Multi Scale Nature of Solar Wind

Download as pdf or txt
Download as pdf or txt
You are on page 1of 136

Living Reviews in Solar Physics (2019) 16:5

https://doi.org/10.1007/s41116-019-0021-0

REVIEW ARTICLE

The multi-scale nature of the solar wind

Daniel Verscharen1,2 · Kristopher G. Klein3 · Bennett A. Maruca4

Received: 9 February 2019 / Accepted: 9 November 2019 / Published online: 9 December 2019
© The Author(s) 2019

Abstract
The solar wind is a magnetized plasma and as such exhibits collective plasma behavior
associated with its characteristic spatial and temporal scales. The characteristic length
scales include the size of the heliosphere, the collisional mean free paths of all species,
their inertial lengths, their gyration radii, and their Debye lengths. The characteristic
timescales include the expansion time, the collision times, and the periods associated
with gyration, waves, and oscillations. We review the past and present research into
the multi-scale nature of the solar wind based on in-situ spacecraft measurements and
plasma theory. We emphasize that couplings of processes across scales are important
for the global dynamics and thermodynamics of the solar wind. We describe methods
to measure in-situ properties of particles and fields. We then discuss the role of expan-
sion effects, non-equilibrium distribution functions, collisions, waves, turbulence, and
kinetic microinstabilities for the multi-scale plasma evolution.

Keywords Solar wind · Spacecraft measurements · Coulomb collisions · Plasma


waves and turbulence · Kinetic instabilities

Contents
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.1 The characteristic scales in the solar wind . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

Electronic supplementary material The online version of this article (https://doi.org/10.1007/s41116-


019-0021-0) contains supplementary material, which is available to authorized users.

B Daniel Verscharen
[email protected]

1 Mullard Space Science Laboratory, University College London, Dorking RH5 6NT, UK
2 Space Science Center, University of New Hampshire, Durham, NH 03824, USA
3 Lunar and Planetary Laboratory and Department of Planetary Sciences, University of Arizona,
Tucson, AZ 85719, USA
4 Bartol Research Institute, Department of Physics and Astronomy, University of Delaware, Newark,
DE 19716, USA

123
5 Page 2 of 136 D. Verscharen et al.

1.2 Global structure of the solar wind . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9


1.3 Categorization of solar wind . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.4 Kinetic properties of the solar wind . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.4.1 Fluid moments and fluid equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
1.4.2 Magnetohydrodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
1.4.3 Standard distributions in solar-wind physics . . . . . . . . . . . . . . . . . . . . . . . 19
1.4.4 Ion properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
1.4.5 Electron properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
1.4.6 Open questions and problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2 In-situ observations of space plasmas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.1 Overview of in-situ solar-wind missions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.2 Thermal-particle instruments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.2.1 Faraday cups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
2.2.2 Electrostatic analyzers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
2.2.3 Mass spectrometers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
2.3 Analyzing thermal-particle measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
2.3.1 Distribution-function imaging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
2.3.2 Moments analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
2.3.3 Fitting model distribution functions . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
2.4 Magnetometers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
2.4.1 Search-coil magnetometers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
2.4.2 Fluxgate magnetometers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
2.4.3 Helium magnetometers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
2.5 Electric-field measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
2.6 Multi-spacecraft techniques . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
3 Coulomb collisions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
3.1 Dimensional analysis of Coulomb collisions . . . . . . . . . . . . . . . . . . . . . . . . . . 47
3.2 Kinetic theory of collisions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
3.2.1 The collision term . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
3.2.2 The Landau collision integral . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
3.2.3 The Coulomb logarithm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
3.2.4 Rosenbluth potentials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
3.2.5 Collisional timescales . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
3.2.6 Coulomb number and collisional age . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
3.3 Observations of collisional relaxation in the solar wind . . . . . . . . . . . . . . . . . . . . 57
3.3.1 Ion collisions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
3.3.2 Electron collisions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
4 Plasma waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
4.1 Plasma waves as self-consistent electromagnetic and particle fluctuations . . . . . . . . . . . 62
4.2 Damping and dissipation mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
4.2.1 Quasilinear diffusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
4.2.2 Entropy cascade and nonlinear phase mixing . . . . . . . . . . . . . . . . . . . . . . 67
4.2.3 Stochastic heating . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
4.3 Wave types in the solar wind . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
4.3.1 Large-scale Alfvén waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
4.3.2 Kinetic Alfvén waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
4.3.3 Alfvén/ion-cyclotron waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
4.3.4 Slow modes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
4.3.5 Fast modes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
5 Plasma turbulence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
5.1 Phenomenology of plasma turbulence in the solar wind . . . . . . . . . . . . . . . . . . . . 79
5.2 Wave turbulence and its composition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
5.3 The concept of critical balance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
5.4 Advanced topics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
5.4.1 Intermittency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
5.4.2 Magnetic reconnection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
5.4.3 Anti-phase-mixing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88

123
The multi-scale nature of the solar wind Page 3 of 136 5

6 Kinetic microinstabilities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
6.1 Wave–particle instabilities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
6.1.1 Temperature anisotropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
6.1.2 Beams and heat flux . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
6.1.3 Multiple sources of free energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
6.2 Wave–wave instabilities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
6.2.1 Parametric-decay instability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
6.2.2 Limits on large-amplitude magnetic fluctuations . . . . . . . . . . . . . . . . . . . . . 101
6.3 The fluctuating-anisotropy effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
7 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
7.1 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
7.2 Future outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
7.3 Broader impact . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105

1 Introduction

The solar wind is a continuous magnetized plasma outflow that emanates from the solar
corona. This extension of the Sun’s outer atmosphere propagates through interplane-
tary space. Its existence was first conjectured based on its interaction with planetary
bodies in the solar system. Although the connection between solar activity and dis-
turbances in the Earth’s magnetic field had been established in the nineteenth century
(Sabine 1851, 1852; Hodgson 1859; Stewart 1861), the connection of these events
with “corpuscular radiation” was not made until the early twentieth century (Birke-
land 1914; Chapman 1917). The arguably first appearance of the notion of a continuous
“swarm of ions proceeding from the Sun” in the literature dates back to a footnote by
Eddington (1910) as an explanation for the observed shape of cometary tails. Later,
Hoffmeister (1943) summarized multiple comet observations and suggested that some
form of solar corpuscular radiation is responsible for the observed lag of comet ion
tails with respect to the heliocentric radius vector (for the link between solar activity
and comet tails, see also Ahnert 1943). Biermann (1951) revisited the relation between
comet tails and solar corpuscular radiation by quantifying the momentum transfer from
the solar wind to cometary ions. He especially noted that the solar radiation pressure
is insufficient to explain the observed structures (Milne 1926) and that the corpuscu-
lar radiation is more variable than the electromagnetic radiation emitted by the Sun.
The origin of the solar corpuscular radiation, however, remained unclear until Parker
(1958) showed that a hot solar corona cannot maintain a hydrostatic equilibrium.
Instead, the pressure-gradient force overcomes gravity and leads to a radial accelera-
tion of the coronal plasma to supersonic velocities, which Parker called “solar wind”
in contrast to a subsonic “solar breeze” (Chamberlain 1961), which was later found to
be unstable (Velli 1994). Soon after this prediction, the solar wind was measured in
situ by spacecraft (Gringauz et al. 1960; Neugebauer and Snyder 1962). For the last
four decades, the solar wind has been monitored almost continuously in situ. Parker’s
underlying concept is the mainstream paradigm for the acceleration of the solar wind,
but many questions remain unresolved. For example, we still have not identified the
mechanisms that heat the solar corona to temperatures orders of magnitude higher
than the photospheric temperature, albeit this discovery was made some 80 years ago

123
5 Page 4 of 136 D. Verscharen et al.

Table 1 The multiple characteristic plasma parameters (top), length scales (middle), and timescales (bottom)
in the solar wind
Symbol Solar wind (Upper) Corona Definition

np , ne 3 cm−3 106 cm−3 Proton and electron number density


Tp , Te 105 K 106 K Proton and electron temperature
B 3 × 10−5 G 1G Magnetic field strength

λmfp,p 3 au 100 Mm Proton collisional mean free path


L 1 au 100 Mm Characteristic size of the system
dp 140 km 230 m Proton inertial length
ρp 160 km 13 m Proton gyration radius
de 3 km 5m Electron inertial length
ρe 2 km 30 cm Electron gyration radius
λp , λe 12 m 7 cm Proton and electron Debye lengths

Πνc 120 d 2h Proton collision time


τ 2.4 d 10 min Expansion time
ΠΩp 26 s 660 μs Proton gyration period
Πωpp 3 ms 5 μs Proton plasma period
ΠΩe 14 ms 360 ns Electron gyration period
Πωpe 70 μs 110 ns Electron plasma period

This table shows typical parameters in the solar wind at 1 au and in the upper solar corona (∼ 100 Mm
above photosphere). For each angular frequency ω, the associated timescale is given by Πω ≡ 2π/|ω|

(Grotrian 1939; Edlén 1943). As we discuss the observed features of the solar wind
in this review, we will encounter further deficiencies in our understanding that require
more detailed analyses beyond Parker’s model. In this process, we will find many
observational facts that models of coronal heating and solar-wind acceleration must
explain in order to achieve a realistic and consistent description of the physics of the
solar wind.
In the first section of this review, we lay out the various characteristic length and
timescales in the solar wind and motivate our thesis that this multi-scale nature defines
the evolution of the solar wind. We then introduce the observed large-scale, global
features and the microphysical, kinetic features of the solar wind as well as the math-
ematical basis to describe the related processes.

1.1 The characteristic scales in the solar wind

Table 1 lists typical values for the characteristic plasma parameters and scales in the
solar wind at 1 au and in the upper solar corona that we introduce and define in this
section. It is important to remember that all of these quantities vary widely in time and
may differ significantly between thermal and superthermal particle populations. We
illustrate the broad range of the characteristic length scales and timescales in Fig. 1.
The solar wind expands to a heliocentric distance of about 90 au, where it transitions
to a subsonic flow by crossing the solar-wind termination shock (Stone et al. 2005;

123
The multi-scale nature of the solar wind Page 5 of 136 5

Fig. 1 Graphical representation of the characteristic length scales (top) and timescales (bottom) in the solar
wind. The bar lengths represent the typical range for each scale given in Table 1. The magenta end of each
bar indicates the typical coronal value, and the cyan end of each bar indicates the typical value at 1 au

Burlaga et al. 2008). Although we do not expound upon the physics of the outer
heliosphere and the interaction of the solar wind with the interstellar medium, this is
the largest spatial scale in the supersonic solar wind. Considering the inner heliosphere
(i.e., the spherical volume centered around the Sun within Earth’s orbit), we identify
the characteristic size of the system as L ∼ 1 au. For a typical radial solar-wind flow
speed Ur in the range of 300 km/s to 800 km/s (Lopez and Freeman 1986), we find an
expansion time of

L
τ∼ ∼ 2.4 d (1)
Ur

for the solar wind from the Sun to 1 au. The Sun’s siderial rotation period at its equator,

τrot ∼ 25 d, (2)

introduces another characteristic global timescale.


