packet that is not affected by PDI. Although the parameters (e.g., δ B / B 0 0.1 , β 0.01 ) used were still away from typical LAPD conditions (e.g., δ B / B 0 10 3 , β 10 4 10 3 ), the study uncovered several limitations of usual periodic boundary condition in addressing AW dynamics in a laboratory setting.

In the present work, we implement wave injection at prescribed locations, similar to the wave injection from an antenna in LAPD experiments, and investigate the resultant PDI dynamics. The basic simulation setup is illustrated in Fig. 1. A typical run involves one grid cell for the x direction and four cells for the y direction. The z direction (i.e., the AW propagation direction) has a size of 9 λ 0 , corresponding to 90 d i , where d i = c / ω p i is the ion inertial length, c is the light speed in a vacuum, and ωpi is the ion plasma frequency. The cell sizes are Δ x = 1 d i , Δ y = 0.1 d i , Δ z = 0.5 d i , and the time step size is Δ t = 0.01 Ω c i 1 , where Ωci is the ion cyclotron frequency and satisfies ω p i / Ω c i = c / v A = 300 with vA being the Alfvén speed. The ions are sampled by 1000 macroparticles per cell. The ion-to-electron temperature ratio, T i / T e , can be varied to approximate different LAPD discharge conditions. A finite resistivity of η = 2 × 10 4 B 0 / e n 0 v A is used to model the spatial wave damping as found in LAPD.53 The wave injection can be prescribed at any locations within the domain and typically involves an up-ramp of 5 T 0 in the temporal profile followed by a long plateau (hundreds of T0), where T 0 = λ 0 / v A .

FIG. 1.

Hybrid simulation of Alfvén wave parametric decay instability using LAPD-like wave injection and wave-plasma parameters. Snapshots of (a) 2D wave pattern and y-averaged (b) wave magnetic fields, (c) velocity fields, and (d) density fluctuation and parallel ion temperature at t = 200 T 0 . (e) and (f) Spatiotemporal evolution of the Bx and density fluctuation, respectively. Notice the AW looks immobile because Bx data are dumped about every wave period. (g) Wave signatures probed at z = 4.58 λ 0 and (h) corresponding Fourier spectra, where the black and red curves refer to forward and backward going waves, respectively.

FIG. 1.

Hybrid simulation of Alfvén wave parametric decay instability using LAPD-like wave injection and wave-plasma parameters. Snapshots of (a) 2D wave pattern and y-averaged (b) wave magnetic fields, (c) velocity fields, and (d) density fluctuation and parallel ion temperature at t = 200 T 0 . (e) and (f) Spatiotemporal evolution of the Bx and density fluctuation, respectively. Notice the AW looks immobile because Bx data are dumped about every wave period. (g) Wave signatures probed at z = 4.58 λ 0 and (h) corresponding Fourier spectra, where the black and red curves refer to forward and backward going waves, respectively.

Close modal

Our simulations essentially model the injection of a plane wave with k = 0 . In this quasi-1D setup, we do not consider the finite transverse scale and large k as found in actual experiments.53,54 Moreover, the ion–neutral collisions present in a partially ionized LAPD plasma41 are not modeled in the current code. Despite the simplifications, our simulations and associated analyses should be of practical interest as we adopt realistic wave and plasma parameters (e.g., wave amplitude, polarization, and plasma beta, temperature) and geometry (e.g., longitudinal plasma size, wave injection, and boundary conditions). Implementation of these features is an important step toward fully modeling LAPD experiments. The wave absorption at boundaries is achieved using field masks as reported in our previous work.52 Notice that the periodic boundary is still used for ion particles. The resulting nonphysical particle recirculation in the simulation domain is delayed by the field mask region, as illustrated by the shaded areas in Fig. 1(b). For typical runs, the mask area is as large as the central unmasked region, such that the particle recirculation effect on wave-plasma interactions in the central region is negligible.

Figure 1 presents a case where strong PDI is observed. A left-hand polarized AW of normalized frequency ω 0 / Ω c i = 0.63 is injected at z = 3.5 λ 0 (green dashed line) by prescribing its magnetic field perturbations as

