We present analysis of about one hundred bipolar structures of positive polarity identified in ten quasi-perpendicular crossings of the Earth's bow shock by the Magnetospheric Multiscale spacecraft. The bipolar structures have amplitudes up to a few tenths of local electron temperature, spatial scales of a few local Debye lengths, and plasma frame speeds of the order of local ion-acoustic speed. We argue that the bipolar structures of positive polarity are slow electron holes, rather than ion-acoustic solitons. The electron holes are typically above the transverse instability threshold, which we argue is due to high values of the ratio ω p e / ω c e between electron plasma and cyclotron frequencies. We speculate that the transverse instability can strongly limit the lifetime of the electron holes, whose amplitude is above a certain threshold, which is only a few mV/m in the Earth's bow shock. We suggest that electron surfing acceleration by large-amplitude electron holes reported in numerical simulations of high-Mach number shocks might not be as efficient in realistic shocks, because the transverse instability strongly limits the lifetime of large-amplitude electron holes at ω p e / ω c e values typical of collisionless shocks in nature.

The Earth's bow shock is a natural laboratory for in situ analysis of plasma processes in supercritical collisionless shock waves,2,31,72,73 in which experimental studies are continuously stimulated by remote observations5,30,63 and simulations1,11,23 of astrophysical shocks. Numerical simulations suggested that one of the viable mechanisms of thermal electron acceleration in high-Mach number astrophysical shocks can be electron surfing acceleration by electrostatic waves produced in a shock transition region.1,12,23,36,55 The electrostatic waves originally proposed to be efficient in electron surfing acceleration are bipolar electrostatic structures with positive polarity of electrostatic potential, known as electron holes, produced in a nonlinear stage of Buneman instability.23,55,56 This instability is excited in the foot region due to the drift between incoming ions and electrons, which appears to compensate the current of ions reflected by the shock. The surfing acceleration mechanism consists of electron trapping into the bipolar structure and electron acceleration by the motional electric field.23,56 The relevance of a specific process revealed by numerical simulations to realistic shocks is not always obvious, because numerical simulations typically cannot be performed at realistic geometry and background plasma parameters, such as ion-to-electron mass ratio and the ratio ω p e / ω c e between electron plasma and cyclotron frequencies, critically affecting nonlinear development of plasma instabilities.9,22,31,40,57 In this study, we present observations of bipolar electrostatic structures of positive polarity in the Earth's bow shock and discuss the efficiency of these structures in electron surfing acceleration at background plasma parameters typical of realistic shocks.

The presence of large-amplitude (up to a few hundreds mV/m) bipolar electrostatic structures in the Earth's bow shock was originally revealed by electric field measurements aboard Wind spacecraft.3,4 The bipolar structures were interpreted in terms of electron holes, hypothetically produced by electron-streaming instabilities, although the polarity of their electrostatic potential could not be determined using Wind spacecraft measurements. Surprisingly, the analysis of a few tens of bipolar structures observed in the Earth's bow shock aboard Cluster and Magnetospheric Multiscale (MMS) spacecraft revealed only bipolar structures of negative polarity,6,21,65,69 which questioned the presence of electron holes in the Earth's bow shock. Quite recently, the analysis of more than 2100 bipolar structures observed aboard Magnetospheric Multiscale spacecraft in ten quasi-perpendicular Earth's bow shock crossings showed that about 95% of the bipolar structures have negative polarity of the electrostatic potential.70 These structures were interpreted in terms of ion holes,70 that is, electrostatic structures produced in a nonlinear stage of ion-streaming instabilities.8,13,20,39,67,68 About 5% of the bipolar structures have positive polarity of the electrostatic potential, but no detailed analysis and interpretation of these structures were performed. The low occurrence rate explains the absence of bipolar structures of positive polarity in the previous studies6,21,65,69 limited to just a few tens of bipolar structures. Of critical importance are the nature and origin of the bipolar structures of positive polarity, as well as the reasons for their low occurrence rate in the Earth's bow shock.

In this paper, we present detailed analysis of about one hundred bipolar structures of positive polarity identified, but not considered, in Ref. 70. We demonstrate that these structures have speeds on the order of ion-acoustic speed, but they are not ion-acoustic solitons and can be only slow electron holes; such electrostatic structures are often observed in the Earth's plasma sheet28,35,47 and magnetopause.19,59 Importantly, the electron holes in the Earth's bow shock violate the criterion of the transverse instability,25,45 which we argue is due to high values of ω p e / ω c e , and can restrict electron hole lifetimes to less than 10 ms. Although the origin of these structures is still elusive, the presented observations demonstrate that at high ω p e / ω c e values, typical of collisionless shocks in nature, the transverse instability substantially restricts electron hole lifetimes and, therefore, is a crucial factor potentially affecting the efficiency of electron surfing acceleration observed in one-dimensional simulations of high-Mach number shocks.23,36,55

We consider 101 bipolar structures of positive polarity identified in a dataset of more than 2100 solitary waves selected among 10 mV/m electric field fluctuations in ten quasi-perpendicular Earth's bow shock crossings.70 We use MMS measurements in burst mode, which include electron moments at 0.03s cadence and ion moments at 0.15s cadence provided by the Fast Plasma Investigation instrument,49 DC-coupled magnetic field at 128 S/s (samples per second) resolution provided by Digital and Analogue Fluxgate Magnetometers,50 AC-coupled electric field at 8192 S/s resolution provided by Axial Double Probe,17 and Spin-Plane Double Probe.34 Note that electric field measurements are also available at 65 kS/s resolution, but not continuously within a burst mode interval.

Figure 1 presents a schematic of axial and spin-plane double probes. The AC-coupled electric field at 8192 S/s resolution is computed using voltage signals V1, V2, V3, and V4 of four voltage-sensitive spherical probes mounted on tips of 60-meter antennas in the spacecraft spin plane and signals V5 and V6 of two probes mounted on tips of roughly 14.6-meter axial antennas along the spin axis. The voltages of the probes are measured with respect to spacecraft and available only at 8192 S/s resolution. The spin plane electric field components in the coordinate system related to the antennas are computed as E 12 = 1.8 · ( V 2 V 1 ) / 2 l 12 and E 34 = 1.8 · ( V 4 V 3 ) / 2 l 34 , while the axial electric field component is E 56 = 1.65 · ( V 6 V 5 ) / 2 l 56 , where l 12 = l 34 = 60 m and l 56 = 14.6 m, and the frequency response factors of spin plane and axial antennas are 1.8 and 1.65. The optimal ratio between frequency response factors of axial and spin plane antennas of about 1.65/1.8 0.9 was determined previously by minimizing the angle between the electric field and propagation direction of ion holes observed in the Earth's bow shock.70 While the ratio of the frequency response factors is around 0.9, the absolute value of the spin plane response factor may be actually around 1.35. The use of that factor instead of 1.8 would not affect any conclusions of this study though.

FIG. 1.

The schematics of voltage-sensitive probes used to compute the electric field aboard MMS spacecraft. The probes 1–4 are in the spacecraft spin plane, and the axial probes 5 and 6 are on the spacecraft spin axis. The coordinate system associated with the probes is indicated in the panel. The figure is from Wang et al.70 Reproduced with permission from Wang et al., “Electrostatic solitary waves in the Earth's bow shock: Nature, properties, lifetimes, and origin,” J. Geophys. Res.: Space Phys. 126, e29357 (2021). Copyright 2000–2022 John Wiley & Sons, Inc. or related companies.

FIG. 1.

The schematics of voltage-sensitive probes used to compute the electric field aboard MMS spacecraft. The probes 1–4 are in the spacecraft spin plane, and the axial probes 5 and 6 are on the spacecraft spin axis. The coordinate system associated with the probes is indicated in the panel. The figure is from Wang et al.70 Reproduced with permission from Wang et al., “Electrostatic solitary waves in the Earth's bow shock: Nature, properties, lifetimes, and origin,” J. Geophys. Res.: Space Phys. 126, e29357 (2021). Copyright 2000–2022 John Wiley & Sons, Inc. or related companies.

Close modal

As a bipolar structure propagates across spacecraft, its electric field induces voltage signals on the voltage-sensitive probes. Time delays between voltage signals of the opposing probes can be used to estimate velocity and, hence, spatial scale of the bipolar structure.58,65,66,70 If the time delay between voltage signals of each pair of opposing probes can be determined reliably (the correlation coefficient between voltage signals exceeds 0.75 and the time delay exceeds 0.06 ms, that is half of the electric field temporal resolution), then the speed Vs of the bipolar structure in the spacecraft rest frame and its direction of propagation k = ( k 12 , k 34 , k 56 ) can be determined as follows:65,66

k i j = V s Δ t i j / l i j , V s = ( Δ t 12 2 / l 12 2 + Δ t 34 2 / l 34 2 + Δ t 56 2 / l 56 2 ) 1 / 2 .
(1)

A solid indicator of a reliable velocity estimate is the smallness of angle θkE between propagation direction and electric field polarization direction, since for locally planar electrostatic structures, these directions must coincide. If the time delay between voltage signals of at least one pair of opposing probes cannot be determined reliably, we assume that k is parallel to unit vector E ̂ along the electric field polarization direction and use reliable time delays to estimate the speed as follows:64,66,70

V s = E ̂ i j l i j / Δ t i j .
(2)

Importantly, to accurately compute E ̂ , we have to compensate amplitude reduction and widening of electric field signals Eij, caused by the fact that the spatial scale of bipolar structures is typically comparable with spatial separation between opposing spin plane probes. The details of the corresponding correction procedure are described in Ref. 70.