In addition to the outer size of the system, a plasma has multiple characteristic
scales due to the interactions of its free charges with electric and magnetic fields. In
a homogeneous and constant magnetic field B0 , a plasma particle with charge q j and
mass m j (where j denotes the particle species) experiences a continuous deflection

123
5 Page 6 of 136 D. Verscharen et al.

of its trajectory due to the Lorentz force. The frequency associated with this helical
motion is given by the gyro-frequency1 (also called the cyclotron frequency)

q j B0
Ωj ≡ , (3)
m jc

where c is the speed of light in vacuum. The timescale for one closed loop around the
magnetic field is then given by the gyro-period ΠΩ j ≡ 2π/|Ω j |. In the solar wind at
1 au, ΠΩp ∼ 26 s and ΠΩe ∼ 14 ms, where the index p represents protons and the
index e represents electrons. On the other hand, in the upper corona (about 100 Mm
above the photosphere), where the magnetic field is much stronger than in the solar
wind, ΠΩp ∼ 660 μs and ΠΩe ∼ 360 ns. Aside from protons, α-particles (i.e., fully
ionized helium atoms) are also dynamically important in the solar wind since they
account for  20% of the mass density.
We define the perpendicular thermal speed as

2kB T⊥ j
w⊥ j ≡ (4)
mj

and the parallel thermal speed as



2kB T j
w j ≡ , (5)
mj

where T⊥ j (T j ) is the temperature of particle species j in the direction perpendicular


(parallel) to B0 and kB is the Boltzmann constant. We define the concept of temper-
atures perpendicular and parallel to B0 in Eqs. (38) and (39). Assuming a thermal
distribution of particles with a perpendicular thermal speed w⊥ j , the characteristic
size of the gyration orbit is given by the gyro-radius

w⊥ j
ρj ≡  . (6)
Ω j 

At 1 au, solar-wind gyro-radii are typically ρp ∼ 160 km and ρe ∼ 2 km. In the upper
corona, the gyro-radii are smaller: ρp ∼ 13 m and ρe ∼ 30 cm.
The plasma frequency

4π n 0 j q 2j
ωp j ≡ , (7)
mj

1 Following the prevalent convention in space plasma physics, we adopt the metric system of Gaussian-cgs
units. The NRL Plasma Formulary (Huba 2016) includes a guide to converting formulæ between cgs and
SI units. In some figures, we plot magnetic field in nT for consistency with the published plots on which
they are based.

123
The multi-scale nature of the solar wind Page 7 of 136 5

where n 0 j is the background number density of species j, corresponds to the charac-


teristic timescale for electrostatic interactions in the plasma: Πωp j ≡ 2π/ωp j . In the
solar wind at 1 au, Πωpp ∼ 3 ms and Πωpe ∼ 70 μs. These timescales are even shorter
in the corona: Πωpp ∼ 5 μs and Πωpe ∼ 110 ns. A reduction of the local electron
number density (e.g., through a spatial displacement of a number of electrons with
respect to the ions) leads to an oscillation of the electrons with respect to the ions, in
which the electrostatic force due to the displaced charge serves as the restoring force.
This plasma oscillation occurs with a frequency ∼ ωpe . In addition, light waves can-
not propagate at frequencies  ωpe in a plasma as the free plasma charges shield the
wave’s electromagnetic fields so that the wave amplitude drops off exponentially with
distance when the wave frequency is  ωpe . The exponential decay length associated
with this shielding is given by the skin-depth de ≡ c/ωpe .
More generally, we define the skin-depth (also called the inertial length) of species
j as
c vA j
dj ≡ = , (8)
ωp j |Ω j |

where
B0
vA j ≡  (9)
4π n 0 j m j

is the Alfvén speed of species j. In the solar wind at 1 au, dp ∼ 140 km and de ∼ 3 km.
In the upper corona, on the other hand, dp ∼ 230 m and de ∼ 5 m. In processes
that occur on length scales greater than dp and timescales greater than ΠΩp , protons
exhibit a magnetized behavior, which means that their trajectory is closely tied to
the magnetic field lines, following a quasi-helical gyration pattern with the frequency
given in Eq. (3). Likewise, electrons exhibit magnetized behavior in processes that
occur on length scales greater than de and timescales greater than ΠΩe .
An important length scale associated with electrostatic effects is the Debye length

kB T j
λj ≡ , (10)
4π n 0 j q 2j

where T j is the (scalar, isotropic) temperature of species j. We note that λp ∼ λe


through much of the heliosphere, which makes the Debye length unique among the
scales we discuss. The total Debye length
⎛ ⎞−1
 1
λD ≡ ⎝ ⎠ (11)
λj
j

is the characteristic exponential decay length for a time-independent global electro-


static potential in a plasma. In the solar wind at 1 au, λp ∼ λe ∼ 12 m, while the
plasma in the upper corona exhibits λp ∼ λe ∼ 7 cm. Collective plasma processes

123
5 Page 8 of 136 D. Verscharen et al.

(i.e., particles behaving as if they only interact with a smooth macroscopic electro-
magnetic field rather than with individual moving charges) become important if the
number of particles within a sphere of radius λD is large,

n 0e λ3D  1, (12)

and if

λD  L. (13)

Equations (12) and (13) guarantee that electrostatic single-particle effects are shielded
by neighboring charges from the surrounding plasma (known as Debye shielding). If
one or both of these conditions are not fulfilled, common plasma-physics methods do
not apply and a material is merely an ionized gas rather than a plasma. The solar wind,
however, satisfies both of these conditions and, therefore, is a plasma.
In addition to these collective plasma length scales and timescales, collisional
effects are associated with their own characteristic scales, which depend on the
type of collisional interaction under consideration (e.g., temperature equilibration or
isotropization) and on different combinations of plasma parameters. We discuss these
effects and the associated timescales in Sect. 3.
Comparing the coronal electron Debye length as the smallest plasma length scale
of the solar wind with the size of the system reveals that the solar wind covers over
twelve orders of magnitude in its characteristic length scales (neglecting length scales
associated with collisions, which can be even greater than L). Similarly, comparing the
corona’s electron plasma period with the solar wind’s expansion time reveals that the
solar wind also covers over twelve orders of magnitude in its characteristic timescales
(again neglecting timescales associated with collisions, which can be even greater than
τ ). These ratios demonstrate the intrinsically multi-scale nature of the solar wind. The
broad range of scales also illustrates the difficulty in treating the solar wind and all
related physics processes numerically since complete numerical simulations would
need to resolve this entire range of scales.
This review describes plasma processes that depend upon or modify the multi-scale
nature of the solar wind. As a truly Living Review, its first edition is limited to small-
scale processes that affect the large-scale evolution of the plasma. In a later major
update, we will describe how large-scale processes affect the small-scale structure of
the plasma such as expansion effects on particle properties, wave reflection and the
creation of turbulence, streaming interactions, mixing from different solar sources in
co-rotating interaction regions, and magnetic focusing effects, as well as the impact
of these processes on global solar-wind modeling. Although every plasma process is
conceivably a multi-scale process, we, by practical necessity, only address the physics
processes we consider most relevant to the multi-scale evolution of the solar wind.
The most prominent processes not covered in this review include detailed discussions
of reconnection (Pontin 2011; Gosling 2012; Paschmann et al. 2013), shock waves
(Balogh et al. 1995; Chashei and Shishov 1997; Lepping 2000; Rice and Zank 2003),
the physics of the outer heliosphere (pick-up ions, energetic neutral atoms, etc., Zank
et al. 1995; Gloeckler and Geiss 1998; Zank 1999; Richardson et al. 2004; McComas

123
The multi-scale nature of the solar wind Page 9 of 136 5

et al. 2012; Zank et al. 2018), interplanetary dust (Krüger et al. 2007; Mann et al. 2010),
interactions with planetary bodies (Grard et al. 1991; Kivelson and Bagenal 2007;
Gardini et al. 2011; Bagenal 2013), eruptive events such as coronal mass ejections
(Zurbuchen and Richardson 2006; Howard and Tappin 2009; Webb and Howard 2012),
solar energetic particles (Ryan et al. 2000; Mikić and Lee 2006; Klein and Dalla 2017),
and (anomalous) cosmic rays (Heber et al. 2006; Potgieter 2008; Giacalone et al. 2012;
Potgieter 2013). We also limit our discussion of minor-ion physics.

1.2 Global structure of the solar wind

At heliocentric distances greater than a few solar radii R , the solar wind’s expansion
is, to first order, radial, which creates large-scale radial gradients in most of the plasma
parameters. For this discussion of the global structure, we concentrate only on long-
term averages of the plasma quantities and neglect their frequent—and, as we will
see later, sometimes comparable to order unity—variations. Figure 2 illustrates these
average quantities as functions of distance in the inner heliosphere and demonstrates
the resulting profiles for the characteristic length scales and timescales. Beyond a
distance of about 10 R , the average radial velocity stays approximately constant.
Continuity under steady-state conditions requires that

∇ · n j U j = 0, (14)
where U j is the bulk velocity of species j. In spherical coordinates and under the
assumption that U j ≈ U jr êr ≈ constant, the average density then decreases ∝ r −2 .
In the acceleration region and in regions of super-radial expansion connected to coronal
holes, continuity requires steeper gradients closer to the Sun as confirmed by white-
light polarization measurements (Cranmer and van Ballegooijen 2005). In addition, the
deceleration of streaming α-particles leads to a small deviation from the r −2 density
profile (Verscharen et al. 2015).
To first order, the average magnetic field follows the Parker spiral in the plane of
the ecliptic (Parker 1958; Levy 1976; Behannon 1978; Mariani et al. 1978, 1979)
as a result of the frozen-in condition of ideal magnetohydrodynamics (MHD; see
Sect. 1.4.2) and the rotation of the Sun. We define
8π n j kB T j
βj ≡ , (15)
B2
where B is the magnetic field, as the ratio between the thermal pressure of species
j and the magnetic pressure. In the solar corona, β j  1, so that the magnetic
field constraints the plasma to co-rotate with the Sun. However, the magnetic field’s
torque on the plasma decreases with distance from the Sun until the plasma outflow
dominates the evolution of the magnetic field and convects the field into interplanetary
space (Weber and Davis 1967). In the Parker model, the Parker angle |φ Br | between
the direction of the magnetic field and the radial direction increases with distance r
from the Sun,

Bφ Ω sin θ
tan φ Br = = (reff − r ) , (16)
Br Upr

123
5 Page 10 of 136 D. Verscharen et al.

