A publishing partnership

The following article is Open access

Local Proton Heating at Magnetic Discontinuities in Alfvénic and Non-Alfvénic Solar Wind

, , , and

Published 2024 March 7 © 2024. The Author(s). Published by the American Astronomical Society.
, , Parker Solar Probe: Insights into the Physics of the Near-Solar Environment Citation C. A. González et al 2024 ApJ 963 148 DOI 10.3847/1538-4357/ad1be5

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

0004-637X/963/2/148

Abstract

We investigate the local proton energization at magnetic discontinuities/intermittent structures and the corresponding kinetic signatures in velocity phase space in Alfvénic (high cross helicity) and non-Alfvénic (low cross helicity) wind streams observed by Parker Solar Probe. By means of the partial variance of increments method, we find that the hottest proton populations are localized around compressible, coherent magnetic structures in both types of wind. Analysis of parallel and perpendicular temperature distributions suggest that the Alfvénic wind undergoes preferential enhancements of T at such structures, whereas the non-Alfvénic wind experiences preferential T enhancements. Although proton beams are present in both types of wind, the proton velocity distribution function displays distinct features. Hot beams, i.e., beams with beam-to-core perpendicular temperature T⊥,b/T⊥,c up to three times larger than the total distribution anisotropy, are found in the non-Alfvénic wind, whereas colder beams are in the Alfvénic wind. Our data analysis is complemented by 2.5D hybrid simulations in different geometrical setups, which support the idea that proton beams in Alfvénic and non-Alfvénic wind have different kinetic properties and different origins. The development of a perpendicular nonlinear cascade, favored in balanced turbulence, allows a preferential relative enhancement of the perpendicular plasma temperature and the formation of hot beams. Cold field-aligned beams are instead favored by Alfvén wave steepening. Non-Maxwellian distribution functions are found near discontinuities and intermittent structures, pointing to the fact that the nonlinear formation of small-scale structures is intrinsically related to the development of highly nonthermal features in collisionless plasmas. Our results contribute to understanding the role of different coherent structures in proton energization and their implication in collisionless energy dissipation processes in space plasmas.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 4.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

Magnetic discontinuities have been observed for a long time throughout the heliosphere (Colburn & Sonett 1966; Burlaga 1991; Parker 1994; Tsurutani & Ho 1999; Vasquez et al. 2007), and particularly in turbulent solar wind streams where discontinuities have been related to intermittent structures (Greco et al. 2008a, 2016). Magnetic discontinuities and intermittent structures may play an important role in energy dissipation and transport, and in particle acceleration (Osman et al. 2010, 2012; Tessein et al. 2013). Several mechanisms have been proposed for solar wind heating, including resonant wave–particle interactions (such as Landau damping and ion-cyclotron resonance; Gary & Borovsky 2008; Sahraoui et al. 2010; Narita & Marsch 2015; Bowen et al. 2022), magnetic pumping (Lichko et al. 2017), stochastic heating (Chen et al. 2001; Johnson & Cheng 2001; Chandran et al. 2010; Vech et al. 2017) and intermittent dissipation (Dmitruk et al. 2004; Karimabadi et al. 2013; Osman et al. 2014). However, solar wind streams differ by the type of fluctuations and turbulence, and which heating mechanisms are favored under different solar wind conditions remains to be understood.

Tangential and rotational discontinuities (TDs and RDs, respectively) are the most common types of discontinuities observed in the solar wind (Neugebauer 2006; Paschmann et al. 2013; Liu et al. 2021). In theory, TDs and RDs can be identified depending on the normal component of the magnetic field and on the changes of the magnetic field strength (Hudson 1970). In this regard, here we refer to compressible fluctuations as those fluctuations that display changes in the magnetic field magnitude. TDs are stationary structures in the plasma frame and have zero normal magnetic field component. These structures represent boundaries between different plasma parcels and are in pressure balance, without restrictions on the variations of magnetic field magnitude across the discontinuity. TDs are commonly found at reconnection sites or, more generally, at the boundaries of flux tubes (Greco et al. 2009; Servidio et al. 2011; Zhdankin et al. 2012). By contrast, RDs have a nonzero normal magnetic field component and a field-aligned flow. RDs are propagating structures that have been associated with phase-steepened Alfvén waves (Tsurutani et al. 1994; Medvedev et al. 1997; Vasquez & Hollweg 2001).

The classification of TDs and RDs is not straightforward, not only due to single-spacecraft limitations on accurate estimations of the normal component but also because discontinuities are non-ideal and can display properties of both TDs and RDs (Neugebauer et al. 1984; Horbury et al. 2001; Knetter et al. 2004; Artemyev et al. 2019). Nevertheless, previous work has shown that different types of magnetic structures are statistically detected in different types of wind. Namely, compressible structures are most likely observed in the slow solar wind, while structures with small compressibility are typically observed in the fast solar wind (Perrone et al. 2016, 2017). The solar wind has been traditionally categorized into slow and fast wind. The "slow wind" generally refers to streams with velocities below 500 km s−1, mostly confined to the equatorial regions of the Sun. By contrast, the "fast wind" has higher speeds, up to 700 km s−1 and above, and it is associated with open field regions on the Sun's surface. Although slow and fast streams are often associated with non-Alfvénic and Alfvénic solar wind (Grappin et al. 1990), recent observations have shown the existence of "slow Alfvenic" wind (D'Amicis et al. 2021). This suggests that the solar wind cannot be categorized based solely on its velocity, and that the content of Alfvénic fluctuations provides a better way to classify wind streams. The Alfvénicity of the wind can be quantified by specific measures, such as the normalized residual energy (the relative energy in kinetic and magnetic fluctuations), the Alfvén ratio (the ratio between kinetic and magnetic fluctuations), and the normalized cross-helicity (the correlation between magnetic and velocity field fluctuations; see, e.g., the discussion in Parashar et al. 2020). In this work, we classify the solar wind based on the average value of cross-helicity. We consider the solar wind as Alfvénic when it contains fluctuations with cross-helicity close to unity. By contrast, the non-Alfvénic wind refers to intervals with cross-helicity approaching zero.

Several studies have shown that plasma temperature enhancements are associated with intermittent structures/discontinuities identified via the partial variance of increment (PVI) method that identifies variations in the magnetic field and strong gradients associated with coherent magnetic structures (Osman et al. 2010; Greco et al. 2018; Qudsi et al. 2020; Sioulas et al. 2022b; Phillips et al. 2023). However, TDs and RDs are fundamentally different types of structures that affect particle energization in different ways, and how compressible (TD-like) and incompressible (RD-like) structures in the solar wind interact with particles ultimately leading to heating (and their velocity-space signatures) remains to be investigated.