δ B = [ δ B x sin ( ω 0 t ) x ̂ + δ B y cos ( ω 0 t ) y ̂ ] χ ( t ) ,
(1)

where δ B x = δ B y = δ B , χ ( t ) is the temporal profile. The background field B0 is along the + z direction. The velocity/electric field perturbations, δv and δE, are automatically generated by the field solver. The finite- ω 0 Ω c i , zero- k planar AW is verified to follow the dispersion35,55 ω 0 k = v A 1 ( ω 0 Ω c i ) 2 and has unequal magnetic/velocity perturbations with the ratio56, R b v δ B / B 0 δ v / v A = ω 0 k v A = 1 ( ω 0 Ω c i ) 2 . The wave amplitude of the present case, δ B / B 0 = 0.01 , is higher than normally achievable with LAPD, but may be realized through, for example, tuning the background magnetic field B0 or AW maser.36 The wave pattern is shown in Figs. 1(a)–1(c) at t = 200 T 0 when the PDI has sufficiently developed; the line plots are averaged over the y direction, on which the results have little dependence because of the plane wave injection. The wave specified by Eq. (1) has no preferential propagation direction, and the field solver generates two identical waves of the same polarization and amplitude that propagate in both ± z directions. These waves start to be damped when they reach the mask regions on both sides. As such, the effective plasma length that the forward wave travels through is L = 2.5  λ 0 = 25 d i as marked in Fig. 1(b); this length is typical for LAPD experiments (e.g., Ref. 42).

One signature of PDI is the excitation of an ion acoustic wave (e.g., displayed as density fluctuations) and associated ion heating, as seen in Fig. 1(d). To better illustrate the full dynamics, Fig. 1(f) shows the space-time evolution of δ ρ / ρ 0 , where the density fluctuations emerge near the wave injection point in just tens T0 and then gradually propagate outward. The density modulation deepens with time, an indication of instability growth. A deep narrow density trough is gradually formed at z = 3.5 λ 0 due to continuous wave injection and localized plasma heating [e.g., the orange curve in Fig. 1(d)]. The density fluctuations propagates forward in the region z > 3.5 λ 0 , consistent with the acoustic wave direction of PDI driven by a forward pump wave. At later times ( t 300 T 0 ), the acoustic wave even forms nonlinear structures propagating at a reduced speed, due to ion trapping (not shown). In PDI, the large density fluctuations further beat with the pump wave, generating a backward daughter AW; the latter overlaps with the pump wave, resulting in the rippling modulation as found in Fig. 1(e).

To gain more insight into the daughter AW, we probe the AW signatures vs time at a fixed location in front of the injection point and further separate the forward/backward waves using the modified Elsasser variables,

Z ± = 1 2 v A ( δ v R b v δ B μ 0 ρ ) ,
(2)

where Z+ and Z refer to the forward wave and the backward wave, respectively. For the present case, R b v 0.78 , and the difference between the magnetic and velocity field components can be clearly seen by comparing Figs. 1(b) and 1(c). The separated waves and corresponding Fourier spectra are presented in Figs. 1(g) and 1(h). Strong bursts of backward daughter AW (red color) due to the PDI are seen at later times ( t 100 T 0 ). Its spectrum shows a prominent peak around 0.96 ω 0 , consistent with the anticipated central frequency of a lower sideband at ω / ω 0 = 1 2 β 0.96 . A broad bandwidth also occurs; this is partly due to the finite bandwidth excitation of PDI at high wave amplitudes52 and also likely enhanced by a spread in β due to nonlinear modification of the plasma density and temperature.

The above example shows, with a simplified plane wave setup, what the PDI may look like if it can indeed be excited in LAPD-like plasmas. Next, using the same setup, we extend the study to a large parameter space. We group the simulations by two parameters, β and δ B / B 0 , which are most relevant in characterizing the PDI. The results are summarized in Fig. 2(a), where typical LAPD conditions correspond to the bottom-left corner (outlined by the dashed square), i.e., β 10 4 10 3 and δ B / B 0 10 3 10 2 . Three ion-to-electron temperature ratios, T i / T e = 0.1 , 0.25 , 1 , are considered and marked in the plot as different colors. Each simulation is represented by a dot in the figure, where open dots refer to simulations with no PDI and filled dots indicate those with PDI. The criteria for PDI onset are based on both density and AW signatures, i.e., whether the peak amplitude of the lower Alfvénic sideband is greater than 5% of the main peak at the fundamental frequency, and the averaged density fluctuation δ ρ / ρ 0 is greater than 5%. We have also checked that, for those open-dot cases, no PDI is found even if the AW is injected over 1000 T 0 .

FIG. 2.

Threshold Alfvén wave amplitude for triggering the parametric decay instability. (a) The parameter space spanned by wave amplitude and plasma beta. Typical LAPD conditions are indicated by the dashed square area in the bottom-left corner. Each dot represents a hybrid simulation. The open dots indicate no PDI found, and the filled dots indicate the opposite. Different colors refer to different ion-to-electron temperature ratios. The solid lines are theoretical predictions. (b1)–(b3) show the ion phase-space distribution z v z , velocity distribution, and spatiotemporal evolution of ion parallel temperature (normalized to initial temperature), respectively, for the case marked as orange cross in (a). The velocity distributions show two instants at t = 0 , 200 T 0 , respectively, for ions within z = [ 4 , 4.5 ] λ 0 only. (c1)–(c3) correspond to the green-cross case marked in (a).