Fig. 2 Characteristic average quantities, length scales, and timescales as functions of distance from the Sun
in the inner heliosphere for typical fast-solar-wind conditions. We calculate these scales based on typical
radial profiles of the solar-wind magnetic-field strength, density, and velocity (shown in the top panel). The
profiles for the magnetic field and the density are taken from Smith et al. (2012) for a radial polar flux
tube. The radial velocity profile then follows from flux conservation, n j U jr /Br = constant. The electron
temperature is taken from a fit to measurements at r < 10 R (Cranmer et al. 1999) and then connected to
a power-law with a power index corresponding to the radial temperature profiles observed with Helios in
the fast solar wind (Štverák et al. 2015). We take Tp ≈ Te for simplicity

where Bφ and Br are the azimuthal and radial components of the magnetic field, Ω
is the angular speed of the Sun’s rotation, θ is the polar angle, and reff is the effective
co-rotation radius. In our sign and coordinate convention, φ Br ≤ 0 if Br > 0 since the
Sun rotates in the + êφ -direction, which differs from Parker’s (1958) original choice.
The radius reff is an auxiliary quantity to describe the heliospheric distance beyond
which the solar wind behaves as if it were co-rotating for r ≤ reff (Hollweg and Lee
1989). Observations indicate that reff ∼ 10 R in the fast wind and reff ∼ 20 R in

123
The multi-scale nature of the solar wind Page 11 of 136 5

the slow wind (Bruno and Bavassano 1997). The Parker angle |φ Br | increases from
0◦ at reff to about 45◦ at r = 1 au. This trend continues into the outer heliosphere as
shown by observations (Thomas and Smith 1980; Forsyth et al. 2002). The magnitude
of the Parker field decreases with distance as

1 + tan2 φ Br
B0 ∝ , (17)
r2

which is ∝ r −2 in the limit tan2 φ Br  1 at small r and ∝ r −1 in the limit tan2 φ Br 


1 at large r . We note that the original Parker model is not completely torque-free,
although a torque-free treatment leads to only minor modifications (Verscharen et al.
2015). Further details about the heliospheric magnetic field can be found in the review
by Owens and Forsyth (2013).

1.3 Categorization of solar wind

Traditionally, the solar wind has been categorized into three groups (Srivastava and
Schwenn 2000):
1. fast wind with bulk velocities between about 500 km/s and 800 km/s,
2. slow wind with bulk velocities between about 300 km/s and 500 km/s, and
3. variable/eruptive events such as coronal mass ejections with speeds from a few
hundreds up to 2000 km/s.
Measurements from the Ulysses spacecraft during solar minimum dramatically
demonstrate that the fast wind emerges predominantly from polar coronal holes and the
slow wind from the streamer belt at the solar equator (Phillips et al. 1995; McComas
et al. 1998b, 2000, 2003; Ebert et al. 2009). The left-hand panel in Fig. 3 illustrates
the clear sector boundary between fast and slow wind during solar minimum. During
solar maximum, however, fast and slow wind emerge from neighboring patches every-
where in the corona. The right-hand panel in Fig. 3 shows that the occurrence of fast
and slow wind streams does not strongly correlate with heliographic latitude during
solar maximum. On average, fast polar wind exhibits both a lower density and less
variation in density than slow wind. The association of different wind streams with
different source regions suggests that the magnetic-field configuration in the corona
plays a crucial role in determining the properties of the wind streams. In addition to
the differences in speed and density, fast and slow wind exhibit further distinguishing
marks. Fast wind, relative to slow wind, generally is more steady, is more Alfvénic
(i.e., it exhibits a higher correlation or anti-correlation between fluctuations in vector
velocity and vector magnetic field; see Sect. 4 and Tu and Marsch 1995), and has a
higher proton temperature (Neugebauer 1976; Wilson et al. 2018). Importantly for its
multi-scale evolution, fast wind is also less collisional (both in terms of the local col-
lisional relaxation times and the cumulative time for collisions to act) than slow wind
(Marsch et al. 1982b; Marsch and Goldstein 1983; Livi et al. 1986; Kasper et al. 2008;
Bourouaine et al. 2011; Ďurovcová et al. 2017), which allows for more kinetic non-
equilibrium features to survive the thermalizing action of Coulomb collisions. Fast

123
5 Page 12 of 136 D. Verscharen et al.

Fig. 3 Ulysses/SWOOP observations of the solar-wind proton radial velocity and density at different helio-
graphic latitudes. The distance from the center in each of these polar plots indicates the velocity (blue) and
density (green). The polar angle represents the heliographic latitude. Since these measurements were taken
at varying distances from the Sun, we compensate for the density’s radial decrease by multiplying n p with
r 2 . The red circle represents Upr = 500 km/s and r 2 n p = 10 au2 cm−3 . The straight red lines indicate the
sector boundaries at ±20◦ latitude. Left panel: Ulysses’ first polar orbit during solar minimum (1990-12-20
through 1997-12-15). Right panel: Ulysses’ second polar orbit during solar maximum (1997-12-15 through
2004-02-22). After McComas et al. (2000) and McComas et al. (2008)

wind, therefore, exhibits more non-Maxwellian structure in its distribution functions


(Marsch 2006; Marsch 2018) as we discuss in the next section.
The elemental composition and the heavy-ion charge states also differ between
fast and slow wind (Bame et al. 1975; Ogilvie and Coplan 1995; von Steiger et al.
1995; Bochsler 2000; von Steiger et al. 2000; Aellig et al. 2001b; Zurbuchen et al.
2002; Kasper et al. 2007, 2012; Lepri et al. 2013). Elements with a low first ionization
potential (FIP) such as magnesium, silicon, and iron exhibit enhanced abundances in
the solar corona and in the solar wind with respect to their photospheric abundances
(Gloeckler and Geiss 1989; Raymond 1999; Laming 2015). Conversely, elements with
a high FIP such as oxygen, neon, and helium have much lower enhancements or even
depletions with respect to their photospheric abundances. This FIP fractionation bias
also varies with wind speed and is generally smaller in fast wind than in slow wind
(Zurbuchen et al. 1999; Bochsler 2007). Since the elemental composition of a plasma
parcel does not change as it propagates through the heliosphere unless it mixes with
neighboring parcels, composition measurements are a reliable method to distinguish
solar-wind source regions. Moreover, studies of heavy ions constrain proposed mod-
els of solar-wind acceleration and heating. For instance, proposed acceleration and
heating scenarios must explain the observed preferential heating of minor ions. In the
solar wind, most heavy ion species i exhibit Ti /Tp ≈ 1.35m i /m p (Tracy et al. 2015;
Heidrich-Meisner et al. 2016; Tracy et al. 2016).
Lately, the traditional classification of wind streams by speed has experienced some
major criticism (e.g., Maruca et al. 2013; Xu and Borovsky 2015; Camporeale et al.
2017). Speed alone does not fully classify the properties of the wind, and there is a
smooth transition in the distribution of wind speeds. At times, fast solar wind shows
properties traditionally associated with slow wind and vice versa, such as collision-

123
The multi-scale nature of the solar wind Page 13 of 136 5

ality, Alfvénicity, FIP-bias, anisotropy, beam structures, etc. Although these atypical
behaviors suggest a false dichotomy between fast and slow wind, we retain the tra-
ditional nomenclature, albeit defining “fast wind” as wind with the typical fast-wind
properties and “slow wind” as wind with the typical slow-wind properties under consid-
eration instead of relying on the flow speeds alone. Nevertheless, we expressly caution
the reader against assuming wind speed alone as a reasonable indication of wind
type.

1.4 Kinetic properties of the solar wind

Kinetic plasma physics describes the statistical properties of a plasma by means of


the particle velocity distribution functions f j (x, v, t) for each plasma species j. We
define and normalize the distribution function so that

f j (x, v, t) d3 x d3 v (18)

represents the number of particles of species j in the phase-space volume d3 x d3 v cen-


tered on the phase-space coordinates (x, v) at time t. The distribution function relates
to the bulk properties (i.e., density, bulk velocity, temperature,...) through its velocity
moments as described in Sect. 1.4.1. A continuous definition of f j is appropriate when
Eq. (12) is fulfilled.
The central equation in kinetic physics is the Boltzmann equation,

∂ fj ∂ fj ∂ fj δ fj
+v· +a· = , (19)
∂t ∂x ∂v δt c

where a is the acceleration of a j-particle due to macroscopic forces, and the right-hand
side describes the temporal change in f j due to particle collisions, which are mediated
by microscopic electric forces among individual particles (see also Sect. 3.2 of this
review; Lifshitz and Pitaevskii 1981). We use the term macroscopic fields to indicate
that these are locally averaged to remove the rapidly fluctuating Coulomb electric
fields due to individual charges, which are responsible for Coulomb collisions. The
applicability of this mean-field approach is a key quality of a plasma and distinguishes
it from other types of ionized gases, in which Eq. (12) is not fulfilled. Without the
collision term, the Boltzmann equation represents a fluid continuity equation for the
density in phase space. It is thus related to Liouville’s theorem and describes the con-
servation of the phase-space density along trajectories in the absence of collisions.2 In
this case, and when using only macroscopic electromagnetic forces in the acceleration

2 We refrain from discussing the multiple ways of deriving the Boltzmann equation such as the closure of the
BBGKY hierarchy (Bogoliubov 1946) or the Klimontovich–Dupree formalism (Dupree 1961; Klimontovich
1967). Instead, we express the Boltzmann equation in terms of Liouville’s theorem and subsume all higher-
order particle interactions in the collision term on the right-hand side of Eq. (19). For more details, see also
Sect. 3.2.