In this work, we address such a problem by investigating proton heating at different types of intermittent structures and/or discontinuities in the inner heliosphere, by using Parker Solar Probe (PSP) data and 2.5D hybrid-kinetic simulations. Toward this goal, we conduct a comparison study between Alfvénic and non-Alfvénic streams and investigate the dependence of temperature anisotropies on Alfvénicity and on the type of magnetic structures. Additionally, hybrid simulations, with simplified geometry settings, are used to investigate what processes are likely to contribute to proton heating and to the generation of the observed proton temperature anisotropy in each type of wind.

This paper is organized as follows. In Section 2, we describe the data and methods, and we report on our data analysis. An overview of the properties of magnetic field fluctuations in Alfvénic and non-Alfvénic wind is given in Section 2.1. In Section 2.2 we discuss the correlation between PVI and proton temperatures and temperature anisotropy, and in Section 2.3 we consider two case studies to show the typical signatures in proton velocity space at different magnetic structures in the two types of wind. Results from numerical simulations and a comparison between numerical outputs and PSP data are reported in Section 3. The summary and discussion are in Section 4.

2. PSP Observations

2.1. Overview of Fluctuation Properties for E6–E9

In this work we have used PSP data from Encounter 6 through Encounter 9 (E6–E9), which occurred from 2020 September 9 to 2021 August 15 covering a range of radial distances from the Sun, 0.07 au < R < 0.263 au (we only considered intervals at radial distances less than 50 solar radii, Rs ). We use magnetic field data from the flux-gate magnetometer at 4 samples cycle−1 resolution (FIELDS; Bale et al. 2016). The 3D proton velocity distribution function (L2) and its moments (L3) at 3 s resolution are obtained by the electrostatic analyzer (SPAN-I; Livi et al. 2022), part of the Solar Wind Electrons Alphas and Protons instrument suite (SWEAP; Kasper et al. 2016). Since SPAN-I is partially obstructed by PSP's thermal protection shield, we use the quasi-thermal noise (QTN) measured by the FIELDS Radio Frequency Spectrometer (Moncuquet et al. 2020) to have a more accurate determination of plasma density. The total parallel and perpendicular proton temperatures, T and T, are obtained by projecting the temperature tensor in the parallel and perpendicular directions with respect to the magnetic field.

For reference, in Figure 1 we provide an overview of E6. The top four panels show the spacecraft radial distance (R), the proton and electron number density (ni,e ), and the magnetic ( b and B = ∣ b ∣) and proton bulk velocity ( V ) fields in the instrument frame (x-component pointing toward the Sun). The fifth panel shows the Alfvénic properties of the wind represented by the normalized cross-helicity σc and residual energy σr of fluctuations given, respectively, by

Equation (1)

where brackets 〈...〉 denote the moving average over a time window, and ${{\boldsymbol{z}}}^{\pm }={\boldsymbol{\delta }}{\boldsymbol{V}}\pm {\boldsymbol{\delta }}{\boldsymbol{b}}/\sqrt{{\mu }_{0}\rho }$ are the Elsässer variables defined by the velocity and magnetic field fluctuations with respect to the mean field and by the QTN mass density ρ assuming quasi-neutrality. In this work we fix the averaging window to 2 hr, which corresponds to several times the correlation time of the magnetic field fluctuations in the inner heliosphere (Parashar et al. 2020; Sioulas et al. 2022b). The cross-helicity measures the relative dominance of inward and outward propagating Alfvénic fluctuations ( z ±), and the residual energy quantifies the partition between the fluctuations' kinetic and magnetic energy. These quantities are not independent and they are geometrically related by the cosine of the angle between the velocity and magnetic field $\cos {\theta }_{\mathrm{vb}}\,=\tfrac{{\boldsymbol{\delta }}{\boldsymbol{V}}\cdot {\boldsymbol{\delta }}{\boldsymbol{b}}}{| {\boldsymbol{\delta }}{\boldsymbol{V}}| | {\boldsymbol{\delta }}{\boldsymbol{b}}| }={\sigma }_{c}/\sqrt{1-{\sigma }_{r}^{2}}$ (Wicks et al. 2013). The bottom panel of Figure 1 shows θvb for E6.

Figure 1.

Figure 1. PSP Encounter 6 data: spacecraft radial distance R (top panel); proton and electron number density ni,e (second panel); magnetic field b , its magnitude B and proton bulk velocity V in the instrument frame (third and fourth panels); normalized cross-helicity σc , residual energy σr , and the cosine of the angle between the velocity and magnetic field fluctuations $\cos {\theta }_{\mathrm{vb}}$ (bottom panel).

Standard image High-resolution image

As will be discussed below, fluctuations during E6 are remarkably Alfvénic, with high cross-helicity intervals (defined by ∣σc ∣ > 0.75) that last several days (particularly during and after perihelion). In addition, a shorter interval of low Alfvénicity is observed in the vicinity of the heliospheric current sheet (2020 September 25 and 26), a complex structure with dominant magnetic energy and nearly zero cross-helicity. This agrees with the observed non-Alfvénic wind reported near the Heliospheric Current Sheet (HCS) in E4 (Chen et al. 2021).

In Figure 2 we show the statistical analysis of Alfvénic properties and compressibility of solar wind fluctuations from E6 through E9. The top panels show the probability density function (PDF) of σc , σr , and $\cos {\theta }_{\mathrm{ub}}$ for each Encounter. Although data display a range of values −1 ≤ σc ≤ 1, intervals of imbalanced fluctuations with ∣σc ∣ > 0.5 can be identified, with peaks of the distribution of σc at ∣σc ∣ ≥ 0.75 (see also Figure 1). In general, fluctuations are magnetically dominated, a trend which is consistent throughout all Encounters (Chen et al. 2020; Shi et al. 2021), with mean value σr ≈ − 0.5 and with only a few intervals being characterized by an excess of kinetic energy. Lastly, the PDF of $\cos {\theta }_{\mathrm{ub}}$ peaks at ±1, corresponding to fluctuations with aligned V and B in all the Encounters. The Alfvénic wind (those intervals with ∣σc ∣ ≥ 0.75) present strong alignment, i.e., $| \cos {\theta }_{\mathrm{vb}}| \sim 0.89$, while the non-Alfvénic wind (those intervals with ∣σc ∣ ≤ 0.25) present a moderate alignment with a broader distribution of angles, having mean value $| \cos {\theta }_{\mathrm{vb}}| \sim 0.57$. In the latter case, such alignment corresponds with the relaxation and suppression of nonlinearities in standard balanced turbulence that have been previously observed in both simulations and solar wind measurements (Boldyrev 2006; Matthaeus et al. 2008, 2012; Chandran et al. 2015).