Figure 2 presents an overview of all ten Earth's bow shock crossings, where 101 bipolar structures of positive polarity were collected. The left panels present magnetic field magnitude profiles across the Earth's bow shock crossings along with the moments of occurrence of the bipolar structures of positive polarity. The fast-magnetosonic Mach number and shock obliqueness are indicated in the panels. The bipolar structures of positive polarity are predominantly observed in the downstream region. More than 70% of the positive polarity structures collected in the ten shocks were observed in three of them crossed on November 2, 2017, and characterized by the lowest fast magnetosonic Mach numbers, MF = 2.5, 2.7, and 2.8. In two of these shocks, the fraction of positive polarity structures among all the selected bipolar structures exceeds 25%, which is significantly higher than the overall average value of 5%. The right panels present a close view at about 10-ms intervals of parallel electric field waveforms measured at 8192 S/s or, if available, at 65 kS/s resolution. These panels show that the bipolar structures have typical peak-to-peak temporal widths of about 1 ms and peak-to-peak parallel electric field amplitudes up to about 100 mV/m. To address the nature of the bipolar structures of positive polarity, we estimated their velocity and other related properties using the interferometry analysis. The interferometry analysis for bipolar structures 1–3 highlighted in Fig. 2 is demonstrated below, and the revealed properties are given in Table I.

FIG. 2.

The overview of ten quasi-perpendicular Earth's bow shock crossings by MMS spacecraft, where the dataset of 101 bipolar structures of positive polarity was collected. The left panels present the magnetic field magnitude profiles (measured at 128 S/s resolution) along with the occurrence moments of the positive polarity structures (green vertical lines). The fast magnetosonic Mach number MF and the angle θBn between shock normal and upstream magnetic field are indicated in the panels. The right panels present waveforms of parallel (magnetic field-aligned) electric field E | | measured at 8192 S/s or, if available, 65 kS/s resolution over about 10-ms intervals indicated in the left panels. The detailed analysis and properties of bipolar structures 1–3 highlighted in the right panels are presented in Figs. 3–5 and Table I.

FIG. 2.

The overview of ten quasi-perpendicular Earth's bow shock crossings by MMS spacecraft, where the dataset of 101 bipolar structures of positive polarity was collected. The left panels present the magnetic field magnitude profiles (measured at 128 S/s resolution) along with the occurrence moments of the positive polarity structures (green vertical lines). The fast magnetosonic Mach number MF and the angle θBn between shock normal and upstream magnetic field are indicated in the panels. The right panels present waveforms of parallel (magnetic field-aligned) electric field E | | measured at 8192 S/s or, if available, 65 kS/s resolution over about 10-ms intervals indicated in the left panels. The detailed analysis and properties of bipolar structures 1–3 highlighted in the right panels are presented in Figs. 3–5 and Table I.

Close modal
TABLE I.

The properties of bipolar structures 1–3 presented in Figs. 3–5: θkB is the angle between propagation direction and local magnetic field, Vs and V s * are spacecraft frame and plasma frame speeds, cIA is local ion-acoustic speed estimated as c I A = [ ( T e + 3 T p ) / m p ] 1 / 2 , L / λ D is the spatial scale in units of local Debye length λD, e Φ 0 / T e is the amplitude of electrostatic potential in units of local electron temperature Te, and k gse is the propagation direction in the GSE system.

No. θkB Vs (km/s) V s * (km/s) cIA (km/s) L / λ D e Φ 0 / T e k gse
24 °   49  20  57  0.1  (0.19, 0.44, 0.88) 
10 °   150  40  100  0.07  (0.51, 0.77, 0.4) 
54 °   503  280  60  0.18  (0.64, −0.38, −0.68) 
No. θkB Vs (km/s) V s * (km/s) cIA (km/s) L / λ D e Φ 0 / T e k gse
24 °   49  20  57  0.1  (0.19, 0.44, 0.88) 
10 °   150  40  100  0.07  (0.51, 0.77, 0.4) 
54 °   503  280  60  0.18  (0.64, −0.38, −0.68) 

Figure 3 presents the interferometry analysis for bipolar structure 1. Panels (a)–(c) demonstrate voltage signals Vi of individual voltage-sensitive probes, measured with respect to the spacecraft, while panel (d) presents three electric field components Eij. The fact that this bipolar structure has positive polarity can be revealed by a simple analysis of E12, V1, and V2 signals. The bipolar structure propagates from probe 2 to probe 1, because signal V 2 leads signal V1, while E12 is first directed from probe 2 to probe 1 and then in the opposite direction. This means that in physical space, E12 has a divergent configuration and, hence, a positive polarity of the electrostatic potential. By analyzing voltage signals of two other pairs of opposing probes in a similar fashion, we reaffirm that the bipolar structure is indeed of positive polarity. The three electric field components in panel (d) have similar bipolar profiles, thereby indicating a locally 1D planar configuration of the bipolar structure. It is worth pointing out that signals E12 and E34 appear wider than signal E56, which is an instrumental effect arising for bipolar structures with spatial scales comparable to the length of spin plane antennas. The associated effect is reduction of the actual amplitude of these electric field components.65,70 After correcting the electric fields (Ref. 70), we use peak-to-peak values of electric field components Eij to determine unit vector E ̂ along the electric field polarization direction. For the bipolar structure in Fig. 3, we found E ̂ = ( 0.36 , 0.37 , 0.86 ) and the corrected electric field El along this direction is shown in panel (e).

FIG. 3.

The interferometry analysis of bipolar structure 1. Panels (a)–(c) present voltage signals of four probes in the spin plane (V1, V2, V3, and V4) and two probes along the spin axis (V5 and V6). The voltage signals of the probes are measured with respect to the spacecraft. The time delays Δ t i j between the voltage signals of the opposing probes (V1 and V 2 , V3 and V 4 , V5 and V 6 ) are indicated in the panels. Bipolar structure 1 exemplifies bipolar structures with reliable time delays between voltage signals of all three pairs of opposing probes (there are 12 such bipolar structures in our dataset). Panel (d) presents the electric field components in the coordinate system related to the probes (Fig. 1): E i j ( V j V i ) / l i j , where lij represent spin plane or axial antenna lengths ( l 12 = l 34 = 60 m, l 56 = 14.6 m). After correcting the electric fields (Sec. II and Ref. 70), we use peak-to-peak amplitudes of Eij to obtain unit vector E ̂ along the electric field polarization direction. Panel (e) presents the electric field El along the polarization direction and the electrostatic potential computed as Φ = E l V s d t , where V s 49 km/s is the spacecraft frame speed of the bipolar structure determined using the time delays Δ t i j . The bottom horizontal axis presents spatial coordinate x = V s d t with x =0 corresponding to the moment of El = 0. In all panels, the dots represent the actual measurements at 8192 S/s resolution, while the solid curves represent spline interpolated data.

FIG. 3.

The interferometry analysis of bipolar structure 1. Panels (a)–(c) present voltage signals of four probes in the spin plane (V1, V2, V3, and V4) and two probes along the spin axis (V5 and V6). The voltage signals of the probes are measured with respect to the spacecraft. The time delays Δ t i j between the voltage signals of the opposing probes (V1 and V 2 , V3 and V 4 , V5 and V 6 ) are indicated in the panels. Bipolar structure 1 exemplifies bipolar structures with reliable time delays between voltage signals of all three pairs of opposing probes (there are 12 such bipolar structures in our dataset). Panel (d) presents the electric field components in the coordinate system related to the probes (Fig. 1): E i j ( V j V i ) / l i j , where lij represent spin plane or axial antenna lengths ( l 12 = l 34 = 60 m, l 56 = 14.6 m). After correcting the electric fields (Sec. II and Ref. 70), we use peak-to-peak amplitudes of Eij to obtain unit vector E ̂ along the electric field polarization direction. Panel (e) presents the electric field El along the polarization direction and the electrostatic potential computed as Φ = E l V s d t , where V s 49 km/s is the spacecraft frame speed of the bipolar structure determined using the time delays Δ t i j . The bottom horizontal axis presents spatial coordinate x = V s d t with x =0 corresponding to the moment of El = 0. In all panels, the dots represent the actual measurements at 8192 S/s resolution, while the solid curves represent spline interpolated data.