123
5 Page 14 of 136 D. Verscharen et al.

term, we obtain the Vlasov equation,

∂ fj ∂ fj qj 1 ∂ fj
+v· + E+ v×B · = 0, (20)
∂t ∂x mj c ∂v

which is the fundamental equation of collisionless kinetic plasma physics. These


macroscopic electric and magnetic fields obey Maxwell’s equations,

∇ · E = 4πρc , (21)
∇ · B = 0, (22)
1 ∂B
∇ ×E=− , (23)
c ∂t

and

4π 1 ∂E
∇ ×B= j+ , (24)
c c ∂t

where the charge density ρc and the current density j are given by integrals over the
distribution functions as
 
ρc = qj f j d3 v (25)
j

and
 
j= qj v f j d3 v. (26)
j

Equations (20) through (26) form a closed set of integro-differential equations in six-
dimensional phase space and time that fully describe the evolution of collisionless
plasma.

1.4.1 Fluid moments and fluid equations

Although the distribution functions f j contain all of the microphysical properties of


the plasma, it is often sufficient to rely on a reduced set of macrophysical parameters
that only depend on time and three-dimensional configuration space (versus time
and six-dimensional phase space). These parameters are called bulk parameters and
correspond to the velocity moments as integrals over the full velocity space of the
distribution function. Certain velocity moments represent named fluid bulk parameters.
For instance, the zeroth velocity moment corresponds to the number density

nj = f j d3 v. (27)

123
The multi-scale nature of the solar wind Page 15 of 136 5

Using n j , the first velocity moment corresponds to the bulk velocity



1
Uj = v f j d3 v, (28)
nj

while the second moment represents the pressure tensor



Pj = m j v − Uj v − U j f j d3 v. (29)

The third moment corresponds to the heat-flux tensor



Qj = m j v − Uj v − Uj v − U j f j d3 v. (30)

For many applications in magnetized-plasma physics, it is useful to choose the


coordinate system to be aligned with the direction b̂ ≡ B/|B| of the magnetic field
and to define the pressure components with respect to the direction of the magnetic
field. In this coordinate system, Equation (30) reduces through contraction to the
perpendicular heat-flux vector

1  
q⊥ j = Q j : I3 − b̂b̂ (31)
2

and the parallel heat-flux vector


 
q j = Q j : b̂b̂ , (32)

where I3 is the three-dimensional unit matrix. We define the double-dot and triple-dot
products in a similar way to the usual dot product as
 
A:B= Ai j B ji and A:̇B = Ai jk Bk ji . (33)
i, j i, j,k

Although higher moments do not give rise to named bulk parameters like these four,
the moment hierarchy can be continued to infinity by multiplying the integrand with
further powers of velocity.
Taking velocity moments of the full Vlasov equation and exploiting the definitions
of the lowest moments above leads to the multi-fluid plasma equations (Barakat and
Schunk 1982; Marsch 2006). The zeroth and first moments of the Vlasov equation are
the continuity equation,

∂n j
+ ∇ · n j U j = 0, (34)
∂t

123
5 Page 16 of 136 D. Verscharen et al.

and the momentum equation,

∂ 1
n jm j + Uj · ∇ Uj = − ∇ · Pj + n jqj E + Uj × B . (35)
∂t c

We define the perpendicular pressure and the parallel pressure as

I3 − b̂b̂
p⊥ j ≡ P j : (36)
2
and
 
p j ≡ P j : b̂b̂ , (37)

respectively, which are related to the temperatures in the directions perpendicular and
parallel to B through
p⊥ j
T⊥ j = (38)
n j kB

and
p j
T j = . (39)
n j kB

We write the perpendicular energy equation as

∂  
+ Uj · ∇ p⊥ j + p⊥ j ∇ · U j + ∇⊥ · U j = b̂b̂ − I3 : τ j · ∇U j
∂t
1 ∂   1  
−∇ · q⊥ j − τ j : + U j · ∇ b̂b̂ − Q j :̇∇ b̂b̂ (40)
2 ∂t 2

and the parallel energy equation as


+ Uj · ∇ p j + p j ∇ · U j + 2∇ · U j = − 2b̂b̂ : τ j · ∇U j
∂t
∂    
−∇ · q j + τ j : + U j · ∇ b̂b̂ + Q j :̇∇ b̂b̂ , (41)
∂t

where

τ j ≡ P j − p⊥ j I3 − p j − p⊥ j b̂b̂ (42)

is the stress tensor,


   
∇⊥ ≡ I3 − b̂b̂ ∇, and ∇ ≡ b̂b̂ ∇. (43)

123
The multi-scale nature of the solar wind Page 17 of 136 5

The hierarchy of moments of the Vlasov equation continues to infinity, and similar
fluid equations exist for the stress tensor, the heat-flux tensor, and all higher-order
moments. However, this gives rise to a closure problem since the nth moment of the
Vlasov equation always includes the (n +1)st moment of the distribution function. For
example, the continuity equation, which is the zeroth moment of the Vlasov equation,
includes the bulk velocity, which corresponds to the first moment of f j . The (n + 1)st
moment of the distribution function, in turn, requires the (n + 1)st moment of the
Vlasov equation as a description of its dynamical evolution. Every fluid model is,
therefore, fundamentally susceptible to a closure problem since the solution of an
infinite chain of non-degenerate equations is formally impossible. For most practical
purposes, the moment hierarchy is thus truncated by expressing a higher-order moment
of f j through lower moments of f j only. Closing the moment hierarchy introduces
limitations on the physics of the problem at hand and deviations in the solutions to
the multi-fluid system of equations from the solutions to the full Vlasov equation. For
example, a typical closure of the moment hierarchy is the assumption of an isotropic
and adiabatic pressure, i.e., P j = p j I3 and p j ∝ n κj , where κ is the adiabatic exponent.
This closure of the momentum equation neglects heat flux and small velocity-space
structure in f j . Therefore, any finite closure is only applicable if the physics of the
problem at hand justifies the neglect of higher-order velocity moments of f j . We note,
for example, that collisions are such a process that can produce conditions under which
higher-order moments are negligible (see Sect. 3).
Assuming only slow changes of the magnetic field compared to ΠΩ j and that
τ j = 0, the second velocity moment of the Vlasov equation (20) leads to the useful
double-adiabatic energy equations (Chew et al. 1956; Whang 1971; Sharma et al.
2006; Chandran et al. 2011),

∂ p⊥ j
njB + Uj · ∇ = − ∇ · q⊥ j − q⊥ j ∇ · b̂ (44)
∂t njB

and
 
n 3j ∂ B 2 p j
+ Uj · ∇ = − ∇ · q j + 2q⊥ j ∇ · b̂. (45)
B2 ∂t n 3j

If we neglect heat flux by setting the right-hand sides of Eqs. (44) and (45) to zero,
we obtain the conservation laws for the double-adiabatic invariants, which are also
referred to as the Chew–Goldberger–Low (CGL) invariants (Chew et al. 1956)

p⊥ j B 2 p j
≈ constant and ≈ constant. (46)
njB n 3j

1.4.2 Magnetohydrodynamics

Magnetohydrodynamics (MHD) is a single-fluid description that results from sum-


ming the fluid equations of all species and defining the moments of the single

123
5 Page 18 of 136 D. Verscharen et al.

magnetofluid as the mass density



ρ≡ m jn j, (47)
j

the bulk velocity

1
U≡ m jn jUj, (48)
ρ
j

and the total scalar pressure

1
P≡ P j : I3 (49)
3
j

under the assumption that P j is isotropic and diagonal. This procedure leads to the
MHD continuity equation,

∂ρ
+ ∇ · (ρU) = 0, (50)
∂t

and the MHD momentum equation,

∂ 1
ρ + U · ∇ U = − ∇ P + (j × B) . (51)
∂t c

The electric-field term from Eq. (35) vanishes under the quasi-neutrality assumption
that ρc from Eq. (25) is negligible, which is justified on scales  λD . Faraday’s law
describes the evolution of the magnetic field as

∂B
= − c∇ × E. (52)
∂t

The electric field follows from the electron momentum equation (35) as the generalized
Ohm’s law,

me ∂ 1 1 1
E= + Ue · ∇ Ue + ∇ · Pe − j×B+ ji × B, (53)
qe ∂t n e qe n e qe c n e qe c

where

ji ≡ j − n e qe Ue (54)

is the ion contribution to the current density. The terms on the right-hand side of
Eq. (53) represent the contributions from electron inertia, the electron pressure gra-
dient (i.e., the ambipolar electric field), the Hall term, and the ion convection term,

123
The multi-scale nature of the solar wind Page 19 of 136 5

respectively. Under the assumptions of quasi-neutrality in a proton–electron plasma


and the negligibility of terms of order m e /m p , we find

1 1 1
E= ∇ · Pe − j × B − U × B. (55)
n e qe n e qe c c

If we furthermore assume small or moderate βe and consider processes occurring on


scales  dp (Chiuderi and Velli 2015), we can neglect the contributions of the electron
pressure gradient and the Hall term to E. We then find the common expression for
Ohm’s law in MHD:
1
E = − U × B. (56)
c
Equations (52) and (56) describe Alfvén’s frozen-in theorem, stating that magnetofluid
bulk motion across field lines is forbidden, since otherwise the infinite resistivity of the
magnetofluid would lead to infinite eddy currents. Instead, the magnetic flux through
a co-moving surface is conserved.3 The assumptions leading to Eq. (56) are fulfilled
for processes on time scales much greater than ΠΩ j and Πωp j as well as on spatial
scales much greater than d j and ρ j . In this limit, the displacement current in Ampère’s
law is also negligible, which allows us to write the current density in Eq. (51) in terms
of the magnetic field:
c
j= ∇ × B. (57)

The MHD equations are often closed with the adiabatic closure relation,

∂ P
+U·∇ = 0, (58)
∂t ρκ

where κ is the adiabatic exponent. The MHD equations are intrinsically scale-free and,
therefore, only valid for processes that do not occur on any of the characteristic plasma
scales of the system introduced in Sect. 1.1. Thus, MHD only applies to large-scale
phenomena that occur
1. on length scales  L,
2. on length scales  max(d j , ρ j ), and
3. on timescales  max(ΠΩ j , Πωp j )
for all j.

1.4.3 Standard distributions in solar-wind physics

Although solar-wind measurements often reveal irregular plasma distribution func-


tions (see Sects. 1.4.4, 1.4.5, as well as Marsch 2012), it is sometimes helpful to

3 Interestingly, the inclusion of the pressure-gradient term from Eq. (55) in Eq. (56) does not affect the
frozen-in condition since it cancels when taking the curl in Eq. (52).