Figure 2.

Figure 2. Fluctuation properties for E6–E9. Top panels: PDF of normalized cross-helicity σc (top-left panel), residual energy σr (middle panel), and the cosine angle of velocity and magnetic field fluctuations $\cos {\theta }_{\mathrm{vb}}$ (top-right panel). Different Encounters correspond to different line styles as indicated in the legend at the top. The shaded blue and yellow areas indicate values ∣σc ∣ > 0.75 and ∣σc ∣ < 0.25, respectively. Bottom panels: PDF of the normalized fluctuations amplitude δ b/∣〈 b 〉∣ (bottom left) and of the ratio of magnetic field and magnetic field strength variances Cb (bottom right) for the Alfvénic wind (blue lines, corresponding to the blue-shaded area in the top-left panel) and for the non-Alfvénic wind (yellow lines, corresponding to the yellow-shaded area). The values in the legend of the bottom panels shows the mean value of each distribution.

Standard image High-resolution image

The bottom panels of Figure 2 show the PDF of the amplitude of magnetic field fluctuations,

Equation (2)

and of the ratio of the variances of the magnetic field and magnetic field strength (Villante & Vellante 1982),

Equation (3)

the latter providing a proxy for magnetic compressibility. In this analysis, we have sorted the solar wind into Alfvénic (intervals with ∣σc ∣ > 0.75, corresponding to the blue-shaded areas in Figure 2, top left) and non-Alfvénic wind (intervals with ∣σc ∣ < 0.25, corresponding to the yellow-shaded area). For the non-Alfvénic wind we have also removed data corresponding to the HCS crossing, where ∣〈 b 〉∣ ≈ 0 (for the time window considered of 2 hr), to avoid fictitious long tails of the PDF.

From Figure 2, bottom-left panel, it can be observed that the distributions of the normalized fluctuations' amplitude are remarkably different between Alfvénic (blue) and non-Alfvénic (yellow) wind. In the former case, relative fluctuations' amplitudes are bounded roughly by δ b/∣〈 b 〉∣ ≲ 2. This is in line with previous observations of Alfvénic fluctuations by Helios and Ulysses, where saturation of amplitudes (although defined differently) was found, a feature that is consistent with spherically polarized fluctuations (Matteini et al. 2018). In the non-Alfvénic case, instead, relative fluctuations' amplitudes are large and there is no constraint on their value (E9 being the only ambiguous case). We interpret the different PDF of the normalized amplitude as a signature of a higher level of compressibility in the non-Alfvénic wind (at the timescale of 2 hr). The bottom-right panel of Figure 2 shows the distribution of Cb for both types of wind. As expected, Cb in the non-Alfvénic wind is systematically larger than in the Alfvénic wind, with the PDF mean in the non-Alfvénic wind reaching up to five times that of the Alfvénic wind.

To summarize, we have considered the statistical properties of fluctuations over a timescale of 2 hr. From this part of our data analysis, we conclude that fluctuations in the Alfvénic (∣σc ∣ > 0.75) and non-Alfvénic (∣σc ∣ < 0.25) wind at distances R < 0.25 au display statistical properties similar to those traditionally observed further away from the Sun, namely, almost incompressible and saturated fluctuations in the Alfvénic wind and highly compressible fluctuations in the non-Alfvénic wind.

2.2. Correlations between PVI and Proton Thermal Properties in Different Types of Wind

To study the connection between proton heating and discontinuities or, more generally, intermittent structures in different plasma conditions, we utilized the PVI method by calculating the following quantity (Greco et al. 2008b):

Equation (4)

where Δ B (t, τ) = b (t + τ) − b (t) is the magnetic field increment vector. The PVI method identifies non-Gaussian features in the magnetic field and events, with values PVI > 3 typically associated with coherent structures representative of intermittent behavior, or discontinuities. Although the PVI defined in Equation (4) captures variations in the magnetic field, it does not distinguish RDs, TDs, or other fluctuations that sometimes display properties of both RDs and TDs (Larosa et al. 2021). For this reason, we also considered the PVI of the magnetic field strength, hereafter called MAG-PVI, by taking the variation of B, i.e., by substituting ΔB(t, τ) ⇒ ΔB(t, τ) = B(t + τ) − B(t) in Equation (4). A comparison of events identified with PVI and MAG-PVI allows us to gain insights on whether compressible or incompressible magnetic structures are statistically associated to local particle energization.

We performed the PVI analysis by using different time lags, τ = 12 s, 3 s, 1 s, to find a compromise between the resolution of the SPAN-i instrument and proton scales. By comparison, the proton gyroperiod has a mean value of τi ≈ 0.1 s over the range of radial distances considered (the gyroperiod is τi ≈ 0.15 s at a distance of R ≈ 50 Rs, and it decreases down to about τi ≈ 0.025 at R ≈ 20 Rs ). By applying the Taylor hypothesis, the lags τ = 12−1 s correspond, on average, to a spatial lag /di ≈ 360–36 and /ρi ≈ 750–75 in units of proton inertial length and gyroradius, respectively.

After constructing the time series of PVI (and MAG-PVI), we computed the normalized PDF of various functions of temperature f(T) conditioned on a given range of PVI (and MAG-PVI) values

Equation (5)

We have calculated the conditional PDF of (i) the total proton temperature (f(T) = T) at a given point in the PVI time series, (ii) the variation of proton temperature over the same scale that PVI has been calculated over (f(T) = T(t + τ) − T(t) ≡ δ T, with τ the time lag of the PVI), (iii) the parallel and perpendicular temperature (f(T) = T, T where ∥ and ⊥ are defined with respect to the local magnetic field direction $\hat{{\boldsymbol{b}}}={\boldsymbol{b}}/B$), and (iv) the temperature anisotropy (f(T) = T/T). We performed the same PVI analysis with τ = 12 s, τ = 3 s, and τ = 1 s, in the latter case by resampling the PVI data to match the SPAN-i resolution. Results vary within the uncertainty of the temperature measurements (usually about 10%) and, therefore, we only show results with τ = 12 s.

In agreement with other studies, we find that the highest values of T and δ T are associated with the highest PVI values in both winds (not shown here; see also, e.g., Osman et al. 2010; Qudsi et al. 2020; Sioulas et al. 2022b), supporting the idea that heating occurs at localized structures. Results of our PVI analysis are reported in Figure 3 for parallel and perpendicular temperatures.