Close modal

The obvious correlations between signals Vi and V j of each pair of the opposing probes enable us to determine corresponding time delays Δ t i j , which are indicated in panels (a)–(c). These time delays are used to estimate the speed of the structure Vs in the spacecraft rest frame, and the direction of propagation k = ( k 12 , k 34 , k 56 ) using Eq. (1). We found propagation direction k = ( 0.32 , 0.34 , 0.88 ) and speed V s 49 km/s. The angle between vectors k and E ̂ is rather small, θ k E 3.2 ° , proving that the bipolar structure indeed has a locally 1D planar configuration. Interestingly, the bipolar structure propagates oblique to the local magnetic field at angle θ k B 24 ° . We use the estimated spacecraft frame velocity Vs to translate the temporal electric field profiles into spatial profiles and to compute the electrostatic potential of the bipolar structure. The spatial coordinate shown in Fig. 3 was computed as x = V s d t with x =0 corresponding to El = 0. The electrostatic potential shown in panel (d) is computed as Φ = E l V s d t . We determine the typical spatial scale L as half of the distance between peaks of electric field El (the actual spatial width is about four times larger). Panel (d) shows that the spatial scale of the bipolar structure is L 24 m or about 4 λ D in units of local Debye length. The amplitude of the electrostatic potential is Φ 0 3.5 V or about 0.1 T e in units of local electron temperature.

The velocity of the bipolar structure in the plasma rest frame is determined as V s * = V s k · U p , where U p is the proton bulk velocity in the spacecraft rest frame measured at the moment closest to the occurrence of the bipolar structure. We found the plasma frame speed V s * 20 km/s, which is of the order of local ion-acoustic speed, c IA 57 km/s. The ion-acoustic speed was estimated as c IA = ( ( T e + 3 T p ) / m p ) 1 / 2 , where mp is the proton mass, Te is the local electron temperature, Tp is the temperature of incoming protons in the upstream region (Wind spacecraft data71), which is considered to be a proxy of the local temperature of incoming protons and cannot be measured aboard MMS spacecraft. Note that this is only an order of magnitude estimate of the actual ion-acoustic speed, because the ion distribution functions are typically not Maxwellian (Sec. III). Thus, the bipolar structure 1 is a Debye-scale positive polarity structure propagating relatively slowly (much slower than thermal electrons) quasi-parallel to local magnetic field.

We identified only 12 bipolar structures with all three time delays Δ t i j determined reliably. For all these bipolar structures, the angle between propagation direction k and electric field E ̂ was within 15°. For 39 bipolar structures in our dataset, we could reliably determine two time delays, while only one time delay could be determined for the remaining 50 bipolar structures. For bipolar structures with one and two reliable time delays, we assumed k parallel to E ̂ and used reliable time delays Δ t i j to estimate Vs using Eq. (2). The difference of 15 ° between k and E ̂ will be used later to estimate uncertainties of our estimates of Vs and other related parameters (spatial scale and amplitude), resulting from our assumption that k is parallel to E ̂ .

Figure 4 presents the interferometry analysis of bipolar structure 2. The voltage signals of the opposing spin plane probes are well correlated and the corresponding time delays, Δ t 12 and Δ t 34 , are indicated in panels (a) and (b). The bipolar structure propagates from probe 2 to probe 1 and from probe 4 to probe 3, while both E12 and E34 are first positive and then negative. This implies that E12 and E34 have divergent configuration and, hence, the bipolar structure has a positive electrostatic potential. The voltage signals V5 and V 6 of the axial probes are not very well correlated, but the electric field E56 has a bipolar profile similar to those of E12 and E34. The similarity of the electric field profiles indicates that the bipolar structure has locally 1D planar configuration. After correcting the electric fields, we found the unit vector along the polarization direction, E ̂ = ( 0.51 , 0.77 , 0.4 ) . Assuming k parallel to E ̂ , we obtain two independent velocity estimates in the spacecraft rest frame, V s ( 1 ) = E ̂ 12 l 12 / Δ t 12 144 km/s and V s ( 2 ) = E ̂ 34 l 34 / Δ t 34 155 km/s, and the averaged value V s 150 km/s. The consistency of independent velocity estimates confirms that the bipolar structure indeed has locally 1D planar configuration. The bipolar structure propagates quasi-parallel to local magnetic field, θ k B 10 ° . Panel (e) shows that the bipolar structure has a positive potential with amplitude Φ 0 8 V 0.07 T e and spatial scale L 48 m 4 λ D . The plasma frame speed V s * = V s k · U p 40 km/s is comparable with local ion-acoustic speed, c IA 100 km/s. Thus, bipolar structure 2 is a slow Debye-scale structure of positive polarity propagating quasi-parallel to local magnetic field.

FIG. 4.

The interferometry analysis of bipolar structure 2. The format of the figure is identical to that of Fig. 3. Bipolar structure 2 exemplifies bipolar structures with reliable time delays between voltage signals of only two pairs of opposing probes. There are 39 such bipolar structures in our dataset.

FIG. 4.

The interferometry analysis of bipolar structure 2. The format of the figure is identical to that of Fig. 3. Bipolar structure 2 exemplifies bipolar structures with reliable time delays between voltage signals of only two pairs of opposing probes. There are 39 such bipolar structures in our dataset.

Close modal

Figure 5 presents interferometry analysis of bipolar structure 3. The voltage signals of all opposing probes are rather well correlated, and all the electric field components Eij have similar bipolar profiles, but only the time delay between V1 and V 2 could be determined reliably. The bipolar structure propagates from probe 2 to probe 1, while E12 is first positive and then negative, indicating positive electrostatic potential. After correcting the electric fields, we found the polarization direction, E ̂ = ( 0.61 , 0.38 , 0.68 ) . Assuming k parallel to E ̂ and using the reliable time delay Δ t 12 , we estimate the spacecraft frame speed, V s = E ̂ 12 l 12 / Δ t 12 503 km/s, and reveal rather oblique propagation to local magnetic field, θ k B 54 ° . The bipolar structure has a positive potential with amplitude Φ 0 6 V 0.18 T e and spatial scale L 80 m 5 λ D . The speed in the plasma rest frame is V s * = V s k · U p 280 km/s that is a few times larger than local ion-acoustic speed, c IA 60 km/s. Thus, bipolar structure 3 is also a slow Debye-scale structure of positive polarity, but propagating oblique to local magnetic field.

FIG. 5.

The interferometry analysis of bipolar structure 3. The format of the figure is identical to that of Fig. 3. Bipolar structure 3 exemplifies bipolar structures with reliable time delays between voltage signals of only one pair of opposing probes. There are 50 such bipolar structures in our dataset.

FIG. 5.

The interferometry analysis of bipolar structure 3. The format of the figure is identical to that of Fig. 3. Bipolar structure 3 exemplifies bipolar structures with reliable time delays between voltage signals of only one pair of opposing probes. There are 50 such bipolar structures in our dataset.

Close modal

Figure 6 presents statistical distributions of amplitudes Φ 0 , spatial scales L and obliqueness θkB of all the 101 positive polarity structures. Panels (a) and (b) show that the bipolar structures have typical amplitudes within a few volts and spatial scales of 10–100 meters. Note that the electrostatic potential of the bipolar structures in their rest frame can be rather well fitted to the Gaussian model:

Φ = Φ 0 exp ( x 2 / 2 L 2 ) ,
(3)

and, therefore, the actual spatial width of the bipolar structures is about four times larger, 40–400 m. Panel (c) shows that about 85% of the bipolar structures propagate quasi-parallel, θ k B 30 ° , to local magnetic field, while about 15% of the bipolar structures propagate obliquely, θ k B 30 ° .

FIG. 6.

The distributions of various parameters of the bipolar structures of positive polarity: (a) amplitude Φ 0 of the electrostatic potential, (b) spatial scale L [see Eq. (3)], and (c) angle θkB between the propagation direction and local magnetic field.

FIG. 6.

The distributions of various parameters of the bipolar structures of positive polarity: (a) amplitude Φ 0 of the electrostatic potential, (b) spatial scale L [see Eq. (3)], and (c) angle θkB between the propagation direction and local magnetic field.

Close modal

Before presenting the results of statistical analysis, let us discuss uncertainties of the estimated parameters of the bipolar structures. The uncertainties of both amplitude Φ 0 and spatial scale L result from the uncertainty of speed Vs in the spacecraft frame. The experimental method to estimate the latter uncertainty was suggested in Ref. 70 and is based on comparison of independent speed estimates obtained using Eq. (2) for bipolar structures with at least two reliable time delays (Figs. 3 and 4). The analysis of bipolar structures of positive polarity with at least two reliable time delays (more than 60% of our dataset) showed that the two independent speed estimates are basically consistent within 30%. In what follows, we use 30% as typical uncertainty of both amplitude Φ 0 and spatial scale L of the bipolar structures.

Figure 7 presents analysis of possible correlations between various parameters of the bipolar structures. In panel (a), we present a scatterplot of normalized amplitudes e Φ 0 / T e vs normalized spatial scales L / λ D ; error bars correspond to the uncertainty of 30%. First, e Φ 0 / T e is in the range from 10 3 to 0.2, while the spatial scales are typically between λD and 10 λ D . There is no distinct correlation between these parameters, but larger spatial scales clearly correspond to larger amplitudes. In panel (b), we compare speeds of the bipolar structures in the plasma rest frame, V s * = V s k · U p , to local ion-acoustic speed estimated as c IA = ( ( T e + 3 T p ) / m p ) 1 / 2 . The plasma frame speeds are in the range from a few to a few hundred km/s. Although the speeds of the bipolar structures are clustered around local ion-acoustic speed, there is no discernible correlation between these quantities. Moreover, the spread of the speeds around local ion-acoustic speed is rather wide, such that a substantial fraction of the bipolar structures has speeds several times smaller or larger than local ion-acoustic speed. The error bars in panel (b) represent uncertainties of V s * (due to uncertainties of k · U p ), which were computed by varying k within 15 ° of the electric field polarization direction. The relative uncertainties are rather large for bipolar structures with velocities below a few tens of km/s, which is a consequence of plasma flow velocities U p being of a few hundred km/s.