123
5 Page 20 of 136 D. Verscharen et al.

invoke closed analytical expressions for the distribution functions in a plasma. In


the following description, we use the cylindrical coordinate system in velocity space
introduced in Sect. 1.4.1 with its symmetry axis to be parallel to b̂.
A gas in thermodynamic equilibrium has a Maxwellian velocity distribution,
 2

nj v − Uj
f M (v) = 3/2 3 exp − , (59)
π wj w 2j

where

2kB T j
wj ≡ (60)
mj

is the (isotropic) thermal speed of species j. Equation (59) has a thermodynamic jus-
tification in equilibrium statistical mechanics based on the Gibbs distribution (Landau
and Lifshitz 1969). An empirically motivated extension of the Maxwellian distribution
is the so-called bi-Maxwellian distribution, which introduces temperature anisotropies
with respect to the background magnetic field yet follows the Maxwellian behavior
on any one-dimensional cut at constant v⊥ or constant v in velocity space:
 2

nj v2 v − U j
f bM (v) = 3/2 2 exp − 2⊥ − , (61)
π w⊥ j w j w⊥ j w2 j

where w⊥ j and w j are the thermal speeds defined in Eqs. (4) and (5). Advanced
methods in thermodynamics such as non-extensive statistical mechanics lead to the
κ-distribution (Tsallis 1988; Livadiotis and McComas 2013; Livadiotis 2017),

 3/2  2
−κ−1
nj 2 Γ (κ + 1) 2 v − Uj
f κ (v) = 3 1+ , (62)
wj π(2κ − 3) Γ (κ − 1/2) 2κ − 3 w 2j

where Γ (x) is the Γ -function (Abramowitz and Stegun 1972) and κ > 3/2. We note
that f κ → f M for κ → ∞. The κ-distribution is characterized by having tails that
are more pronounced for smaller κ (i.e., the kurtosis of the distribution increases as κ
decreases). Analogous to the bi-Maxwellian is the bi-κ-distribution,

 3/2
nj 2 Γ (κ + 1)
f bκ (v) =
w⊥
2 w
j j
π(2κ − 3) Γ (κ − 1/2)
  2
−κ−1
2 v⊥
2 v − U j
× 1+ + . (63)
2κ − 3 w⊥ 2
j w2 j

123
The multi-scale nature of the solar wind Page 21 of 136 5

Fig. 4 Illustration of ion (left) and electron (right) kinetic features in the solar wind. We show cuts through
the distribution function along the direction of the magnetic field. We normalize the distribution functions
to the maxima of the proton and electron distribution functions, respectively. We normalize the parallel
velocity to the thermal speed of the proton and electron core components, wc,p and wc,e , respectively. We
note that wc,p  wc,e . The gray curves show the underlying core distribution alone. The distributions are
shown in the reference frames in which the core distribution is at rest

In the following sections, we will encounter observed distribution functions and rec-
ognize some of the uses and limitations of these analytical expressions.

1.4.4 Ion properties

In-situ spacecraft instrumentation has been measuring ion and electron velocity distri-
butions for decades (see Sect. 2.2). Figure 4 summarizes some of the observed features
in ion and electron distribution functions schematically.
These observations show that proton distributions often deviate from the Maxwellian
equilibrium distribution given by Eq. (59). For instance, proton distributions often dis-
play a field-aligned beam: a second proton component streaming faster than the proton
core component along the direction of the magnetic field with a relative speed  vAp
(Asbridge et al. 1974; Feldman et al. 1974b; Marsch et al. 1982b; Goldstein et al.
2000; Tu et al. 2004; Alterman et al. 2018). In Fig. 4 (left), the proton beam is shown
in green as an extension of the distribution function toward greater v . Protons also
show temperature anisotropies with respect to the magnetic field (Hundhausen et al.
1967a, b; Marsch et al. 1981; Kasper et al. 2002; Marsch et al. 2004; Hellinger et al.
2006; Bale et al. 2009; Maruca et al. 2012), which manifest in unequal diagonal ele-
ments of P j in Eq. (29). Figure 5 shows isosurfaces of f p based on measurements
from the Helios spacecraft. The background magnetic field is vertically aligned, and
the color-coding represents the distance of the isosurfaces from the center-of-mass
velocity. A standard Maxwellian distribution would be a monochromatic sphere in
these diagrams. Instead, we see that the proton distribution is anisotropic. The exam-
ple on the left-hand side shows an extension of the isosurface along the magnetic-field
direction, which indicates the proton-beam component. Almost always, the proton
beam is directed away from the Sun and along the magnetic-field axis.4 This observa-

4 The proton beam may be directed toward the Sun or be bi-directional if the local radial component of the
magnetic field changed its sign during the passage of the plasma parcel from the Sun to the location of the
measurement.

123
5 Page 22 of 136 D. Verscharen et al.

Fig. 5 Interpolated isosurfaces in velocity space of two proton distribution functions measured by Helios
2. The arrow B0 indicates the direction of the local magnetic field. The color-coding represents the distance
of the isosurface from the center-of-mass velocity. Left: measurement from 1976-02-04 at 10:21:43 UTC.
The center-of-mass velocity is 478 km/s. The elongation along the magnetic-field direction represents the
proton beam. Right: measurement from 1976-04-16 at 07:50:54 UTC. The center-of-mass velocity is 768
km/s. The oblate structure of the distribution function represents a temperature anisotropy with T⊥p > Tp .
These distribution functions are available as animations in the online supplementary material

tion suggests that the beam represents a preferentially accelerated proton component.
The existence of this beam thus puts a major observational constraint on potential
mechanisms for solar-wind heating and acceleration, which must generate this almost
ubiquitous feature in f p . In the example on the right-hand side of Fig. 5, the isosurface
is spread out in the directions perpendicular to the magnetic field, which indicates that
T⊥p > Tp . Although the plasma also exhibits periods with T⊥p < Tp , the predom-
inance of cases with T⊥p > Tp in the fast wind in the inner heliosphere (Matteini
et al. 2007) suggests an ongoing heating mechanism in the solar wind that counter-acts
the double-adiabatic expansion quantified in Eqs. (44) and (45). The double-adiabatic
expansion alone would create T⊥p  Tp in the inner heliosphere when we neglect
the action of heat flux and collisions on protons. Therefore, only heating mechanisms
that explain the observed anisotropies with T⊥p > Tp in the solar wind (and possibly
also in the corona; see Kohl et al. 2006) are successful candidates for a complete
description of the physics of the solar wind.
The colors on the isosurfaces in Fig. 5 illustrate that the bulk velocity of the proton
distribution function differs significantly from the center-of-mass velocity. This is
mostly due to the α-particles in the solar wind (Ogilvie 1975; Asbridge et al. 1976;
Marsch et al. 1982a; Neugebauer et al. 1994, 1996; Steinberg et al. 1996; Reisenfeld
et al. 2001; Berger et al. 2011; Gershman et al. 2012; Bourouaine et al. 2013). Although
their number density is small (n α  0.05n p ), their mass density corresponds to about
20% of the proton mass density. We often observe the α-particles, like the proton
beam, to drift with respect to the proton core along the magnetic-field direction and
away from the Sun with a typical drift speed  vAp . In Fig. 4 (left), the α-particles

123
The multi-scale nature of the solar wind Page 23 of 136 5

are shown as a separate shifted distribution in red, centered around the α-particle drift
speed.
The solar wind also exhibits anisothermal behavior; i.e., not all plasma species
have equal temperatures (Formisano et al. 1970; Feldman et al. 1974a; Bochsler et al.
1985; Cohen et al. 1996; von Steiger and Zurbuchen 2002, 2006). The α-particles
often show Tα  4Tp (Kasper et al. 2007, 2008, 2012). Electrons are typically
colder than protons in the fast solar wind but hotter than protons in the slow solar
wind (Montgomery et al. 1968; Hundhausen 1970; Newbury et al. 1998). As stated in
Sect. 1.2, heavy-ion-to-proton temperature ratios are typically greater than the corre-
sponding heavy-ion-to-proton mass ratios for almost all observable ions in the solar
wind. Like the other kinetic features, solar-wind heating and acceleration models are
only fully successful if they explain the observed anisothermal behavior.
All of these non-equilibrium features (temperature anisotropies, beams, drifts, and
anisothermal behavior) are less pronounced in the slow solar wind than in the fast
wind, which is typically attributed to the greater collisional relaxation rates and the
longer expansion times in the slow wind (see Sect. 3.3). These non-equilibrium features
reflect the multi-scale nature of the solar wind, since they are driven by a combination
of large-scale expansion effects, local kinetic processes, and the feedback of small-
scale processes on the large-scale evolution.

1.4.5 Electron properties

Although the mass of an electron is much less than the mass of a proton (m e /m p ≈
1/1836), and the electrons’ contribution to the total solar-wind momentum flux
is insignificant, electrons do affect the large-scale evolution of the solar wind
(Montgomery 1972; Salem et al. 2003). As the most abundant particle species,
they guarantee quasi-neutrality: ρc ≈ 0 and j ≈ 0 at length scales  λe and
timescales  Πωpe . Due to their small mass, they are highly mobile and have a
much greater thermal speed than the protons, leading to their subsonic behavior
(i.e., Ue  we ). Their momentum balance in Eq. (35) is dominated by their pressure
gradient and electromagnetic forces. Through these contributions, the electrons create
an ambipolar electrostatic field in the expanding solar wind. This field is the central
underlying acceleration mechanism of exospheric models (see Sect. 3.1; Lemaire and
Scherer 1973; Maksimovic et al. 2001). Parker’s (1958) solar-wind model does not
explicitly invoke an ambipolar electrostatic field. Nevertheless, the electron contribu-
tion to the pressure gradient in Parker’s MHD equation of motion is equivalent to the
ambipolar electric field that follows from Eq. (35) for electrons in the limit m e → 0
(Velli 1994, 2001).
Although electrons typically have greater collisional relaxation rates than ions, they
exhibit a number of characteristic kinetic non-equilibrium features, which, as for the
ions, are more pronounced in the fast solar wind. Most notably, the electron distribution
often consists of three distinct components (Feldman et al. 1975; Pilipp et al. 1987a, b;
Hammond et al. 1996; Maksimovic et al. 1997; Fitzenreiter et al. 1998):
– a thermal core, which mostly follows a Maxwellian distribution and has a thermal
energy of ∼ 10 eV—blue in Fig. 4 (right);

123
5 Page 24 of 136 D. Verscharen et al.