Figure 3.

Figure 3. Conditional PDF with respect to PVI (top panels) and MAG-PVI (bottom panels) of proton temperatures using data from E6 through E9 for a time lag τ = 12 s: PDF of parallel proton temperature T (left panels); PDF of perpendicular proton temperature T (middle panels); PDF of proton temperature anisotropy T/T (right panels). Solid lines correspond to Alfvénic wind and dashed lines to non-Alvénic wind. Different color lines show PDFs that corresponds to different ranges of PVI (MAG-PVI): black lines (0 < PVI ≤ 1.5), blue lines (1.5 < PVI ≤ 3.0), and red lines (PVI > 3.0). The numbers on the legends indicate the mean value of each distribution.

Standard image High-resolution image

Figure 3 shows the conditional PDF with respect to PVI (top panels) and MAG-PVI (bottom panels) of T (left), T (middle), and T/T (right). The PDFs have been obtained by combining data from all Encounters (E6–E9) and by using the cross-helicity to sort Alfvénic (∣σc ∣ > 0.75, solid lines) and non-Alfvénic (∣σc ∣ < 0.25, dashed lines) intervals. Red, blue, and black colors correspond to different ranges of PVI (MAG-PVI), which we split into 0 < PVI ≤ 1.5, 1.5 < PVI ≤ 3.0, and PVI > 3.0, respectively. The legend in each panel reports the mean value of each distribution. Finally, E6 was also analyzed individually, but we found the same trends as those shown in Figure 3 and discussed below.

As can be seen from the PDFs of T/T (Figure 3, right panels), the non-Alfvénic wind is more anisotropic than Alfvénic wind and, on average, ${({T}_{\perp }/{T}_{\parallel })}_{\mathrm{nonAlfv}}\lt {({T}_{\perp }/{T}_{\parallel })}_{{Alfv}}\lt 1$. As we will see, such an anisotropy is due to the ubiquitous presence of proton beams. While proton beams have been commonly observed at all radial distances in the Alfvénic wind before PSP (e.g., Marsch 2012), the frequent observation of field-aligned beams in the non-Alfvénic wind (including the so-called "hammerhead" distributions; Verniero et al. 2020, 2022) is one of the unexpected results of PSP. Further inspection of the temperature anisotropy PDFs shows that Alfvénic and non-Alfvénic wind display an opposite correlation between PVI (and MAG-PVI) values and T/T. For the PVI case (Figure 3, top-right panel), the mean values of the non-Alfvénic wind anisotropy lie in the range 0.71 ≲ T/T ≲ 0.73 with increasing PVI (dashed lines). In the Alfvénic wind (solid lines), we find instead 0.88 ≲ T/T ≲ 0.97 with decreasing PVI. The MAG-PVI case (Figure 3, bottom-right panel) shows similar values of anisotropy and trends.

Insights on parallel and perpendicular temperatures can be found by inspecting the PDFs of T and T (Figure 3, left and middle panels). Interestingly, we find that the ensembles of events selected through each MAG-PVI value have consistently higher average temperatures than the ensembles of events selected with the traditional PVI values. In Table 1 we report the range of the average T and T of the distributions shown in Figure 3, bottom panels, spanned as MAG-PVI increases. To quantify the trends of perpendicular and parallel temperature with MAG-PVI, the relative average temperature enhancement between the smallest and the largest PVI range is also reported in the table. We find that the Alfvénic wind shows a relative enhancement of parallel temperature larger than in the perpendicular temperature, as larger PVI values are considered. The opposite trend is found in the non-Alfvénic wind. This explains why the anisotropy T/T shifts away from (toward) unity as PVI values increase for the Alfvénic (non-Alfvénic) wind. Relative variations of temperature at intermittent structures have been found to be about 10% (Sioulas et al. 2022b), so that we expect parallel and perpendicular relative variations to be of that order.

Table 1. Average Parallel and Perpendicular Temperatures Shown in Figure 3 Conditioned on the MAG-PVI

Wind type T range (×105 K) T range(×105 K)ΔT/T ΔT/T
Alfvénic4.8 < T < 5.84.6 < T < 5.420%17%
non-Alfvénic4.1 < T < 4.62.6 < T < 3.112%19%

Note. The temperature values are rounded to the first significant figure based on an uncertainty of 10%. The last two columns report the relative variation of parallel and perpendicular average temperature as MAG-PVI varies from the lowest to the largest range of values.

Download table as:  ASCIITypeset image

From this analysis we conclude, first, that structures with the largest variations in B (magnetic compressible structures) contribute to the highest T and T in both types of wind, suggesting that the largest temperature enhancements occur in compressible localized structures also in the Alfvénic wind. Second, analysis of the distribution of T and T suggests that the Alfvénic wind undergoes both parallel and perpendicular local temperature enhancements, with a slightly a preferential enhancement of T, whereas the non-Alfvénic wind undergoes a preferential local enhancement of T.

2.3. Proton Kinetic Features at Different Magnetic Structures

To investigate the local energization of protons at discontinuities/intermittent structures and kinetic features of the distribution function in velocity space, we manually inspected 1 hr intervals during quiet solar wind and without switchbacks. We selected two subintervals that characterize the Alfvénic and the non-Alfvénic wind (according to high and low cross-helicity), respectively. The two selected subintervals occurred during E6 at approximately the same radial distance from the Sun (30 RS ). Different fields and particle quantities are reported in Figure 4, where panels on the left correspond to the non-Alfvénic wind and those on the right to the Alfvénic wind.

Figure 4.

Figure 4. Data from two intervals during E6. The non-Alfvénic interval is shown on the left panels and the Alfvénic one on the right. From top to bottom: proton mass density (ρ [kg m−3]); magnetic field ( b and B [nT]); proton bulk velocity ( V [km s−1]); parallel and perpendicular temperature (T and T [eV]); temperature anisotropy (T/T) and plasma beta (β). The last two panels show fluctuations properties (σr , σc , and $\cos {\theta }_{\mathrm{vb}}$) and the PVI values.

Standard image High-resolution image

The non-Alfvénic subinterval occurred near the heliospheric current sheet crossing, on 2020 September 25 from 03:10:00 to 04:10:00 UT, with an average cross-helicity σc ≃ −0.25. This interval presents many discontinuities with large PVI values. The structures are substantially compressible with variations of B and β enhancements.