FIG. 2.

Threshold Alfvén wave amplitude for triggering the parametric decay instability. (a) The parameter space spanned by wave amplitude and plasma beta. Typical LAPD conditions are indicated by the dashed square area in the bottom-left corner. Each dot represents a hybrid simulation. The open dots indicate no PDI found, and the filled dots indicate the opposite. Different colors refer to different ion-to-electron temperature ratios. The solid lines are theoretical predictions. (b1)–(b3) show the ion phase-space distribution z v z , velocity distribution, and spatiotemporal evolution of ion parallel temperature (normalized to initial temperature), respectively, for the case marked as orange cross in (a). The velocity distributions show two instants at t = 0 , 200 T 0 , respectively, for ions within z = [ 4 , 4.5 ] λ 0 only. (c1)–(c3) correspond to the green-cross case marked in (a).

Close modal

It is seen from Fig. 2(a) that the AW amplitude must exceed a threshold to trigger PDI in the system and the threshold amplitudes increase with β and T i / T e . In general, parametric instabilities occur when the pump amplitude exceeds a threshold determined by damping effects, and both the AW waves and acoustic waves may experience damping. In LAPD experiments, AW spatial damping is due to electron–ion collisions and/or AW Landau damping; both may be varied by scanning parameters such as the electron temperature.57 In the present simulations, the spatial damping of pump AWs due to electron-ion collisions is characterized by the finite η. While the resultant AW spatial damping is found to cause slightly weaker PDI development (compared to cases without spatial damping, i.e., η 0 ), its effect is negligible. Meanwhile, the acoustic modes in LAPD may be damped by ion–neutral collisions or collisionless Landau damping; only the latter is modeled in the current code. Therefore, the threshold AW amplitudes in the present investigation should be determined by the balance between the PDI growth and Landau damping.

The PDI growth rate may be approximated in the low-beta, low-amplitude limit15 as

γ g / ω 0 0.5 ( δ B / B 0 ) β 1 / 4 .
(3)

It should be emphasized that this normalized growth rate was obtained in the low-frequency single-fluid limit ( ω 0 / Ω c i 0 ), but it is found to approximate our finite-frequency ( ω 0 / Ω c i 0.6 ) cases well; we have checked against two-fluid growth rates22 and found that noticeable differences start to appear only when δ B / B 0 > 10 2 for typical LAPD beta values. On the other hand, we take the Landau damping rate, γ d / ω T i / T e , for the Cauchy ion distribution,58 where ω is the acoustic wave frequency. Using the Cauchy distribution greatly simplifies the calculation of Landau damping and the Cauchy distribution reasonably approximates the actual Gaussian distribution in the ion velocity range of | v i / v t h , i | < 1.5 , where v t h , i = k B T i / M i is the ion thermal speed with Mi being the ion mass and kB the Boltzmann constant (though the Cauchy distribution has a heavier tail at larger vi). These simplifications are justified as we aim at providing a physical interpretation for the threshold-amplitude behavior, instead of an exact numerical prediction; the latter would require considering effects neglected by the simplifying assumptions made in Sec. II. In the low-beta regime of PDI, the acoustic wave frequency satisfies ω / ω 0 2 β . Cast in units of ω0, the Landau damping rate takes the form

γ d / ω 0 2 β T i / T e .
(4)

In order to trigger PDI, one needs to satisfy γ g > γ d , which is translated into a threshold for the AW amplitude in a collisionless plasma,

δ B / B 0 > 4 β 3 / 4 T i / T e .
(5)

This formula is appended as color lines in Fig. 2(a) for different T i / T e , which show reasonable agreement with the simulation results, specifically in the LAPD parameter regime. The agreement confirms that the threshold is determined by the competition between PDI growth and Landau damping in our simulations. Importantly, these curves cross the bottom-left corner, an indication that the excitation of PDI might be within reach at LAPD if effects neglected by our model (see discussion in Sec. V) do not significantly modify the balance between growth and damping.