Fig. 6 Electron velocity distribution function measured by Helios 2 in the fast solar wind at a heliocentric
distance of 0.29 au on 1976-04-18 at 23:38:35 UTC. Left: isocontours of the distribution in a field-aligned
coordinate system. Right: a cut through the distribution function along the magnetic-field direction. The
red dashed curve shows a Maxwellian fit to the core of the distribution function. The strahl is clearly visible
as an enhancement in the distribution function at v > 0

– a non-thermal halo, which mostly follows a κ-distribution, manifests as enhanced


high-energy tails in the electron distribution, and has a thermal energy of 
80 eV—green in Fig. 4 (right); and
– a strahl,5 which is a field-aligned beam of electrons and usually travels in the
anti-Sunward direction with a bulk energy  100 eV—red in Fig. 4 (right).
The core typically includes ∼ 95% of the electrons. It sometimes displays a tem-
perature anisotropy (Serbu 1972; Phillips et al. 1989b; Štverák et al. 2008) and a
relative drift with respect to the center-of-mass frame (Bale et al. 2013). A recent
study suggests that a bi-self-similar distribution, which forms through inelastic parti-
cle scattering, potentially describes the core distribution better than a bi-Maxwellian
distribution (Wilson et al. 2019).
The strahl probably results from a more isotropic distribution of superthermal elec-
trons in the corona that has been focused by the mirror force in the nascent solar wind
(Owens et al. 2008), explaining the anti-Sunward bulk velocity of the strahl in the
solar-wind rest frame. As with the ion beams, a Sunward or bi-directional electron
strahl can occur when the magnetic-field configuration changes during the plasma’s
passage from the Sun (Gosling et al. 1987; Owens et al. 2017). Figure 6 shows an
example of an electron velocity distribution function measured in the solar wind. This
distribution exhibits a significant strahl at v > 0 but shows no clear halo component.
We reiterate our paradigm that all successful solar-wind acceleration and heating sce-
narios must account for the observed kinetic structure of the solar wind, including these
features in the electron distributions. At highest energies  2 keV, a nearly isotropic
superhalo of electrons exists; however, its number density is very small compared
to the densities of the other electron species ( 10−5 cm−3 at 1 au), and its origin
remains poorly understood (Lin 1998; Wang et al. 2012; Yang et al. 2015; Tao et al.
2016).
Observations of the superthermal electrons (i.e., strahl and halo) reveal that (n s +
n h )/n e remains largely constant with heliocentric distance, where n s is the strahl
5 From strahl—the German word for “beam”.

123
The multi-scale nature of the solar wind Page 25 of 136 5

density and n h is the halo density. Conversely, n s /n e decreases with distance from the
Sun while n h /n e increases (Maksimovic et al. 2005; Štverák et al. 2009; Graham et al.
2017). Various processes have been proposed to explain this phenomenon, most of
which involve the scattering of strahl electrons into the halo (Vocks et al. 2005; Gary
and Saito 2007; Pagel et al. 2007; Saito and Gary 2007; Owens et al. 2008; Anderson
et al. 2012; Gurgiolo et al. 2012; Landi et al. 2012; Verscharen et al. 2019a).
Locally, electrons often show isothermal behavior (i.e., having a polytropic index of
one) due to their large field-parallel mobility. Globally, their non-thermal distribution
functions carry a large heat flux according to Eq. (30) into the heliosphere (Feldman
et al. 1976; Scime et al. 1995). Observations of large-scale electron temperature profiles
suggest that the electron heat flux, rather than local heating, dominates their tempera-
ture evolution (Pilipp et al. 1990; Štverák et al. 2015). These energetic considerations
also reveal that a combination of processes regulate the heat flux of the distribution.
Collisions and collective kinetic processes such as microinstabilities are the prime
candidates for explaining electron heat-flux regulation (see Sects. 3.3.2, 6.1.2; Scime
et al. 1994, 1999, 2001; Bale et al. 2013; Lacombe et al. 2014).

1.4.6 Open questions and problems

The major outstanding science questions in solar-wind physics require a detailed


understanding of the interplay between the multi-scale nature and the observed kinetic
features of the solar wind. This theme applies to the coronal and solar-wind heating
problem as well as the overall energetics of the inner heliosphere. We remind ourselves
that any answer to the heating problem must be consistent with multiple detailed obser-
vational constraints as we have seen in the previous sections.
The observed temperature profiles and overall particle energetics of ions and elec-
trons are consequences of the complex interactions of global heat flux, Coulomb
collisions (Sect. 3), local wave action (Sect. 4), turbulent heating (Sect. 5), microinsta-
bilities (Sect. 6), and double-adiabatic expansion (Mihalov and Wolfe 1978; Feldman
et al. 1979; Gazis and Lazarus 1982; Marsch et al. 1983, 1989; Pilipp et al. 1990;
McComas et al. 1992; Gazis et al. 1994; Issautier et al. 1998; Maksimovic et al. 2000;
Matteini et al. 2007; Cranmer et al. 2009; Hellinger et al. 2011; Le Chat et al. 2011;
Hellinger et al. 2013; Štverák et al. 2015). We still lack a detailed physics-based under-
standing of the majority of these processes, and the quantification of these processes
and their role for the overall energetics of the solar wind remains one of the most
outstanding science problems in space research.
Observed temperature profiles (including anisotropies) are some of the central mes-
sengers about the overall solar-wind energetics, apart from velocity profiles. Figure 7
illustrates the radial evolution of the proton and electron temperatures in the direc-
tions perpendicular and parallel to the magnetic field and separated by fast and slow
wind. We also show the expected temperature profiles under the assumption that the
evolution follows the double-adiabatic (CGL) expansion according to Eqs. (44) and
(45) only. All of the measured temperature profiles deviate from the CGL profiles to
some degree, and this trend continues at greater heliocentric distances (Cranmer et al.
2009). Explaining these deviations lies at the heart of the challenge to explain coronal
and solar-wind heating and acceleration.

123
5 Page 26 of 136 D. Verscharen et al.

Fig. 7 Temperature profiles in the inner heliosphere for fast (left) and slow (right) wind. We show radial
power-law fits to proton-temperature measurements separated by fast (700 km/s ≤ Upr ≤ 800 km/s) and
slow (300 km/s ≤ Upr ≤ 400 km/s) solar-wind conditions from Hellinger et al. (2013). Likewise, we
show radial power-law fits to electron-temperature measurements separated by fast (Upr ≥ 600 km/s) and
slow (Upr ≤ 500 km/s) solar-wind conditions from Štverák et al. (2015). The thin-dashed lines indicate
the CGL temperature profiles according to Eqs. (44) and (45), where we set the right-hand sides of both
equations to zero and determine the magnetic field through Eqs. (16) and (17) using n j ∝ 1/r 2 , θ = 90◦ ,
reff = 10 R , and Upr = 500 km/s

We intend this review to give an overview over the relevant multi-scale processes
in the solar wind. In the near future, data from the Parker Solar Probe (Fox et al.
2016) and Solar Orbiter (Müller et al. 2013) spacecraft will provide us with detailed
observations of the local and global properties of the solar wind at different distances
from the Sun. These groundbreaking observations will help us to quantify the roles of
the multi-scale processes described in this review.
Section 2 describes the methods to measure solar-wind particles and fields in situ.
In Sect. 3, we discuss the effects of collisions on the multi-scale evolution of the
solar wind. Section 4 introduces waves, and Sect. 5 introduces turbulence as mecha-
nisms that affect the local and global plasma behavior. We describe the role of kinetic
microinstabilities and parametric instabilities in Sect. 6. In Sect. 7, we summarize this
review and consider future developments in the study of the multi-scale evolution of
the solar wind.

2 In-situ observations of space plasmas

Observations of space plasmas can be roughly divided into two categories: remote and
in-situ. Remote observations include both measurements of the plasma’s own emis-
sions (e.g., radio waves, visible light, and X-ray photons) as well as measurements of
the effects that the plasma has on emissions from other sources (e.g., Faraday rotation
and absorption lines). In this way, regions such as the chromosphere that are inac-
cessible to spacecraft can still be studied. Additionally, imaging instruments such as
coronagraphs provide information on the global structure of space plasma. Neverthe-
less, due to limited spectral and angular resolution, these instruments cannot provide
information on all of the small-scale processes at work within the plasma. Remote
observations also only offer limited information on three-dimensional phenomena. If

123
The multi-scale nature of the solar wind Page 27 of 136 5

the observed plasma is optically thick (e.g., the photosphere in visible light), its interior
cannot be probed; if it is optically thin (e.g., the corona in EUV), remote observations
suffer from the effects of line-of-sight integration.
In contrast, in-situ observations provide detailed information on microkinetic pro-
cesses in space plasmas. Spacecraft carry in-situ instruments into the plasma to directly
detect its particles and fields and thereby to provide small-scale observations of
localized phenomena. Although an in-situ instrument only detects the plasma in its
immediate vicinity, statistical studies of ensembles of measurements have provided
remarkable insights into how small-scale processes affect the plasma’s large-scale
evolution.
This section briefly overviews both the capabilities and the limitations of instru-
ments used to observe the solar wind in situ. Although a full treatment of the subject is
beyond the scope of this review, a basic understanding of these instruments is essential
for the proper scientific analysis of their measurements. Section 2.1 highlights some
significant heliospheric missions. Two sections are dedicated to in-situ observations
of thermal ions and electrons: Sect. 2.2 overviews the instrumentation, and Sect. 2.3
addresses the analysis of particle data. Sections 2.4 and 2.5 respectively discuss the
in-situ observation of the solar wind’s magnetic and electric fields. Section 2.6 presents
a short description of multi-spacecraft techniques.

2.1 Overview of in-situ solar-wind missions

In-situ plasma instruments were among the first to be flown on spacecraft. Gringauz
et al. (1960) used data from Luna 1, Luna 2, and Luna 3, which at the the time
were known as the Cosmic Rockets, to report the first detection of super-sonic solar-
wind ions as predicted by Parker (1958). These observations were soon confirmed by
Neugebauer and Snyder (1962), who used in-situ measurements from Mariner 2 en
route to Venus.
Since then, numerous spacecraft have carried in-situ instruments throughout the
heliosphere to observe the solar wind’s particles and fields. Table 2 lists a selection of
these missions grouped as completed, active, and future missions. The column “Radial
Coverage” lists the ranges of heliocentric distance for which in-situ data are available,
which are presented graphically in Fig. 8. Currently, Voyager 1 (Kohlhase and Penzo
1977) is the most distant spacecraft from the Sun—a superlative that it will continue
to hold for the foreseeable future. Helios 2 (Porsche 1977) held for several decades
the record for closest approach to the Sun, but, in late 2018, Parker Solar Probe (Fox
et al. 2016) achieved a substantially closer perihelion.