The Alfvénic interval occurred on 2020 September 30 between 01:00:00 and 02:00:00 UT with an overall $\left|{\sigma }_{c}\right|\simeq 0.88$. The fluctuations show the typical velocity field is highly correlated with magnetic field that characterizes wave-like fluctuations with rotation of $\cos {\theta }_{\mathrm{ub}}$ near small-scale structures. In general, the variations of the magnitude of the magnetic field and the PVI values are smaller compared to the discontinuities in the non-Alfvénic wind. In both cases, however, the most substantial variations in proton temperature occur within large PVI and MAG-PVI structures and correspond to a net enhancement of T/T, as can be seen by comparing the fourth or fifth panel and the bottom panel.

The signatures in proton velocity distribution function (VDF) of local heating for these two representative subintervals are given in Figure 5, where we show the reduced VDF, $f(t,{v}_{\parallel }),f(t,{v}_{{\perp }_{\mathrm{1,2}}})$ for the non-Alfvénic (left panels) and for the Alfvénic (right panels) subintervals described above (and reported in Figure 4). We also show the reconstructed 3D VDFs in the plane $\{({v}_{\parallel },{v}_{{\perp }_{1}},{v}_{{\perp }_{2}})\}$, and the VDF projection into the $({v}_{\parallel },{v}_{{\perp }_{\mathrm{1,2}}})$ planes, at the bottom panels for each type of wind. The reduced VDF has been obtained by first transforming the initial 3D energy distribution function from the {E, θ, φ} space (energy, elevation, and azimuth) to velocity space in the field-aligned coordinate system $\{{v}_{\parallel },{v}_{{\perp }_{1}},{v}_{{\perp }_{2}}\}$. Details can be found in Appendix A.

Figure 5.

Figure 5. The reduced proton VDF as a function of parallel (f(t, v)) and perpendicular ($f(t,{v}_{{\perp }_{\mathrm{1,2}}})$) velocity components for the non-Alfvénic (left column) and Alfvénic (right column) subintervals. The bottom panels show two examples of the reconstructed 3D VDF $f({v}_{\parallel },{v}_{{\perp }_{1}},{v}_{\perp ,2})$ and its 2D projections, $f({v}_{\parallel },{v}_{{\perp }_{\mathrm{1,2}}})$, at the time of an event with PVI > 3 in each type of wind.

Standard image High-resolution image

As can be seen from Figure 5, the proton VDF shows strong deviations from local thermal equilibrium and gyrotropy in the vicinity of large PVI values. We observe that protons have enhanced perpendicular velocities in those regions, and parallel propagating proton beams around the local Alfvén speed are observed near discontinuities in both Alfvénic and non-Alfvénic intervals. This can be seen by inspecting Figure 6, which shows a scatterplot of the beam-to-core drift velocity normalized by Alfvén speed versus ${({T}_{\perp ,b}/{T}_{\perp ,c})}^{* }$, which is the ratio of the perpendicular temperature of the beam and the core (T⊥,b /T⊥,c ) normalized by the total (beam+core) T/T values. This normalized quantity considers the strength of the beam parallel to the mean field and it provides a measure of the deformation of the distribution function. To help intuition, we have split the scatterplot into four quadrants that we use as a reference to characterize the shape of the VDF. The upper-right quadrant corresponds to distributions with a beam and T⊥,b > T⊥,c (i.e., "hammerhead" distributions); the lower-right quadrant corresponds to distributions with a beam and T⊥,b /T⊥,c < (T/T), thus indicating, in the most extreme case, a narrowly focused beam in the vv plane. On the left side, the beam and core are not well separated, and one can attribute those data points to distributions with no beams. A sketch of the shape of the distribution as one moves above and below the line ${({T}_{{\perp }_{b}}/{T}_{{\perp }_{c}})}^{* }=1$ is also presented on the bottom-right corner of Figure 6. Details on the fitting method can be found in Appendix B. For reference, the crosses show data points at t = 03:35:39 UT (non-Alfvénic wind) where ${({T}_{\perp ,b}/{T}_{\perp ,c})}^{* }=2.81$ and at t = 01:22:33 UT where ${({T}_{\perp ,b}/{T}_{\perp ,c})}^{* }=1.16$. Figure 5 suggests that beams in the Alfvénic and non-Alfvénic wind have different origins and display distinct properties.

Figure 6.

Figure 6. Scatterplot of the beam-to-core drift velocity normalized by Alfvén speed vs. the ratio of the perpendicular temperature of the beam and the core for the same Alfvenic and non-Alfvenic intervals presented before. The black crosses in the blue and red contours correspond to the local values at t = 03:35:39 UT and t = 01:22:33 UT, respectively.

Standard image High-resolution image

3. Hybrid Simulations

3.1. Initial Conditions

We performed hybrid-kinetic simulations to investigate proton heating at different types of magnetic structures under different plasma conditions and compared our results with observations. We considered a quasi-neutral hybrid plasma model with massless and isothermal electrons, while protons are modeled as kinetic particles using the particle-in-cell method (Matthews 1994; Franci et al. 2018). We adopted the following normalization: lengths are normalized to the proton inertial length di = c/ωp with ${\omega }_{p}={(4\pi {{ne}}^{2}/{m}_{i})}^{1/2}$, the proton plasma frequency, the time is normalized to the inverse of the proton gyrofrequency ${\omega }_{{ci}}^{-1}={({{eB}}_{0}/{m}_{i}c)}^{-1}$, and velocities are normalized to the Alfvén speed ${v}_{A}={B}_{0}/{(4\pi {n}_{0}{m}_{i})}^{1/2}$ with B0, the mean magnetic field. The plasma beta for both ions and electrons is defined as ${\beta }_{p,e}=8\pi {n}_{0}{T}_{p,e}/{B}_{0}^{2}$. To avoid energy accumulation at the grid scale, we have included explicit resistivity with a corresponding dissipation length (ld ) related to the Reynolds number and the box size (L) through ${R}_{e}\sim {(L/{l}_{d})}^{4/3}$, which is chosen to be greater than the grid size.