A closer comparison between Eq. (5) and the simulation data in Fig. 2 shows a few prominent differences (notice the log-scale axes). First, the simulations show lower threshold amplitudes, especially when T i / T e is small. Physically, Te determines the phase speed of acoustic wave and Ti determines the ion thermal speeds. When T i / T e 1 , the phase speed is much larger than the thermal speed, and the Cauchy distribution used for Eq. (4) tends to overestimate the damping rate because of a heavier tail in the velocity distribution. Therefore, the theoretical curves tend to overestimate the required AW amplitude when T i / T e 1 . Second, the simulations also show threshold amplitudes below the theoretical curves in the higher-beta, higher-amplitude regime, i.e., the upper-right corner of Fig. 2(a). In this regime, despite that Eq. (3) tends to slightly overestimate the growth rate, the much stronger acoustic mode damping causes significant plasma heating which may greatly reduce the Landau damping rate. To see this clearly, we compare the plasma heating results of two example cases in Figs. 2(b) and 2(c), which are marked in Fig. 2(a) as orange/green crosses, respectively. The orange-cross case ( β = 5 × 10 4 , δ B / B 0 = 1 × 10 2 , T i / T e = 0.25 ) represents the low-amplitude regime and mostly has regular ion phase-space [Fig. 2(b1)] and parallel temperature distributions [Fig. 2(b3)] and negligible modification to the local ion distribution function [Fig. 2(b2)] (for particles within sub-wavelength scale z = [ 4 , 4.5 ] λ 0 ). Notice that nonlinear phase-space islands form at the late stage ( t = 200 T 0 ) for this case which sits above the threshold curve. For wave amplitudes around/below the threshold curve and/or earlier linear stage of PDI, the phase-space distribution is very linear and Eq. (4) is a good approximation to describe the Landau damping. In contrast, the green-cross case ( β = 3 × 10 3 , δ B / B 0 = 5 × 10 2 , T i / T e = 1 ) represents the high-amplitude regime and has significantly stronger heating [Fig. 2(c3)], which causes large distortions in the local phase-space [Fig. 2(c1)] and velocity distributions [Fig. 2(c2)].

FIG. 3.

Threshold Alfvén wave frequency for triggering the parametric decay instability. (a) and (b) Spatiotemporal evolutions of the density fluctuation for case ω 0 / Ω c i = 0.42 , 0.21 , respectively. The other parameters are the same as for Fig. 1. (c) The same as (b) but for an enlarged simulation domain, three times bigger in L. (d) Maximum of the Fourier spectral amplitude of Ez fluctuation for different ω 0 / Ω c i , where the vertical dashed line marks the threshold frequency predicted by Eq. (7). (e) Theoretical estimate of the threshold wave frequency over the full parameter space spanned by wave amplitude and plasma beta [i.e., Eq. (7)], where L = 25 d i and the appended lines represent contours at specific ω 0 / Ω c i .

FIG. 3.

Threshold Alfvén wave frequency for triggering the parametric decay instability. (a) and (b) Spatiotemporal evolutions of the density fluctuation for case ω 0 / Ω c i = 0.42 , 0.21 , respectively. The other parameters are the same as for Fig. 1. (c) The same as (b) but for an enlarged simulation domain, three times bigger in L. (d) Maximum of the Fourier spectral amplitude of Ez fluctuation for different ω 0 / Ω c i , where the vertical dashed line marks the threshold frequency predicted by Eq. (7). (e) Theoretical estimate of the threshold wave frequency over the full parameter space spanned by wave amplitude and plasma beta [i.e., Eq. (7)], where L = 25 d i and the appended lines represent contours at specific ω 0 / Ω c i .

Close modal

Aside from the requirement on AW amplitude, another important parameter for consideration in experiments is what wave frequencies one should use. This is particularly relevant to the LAPD plasma which is bounded and normally contains a few AW wavelengths.53 It was argued59 that the absolute instability develops in a bounded unstable region if the size of the system, L, satisfies

γ g L / | V 1 V 2 | 1 / 2 > π / 2 ,
(6)

where V1, V2 are the group velocity of the two daughter waves, respectively. Physically, this means the system must be large enough for the instability to emerge before the daughter waves convect away. To simplify the calculation, we neglect the Landau damping which effectively means the following analysis is applicable only for small T i / T e or more precisely, for the damping rate being small as compared to the growth rate. In terms of the PDI investigated here, one has V 1 = d ω 0 d k and V 2 = c s , where cs is the acoustic wave speed. Making use of the AW dispersion relation k d i = ω 0 / Ω c i 1 ( ω 0 / Ω c i ) 2 , one has d ω 0 d k = 1 ( ω 0 / Ω c i ) 2 1 + k 2 d i 2 v A = [ 1 ( ω 0 / Ω c i ) 2 ] 3 / 2 v A . The acoustic wave speed takes the form c s = γ e k B T e + γ i k B T i M i = 5 β / 6 v A , where γ e = γ i = 5 / 3 are the electron/ion polytropic indices and we have used β = β e +  β i = ( 1 + T e / T i ) β i and β i = 2 ( v t h , i / v A ) 2 . Notice that the form of cs is independent of the temperature ratio T i / T e for a fixed β. Substituting the expressions for V1, V2 into Eq. (6), the requirement on the system size is translated to a threshold for the wave frequency,