FIG. 7.

(a) A scatterplot between amplitudes e Φ 0 of the bipolar structures normalized to local electron temperature Te and spatial scales L normalized to local Debye length λD, the error bars correspond to the uncertainty of 30%; (b) a scatterplot between the plasma frame speeds V s * of the bipolar structures and local ion-acoustic speed c IA = ( ( T e + 3 T p ) / m p ) 1 / 2 , where Te is local electron temperature, Tp is the proton temperature in the upstream region (Wind spacecraft data71), and mp is the proton mass. The velocities in the plasma rest frame were computed as V s * = V s k · U p , where Vs and U p are velocities of bipolar structures and of local plasma flow in the spacecraft rest frame, and k is the propagation direction of the bipolar structure. The error bars in panel (b) are uncertainties due to 15 ° uncertainty of the propagation direction k and due to the uncertainty of 30% in Vs itself.

FIG. 7.

(a) A scatterplot between amplitudes e Φ 0 of the bipolar structures normalized to local electron temperature Te and spatial scales L normalized to local Debye length λD, the error bars correspond to the uncertainty of 30%; (b) a scatterplot between the plasma frame speeds V s * of the bipolar structures and local ion-acoustic speed c IA = ( ( T e + 3 T p ) / m p ) 1 / 2 , where Te is local electron temperature, Tp is the proton temperature in the upstream region (Wind spacecraft data71), and mp is the proton mass. The velocities in the plasma rest frame were computed as V s * = V s k · U p , where Vs and U p are velocities of bipolar structures and of local plasma flow in the spacecraft rest frame, and k is the propagation direction of the bipolar structure. The error bars in panel (b) are uncertainties due to 15 ° uncertainty of the propagation direction k and due to the uncertainty of 30% in Vs itself.

Close modal

Thus, the bipolar structures of positive polarity observed in the Earth's bow shock are Debye-scale structures propagating typically quasi-parallel to local magnetic field, though highly oblique structures are present too, at plasma frame speeds of the order of local ion-acoustic speed. These structures cannot be ion-acoustic solitons, because at small amplitudes theory predicts that ion-acoustic solitons must have their spatial scale strictly related to amplitude, e Φ 0 / T e ( λ D / L ) 2 (see Refs. 37, 52, and 61). It would be interesting to test the applicability of this scaling relation at the observed amplitudes using the full Sagdeev potential approach.32,33 The only other interpretation is that these bipolar structures are electron holes, purely kinetic nonlinear structures (Bernstein–Green–Kruskal modes7) whose existence is due to a depressed phase space density of electrons trapped within the positive electrostatic potential.16,24,44,54 Note though that in contrast to widely reported electron holes with speeds of a fraction of electron thermal speed (e.g., Refs. 18, 43, 48, and 60), electron holes in the Earth's bow shock are slow structures. Similar slow electron holes have been recently reported in the Earth's plasma sheet28,29,35,47 and at the Earth's magnetopause.19,59 Slow electron holes in these space plasma regions are hypothetically produced by Buneman or two-stream electron instabilities,35,46 but operation of these instabilities has not been revealed by in situ particle measurements yet. In Sec. III, we present ion and electron velocity distribution functions (VDF) associated with the slow electron holes observed in the Earth's bow shock.

Previous studies suggested that slow electron holes can be produced by Buneman or two-stream electron instabilities (see, e.g., simulations in Refs. 10 and 14 and observations in Refs. 35 and 46). In addition, Kamaletdinov et al.28 have recently shown that whatever is the origin of slow electron holes, they can avoid self-acceleration reported in simulations15,74 only if their speed is around the local minimum of a double-humped velocity distribution function of background ions (see also detailed theory in Ref. 26). Otherwise, the interaction with ions accelerates an initially slow electron–hole to velocities much larger than ion thermal speed.27 To address the origin and stability of speed of the slow electron holes in the Earth's bow shock, we inspected electron and ion velocity distribution functions (VDF) observed around them. Since the electron VDFs have 30-ms resolution and the typical electron gyroperiod is shorter than a few milliseconds, we present only gyrophase-averaged electron VDFs. The ion VDFs have 150 ms resolution (much smaller than typical ion gyroperiod), and therefore, a more detailed analysis of ion VDFs is presented.

Figure 8 demonstrates electron VDFs in the spacecraft rest frame observed around bipolar structures 1–3. We present phase space densities of electrons averaged over pitch angles in the ranges ( 0 ° , 15 ° ) , ( 75 ° , 105 ° ) , and ( 165 ° , 180 ° ) . For bipolar structures 1 and 2, we can see electron beams at energies of a few tens of eV propagating parallel or anti-parallel to local magnetic field. We inspected similar phase space densities measured around all other slow electron holes in our dataset and found electron beams for about 25% of them. It turned out however that instabilities potentially excited by these beams are unlikely to result in formation of the observed electron holes, because the electron holes typically, though not always, propagated opposite to these electron beams. For about 75% of the electron holes, we did not observe any electron beams, as in the case of bipolar structure 3.

FIG. 8.

The electron distribution functions measured at 30-ms resolution at the moments closest to the observations of bipolar structures 1–3 (Figs. 3–5). Panels (a)–(c) present averaged phase space densities (PSD) of electrons with pitch angles within ( 0 ° , 15 ° ) , ( 75 ° , 105 ° ) , and ( 165 ° , 180 ° ) , labeled, respectively, as “ǁ,” “ ,” and “ǁ.”

FIG. 8.

The electron distribution functions measured at 30-ms resolution at the moments closest to the observations of bipolar structures 1–3 (Figs. 3–5). Panels (a)–(c) present averaged phase space densities (PSD) of electrons with pitch angles within ( 0 ° , 15 ° ) , ( 75 ° , 105 ° ) , and ( 165 ° , 180 ° ) , labeled, respectively, as “ǁ,” “ ,” and “ǁ.”

Close modal

Figure 9 presents ion VDFs in the spacecraft rest frame associated with bipolar structures 1–3. All the distribution functions are in the spacecraft rest frame and in the Geocentric Solar Ecliptic (GSE) coordinate system, whose z axis is perpendicular to the ecliptic plane, while the x axis is directed toward the Sun. Note that the coordinate system related to electric field antennas (Fig. 1) differs from the GSE system by a rotation around the z axis. The propagation direction of bipolar structure 1 in the GSE coordinate system is almost within the yz plane, k gse = ( 0.19 , 0.44 , 0.88 ) . Panel (a) presents a reduced 2D distribution function of background ions, VDF d v x , and indicates the projection of the velocity of bipolar structure 1 onto the v y v z plane. Also, panel (d) presents a 1D ion distribution function reduced over two velocities perpendicular to propagation direction k gse . We can see that bipolar structure 1 is accompanied by a single-humped 1D ion distribution and its speed is at a positive slope of this distribution. Panels (b) and (e) present similarly computed reduced ion distribution functions for bipolar structure 2, whose propagation direction was predominantly in the xy plane, k gse = ( 0.51 , 0.77 , 0.40 ) . The 1D reduced distribution is clearly double-humped with about 300 km/s separation between the two beam-like populations. The speed of bipolar structure 2 is at the local minimum of the double-humped distribution that is exactly what is required for slow electron holes to avoid acceleration due to interaction with ions.26,28 Panels (c) and (f) present ion distributions associated with bipolar structure 3, whose propagation direction was predominantly in the xz plane, k gse = ( 0.64 , 0.38 , 0.68 ) . The 1D ion distribution is double-humped with two beam-like populations separated by about 1000 km/s; the speed of bipolar structure 3 is between the beam-like populations, but much closer to the bulk velocity of one of them.

FIG. 9.

The ion distribution functions measured at 150-ms resolution at the moments closest to the observations of bipolar structures 1–3. All velocities are given in the spacecraft rest frame and GSE coordinate system. Panels (a)–(c) present 2D reduced ion distributions: VDF d v x , VDF d v z , and VDF d v y . The reduction direction was chosen based on the propagation direction k gse of a bipolar structure (see Table I and Sec. III for details). Panels (d)–(f) present 1D reduced distributions, where the reduction was done over two velocity components perpendicular to propagation direction k gse . Magenta arrow in panels (a)–(c) corresponds to the velocity projection of a bipolar structure onto the corresponding plane. Vertical magenta lines in panels (d)–(f) represent the speed of a bipolar structure along propagation direction k gse .

FIG. 9.

The ion distribution functions measured at 150-ms resolution at the moments closest to the observations of bipolar structures 1–3. All velocities are given in the spacecraft rest frame and GSE coordinate system. Panels (a)–(c) present 2D reduced ion distributions: VDF d v x , VDF d v z , and VDF d v y . The reduction direction was chosen based on the propagation direction k gse of a bipolar structure (see Table I and Sec. III for details). Panels (d)–(f) present 1D reduced distributions, where the reduction was done over two velocity components perpendicular to propagation direction k gse . Magenta arrow in panels (a)–(c) corresponds to the velocity projection of a bipolar structure onto the corresponding plane. Vertical magenta lines in panels (d)–(f) represent the speed of a bipolar structure along propagation direction k gse .