We have considered two different setups in 2.5D geometry characterized by two different types of turbulent fluctuations resembling non-Alfvénic and Alfvénic wind, namely, (i) a decaying 2D balanced turbulence simulation, and (ii) a wave simulation with an initial parallel propagating, Alfvénic broadband spectrum. For case (i), hereafter referred to as the "turbulence simulation," we considered an initial large-amplitude fluctuation with σc ≈ 0, in energy equipartition and with an out-of-plane guide field ${{\boldsymbol{B}}}_{0}={B}_{0}\hat{{\boldsymbol{z}}}$. The initial condition consists of large-amplitude perturbations of perpendicular wave modes in the range 0.196 < k di < 0.49 with random phases, similar to previous work (Servidio et al. 2012; Franci et al. 2015; Cerri & Califano 2017). For case (ii), referred to as the "wave-like simulation," we considered an in-plane mean field ${{\boldsymbol{B}}}_{0}={B}_{0}\hat{{\boldsymbol{x}}}$, and the initial perturbation satisfies the Walèn relation in the dispersionless limit δ u = −(ω0/k0)δ b . The wave frequency ω0 is determined from the normalized dispersion relation ${k}_{0}^{2}={\omega }_{0}^{2}/(1-{\omega }_{0})$ for left-handed polarized parallel propagation waves. This initial condition corresponds to a broadband Alfvénic fluctuation composed of parallel modes in the range 0.196 < k di < 0.49 (see also Malara et al. 2000, Matteini et al. 2010, González et al. 2021). In both cases, an initially isotropic and Maxwellian plasma with proton beta βi = 0.5 and equal ion and electron temperature Ti /Te = 1 are considered. The guide field is B0 = 1 and the same rms of the magnetic fluctuations δ brms/B0 = 0.63 is chosen for both simulations. We adopt periodic boundary conditions and fix a box of side L = 128 di by using 10242 grid points with mesh size 0.125 di and 8000 particles per cell.

Because of the different geometry adopted, interactions between fields and particles are fundamentally different in the two setups (Gary et al. 2020). Naturally, our simulations are not meant to represent realistically the two types of wind (both requiring 3D fields), but rather to isolate different processes that might be dominant in each type of wind and that bear specific signatures in phase space. In Section 3.3 we discuss the correlation between anisotropy and PVI, and the signatures in phase space of proton heating/acceleration in correspondence with large PVI values.

3.2. Overview of Numerical Results

In the turbulence simulation, once the system reaches the fully developed turbulent state, the energy stored in the fields is progressively converted into thermal energy, resulting in an average (over the spatial domain) preferential perpendicular heating with 〈T/Tbox = 1.37 at $t=120{\omega }_{{ci}}^{-1}$. The wave-like simulation, by contrast, displays a strong enhancement of the spatially averaged T, and the temperature anisotropy reaches 〈T/Tbox = 0.6 at $t=180{\omega }_{{ci}}^{-1}$. 4

These different heating processes in the two setups are expected, since the turbulence simulation leads to a well-developed magnetic energy spectrum perpendicular (in k space) to the guide field, allowing for more channels of particle heating, such as stochastic heating and particle scattering at current sheets (Cerri et al. 2021; Sioulas et al. 2022a). Those processes are suppressed in the wave-like simulation due to the different geometry adopted. That said, in the wave-like simulation, the available energy carried by the wave is converted into kinetic and thermal energy via the rapid disruption of the wave packet mediated by wave steepening along the guide field (i.e., through the formation of gradients parallel to B 0); this leads to both the formation of a field-aligned proton beam at about the local Alfvén speed, and local perpendicular heating at the steepened edges (González et al. 2021). This mechanism is in turn suppressed in the turbulence simulation. Figure 7 shows the contour plot of Jz for the two setups to show the different types of structures that form nonlinearly.

Figure 7.

Figure 7. Contour plot of the out-of-plane current density Jz for the turbulence simulation (left) and the wave-like simulation (right). The black dot shows the location at which the time series shown in Figure 8 have been taken.

Standard image High-resolution image

To make a comparison with PSP data, in Figure 8 we show the time series of fields and plasma quantities in the two setups. The time series have been taken at the location marked by the black dot in Figure 7. From top to bottom, Figure 8 shows the time series of n, B and ∣B∣, u , the electric field e , T, T and β, $\cos {\theta }_{\mathrm{vb}}$, σr and σc , and the PVI values (see Section 3.3 for details on the PVI analysis). The turbulence simulation (left panels) is characterized on average by ∣σc ∣ < 0.5 and σr < 0 (predominantly). The wave-like simulation (right panel) has σr ≈ 0 and ∣σc ∣ ≈ 1. As can be seen, temperature enhancements occur in correspondence with large PVI values in both cases. In the turbulence case, those large PVI values are associated with structures such as flux tubes, current sheets, and small-scale plasmoids. These structures have a strong perpendicular electric field, and some also display large out-of-plane electric and magnetic fields (along the z-axis) with strong compressibility, mediating particle acceleration and heating (Dmitruk et al. 2004; Wan et al. 2015; Comisso & Sironi 2022). As discussed above, in the wave-like simulation large PVI values are associated with the steepened edges of Alfvénic fluctuations. Also, rapid and large variations of $\cos {\theta }_{\mathrm{ub}}$ near these structures can be identified (sixth panel), similar to what is observed during the Alfvénic subinterval presented in Section 2.3. In contrast to the magnetic structures in the turbulent simulation, the steepened edges are mostly rotational discontinuities with an embedded compressive component (Matteini et al. 2010; González et al. 2021).

Figure 8.

Figure 8. Single-point time evolution of fields from the turbulence simulation (left panels) and wave-like simulation (right panels). From top to bottom: proton density, magnetic field, proton bulk velocity, electric field, parallel and perpendicular temperature, temperature anisotropy, and plasma beta. Finally, the last two panels show σc , σr and $\cos {\theta }_{\mathrm{vb}}$, and the PVI values.

Standard image High-resolution image

3.3. Correlations between Magnetic Structures and Proton Kinetic Features

The PDFs conditioned on PVI and MAG-PVI are shown in Figure 9 in the same format as Figure 3. To obtain the PDFs, we used a similar methodology to that employed for PSP data and we have taken the time series of plasma quantities at a resolution of $0.1{\omega }_{{ci}}^{-1}$ at 104 fixed equidistant points in the simulation domain. Because of computational limitations, our simulations are not long enough to allow us to take a time lag similar to that in the data analysis. That said, we have sufficient time resolution to explore fluctuations around the gyroperiod ($\tau =1{\omega }_{{ci}}^{-1}$). Therefore, we fixed the time lag equal to the proton gyroperiod to select coherent kinetic-scale structures, despite our inability to directly probe this scale with in situ data.

Figure 9.