F ( ω 0 / Ω c i ) ω 0 / Ω c i [ 1 ( ω 0 / Ω c i ) 2 ] 3 / 4  ⪆  π ( 5 / 6 ) 1 / 4 L / d i β δ B / B 0 ,
(7)

where we have used the identity c / v A = ω p i / Ω c i and F ( ω 0 / Ω c i ) is a monotonic increasing function of ω 0 / Ω c i . Using ω 0 = 2 π v A 1 ( ω 0 / Ω c i ) 2 / λ (with λ = 2 π / k ), the frequency threshold can be recast as

L λ  ⪆  0.48 β δ B / B 0 [ 1 ( ω 0 / Ω c i ) 2 ] 1 / 4 ,
(8)

which demands sufficient number of wavelengths to be present in the plasma region.

To verify the threshold frequency requirement, we have performed a series of simulations with β = 5 × 10 4 , δ B / B 0 = 1 × 10 2 , T i / T e = 0.25 , the same as the case of Fig. 1, but having different ω 0 / Ω c i . The case of Fig. 1(f) corresponds to ω 0 / Ω c i 0.63 , and we also select two other cases of ω 0 / Ω c i 0.42 , 0.21 and present their space-time density fluctuations in Figs. 3(a) and 3(b), respectively. It indeed shows that, as the pump frequency drops, the density fluctuations become weakened and eventually disappear for the case ω 0 / Ω c i 0.21 . Note that, for the present wave and plasma parameters, the threshold frequency predicted from Eq. (7) is ω 0 / Ω c i > 0.26 , consistent with the above simulations. The agreement is better illustrated by Fig. 3(d), which plots the maximum strength of acoustic wave (in terms of longitudinal Ez field) for a series of ω 0 / Ω c i . A clear threshold behavior around ω 0 / Ω c i 0.3 is seen, below which the wave strength is at the noise level and above which a sharp increase in the wave strength is found.

For the case with lowest frequency ω 0 / Ω c i 0.21 , we find in simulation L / λ 0.83 , meaning the plasma length is shorter than one Alfvén wavelength. Equation (8) demands L / λ > 1.1 for the parameters used, in order to see PDI. To verify Eq. (8), we keep the same low frequency ω 0 / Ω c i 0.21 but make L three times bigger, i.e., L / λ 2.5 . As shown in Fig. 3(c), the density fluctuations indeed reappear in the larger system.

Using Eq. (7) and L = 25 d i , the dependence of threshold frequency on the broad parameter space spanned by β and δ B / B 0 is shown in Fig. 3(e). The contours show the dependencies for a few constant ω 0 / Ω c i . Generally, the onset of PDI demands higher threshold frequencies toward the lower-right corner of the parameter space and vice versa. Because the threshold frequency in Eq. (7) is inversely proportional to the wave amplitude, the threshold frequency increases quickly with β at small wave amplitudes. One should avoid approaching ω 0 / Ω c i = 1 (bottom-right corner) where strong ion cyclotron resonances for a left-hand wave will be invoked. This plot would be useful for guiding the choice of wave frequency in future experiments.

We have so far considered two critical parameters, AW amplitude δ B / B 0 and frequency ω 0 / Ω c i , and studied their dependencies on plasma beta, temperature, and size to excite PDI in a bounded plasma with LAPD-like wave injection. However, there are several other factors that need to be considered for optimizing PDI experiments.

First, in presenting the threshold δ B / B 0 vs β (i.e., Fig. 2), the background field B0 is involved in both δ B / B 0 and β, rendering B0 a new degree of freedom. How to choose B0 is of practical importance in LAPD experiments. To clarify this point, we introduce the parameter ξ which is defined as the ratio between PDI growth and Landau damping, i.e.,

ξ = γ g γ d = 1 4 δ B B 0 ( B 0 2 / 2 μ 0 ) 3 / 4 ( n 0 K T e + n 0 K T i ) 3 / 4 T e T i B 0 .
(9)

The B 0 dependence suggests that PDI favors higher-B0 regime, which corresponds to the lower-beta, lower-normalized-amplitude regime.