Close modal

We inspected ion VDFs measured around all the slow electron holes in our dataset and found that about 15% of them are double-humped, which look similar to those presented in panels (b) and (e) and (c) and (f). The speed of the corresponding electron holes is either at the local minimum or in between the beam-like populations. However, 85% of the electron holes are associated with single-humped ion distributions similar to that in panels (a) and (d).

We have shown that bipolar structures of positive polarity in the Earth's bow shock are Debye-scale structures with plasma frame speeds of the order of local ion-acoustic speed. The electrostatic structures are most likely slow electron holes, because the scaling relation between spatial scale and amplitude, expected for small-amplitude ion-acoustic solitons, is not observed [Fig. 7(a)]. Note though that the applicability of this scaling relation at the observed amplitudes needs to be tested in the future using the full Sagdeev potential approach.32,33 Electron holes have no constraints or correlations expected for ion-acoustic solitons, because of a free parameter controlling the phase space density depletion of trapped electrons.16,24,44 Slow electron holes reduce to classical ion-acoustic solitons when there is no phase space density depletion of the trapped electron population.51,61 We expect that slow electron holes are produced by Buneman10,46 or two-stream41,46 instabilities, but we could not reveal any of them in the Earth's bow shock (Sec. III). The most likely reason is that the saturation time of these instabilities is much shorter than 30-ms resolution of electron VDF measurements. Once produced, the electron holes have a finite lifetime, either because of their internal instabilities or dissipation processes. Clearly, the lifetime τL should be larger than typical bounce period of trapped electrons, τ L 2 π / ω b , where

ω b = 1 L ( e Φ 0 m e ) 1 / 2 = ω p e λ D L ( e Φ 0 T e ) 1 / 2 .
(4)

The statistical distribution of bounce periods τ b = 2 π / ω b in Fig. 10(b) shows that the lifetime of the slow electron holes in the Earth's bow shock should be at least about 1 ms. The experimental measurements also indicate that the lifetime should be at least 1 ms, because it cannot be smaller than typically observed 1 ms temporal width of the electron holes (Fig. 2). We show below that the slow electron holes in the Earth's bow shock are highly likely unstable to the transverse instability,25,45 which can determine the upper threshold on their lifetime.

FIG. 10.

Panel (a) compares typical bounce frequency ωb of electrons trapped within the bipolar structures to local electron cyclotron frequency ωce. The bounce frequency was computed using Eq. (4), and the error bars represent the uncertainty of 15%. The dashed line in panel (a) represents the transverse instability threshold, ω b ω c e . Electron holes with ω b ω c e are susceptible to the transverse instability.25,45 Panel (b) presents the distribution of bounce periods τ b = 2 π / ω b of electrons trapped within the bipolar structures.

FIG. 10.

Panel (a) compares typical bounce frequency ωb of electrons trapped within the bipolar structures to local electron cyclotron frequency ωce. The bounce frequency was computed using Eq. (4), and the error bars represent the uncertainty of 15%. The dashed line in panel (a) represents the transverse instability threshold, ω b ω c e . Electron holes with ω b ω c e are susceptible to the transverse instability.25,45 Panel (b) presents the distribution of bounce periods τ b = 2 π / ω b of electrons trapped within the bipolar structures.

Close modal

The multi-dimensional simulations of electron-streaming instabilities showed that electron holes disappear rather quickly in unmagnetized plasma,42 while a sufficiently strong background magnetic field can provide electron–hole stability for thousands of ω p e 1 (Refs. 38 and 62). The detailed analysis in Refs. 45 and 25 showed that in weakly magnetized plasma electron holes disappear due to the transverse instability, and demonstrated that for a stable propagation, the electron cyclotron frequency ωce should roughly exceed typical bounce frequency ωb of trapped electrons. Figure 10(a) shows that for almost all slow electron holes in the Earth's bow shock, we have the opposite relation, ω b ω c e , which means the electron holes are susceptible to the transverse instability. In this regime, the lifetime of the electron holes is only several bounce periods of trapped electrons, that is, τ L 10 τ b (Refs. 25, 27, and 45). This implies the slow electron holes in the Earth's bow shock cannot live longer than about 10 ms. Note that uncertainties of the bounce frequency estimates in Fig. 10(a) are within 15%, because ω b ( Φ 0 / L 2 ) L 1 / 2 and the uncertainties of spatial scales L are within 30%.

A finite transverse width of the electron holes can certainly reduce the crucial effect of the transverse instability. The transverse width exceeds the spatial scale L 10 100 m by a factor of ten at least, since electric fields transverse to the polarization direction are by a factor of ten smaller than those parallel to the polarization direction (not shown here), but does not exceed 10 km, since different MMS spacecraft being spatially separated by about 10 km do not observe the same electron holes. Even though a finite transverse width may affect the transverse instability, there are still several additional strong indications that the lifetime of the slow electron holes in the Earth's bow shock is very limited. The first evidence for a relatively short lifetime is that more than 85% of the slow electron holes are typically associated with single-humped ion distribution functions (Sec. III). The fact that the electron holes were observed as slow implies they have not had enough time to be accelerated due to interaction with ions. Since the expected acceleration timescale is comparable with the bounce period τb (Ref. 27), the electron–hole lifetime should be less than a few τb. The second evidence is that about 20% of the electron holes propagate oblique to local magnetic field, θ k B 30 ° (Fig. 6). The previous observations28,35,48,59,60 and numerical simulations38,62 reported only electron holes propagating roughly parallel to local magnetic field. In the case of a long lifetime, τ L 2 π / ω c e , we expect quasi-parallel propagation, since only in this case a depletion of the phase space density of trapped electrons, required for the electron hole existence,16,24,44 can be sustained. The oblique propagation implies the lifetime of such electron holes is not very different from the electron cyclotron period, τ L 2 π / ω c e , that is of the order of τb again [Fig. 8(a)].

In contrast to the slow electron holes in the Earth's bow shock, similar electron holes in the Earth's plasma sheet and magnetopause were found to be stable to the transverse instability.35,59 We believe the reason for this difference is due to different values of ω p e / ω c e , which is below a few tens in the previous studies,35,59 but on the order of one hundred in the Earth's bow shock. The electron holes are most likely produced by some electron-streaming instability, whose fastest growth rate γ is proportional to local electron plasma frequency, γ ω p e . The saturation of the instability occurs when the amplitude of electric field fluctuations becomes sufficiently large so that the bounce frequency of trapped electrons becomes comparable to the growth rate, that is, ω b γ (Ref. 53). Therefore, we expect that ω b / ω c e ω p e / ω c e and electron holes observed at larger values of ω p e / ω c e are more likely to be susceptible to the transverse instability. This argument is supported by Lotekar et al.,35 who showed that in the Earth's plasma sheet electron holes observed at ω p e / ω c e 10 are more likely to be around or even above the transverse instability threshold than electron holes observed at ω p e / ω c e 10 . We conclude that the slow electron holes observed in the Earth's bow shock are typically above the transverse instability threshold because ω p e / ω c e 100 and this instability may limit their lifetime to below a few electron bounce periods, τ L 10 τ b .

We should stress that, in principle, electron holes in the Earth's bow shock can have longer lifetimes if their amplitude is sufficiently small. The transverse instability criterion, ω b ω c e , requires that the electric field amplitude E0 of stable electron holes should be below a threshold given as follows:

E 0 T e e λ D L λ D ω c e 2 ω p e 2 .
(5)

Assuming typical background parameters, T e 10 100 eV, λ D 10 m, and ω p e / ω c e 100 , we find that Debye-scale electron holes can be stable only if they have amplitudes below a few mV/m. The bipolar structures considered in Ref. 70 and in this paper were selected among electric field fluctuations with amplitudes  10 mV/m. Thus, we cannot rule out that there are plenty of small-amplitude electron holes, because their lifetime is not strongly restricted by the transverse instability.

What are the implications of all these findings for the electron surfing acceleration in collisionless shock waves? Although the Earth's bow shock has Mach numbers much smaller than those of typically simulated astrophysical shocks, the in situ observations can still contribute to understanding of various processes revealed in simulations. The early 1D simulations showed that large-amplitude electron holes can be produced in the shock transition region of a high Mach number shock by Buneman instability and can result in efficient electron acceleration via the surfing mechanism.23,36,55 The larger amplitudes of these electron holes facilitate the efficiency of electron acceleration. The transverse instability cannot develop in 1D simulations and, thus, does not restrict the efficiency of the acceleration process. By contrast, the observations in the Earth's bow shock demonstrate that the lifetime of large-amplitude electron holes can be limited by the transverse instability. Under realistic values of parameter ω p e / ω c e , only rather small-amplitude electron holes may have lifetimes longer than a few bounce periods. Since the transverse instability is a multi-dimensional effect, 1D simulations overestimate the contribution of large-amplitude electron holes to the surfing acceleration. Multi-dimensional simulations may overestimate their contributions too, because the threshold on electron hole amplitude depends on ω p e / ω c e , whose values in realistic shocks are typically much larger than in simulations. Based on the observations in the Earth's bow shock, we conclude that electron holes might be not that efficient in electron acceleration in realistic shocks, because the transverse instability strongly limits the lifetime of large-amplitude electron holes at realistic ω p e / ω c e values.