2.2 Thermal-particle instruments

Thermal particles constitute the most abundant but lowest-energy particles in solar-
wind plasma. Although no formal definition exists, the term commonly refers to
particles whose energies are within several (“a few”) thermal widths of the plasma’s
bulk velocity. We define these as protons with energies  10 keV and electrons with

123
5

Table 2 Select heliospheric missions: completed, active, and future

Mission Years activea Radial coverageb (au) Source

123
Luna 1, 2, and 3 1959–1959 ≈ 1.0c NSSDC; Johnson (1979)
Mariner 2 1962–1962 0.866–1.003 COHOWeb
Page 28 of 136

Pioneer 6 1965–1971 0.814–0.984 COHOWeb


Pioneer 7 1966–1968 1.010–1.126 COHOWeb
Pioneer 10 1972–1995 0.99–63.04 CDAWeb (PIONEER10_COHO1HR_MERGED_MAG_PLASMA)
Pioneer 11 1973–1992 1.00–36.26 CDAWeb (PIONEER11_COHO1HR_MERGED_MAG_PLASMA)
Pioneer Venus 1978–1992 0.72–0.73 CDAWeb (PIONEERVENUS_COHO1HR_MERGED_MAG_PLASMA)
ISEE-3 (ICE) 1978–1990 0.93–1.03 CDAWeb (ISEE-3_MAG_1MIN_MAGNETIC_FIELD)
Helios 1 1974–1981 0.31–0.98 CDAWeb (HELIOS1_COHO1HR_MERGED_MAG_PLASMA)
Helios 2 1976–1980 0.29–0.98 CDAWeb (HELIOS2_COHO1HR_MERGED_MAG_PLASMA)
Ulysses 1990–2009 1.02–5.41 CDAWeb (UY_COHO1HR_MERGED_MAG_PLASMA)
Cassini 1997–2017 0.67–10.07 COHOWeb; OMNIWeb Plus (helio1day)
STEREO B 2006–2014 1.00–1.09 CDAWeb (STB_COHO1HR_MERGED_MAG_PLASMA)

Voyager 1 1977– 1.01–140.71d CDAWeb (VOYAGER1_COHO1HR_MERGED_MAG_PLASMA)


Voyager 2 1977– 1.00–118.91d CDAWeb (VOYAGER2_COHO1HR_MERGED_MAG_PLASMA)
Wind 1994– 0.972–1.017 CDAWeb (WI_OR_PRE)
SOHO 1995– 0.972–1.011 CDAWeb (SO_OR_PRE)
ACE 1997– 0.973–1.010 CDAWeb (AC_OR_SSC)
New Horizons 2006– 11.268–42.775d CDAWeb (NEW_HORIZONS_SWAP_VALIDSUM)
STEREO A 2006– 0.96–0.97 CDAWeb (STA_COHO1HR_MERGED_MAG_PLASMA)
DSCOVR 2015– 0.973–1.007 CDAWeb (DSCOVR_ORBIT_PRE)
PSP 2018– 0.0459–0.25e,f Fox et al. (2016)
D. Verscharen et al.
Table 2 continued

Mission Years activea Radial coverageb (au) Source

Solar Orbiter 2020g,h 0.28–1.2e Müller et al. (2013)


IMAP 2024g 0.973–1.007i NASA Release 18-046
a Year of launch to final year (with non-fill data) in cited dataset
b Incomplete for some missions due to data gaps
c Exact range not available
d Distance still increasing; values on 2018-01-01 (Voyager 1), 2018-10-26 (Voyager 2), or 2018-10-31 (New Horizons)
e Anticipated radial coverage
f Perihelion of first three orbits: 0.163 au
The multi-scale nature of the solar wind

g Anticipated launch date


h https://www.esa.int/Our_Activities/Space_Science/Solar_Orbiter, accessed 2019-09-10
i Approximate radial coverage of the first Lagrangian point of the Earth–Sun system
Page 29 of 136

123
5
5 Page 30 of 136 D. Verscharen et al.

Fig. 8 Radial coverage of select heliospheric missions based on Table 2. Colors indicate the status of each
mission: completed (blue), active (green), and future (red). The colored bar for each mission does not reflect
any data gaps that may be present in its dataset(s). Mixed coloring has been used for PSP to reflect that, while
the mission is active, final radial coverage has not yet been achieved. Red arrows indicate that the radial
coverages of Voyager 1 and 2 and New Horizons are still increasing. Vertical lines indicate the semi-major
axes of the eight planets (black) and the dwarf planets Ceres, Pluto, and Eris (gray)

energies  100 eV under typical solar-wind conditions at 1 au. We note, however,


that most thermal-particle instruments cover a wider range of energies.
Although particle moments such as density, bulk velocity, and temperature are
useful quantities for characterizing the plasma, these parameters generally cannot
be measured directly. Instead, thermal-particle instruments measure particle spectra,
which give the distribution of particle energies in various directions. These spectra
must then be analyzed to derive values for the particle moments (see Sect. 2.3).
This section focuses on the basic design and operation of three types of thermal-
particle instruments: Faraday cups, electrostatic analyzers (ESAs), and mass spec-
trometers. Since particle acceleration beyond thermal energies is outside of the scope
of this review, we do not address instruments for measuring higher-energy particles.
Some other techniques and instruments exist for measuring thermal particles in
solar-wind plasma, but we omit extensive discussion of these since they generally
provide limited information about the phase-space structure of particle distributions.
For example, an electric-field instrument can be used to infer some electron properties
(especially density; see Sect. 2.5). Likewise Langmuir probes provide some electron
moments (Mott-Smith and Langmuir 1926). A series of bias voltages is applied to a
Langmuir probe relative either to the spacecraft or to another Langmuir probe. The
electron density and temperature can then be inferred from measurements of current

123
The multi-scale nature of the solar wind Page 31 of 136 5

Fig. 9 Simplified cross-sectional diagram of a Faraday cup for observing ions. The cup’s aperture is on
the right, its collector plate is on the left, and its three grids are indicated by dashed lines. A square-wave
voltage, E = E0 ± ΔE/2 > 0, is applied to the middle grid, which is known as the modulator. Blue
arrows indicate inflowing j-ions. Depending on vz , the normal component of the ion’s velocity, it is either
always accepted by the modulator (high speed), always rejected (low speed), or only accepted when the
modulator’s voltage is low (intermediate speed). The accepted ions produce a current at the collector plate,
which the detection system amplifies, demodulates, and integrates to measure, in effect, the current from
only the intermediate-speed ions according to Eq. (66)

at each bias voltage. The Cassini spacecraft included a spherical Langmuir probe
(Gurnett et al. 2004) along with other plasma instruments (Young et al. 2004).

2.2.1 Faraday cups

Faraday cups rank among the earliest instruments for studying space plasmas. His-
torically noteworthy examples include the charged-particle traps on Luna 1, Luna 2,
and Luna 3 (Gringauz et al. 1960) and the Solar Plasma Experiment on Mariner 2
(Neugebauer and Snyder 1962), which provided the first in-situ observations of the
solar wind’s supersonic ions. Since then, Faraday cups on Pioneer 6 and Pioneer 7
(Lazarus et al. 1966, 1968), Voyager 1 and 2 (Bridge et al. 1977), Wind (Ogilvie
et al. 1995), and DSCOVR (Aellig et al. 2001a) have continued to observe solar-wind
particles.
As depicted in Fig. 9, a Faraday cup consists of a grounded metal structure with an
aperture. A typical Faraday cup has a somewhat “squat” geometry with a wide aperture
so that it accepts incoming particles from a wide range of directions. For example,
the full-width half-maximum field of view of each of the Wind/SWE Faraday cups
is about 105◦ . At the back of the cup is a metal collector plate, which receives the
current I of the inflowing charged particles.
Figure 9 shows three of the fine mess grids that are placed between a Faraday cup’s
aperture and collector. The inner and outer grids are electrically grounded. A voltage E
is applied to the middle grid, known as the modulator, to restrict the ability of particles

123
5 Page 32 of 136 D. Verscharen et al.

to reach the collector. We define ẑ to indicate the direction into the Faraday cup so
that −ẑ is the cup’s look direction. Consider a j-particle of mass m j and charge q j
that enters the cup with a velocity v. For a modulator voltage E, the particle can only
reach the collector if the normal component of its velocity, vz = v · ẑ, is greater than
the cutoff speed
⎧

⎨ 2qj E
(c) if q j E > 0
v j (E) ≡ mj . (64)


0 else

When E and q j have opposite signs, the modulator places no restriction on the particle’s
ability to reach the collector.
Typically, the modulator is not kept at a constant voltage but rather alternated
between two voltages:

ΔE
E = E0 ± , (65)
2
where E0 is the offset and ΔE is the peak-to-peak amplitude. In this configuration, the
detector circuit is designed to use synchronous detection to measure the difference in
the collector current between the two states:

ΔE ΔE
ΔI (E0 , ΔE) = I E0 − − I E0 + . (66)
2 2

Essentially, ΔI is the current from particles whose velocities are sufficient for them
to reach the collector when the modulator voltage is low but not when it is high. This
method suppresses contributions to the collector current that do not vary with the
modulator voltage. These contributions include the signal from any particle species
with a charge opposite that of the modulator since, per Eq. (64), the modulator does
not restrict the inflow of such particles. This method also mitigates the effects of
photoelectrons, which are liberated from the collector by solar UV photons and whose
signal can exceed that of solar-wind particles by orders of magnitude (Bridge et al.
1960).
A set of E0 and ΔE values defines a voltage window. By measuring the differential
current ΔI for a series of these, a Faraday cup produces an energy distribution of
solar-wind particles. The size and number of voltage windows determine the spectral
resolution and range, which, for many Faraday cups, can be adjusted in flight to
accommodate changing plasma conditions. Since a Faraday cup is simply measuring
current, its detector electronics often exhibit little degradation with time. For example,
Kasper et al. (2006) demonstrate that the absolute gain of each of the Wind/SWE
Faraday cups (Ogilvie et al. 1995) drifts  0.5% per decade.
Various approaches exist to use Faraday cups to measure the direction of inflow-
ing particles, which is necessary for inferring parameters such as bulk velocity and
temperature anisotropy. The Voyager/PLS investigation (Bridge et al. 1977) and the
BMSW solar-wind monitor on SPECTR-R (Šafránková et al. 2008) include multiple

123
The multi-scale nature of the solar wind Page 33 of 136 5

Faraday cups pointed in different directions. DSCOVR/PlasMag (Aellig et al. 2001a)


has only a single Faraday cup but multiple collector plates: a split collector. Each
collector is off-axis from the aperture and thus has a slightly different field of view.
Pioneer 6, Pioneer 7 (Lazarus et al. 1966, 1968), and Wind (Ogilvie et al. 1995) are
spinning spacecraft, so their Faraday cups make measurements in various directions
as the spacecraft rotate.
A Faraday cup’s response function is a mathematical model for what the instrument
measures under different plasma conditions: i.e., an expression for ΔI as a function of
the particle distribution functions. For simplicity, we initially consider only one particle
species j and assume that the distribution function f j is, during the measurement cycle,
a function of v only. The number density of j-particles in a phase-space volume d3 v
centered on v is

dn j = f j (v) d3 v. (67)

The current that the Faraday cup measures from the particles in this volume is

dI j = q j vz A(θ, φ) dn j = q j vz A(θ, φ) f j (v) d3 v, (68)

where (v, θ, φ) are the spherical coordinates of v, and A(θ, φ) is the Faraday cup’s
effective collecting area as a function of particle-inflow direction.6 If the modulator
voltage spans the voltage window E0 ± ΔE/2, then the contribution of all j-particles
to the measured differential current is

(c)
v j (E0 +ΔE /2)
  ∞ ∞
ΔI j = dI j = q j dvz vz dv y dvx A(θ, φ) f j (v). (69)
(c)
v j (E0 −ΔE /2) −∞ −∞

Since a Faraday cup cannot distinguish current from different types of particles, the
measured current is

ΔI = ΔI j , (70)
j

where the sum is carried out over all particle species in the plasma.
Equations (69) and (70) provide the general form of the response function of a
Faraday cup. Section 2.3 overviews the process of inverting the response function to
determine the particle moments from a measured particle spectrum.