Second, the ion-to-electron temperature ratio has been shown as an important parameter [Fig. 2(a)]. Recently, an upgrade was made in LAPD including the installation of a new L a B 6 cathode, which increased the accessible temperature ranges for both electrons (1–15 eV) and ions (sub-eV to 10 eV) and may also allow more flexibility in the temperature ratios. More importantly, the high electron temperatures will allow us to achieve higher pump AW amplitudes by minimizing both electron-ion collisional damping and AW Landau damping.57 

Third, actual LAPD experiments usually involve a finite wave scale in the transverse direction, which is determined by the antenna size.53 The source size is typically much smaller than λ , but the wave remains collimated over the relatively small plasma dimension due to insignificant geometric attenuation of AWs,54 which may justify the use of the quasi-1D setup in our work. However, the finite k ρ s 0.2 (with ρs being the ion sound gyroradius) arising from the source size could be an important factor that may potentially complicate the PDI physics.42 A detailed theoretical understanding of the finite k effect on PDI has not yet been developed.

Finally, the ion–neutral collisions present in the LAPD may enhance the damping of the acoustic wave, raising the threshold curves in Fig. 2(a), i.e., requiring higher AW amplitudes. We are currently implementing 3D wave injection of finite source sizes and ion–neutral collisions in the H3D code. Studies of PDI with these additional features will be presented in a separate work.

It is also worth noting that the current study is limited to circularly polarized wave injection, which is an exact solution to MHD equations and routinely adopted by LAPD using the rotating magnetic field antenna.53 By launching a linear polarized wave, qualitatively the same threshold behaviors for triggering PDI are also found. Quantitatively, the interaction is complicated by the group velocity dispersion between the left- and right-hand waves, of which the linear wave is composed. This dispersion effect is especially significant for the high wave frequencies ω 0 / Ω c i studied here. A systematic study of linear polarized injection is left for future.

In summary, we have presented novel hybrid simulation capabilities to model continuous wave injection in a bounded plasma with absorption wave boundaries and realistic wave-plasma parameters. Such simulations could be a useful tool to investigate a variety of LAPD AW experiments such as beat wave excitation41 and turbulence generation.40 Here, we have focused on the PDI of nonlinear AWs, a fundamental process in a magnetized plasma. Using parameters that resemble LAPD conditions, our simulations and accompanied theoretical analyses have provided important insight into the requirements for exciting PDI in a LAPD-like plasma, including the threshold values for AW amplitude and frequency. Other factors not included in the present simplified model (plane wave injection, collisionless plasma) are also briefly discussed. The current work is a critical step toward direct hybrid simulation of LAPD plasmas. With further development of the missing features, the simulation may provide quantitative guidance for future laboratory demonstration of the fundamental PDI process. The understanding acquired for relevant physical processes could have significant implications in space and astrophysical plasmas.

This work was supported by a National Science Foundation and Department of Energy Partnership in Basic Plasma Science and Engineering program under the Grant No. DE-SC0021237. X.F. was also supported by the Los Alamos National Laboratory/Laboratory Directed Research and Development program and DOE/Office of Fusion Energy Sciences. S.D. was also supported by the National Aeronautics and Space Administration (NASA) Grant No. 80NSSC18K1235. F.L. and X.F. acknowledge the Texas Advanced Computing Center (TACC) at The University of Texas at Austin and the National Energy Research Scientific Computing Center (NERSC) for providing HPC and visualization resources. NERSC is supported by the Office of Science of the U.S. Department of Energy under Contract No. DE-AC02–05CH11231.

The authors have no conflicts to disclose.

Feiyu Li: Investigation (lead); Writing – original draft (lead). Xiangrong Fu: Investigation (supporting); Writing – review & editing (supporting). Seth Dorfman: Investigation (supporting); Writing – review & editing (supporting).