Figure 9. Conditional PDF with respect to PVI (top panels) and MAG-PVI (bottom panels) of proton temperatures using simulation data: PDF of parallel proton temperature T (in normalized units; left panels); PDF of perpendicular proton temperature T (middle panels); PDF of proton temperature anisotropy T/T (right panels). The solid lines correspond to the wave-like simulation and the dashed lines to the turbulence simulation. Different color lines show PDFs that correspond to different ranges of PVI and MAG-PVI: black lines (0 < PVI ≤ 1.5), blue lines (1.5 < PVI ≤ 3.0), and red lines (PVI > 3.0). The numbers on the legends are the mean value of each distribution.

Standard image High-resolution image

In general, there is a positive correlation between PVI values with T and δ T (not shown), indicating that the hottest protons are found locally near small-scale structures (in this case represented by PVI > 3 values). Although the turbulence simulation is hotter than the wave-like simulation (this is true both locally at discontinuities and globally over the simulation domain), protons remain close to Maxwellian in the turbulence simulation. By contrast, in the wave-like simulation, particles develop large anisotropies localized at small-scale structures. In that case such a local anisotropy, reaching an average value of T/T ≈ 2 for PVI > 3, is caused by strong particle pitch-angle scattering at the steepened fronts, resulting in effective local proton perpendicular energization (González et al. 2021; Malara et al. 2021; Gonzalez et al. 2023). Since this effect is localized at the steepened edges, the background temperature (represented by the lower PVI range) and the spatially averaged temperatures over the entire domain display an opposite anisotropy, T/T < 1, as discussed previously, due to the formation of a field-aligned beam that fills the entire simulation domain.

Some differences between PVI (Figure 9, top panels) and MAG-PVI (Figure 9, bottom panels) results can be noticed. In the wave-like simulation, the magnetic structures correspond essentially to rotational discontinuities with relatively small variations in the magnitude of B and, thus, the correlation with the various PVI ranges is reduced in the MAG-PVI case. This can be seen from the relative temperature variations between the smallest and the largest range of PVIs, showing the aforementioned trend (ΔT/T ≃ 3% and ΔT/T ≃ 59%). For the turbulence simulation, on the contrary, a higher relative temperature increase is found, similar to observations (ΔT/T ≃ 6% and ΔT/T ≃ 8%).

In the top two panels of Figure 10 we show the gyroaveraged reduced perpendicular and parallel VDF (f(t, v) and f(t, v), respectively, and in the bottom panels we show the reduced VDF f(v, v) at the time indicated by the dashed line, marking the occurrence of an event with PVI > 3 in each case. The left column corresponds to the turbulence simulation and the right to the wave-like. As can be seen, magnetic discontinuities/intermittent structures are characterized by non-Maxwellian features, namely temperature anisotropies and beams at the Alfvén speed. However, the turbulence simulation undergoes perpendicular heating in both the core and the beams, so that TT locally and, on average, and beams are "hot." The wave-like simulation instead displays a strong perpendicular heating of the core, localized at the discontinuities, and a cold beam at the Alfvén speed (more focused in the perpendicular direction than in the turbulence setup).

Figure 10.

Figure 10. Single-point proton VDF from simulations. The reduced proton VDF of perpendicular ($f(t,{v}_{{\perp }_{1},{\perp }_{2}},t)$) and parallel f(t, v) of proton velocity for the turbulence simulation (left panels) and wave-like simulation (right panels). The bottom panels show the reduced VDF f(v, v) at the time indicated by the red dashed line.

Standard image High-resolution image

To summarize, our simulations show that, in analogy with PSP data, higher temperatures and temperature variations are found at higher PVI values in both setups (both winds), and that colder beams (as found in Alfvénic wind) are associated with the steepening of Alfvénic fluctuations while hot beams (as found in non-Alfvénic wind) are generated in low cross-helicity turbulence. The development of a standard turbulent cascade also favors preferential perpendicular heating due to different mechanisms such as stochastic heating (as in the non-Alfvénic wind). Even though our simulations highlight the different origins and properties of proton beams in the two types of setups/winds and qualitatively reproduce observed properties of non-Alfvénic wind, they do not reflect all the local trends of temperatures at magnetic structures observed in the Alfvénic wind. In particular, the PVI-temperature analysis in the wave-like simulation shows a much larger relative increase of perpendicular temperature than observed (larger than the increase in parallel temperature), and compressibility of the Alfvénic wind at small scales is also not well reproduced numerically. These discrepancies are likely due to the geometry contsraints. However, we also mention the possibility that instrument resolution may also play a role in underestimating strong temperature variations at small scales.

Before concluding we also remark that our simulations do not reproduce differences between the two types of winds in their average properties, such as the fact that the Alfvénic wind is hotter, and that the global (average) anisotropy is T/T < 1. This is expected since our setups cannot reproduce fully 3D dynamics. Furthermore, such differences might well be related to different coronal origins and/or a different radial evolution of Alfvénic and non-Alfvénic wind, an aspect that is not addressed in this work.

4. Summary and Conclusions

We have studied the correlation between proton temperatures and magnetic discontinuities/ intermittent structures in different solar wind turbulence conditions (high and low cross-helicity, i.e., Alfvénic and non-Alfvénic wind) by using PSP observations from E6–E9 and hybrid-kinetic simulations. Our main results and conclusions are summarized as follows:

  • 1.  
    At large (MHD) scales, the Alfvénic wind is much less compressible than the non-Alfvénic wind (see Figure 2). However, our PVI and MAG-PVI analyses show that the hottest protons are localized at kinetic-scale compressible structures in both types of wind (see Figure 3).
  • 2.  
    There is a statistical correlation between the highest proton total temperature and coherent structures (quantified by PVI values), consistent with previous studies (Qudsi et al. 2020; Sioulas et al. 2022b). Furthermore, analysis of the temperature distributions suggests that the Alfvénic wind undergoes a preferential enhancement of T as larger PVI values are considered, whereas the non-Alfvénic wind shows a preferential enhancement of T (see Figure 3).
  • 3.  
    Proton beams are ubiquitous in both types of wind (leading to an average anisotropy T > T). However, local kinetic features of proton VDFs differ in the two winds (see Figures 5 and 6). The non-Alfvénic wind is characterized by "hot" beams (${T}_{\perp ,b}/{T}_{\perp ,c}^{* }\gtrsim 2$) resembling the "hammerhead" distributions. The Alfvénic wind is characterized by "colder" beams (${T}_{\perp ,b}/{T}_{\perp ,c}^{* }\lesssim 1$).
  • 4.  
    We find some similarities between the hybrid-kinetic simulations and in situ measurements despite the limitations due to the reduced geometry adopted and the different resolution between simulations and data. The field-aligned proton beams that develop in our simulations display distinct features, supporting the idea that proton beams in Alfvénic and non-Alfvénic wind have different properties and different origins. Simulations suggest that the development of a perpendicular cascade, favored in balanced turbulence, allows a preferential relative enhancement of T, and the formation of hot beams via nonlinear dynamics and reconnection. By contrast, cold field-aligned beams are favored by Alfvén wave steepening (see Figure 9 and Figure 10).