The work of S.R.K. (observation and data processing, visualization) and E.V.Y. (theoretical interpretation) was supported by the Russian Science Foundation through Grant No. 19-12-00313. The work of F.S.M., I.Y.V., and R.W. was supported by NASA MMS Guest Investigator Grant No. 80NSSC21K0730. The work of A.V.A. was supported by NASA Heliophysics Supporting Research Grant No. 80NSSC20K1325. I.Y.V. also thanks the International Space Science Institute (ISSI), Bern, Switzerland for support. We would like to thank Ian Hutchinson, Konrad Steinvall, and Yuri Khotyaintsev for their valuable contributions to this study.

The authors have no conflicts to disclose.

Sergey Railievich Kamaletdinov: Formal analysis (lead); Investigation (equal); Software (lead); Visualization (equal). Ivan Yurievich Vasko: Conceptualization (lead); Writing – original draft (equal); Writing – review & editing (lead). Rachel Wang: Data curation (supporting); Writing – review & editing (supporting). Anton Artemyev: Data curation (supporting); Writing – review & editing (supporting). Egor Yushkov: Writing – review & editing (supporting). Forrest Mozer: Writing – review & editing (supporting).

The data that support the findings of this study are openly available in MMS data repository, Ref. 75.

1.
T.
Amano
and
M.
Hoshino
, “
Nonlinear evolution of Buneman instability and its implication for electron acceleration in high Mach number collisionless perpendicular shocks
,”
Phys. Plasmas
16
,
102901
(
2009
).
2.
T.
Amano
,
T.
Katou
,
N.
Kitamura
,
M.
Oka
,
Y.
Matsumoto
,
M.
Hoshino
,
Y.
Saito
,
S.
Yokota
,
B. L.
Giles
,
W. R.
Paterson
,
C. T.
Russell
,
O.
Le Contel
,
R. E.
Ergun
,
P. A.
Lindqvist
,
D. L.
Turner
,
J. F.
Fennell
, and
J. B.
Blake
, “
Observational evidence for stochastic shock drift acceleration of electrons at the Earth's bow shock
,”
Phys. Rev. Lett.
124
,
065101
065110
(
2020
).
3.
S. D.
Bale
,
A.
Hull
,
D. E.
Larson
,
R. P.
Lin
,
L.
Muschietti
,
P. J.
Kellogg
,
K.
Goetz
, and
S. J.
Monson
, “
Electrostatic turbulence and Debye-scale structures associated with electron thermalization at collisionless shocks
,”
Astrophys. J. Lett.
575
,
L25
L28
(
2002
).
4.
S. D.
Bale
,
P. J.
Kellogg
,
D. E.
Larsen
,
R. P.
Lin
,
K.
Goetz
, and
R. P.
Lepping
, “
Bipolar electrostatic structures in the shock transition region: Evidence of electron phase space holes
,”
Geophys. Res. Lett.
25
,
2929
2932
, (
1998
).
5.
A.
Bamba
,
R.
Yamazaki
,
M.
Ueno
, and
K.
Koyama
, “
Small-scale structure of the SN 1006 shock with Chandra observations
,”
Astrophys. J.
589
,
827
837
(
2003
).
6.
R.
Behlke
,
M.
André
,
S. D.
Bale
,
J. S.
Pickett
,
C. A.
Cattell
,
E. A.
Lucek
, and
A.
Balogh
, “
Solitary structures associated with short large-amplitude magnetic structures (SLAMS) upstream of the Earth's quasi-parallel bow shock
,”
Geophys. Res. Lett.
31
,
L16805
, (
2004
).
7.
I. B.
Bernstein
,
J. M.
Greene
, and
M. D.
Kruskal
, “
Exact nonlinear plasma oscillations
,”
Phys. Rev.
108
,
546
550
(
1957
).
8.
S.
Børve
,
H. L.
Pécseli
, and
J.
Trulsen
, “
Ion phase-space vortices in 2.5-dimensional simulations
,”
J. Plasma Phys.
65
,
107
129
(
2001
).
9.
A.
Bret
and
M. E.
Dieckmann
, “
How large can the electron to proton mass ratio be in particle-in-cell simulations of unstable systems?
,”
Phys. Plasmas
17
,
032109
(
2010
).
10.
J.
Büchner
and
N.
Elkina
, “
Anomalous resistivity of current-driven isothermal plasmas due to phase space structuring
,”
Phys. Plasmas
13
,
082304
(
2006
).
11.
P. J.
Cargill
and
K.
Papadopoulos
, “
A mechanism for strong shock electron heating in supernova remnants
,”
Astrophys. J. Lett.
329
,
L29
(
1988
).
12.
M. E.
Dieckmann
,
S. C.
Chapman
,
K. G.
McClements
,
R. O.
Dendy
, and
L. O.
Drury
, “
Electron acceleration due to high frequency instabilities at supernova remnant shocks
,”
Astron. Astrophys.
356
,
377
388
(
2000
).
13.
M. E.
Dieckmann
,
G.
Sarri
,
D.
Doria
,
M.
Pohl
, and
M.
Borghesi
, “
Modification of the formation of high-Mach number electrostatic shock-like structures by the ion acoustic instability
,”
Phys. Plasmas
20
,
102112
(
2013
).
14.
J. F.
Drake
,
M.
Swisdak
,
C.
Cattell
,
M. A.
Shay
,
B. N.
Rogers
, and
A.
Zeiler
, “
Formation of electron holes and particle energization during magnetic reconnection
,”
Science
299
,
873
877
(
2003
).
15.
B.
Eliasson
and
P. K.
Shukla
, “
Dynamics of electron holes in an electron oxygen-ion plasma
,”
Phys. Rev. Lett.
93
,
045001
(
2004
).
16.
B.
Eliasson
and
P. K.
Shukla
, “
Formation and dynamics of coherent structures involving phase-space vortices in plasmas
,”
Phys. Rep.
422
,
225
290
(
2006
).
17.
R. E.
Ergun
,
S.
Tucker
,
J.
Westfall
,
K. A.
Goodrich
,
D. M.
Malaspina
,
D.
Summers
,
J.
Wallace
,
M.
Karlsson
,
J.
Mack
,
N.
Brennan
,
B.
Pyke
,
P.
Withnell
,
R.
Torbert
,
J.
Macri
,
D.
Rau
,
I.
Dors
,
J.
Needell
,
P. A.
Lindqvist
,
G.
Olsson
, and
C. M.
Cully
, “
The axial double probe and fields signal processing for the MMS mission
,”
Space Sci. Rev.
199
,
167
188
(
2016
).
18.
W.
Fox
,
M.
Porkolab
,
J.
Egedal
,
N.
Katz
, and
A.
Le
, “
Observations of electron phase-space holes driven during magnetic reconnection in a laboratory plasma
,”
Phys. Plasmas
19
,
032118
(
2012
).
19.
D. B.
Graham
,
Y. V.
Khotyaintsev
,
A.
Vaivads
, and
M.
André
, “
Electrostatic solitary waves with distinct speeds associated with asymmetric reconnection
,”
Geophys. Res. Lett.
42
,
215
224
, (
2015
).
20.
P.
Guio
,
S.
Børve
,
L. K. S.
Daldorff
,
J. P.
Lynov
,
P.
Michelsen
,
H. L.
Pécseli
,
J. J.
Rasmussen
,
K.
Saeki
, and
J.
Trulsen
, “
Phase space vortices in collisionless plasmas
,”
Nonlinear Process. Geophys.
10
,
75
86
(
2003
).
21.
Y.
Hobara
,
S. N.
Walker
,
M.
Balikhin
,
O. A.
Pokhotelov
,
M.
Gedalin
,
V.
Krasnoselskikh
,
M.
Hayakawa
,
M.
André
,
M.
Dunlop
,
H.
RèMe
, and
A.
Fazakerley
, “
Cluster observations of electrostatic solitary waves near the Earth's bow shock
,”
J. Geophys. Res.: Space Phys.
113
,
A05211
, (
2008
).
22.
J.
Hong
,
E.
Lee
,
K.
Min
, and
G. K.
Parks
, “
Effect of ion-to-electron mass ratio on the evolution of ion beam driven instability in particle-in-cell simulations
,”
Phys. Plasmas
19
,
092111
(
2012
).
23.
M.
Hoshino
and
N.