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
H.
Alfvén
,
Nature
150
,
405
(
1942
).
2.
J. M.
Wilcox
,
F. I.
Boley
, and
A. W.
De Silva
,
Phys. Fluids
3
,
15
(
1960
).
3.
A.
Hasegawa
and
C.
Uberoi
, “
Alfvén wave: DOE critical review series
,” DOE/TIC-11197 (Technical Information Center, U.S. Department of Energy, Springfield,
1982
).
4.
W.
Gekelman
,
J. Geophys. Res.
104
,
14417
, (
1999
).
5.
B.
De Pontieu
,
S.
McIntosh
,
M.
Carlsson
,
V.
Hansteen
,
T.
Tarbell
,
C.
Schrijver
,
R.
Shine
,
S.
Tsuneta
,
Y.
Katsukawa
,
K.
Ichimoto
 et al,
Science
318
,
1574
(
2007
).
6.
L.
Chen
and
F.
Zonca
,
Rev. Mod. Phys.
88
,
015008
(
2016
).
7.
R. H.
Kraichnan
,
Phys. Fluids
8
,
1385
(
1965
).
8.
C.
Kletzing
,
J. Geophys. Res.
99
,
11095
, (
1994
).
9.
M.
Velli
and
F.
Pruneti
,
Plasma Phys. Controlled Fusion
39
,
B317
(
1997
).
10.
B.
Inhester
,
J. Geophys. Res.
95
,
10525
, (
1990
).
11.
L.
Del Zanna
,
M.
Velli
, and
P.
Londrillo
,
Astron. Astrophys.
367
,
705
(
2001
).
12.
P. H.
Yoon
and
T.-M.
Fang
,
Plasma Phys. Controlled Fusion
50
,
085007
(
2008
).
13.
J.
Lee
,
W.
Peebles
,
E.
Jaeger
,
E.
Doyle
,
N.
Luhmann
, Jr.
,
C.
Petty
,
R.
Pinsker
,
R.
Prater
, and
T.
Rhodes
,
Phys. Rev. Lett.
80
,
2330
(
1998
).
14.
T.
Oosako
,
Y.
Takase
,
A.
Ejiri
,
Y.
Nagashima
,
Y.
Adachi
,
H.
Kasahara
,
T.
Yamada
,
O.
Watanabe
,
H.
Tojo
,
S.
Kainaga
 et al,
Nucl. Fusion
49
,
065020
(
2009
).
15.
R. Z.
Sagdeev
and
A. A.
Galeev
,
Nonlinear plasma theory
(
Benjamin
,
New York
,
1969
).
16.
A.
Hasegawa
and
L.
Chen
,
Phys. Rev. Lett.
36
,
1362
(
1976
).
17.
N.
Derby
, Jr.
,
Astrophys. J.
224
,
1013
(
1978
).
18.
M. L.
Goldstein
,
Astrophys. J.
219
,
700
(
1978
).
19.
H.
Wong
and
M.
Goldstein
,
J. Geophys. Res.
91
,
5617
, (
1986
).
20.
M.
Longtin
and
B. Ö.
Sonnerup
,
J. Geophys. Res.
91
,
6816
, (
1986
).
21.
J. V.
Hollweg
,
R.
Esser
, and
V.
Jayanti
,
J. Geophys. Res.
98
,
3491
, (
1993
).
22.
J. V.
Hollweg
,
J. Geophys. Res.
99
,
23431
, (
1994
).
23.
A.
Van Ballegooijen
,
M.
Asgari-Targhi
,
S.
Cranmer
, and
E.
DeLuca
,
Astrophys. J.
736
,
3
(
2011
).
24.
X.
Fu
,
F.
Guo
,
H.
Li
, and
X.
Li
,
Astrophys. J.
890
,
161
(
2020
).
25.
X.
Fu
,
H.
Li
,
F.
Guo
,
X.
Li
, and
V.
Roytershteyn
,
Astrophys. J.
855
,
139
(
2018
).
26.
G. A.
Gary
,
Sol. Phys.
203
,
71
(
2001
).
27.
T.
Matsumoto
and
T. K.
Suzuki
,
Mon. Not. R. Astron. Soc.
440
,
971
(
2014
).
28.
M.
Shoda
,
T.
Yokoyama
, and
T. K.
Suzuki
,
Astrophys. J.
860
,
17
(
2018
).
29.
V.
Réville
,
A.
Tenerani
, and
M.
Velli
,
Astrophys. J.
866
,
38
(
2018
).
30.
M.
Shi
,
H.
Li
,
C.
Xiao
, and
X.
Wang
,
Astrophys. J.
842
,
63
(
2017
).
31.
T. A.
Bowen
,
S.
Badman
,
P.
Hellinger
, and
S. D.
Bale
,
Astrophys. J. Lett.
854
,
L33
(
2018
).
32.
S. R.
Spangler
,
J. A.
Leckband
, and
I. H.
Cairns
,
Phys. Plasmas
4
,
846
(
1997
).
33.
Y.
Narita
,
K. H.
Glassmeier