Additionally, we have shown for the first time 3D proton VDFs from PSP displaying non-Maxwellian and nongyrotropic features near discontinuities, warranting a more general approach to fit VDFs than the widely adopted fit to bi-Maxwellians. Furthermore, non-Maxwellian and nongyrotropic proton VDFs around discontinuities/intermittent structures are found in both winds, confirming that nonlinearities and strong deviations from nonthermal distributions are intrinsically related in collisionless plasmas (Valentini et al. 2014), resulting in a universal heating channel regardless of the Alfvénic properties of the solar wind. In conclusion, this work contributes to understanding the distinctive role of coherent structures in heating collisionless plasmas. To gain a comprehensive understanding, our simulation results should be extended to include three-dimensional effects and encompass a broader range of initial Alfvénic properties rather than the extreme cases considered here (σc ≃ 0 and σc ≃ −1). This will be the subject of future investigations.

Acknowledgments

This research was supported by NASA grant No. 80NSSC18K1211 and NSF CAREER award 2141564. We acknowledge the Parker Solar Probe (PSP) mission for the use of the data publicly available at https://spdf.gsfc.nasa.gov/, the NASA Space Physics Data Facility, and the Texas Advanced Computing Center (TACC) at The University of Texas at Austin for providing HPC resources that have contributed to the research results reported within this paper. Simulations have been run on the Frontera supercomputer http://www.tacc.utexas.edu. J.L.V. acknowledges support from NASA PSP-GI 80NSSC23K0208 and NASA LWS 80NSSC22K1014.

Appendix A: VDF Interpolation Method

To visualize the proton velocity distribution in velocity space we mapped the initial 3D energy distribution function from energy, elevation, and azimuthal angle coordinates (E = 1/2mv2, θ, ϕ) into velocity-space coordinates in the instrument frame (vx , vy , vz ) using transformation from spherical to Cartesian coordinates:

with v the amplitude of the velocity field. The vz component is defined as the rotational symmetry axis of the instrument, the vx direction points toward the Sun, and the vy component is then defined orthogonal to them. However, the results presented in Section 2 are shown in the field-aligned frame, which was obtained by applying a rotation matrix to the proton velocity vectors in the instrument frame:

The above expression is obtained by using the Euler–Rodriguez formula that corresponds to a rotation by an angle ψ about an axis defined by the unit vector $\hat{{\boldsymbol{k}}}$. Here, we chose this axis to be defined as $\hat{{\boldsymbol{k}}}=[0,-\hat{{b}_{z}},\hat{{b}_{y}}]$ and ψ the angle between the instrument frame pointing to the Sun and the magnetic field vector $\hat{{\boldsymbol{b}}}$. Further details can be found in Woodham et al. (2021). To obtain a reduced representation (1D/2D) of the integrated velocity distribution function that accurately accounts for the different velocity ranges at different planes, we used 3D linear interpolation on a mesh with a fixed velocity range. This method ensures the proper integration along the preferred component that appropriately adds up the number of counts in each velocity bin. The interpolation method is commonly used to approximate a function from a set of discrete data and it returns a callable function that can be used to evaluate the interpolated function at any point within a defined interval. We employed the cubic interpolation method from the griddata function in the scipy.interpolate module (Virtanen et al. 2020). For illustration, Figure A1 presents a 3D render of the original and the interpolated proton VDF obtained using the interpolation code (Gonzalez 2023).

Figure A1.

Figure A1. Illustration of the interpolation method. Several VDFs that correspond to different measurements in azimuthal angles (represented by the range of colors from yellow to purple) and for different elevation angles (arranged from left to right). For each VDF, the original signal (solid lines) and the interpolated signal (dashed lines) are shown.

Standard image High-resolution image

Appendix B: Fitting Method

To obtain additional information on the proton VDFs, the SPAN-I L2 sf00 data product was fit to a sum of two 3D bi-Maxwellians, one for the "core" and one for "beam," following a similar procedure described in Verniero et al. (2020). Note that the "beam" was constrained to lie parallel to the mean magnetic field and the "core" was labeled as the peak in phase-space density. Since SPAN-I is partially obstructed by PSP's thermal protection shield, only partial moments of the VDF are obtained. We therefore only include fits that are at least 80% in the field of view (FOV). We quantify this by first computing moments of the distribution from a VDF reconstructed from the fitting parameters. Next, we take the sum of the fitted beam and core densities and divide by the reconstructed moment density. To exceed an FOV threshold of 80%, this number must exceed 0.8.

Figure A2.

Figure A2. Global evolution of the simulations. The three-panel plot on the left shows the temporal evolution of the mean amplitude of magnetic and velocity field fluctuations, respectively, and the bottom panel shows the temporal evolution of mean parallel and perpendicular proton temperatures. Results are for the turbulence (solid lines) and the wave-like (dashed lines) simulations. The two-column plot on the right shows the temporal evolution of the proton velocity distribution function averaged over the entire simulation domain. The left column presents results for the turbulence simulation, while the right column shows results for the wave simulation.

Standard image High-resolution image

Figure A3. Animation of Turbulence simulation. It shows the time evolution of the Jz component of the current density in the simulations. The black dot represents the location where we obtained the time series presented on the left column of Figure 8. The animation shows the z-component of the current density Jz for the turbulence simulation during 300 ${\omega }_{{ci}}^{-1}$. The real-time duration of the animation is 90 s.

(An animation of this figure is available.)

Video Standard image High-resolution image

Figure A4. Animation of the Wave simulation. It shows the time evolution of the Jz component of the current density in the simulations. The black dot represents the location where we obtained the time series presented on the right column of Figure 8. The animation shows the z-component of the current density Jz for the wave simulation during 300 ${\omega }_{{ci}}^{-1}$. The real-time duration of the animation is 90 s.

(An animation of this figure is available.)

Video Standard image High-resolution image

Appendix C: Supplemental Material

Additional figures and animations (Figures A2-A4) supporting the results presented in the manuscript are shown in this appendix.

Footnotes

  • 4  

    Additional plots and movies showing the global dynamics and overall proton heating in the simulations can be found in Appendix C.

Please wait… references are loading.
10.3847/1538-4357/ad1be5