Shimada
, “
Nonthermal electrons at high Mach number shocks: Electron shock surfing acceleration
,”
Astrophys. J.
572
,
880
887
(
2002
).
24.
I. H.
Hutchinson
, “
Electron holes in phase space: What they are and why they matter
,”
Phys. Plasmas
24
,
055601
(
2017
).
25.
I. H.
Hutchinson
, “
Kinematic mechanism of plasma electron hole transverse instability
,”
Phys. Rev. Lett.
120
,
205101
(
2018
).
26.
I. H.
Hutchinson
, “
How can slow plasma electron holes exist?
,”
Phys. Rev. E
104
,
015208
(
2021
).
27.
I. H.
Hutchinson
, “
Overstability of plasma slow electron holes
,”
J. Plasma Phys.
88
,
555880101
(
2022
).
28.
S. R.
Kamaletdinov
,
I. H.
Hutchinson
,
I. Y.
Vasko
,
A. V.
Artemyev
,
A.
Lotekar
, and
F.
Mozer
, “
Spacecraft observations and theoretical understanding of slow electron holes
,”
Phys. Rev. Lett.
127
,
165101
(
2021
).
29.
Y. V.
Khotyaintsev
,
A.
Vaivads
,
M.
André
,
M.
Fujimoto
,
A.
Retinò
, and
C. J.
Owen
, “
Observations of slow electron holes at a magnetic reconnection site
,”
Phys. Rev. Lett.
105
,
165002
(
2010
).
30.
K.
Koyama
,
R.
Petre
,
E. V.
Gotthelf
,
U.
Hwang
,
M.
Matsuura
,
M.
Ozaki
, and
S. S.
Holt
, “
Evidence for shock acceleration of high-energy electrons in the supernova remnant SN1006
,”
Nature
378
,
255
258
(
1995
).
31.
V.
Krasnoselskikh
,
M.
Balikhin
,
S. N.
Walker
,
S.
Schwartz
,
D.
Sundkvist
,
V.
Lobzin
,
M.
Gedalin
,
S. D.
Bale
,
F.
Mozer
,
J.
Soucek
,
Y.
Hobara
, and
H.
Comisel
, “
The dynamic quasiperpendicular shock: Cluster discoveries
,”
Space Sci. Rev.
178
,
535
598
(
2013
).
32.
G. S.
Lakhina
,
S.
Singh
,
R.
Rubia
, and
S.
Devanandhan
, “
Electrostatic solitary structures in space plasmas: Soliton perspective
,”
Plasma
4
,
681
731
(
2021
).
33.
G. S.
Lakhina
,
S. V.
Singh
,
R.
Rubia
, and
T.
Sreeraj
, “
A review of nonlinear fluid models for ion-and electron-acoustic solitons and double layers: Application to weak double layers and electrostatic solitary waves in the solar wind and the lunar wake
,”
Phys. Plasmas
25
,
080501
(
2018
).
34.
P. A.
Lindqvist
,
G.
Olsson
,
R. B.
Torbert
,
B.
King
,
M.
Granoff
,
D.
Rau
,
G.
Needell
,
S.
Turco
,
I.
Dors
,
P.
Beckman
,
J.
Macri
,
C.
Frost
,
J.
Salwen
,
A.
Eriksson
,
L.
Åhlén
,
Y. V.
Khotyaintsev
,
J.
Porter
,
K.
Lappalainen
,
R. E.
Ergun
,
W.
Wermeer
, and
S.
Tucker
, “
The spin-plane double probe electric field instrument for MMS
,”
Space Sci. Rev.
199
,
137
165
(
2016
).
35.
A.
Lotekar
,
I. Y.
Vasko
,
F. S.
Mozer
,
I.
Hutchinson
,
A. V.
Artemyev
,
S. D.
Bale
,
J. W.
Bonnell
,
R.
Ergun
,
B.
Giles
,
Y. V.
Khotyaintsev
,
P. A.
Lindqvist
,
C. T.
Russell
, and
R.
Strangeway
, “
Multisatellite MMS analysis of electron holes in the Earth's magnetotail: Origin, properties, velocity gap, and transverse instability
,”
J. Geophys. Res.: Space Phys.
125
,
e2020JA028066
, (
2020
).
36.
K. G.
McClements
,
M. E.
Dieckmann
,
A.
Ynnerman
,
S. C.
Chapman
, and
R. O.
Dendy
, “
Surfatron and stochastic acceleration of electrons at supernova remnant shocks
,”
Phys. Rev. Lett.
87
,
255002
(
2001
).
37.
J. F.
McKenzie
, “
The ion-acoustic soliton: A gas-dynamic viewpoint
,”
Phys. Plasmas
9
,
800
805
(
2002
).
38.
T.
Miyake
,
Y.
Omura
,
H.
Matsumoto
, and
H.
Kojima
, “
Two-dimensional computer simulations of electrostatic solitary waves observed by Geotail spacecraft
,”
J. Geophys. Res.
103
,
11841
11850
, (
1998
).
39.
Q.
Moreno
,
M. E.
Dieckmann
,
D.
Folini
,
R.
Walder
,
X.
Ribeyre
,
V. T.
Tikhonchuk
, and
E.
d'Humières
, “
Shocks and phase space vortices driven by a density jump between two clouds of electrons and protons
,”
Plasma Phys. Controlled Fusion
62
,
025022
(
2020
).
40.
Q.
Moreno
,
M. E.
Dieckmann
,
X.
Ribeyre
,
S.
Jequier
,
V. T.
Tikhonchuk
, and
E.
d'Humières
, “
Impact of the electron to ion mass ratio on unstable systems in particle-in-cell simulations
,”
Phys. Plasmas
25
,
062125
(
2018
).
41.
R. L.
Morse
and
C. W.
Nielson
, “
Numerical simulation of warm two-beam plasma
,”
Phys. Fluids
12
,
2418
2425
(
1969
).
42.
R. L.
Morse
and
C. W.
Nielson
, “
One-, two-, and three-dimensional numerical simulation of two-beam plasmas
,”
Phys. Rev. Lett.
23
,
1087
1090
(
1969
).
43.
F. S.
Mozer
,
O. V.
Agapitov
,
A.
Artemyev
,
J. F.
Drake
,
V.
Krasnoselskikh
,
S.
Lejosne
, and
I.
Vasko
, “
Time domain structures: What and where they are, what they do, and how they are made
,”
Geophys. Res. Lett.
42
,
3627
3638
, (
2015
).
44.
F. S.
Mozer
,
O. V.
Agapitov
,
B.
Giles
, and
I.
Vasko
, “
Direct observation of electron distributions inside millisecond duration electron holes
,”
Phys. Rev. Lett.
121
,
135102
(
2018
).
45.
L.
Muschietti
,
I.
Roth
,
C. W.
Carlson
, and
R. E.
Ergun
, “
Transverse instability of magnetized electron holes
,”
Phys. Rev. Lett.
85
,
94
97
(
2000
).
46.
C.
Norgren
,
M.
André
,
D. B.
Graham
,
Y. V.
Khotyaintsev
, and
A.
Vaivads
, “
Slow electron holes in multicomponent plasmas
,”
Geophys. Res. Lett.
42
,
7264
7272
, (
2015
).
47.
C.
Norgren
,
M.
André
,
A.
Vaivads
, and
Y. V.
Khotyaintsev
, “
Slow electron phase space holes: Magnetotail observations
,”
Geophys. Res. Lett.
42
,
1654
1661
, (
2015
).
48.
C.
Norgren
,
D. B.
Graham
,
M. R.
Argall
,
K.
Steinvall
,
M.
Hesse
,
Y. V.
Khotyaintsev
,
A.
Vaivads
,
P.
Tenfjord
,
D. J.
Gershman
,
P. A.
Lindqvist
,
J. L.
Burch
, and
F.
Plaschke
, “
Millisecond observations of nonlinear wave-electron interaction in electron phase space holes
,”
Phys. Plasmas
29
,
012309
(
2022
).
49.
C.
Pollock
,
T.
Moore
,
A.
Jacques
,
J.
Burch
,
U.
Gliese
,
Y.
Saito
,
T.
Omoto
,
L.
Avanov
,
A.
Barrie
,
V.
Coffey
,
J.
Dorelli
,
D.
Gershman
,
B.
Giles
,
T.
Rosnack
,
C.
Salo
,
S.
Yokota
,
M.
Adrian
,
C.
Aoustin
,
C.
Auletti
,
S.
Aung
,
V.
Bigio
,
N.
Cao
,
M.
Chandler
,
D.
Chornay
,
K.
Christian
,
G.
Clark
,
G.
Collinson
,
T.
Corris
,
A.
De Los Santos
,
R.
Devlin
,
T.
Diaz
,
T.
Dickerson
,
C.
Dickson
,
A.
Diekmann
,
F.
Diggs
,
C.
Duncan
,
A.
Figueroa-Vinas
,
C.
Firman
,
M.
Freeman
,
N.
Galassi
,
K.
Garcia
,
G.
Goodhart
,
D.
Guererro
,
J.
Hageman
,
J.
Hanley
,
E.
Hemminger
,
M.
Holland
,
M.
Hutchins
,
T.
James
,
W.
Jones
,
S.
Kreisler
,
J.
Kujawski
,
V.
Lavu
,
J.
Lobell
,
E.
LeCompte
,
A.
Lukemire
,
E.
MacDonald
,
A.
Mariano
,
T.
Mukai
,
K.
Narayanan
,
Q.
Nguyan