,
M.
Fränz
,
Y.
Nariyuki
, and
T.
Hada
,
Nonlinear Processes Geophys.
14
,
361
(
2007
).
34.
S.
Dorfman
,
H.
Hietala
,
P.
Astfalk
, and
V.
Angelopoulos
,
Geophys. Res. Lett.
44
,
2120
, (
2017
).
35.
W.
Gekelman
,
S.
Vincena
,
D.
Leneman
, and
J.
Maggs
,
J. Geophys. Res.
102
,
7225
, (
1997
).
36.
J.
Maggs
and
G.
Morales
,
Phys. Rev. Lett.
91
,
035004
(
2003
).
37.
S.
Vincena
and
W.
Gekelman
,
Phys. Plasmas
13
,
064503
(
2006
).
38.
D.
Auerbach
,
T.
Carter
,
S.
Vincena
, and
P.
Popovich
,
Phys. Rev. Lett.
105
,
135005
(
2010
).
39.
D.
Auerbach
,
T.
Carter
,
S.
Vincena
, and
P.
Popovich
,
Phys. Plasmas
18
,
055708
(
2011
).
40.
G.
Howes
,
D.
Drake
,
K.
Nielson
,
T.
Carter
,
C.
Kletzing
, and
F.
Skiff
,
Phys. Rev. Lett.
109
,
255001
(
2012
).
41.
S.
Dorfman
and
T.
Carter
,
Phys. Rev. Lett.
110
,
195001
(
2013
).
42.
S.
Dorfman
and
T.
Carter
,
Phys. Rev. Lett.
116
,
195002
(
2016
).
43.
T.
Terasawa
,
M.
Hoshino
,
J.-I.
Sakai
, and
T.
Hada
,
J. Geophys. Res.
91
,
4171
, (
1986
).
44.
B. J.
Vasquez
,
J. Geophys. Res.
100
,
1779
, (
1995
).
45.
J. A.
Araneda
,
E.
Marsch
, and
A. F.
Vinas
,
J. Geophys. Res.
112
,
A04104
, (
2007
).
46.
L.
Matteini
,
S.
Landi
,
L.
Del Zanna
,
M.
Velli
, and
P.
Hellinger
,
Geophys. Res. Lett.
37
,
L20101
, (
2010
).
47.
C.
González
,
A.
Tenerani
,
M.
Velli
, and
P.
Hellinger
,
Astrophys. J.
904
,
81
(
2020
).
48.
Y.
Nariyuki
,
S.
Matsukiyo
, and
T.
Hada
,
New J. Phys.
10
,
083004
(
2008
).
49.
S.
Tanaka
,
T.
Ogino
, and
T.
Umeda
,
J. Geophys. Res.
112
,
A10110
, (
2007
).
50.
A.
Tenerani
,
M.
Velli
,
L.
Matteini
,
V.
Réville
,
C.
Shi
,
S. D.
Bale
,
J. C.
Kasper
,
J. W.
Bonnell
,
A. W.
Case
,
T. D.
De Wit
 et al,
Astrophys. J., Suppl. Ser.
246
,
32
(
2020
).
51.
H.
Karimabadi
,
H.
Vu
,
D.
Krauss-Varban
, and
Y.
Omelchenko
,
Numer. Modell. Space Plasma Flows
,
359
,
257
(
2006
).
52.
F.
Li
,
X.
Fu
, and
S.
Dorfman
,
Astrophys. J.
924
,
33
(
2022
).
53.
A.
Gigliotti
,
W.
Gekelman
,
P.
Pribyl
,
S.
Vincena
,
A.
Karavaev
,
X.
Shao
,
A. S.
Sharma
, and
D.
Papadopoulos
,
Phys. Plasmas
16
,
092106
(
2009
).
54.
A.
Karavaev
,
N.
Gumerov
,
K.
Papadopoulos
,
X.
Shao
,
A.
Sharma
,
W.
Gekelman
,
Y.
Wang
,
B.
Van Compernolle
,
P.
Pribyl
, and
S.
Vincena
,
Phys. Plasmas
18
,
032113
(
2011
).
55.
K.
Stasiewicz
,
P.
Bellan
,
C.
Chaston
,
C.
Kletzing
,
R.
Lysak
,
J.
Maggs
,
O.
Pokhotelov
,
C.
Seyler
,
P.
Shukla
,
L.
Stenflo
 et al,
Space Sci. Rev.
92
,
423
(
2000
).
56.
J. V.
Hollweg
,
J. Geophys. Res.
104
,
14811
, (
1999
).
57.
D.
Thuecks
,
C.
Kletzing
,
F.
Skiff
,
S.
Bounds
, and
S.
Vincena
,
Phys. Plasmas
16
,
052110
(
2009
).
58.
D. A.
Gurnett
and
A.
Bhattacharjee
,
Introduction to Plasma Physics: With Space and Laboratory Applications
(
Cambridge University Press
,
2005
).
59.
D.
Pesme
,
G.
Laval
, and
R.
Pellat
,
Phys. Rev. Lett.
31
,
203
(
1973
).