,
M.
Onizuka
,
W.
Paterson
,
S.
Persyn
,
B.
Piepgrass
,
F.
Cheney
,
A.
Rager
,
T.
Raghuram
,
A.
Ramil
,
L.
Reichenthal
,
H.
Rodriguez
,
J.
Rouzaud
,
A.
Rucker
,
Y.
Saito
,
M.
Samara
,
J. A.
Sauvaud
,
D.
Schuster
,
M.
Shappirio
,
K.
Shelton
,
D.
Sher
,
D.
Smith
,
K.
Smith
,
S.
Smith
,
D.
Steinfeld
,
R.
Szymkiewicz
,
K.
Tanimoto
,
J.
Taylor
,
C.
Tucker
,
K.
Tull
,
A.
Uhl
,
J.
Vloet
,
P.
Walpole
,
S.
Weidner
,
D.
White
,
G.
Winkert
,
P. S.
Yeh
, and
M.
Zeuch
, “
Fast plasma investigation for magnetospheric multiscale
,”
Space Sci. Rev.
199
,
331
406
(
2016
).
50.
C. T.
Russell
,
B. J.
Anderson
,
W.
Baumjohann
,
K. R.
Bromund
,
D.
Dearborn
,
D.
Fischer
,
G.
Le
,
H. K.
Leinweber
,
D.
Leneman
,
W.
Magnes
,
J. D.
Means
,
M. B.
Moldwin
,
R.
Nakamura
,
D.
Pierce
,
F.
Plaschke
,
K. M.
Rowe
,
J. A.
Slavin
,
R. J.
Strangeway
,
R.
Torbert
,
C.
Hagen
,
I.
Jernej
,
A.
Valavanoglou
, and
I.
Richter
, “
The magnetospheric multiscale magnetometers
,”
Space Sci. Rev.
199
,
189
256
(
2016
).
51.
K.
Saeki
and
H.
Genma
, “
Electron-hole disruption due to ion motion and formation of coupled electron hole and ion-acoustic soliton in a plasma
,”
Phys. Rev. Lett.
80
,
1224
1227
(
1998
).
52.
R. Z.
Sagdeev
, “
Cooperative phenomena and shock waves in collisionless plasmas
,”
Rev. Plasma Phys.
4
,
23
(
1966
).
53.
R. Z.
Sagdeev
and
A. A.
Galeev
,
Nonlinear Plasma Theory
(IOP Publishing,
1969
).
54.
H.
Schamel
, “
Hole equilibria in Vlasov–Poisson systems: A challenge to wave theories of ideal plasmas
,”
Phys. Plasmas
7
,
4831
4844
(
2000
).
55.
H.
Schmitz
,
S. C.
Chapman
, and
R. O.
Dendy
, “
Electron preacceleration mechanisms in the foot region of high Alfvénic Mach number shocks
,”
Astrophys. J.
579
,
327
336
(
2002
).
56.
N.
Shimada
and
M.
Hoshino
, “
The dynamics of electron-ion coupling in the shock transition region
,”
Phys. Plasmas
10
,
1113
1119
(
2003
).
57.
N.
Shimada
and
M.
Hoshino
, “
Electron heating and acceleration in the shock transition region: Background plasma parameter dependence
,”
Phys. Plasmas
11
,
1840
1849
(
2004
).
58.
K.
Steinvall
,
Y. V.
Khotyaintsev
, and
D. B.
Graham
, “
On the applicability of single-spacecraft interferometry methods using electric field probes
,”
J. Geophys. Res.: Space Phys.
127
,
e2021JA030143
, (
2022
).
59.
K.
Steinvall
,
Y. V.
Khotyaintsev
,
D. B.
Graham
,
A.
Vaivads
,
P. A.
Lindqvist
,
C. T.
Russell
, and
J. L.
Burch
, “
Multispacecraft analysis of electron holes
,”
Geophys. Res. Lett.
46
,
55
63
, (
2019
).
60.
Y.
Tong
,
I.
Vasko
,
F. S.
Mozer
,
S. D.
Bale
,
I.
Roth
,
A. V.
Artemyev
,
R.
Ergun
,
B.
Giles
,
P. A.
Lindqvist
,
C. T.
Russell
,
R.
Strangeway
, and
R. B.
Torbert
, “
Simultaneous multispacecraft probing of electron phase space holes
,”
Geophys. Res. Lett.
45
,
11,513
11,519
, (
2018
).
61.
M. Q.
Tran
, “
Ion acoustic solitons in a plasma: A review of their experimental properties and related theories
,”
Phys. Scr.
20
,
317
327
(
1979
).
62.
T.
Umeda
,
Y.
Omura
,
T.
Miyake
,
H.
Matsumoto
, and
M.
Ashour-Abdalla
, “
Nonlinear evolution of the electron two-stream instability: Two-dimensional particle simulations
,”
J. Geophys. Res.: Space Phys.
111
,
A10206
, (
2006
).
63.
R. J.
van Weeren
,
H. J. A.
Röttgering
,
M.
Brüggen
, and
M.
Hoeft
, “
Particle acceleration on megaparsec scales in a merging galaxy cluster
,”
Science
330
,
347
(
2010
).
64.
I. Y.
Vasko
,
F.
Mozer
,
S.
Bale
, and
A.
Artemyev
, “
Ion-acoustic waves in a quasi-perpendicular earth's bow shock
,”
Geophys. Res. Lett.
49
,
e2022GL098640
, (
2022
).
65.
I. Y.
Vasko
,
F. S.
Mozer
,
V. V.
Krasnoselskikh
,
A. V.
Artemyev
,
O. V.
Agapitov
,
S. D.
Bale
,
L.
Avanov
,
R.
Ergun
,
B.
Giles
,
P. A.
Lindqvist
,
C. T.
Russell
,
R.
Strangeway
, and
R.
Torbert
, “
Solitary waves across supercritical quasi-perpendicular shocks
,”
Geophys. Res. Lett.
45
,
5809
5817
, (
2018
).
66.
I. Y.
Vasko
,
R.
Wang
,
F. S.
Mozer
,
S. D.
Bale
, and
A. V.
Artemyev
, “
On the nature and origin of bipolar electrostatic structures in the Earth's bow shock
,”
Front. Phys.
8
,
156
(
2020
).
67.
A. R.
Vazsonyi
,
K.
Hara
, and
I. D.
Boyd
, “
Non-monotonic double layers and electron two-stream instabilities resulting from intermittent ion acoustic wave growth
,”
Phys. Plasmas
27
,
112303
(
2020
).
68.
R.
Wang
,
I. Y.
Vasko
,
A. V.
Artemyev
,
L. C.
Holley
,
S. R.
Kamaletdinov
,
A.
Lotekar
, and
F. S.
Mozer
, “
Multisatellite observations of ion holes in the Earth's plasma sheet
,”
Geophys. Res. Lett.
49
,
e2022GL097919
, (
2022
).
69.
R.
Wang
,
I. Y.
Vasko
,
F. S.
Mozer
,
S. D.
Bale
,
A. V.
Artemyev
,
J. W.
Bonnell
,
R.
Ergun
,
B.
Giles
,
P. A.
Lindqvist
,
C. T.
Russell
, and
R.
Strangeway
, “
Electrostatic turbulence and Debye-scale structures in collisionless shocks
,”
Astrophys. J. Lett.
889
,
L9
(
2020
).
70.
R.
Wang
,
I. Y.
Vasko
,
F. S.
Mozer
,
S. D.
Bale
,
I. V.
Kuzichev
,
A. V.
Artemyev
,
K.
Steinvall
,
R.
Ergun
,
B.
Giles
,
Y.
Khotyaintsev
,
P. A.
Lindqvist
,
C. T.
Russell
, and
R.
Strangeway
, “
Electrostatic solitary waves in the Earth's bow shock: Nature, properties, lifetimes, and origin
,”
J. Geophys. Res.: Space Phys.
126
,
e2021JA029357
, (
2021
).
71.
L. B.
Wilson, III
,
A. L.
Brosius
,
N.
Gopalswamy
,
T.
Nieves-Chinchilla
,
A.
Szabo
,
K.
Hurley
,
T.
Phan
,
J. C.
Kasper
,
N
,
Lugaz
,
I. G.
Richardson
,
C. H. K.
Chen
,
D.
Verscharen
,
R. T.
Wicks
, and
J. M.
TenBarge
, “
A quarter century of wind spacecraft discoveries
,”
Rev. Geophys.
59
,
e2020RG000714
, (
2021
).
72.
L. B.
Wilson
,
L.
Chen
, and
V.
Roytershteyn
, “
The discrepancy between simulation and observation of electric fields in collisionless shock
,”
Front. Astron. Space Sci.
7
,
592634
(
2021
).
73.
L. B.
Wilson
,
D. G.
Sibeck
,
A. W.
Breneman
,
O.
Le Contel
,
C.
Cully
,
D. L.
Turner
,
V.
Angelopoulos
, and
D. M.
Malaspina
, “
Quantified energy dissipation rates in the terrestrial bow shock: 2. Waves and dissipation
,”
J. Geophys. Res.: Space Phys.
119
,
6475
6495
, (
2014
).
74.
C.
Zhou
and
I. H.
Hutchinson
, “
Dynamics of a slow electron hole coupled to an ion-acoustic soliton
,”
Phys. Plasmas
25
,
082303
(
2018
).
75.
MMS Science Data Center, see https://lasp.colorado.edu/mms/sdc/public/ for “
Instruments, Software
.”