6 Typically, the function A(θ, φ) is calculated from the Faraday cup’s geometry and/or is measured in
ground testing. The value of A(θ, φ) is generally largest for θ = 0, when particles flow straight into the
cup, and then falls off as θ increases and less of the collector is “illuminated” by inflowing particles. If a
Faraday cup has an asymmetric shape and/or multiple collectors, A(θ, φ) will also depend on φ.

123
5 Page 34 of 136 D. Verscharen et al.

Fig. 10 Simplified cross-sectional diagram of a top-hat style electrostatic analyzer (ESA). The aperture
is shown on the upper left and right, and can provide up to 360◦ of coverage of azimuth φ. In contrast,
only particles within a limited range of elevation θ are able to pass through the curved collimator plates
and reach the detector. A DC voltage E is sustained between the plates and sets the sign and value of the
target energy per charge K /q j for incoming particles. The spacing between the collimator plates defines
the width of the energy windows

2.2.2 Electrostatic analyzers

Like Faraday cups, electrostatic analyzers (ESAs) have a long history of use in the
observation of thermal particles in the solar wind. Though ESAs are substantially
more complex than Faraday cups, they enable much more direct and detailed studies
of distribution functions (see Sect. 2.3.1). Additionally, they can be combined with
mass spectrometers (see Sect. 2.2.3) to directly probe the ion composition of the
plasma.
Figure 10 shows a simplified cross-section of the common top-hat design for an
ESA (Carlson et al. 1983). Such a device consists of two hemispherical shells that are
nested concentrically so as to leave a narrow gap between them. Particles enter via a
hole in the top of the larger hemisphere and are then subjected to the electric field that
is created by maintaining a DC voltage E between the two hemispheres. The value of
E and the curvature and spacing of the hemispheres define an energy-per-charge range
for an incoming particle to reach the detectors at the base of the hemispheres. If an
incoming particle has a kinetic energy K and charge q j , it can only reach the detectors
if the ratio K /q j falls within that range. To generate a particle spectrum, E is swept
through a series of values. The range of particle energies is set by the range of E values,
which, on most ESAs, can be adjusted in flight. Nevertheless, the width of an ESA’s
energy window ΔK /K 0 is fixed geometrically by the spacing between its collimator
plates. In contrast, the width of a Faraday cups’ energy window is adjustable in flight
since it is set by a voltage range according to Eq. (65).
An ESA’s detectors are typically arranged around the base of the hemispheres.
While Faraday cups detect incoming particles by measuring their net current, an
ESA’s detectors usually count particle cascades generated by the strikes from indi-
vidual particles. Such detectors would be impractical for a Faraday cup because they
would be overwhelmed by solar UV photons. On a top-hat ESA, the tight spacing

123
The multi-scale nature of the solar wind Page 35 of 136 5

of the deflectors and a low-albedo coating7 on their surfaces ensure that very few
photons reach the detectors. Each of the detectors is typically some type of electron
multiplier, which uses an electrostatic potential in such a way that a strike by a sin-
gle charged particle produces a cascade of electrons, which can then be registered.
Channel electron multipliers (CEMs) were used for ACE/SWEPAM (McComas et al.
1998a), while micro-channel plates (MCPs) were used for Wind/3DP (Lin et al. 1995)
and STEREO/IMPACT/SWEA (Sauvaud et al. 2008). Both CEM and MCP detectors
require more complex calibration than is needed for a Faraday cup. For example, after
each particle strike, an electron multiplier experiences a dead time, during which the
electron cascade is in progress and the detector cannot respond to another particle.
Furthermore, electron multipliers (and MCPs in particular) often exhibit significant
degradation in their efficiency with time.
A typical top-hat ESA has a fan-beam field of view. The size and number of detectors
define its azimuthal resolution and coverage, and ESAs can be designed with up to 360◦
of φ-coverage. In contrast, most ESAs only sample particles over a limited range of
elevation θ , and a number of strategies have been employed to provide θ -coverage. The
ESAs in the Helios plasma investigation (Schwenn et al. 1975; Rosenbauer et al. 1977)
and in Wind/3DP (Lin et al. 1995) were designed to rely on spacecraft spin to sweep
their fan beams. Although the Cassini spacecraft was three-axis stabilized, its CAPS
instrument suite was mounted on an actuator, which a motor rotated through about 180◦
of azimuth every 3 min (Young et al. 2004). The MAVEN spacecraft is likewise three-
axis stabilized, but its SWIA instrument (Halekas et al. 2015) incorporated a second
set of electrostatic deflectors to effectively steer its fan beam by adjusting the path of
ions entering the top hat. Finally, the unique design of MESSENGER/FIPS (Andrews
et al. 2007) moved beyond the top hat to give that instrument wide θ -coverage (versus
a fan beam) but reduced aperture size.
For any given value of E, each ESA detector essentially has its own effective collect-
ing area A j (K , θ, φ), which depends on the energy K = m j v 2 /2 and direction (θ, φ)
of incoming j-particles. The number of j-particles detected from an infinitesimal
volume d3 v of phase-space during a time interval Δt is

dN j = Δt v A j (K , θ, φ) dn j , (71)

where dn j is the number density of j-particles in d3 v. Substituting Eq. (67) and


converting to spherical coordinates gives
2 Δt
dN j = A j (K , θ, φ) f j (K , θ, φ)K sin θ dK dθ dφ, (72)
m 2j

where f j has been parameterized in energy and direction rather than vector velocity.
The total number of j-particles detected in Δt is
 ∞ π 2π
2 Δt
ΔN j = dN j = 2 dK K dθ sin θ dφ A j (K , θ, φ) f j (K , θ, φ). (73)
mj
0 0 0

7 For example, gold black was used on the Wind/3DP ESAs (Lin et al. 1995).

123
5 Page 36 of 136 D. Verscharen et al.

Formally, the integrals in Eq. (73) are carried out over all energies and directions (i.e.,
all of phase space) but most ESAs are designed so that a given detector is only sensitive
to particles from a relatively narrow range of energies and directions. Consequently,
the detector’s effective collecting area is often approximated as

⎨ A0 if |K − K 0 | < ΔK , |θ − θ0 | < Δθ, |φ − φ0 | < Δφ
A j (K , θ, φ) ≈ sin θ0 ,

0 else
(74)

where A0 is the nominal collecting area, (θ0 , φ0 ) is the look direction, Δθ and Δφ set
the field of view, and K 0 and ΔK set the energy range of j-particles. Using Eq. (74)
and assuming that ΔK , Δθ , and Δφ are small relative to variations in f j (K , θ, φ),
we approximate Eq. (73) as

2 A0 K 0 2K 02
ΔN j ≈ Δt ΔK Δθ Δφ f j (K 0 , θ0 , φ 0 ) ≈ G f j (K 0 , θ0 , φ0 ), (75)
m 2j m 2j

where
ΔK
G ≡ A0 Δt Δθ Δφ (76)
K0

is known as the geometric factor. ESAs are often designed and operated in such a way
that G is approximately constant.
If an ESA does not have any mass-spectrometry capability (see Sect. 2.2.3), then
each of its detectors measures the count of all particles of any species that reach it.
Thus, the measured quantity is

ΔN = ΔN j , (77)
j

where the sum is carried out over all particle species j.


Equations (73) and (77) specify the response function of a top-hat ESA. A particle
spectrum from such an instrument consists of a set of measured ΔN -values made over
various E-values and in various directions. Section 2.3 describes how the response
function can be used to extract information about particle distribution functions from
a measured spectrum.

2.2.3 Mass spectrometers

As noted above, neither a Faraday cup nor an ESA can, on its own, directly distinguish
among different ion species: they simply measure the current and counts, respectively,
of the incoming particles. A limited composition analysis, though, is still possible
because the voltage E needed for either type of instrument to detect a j-particle of
speed v is proportional to m j /q j . Though relative drift is often observed among

123
The multi-scale nature of the solar wind Page 37 of 136 5

different particle species in the solar wind, it generally remains far less than the bulk
speed (see Sect. 1.4.4). Thus, in a particle spectrum, the signals from different particle
species appear shifted by their mass-to-charge ratios. By separately analyzing these
signals (see Sect. 2.3), values can be inferred for the moments of the various particle
species.
This strategy does have significant limitations. First, it provides no mechanism
for distinguishing ions with the same mass-to-charge ratio (e.g., 12 C3+ and 16 O4+ ).
Second, even when particle species have distinct mass-to-charge ratios, ambiguity
can still arise from the overlap of their spectral signal. For example, the mass-to-
charge ratios of protons and α-particles differ enough that values for their moments
can often be derived for both species from Faraday-cup (e.g., Kasper 2002, Chapter
4) and ESA (e.g., Marsch et al. 1982b) spectra. Nevertheless, the α-particle signal
can suffer confusion with minor ions (e.g., Bame et al. 1975), and, especially at low
Mach numbers, the proton and α-particle signals can almost completely overlap (e.g.,
Maruca 2012, Sect. 3.3).
A mass spectrometer is required to achieve the most accurate measurements of
solar-wind composition (see also the more complete review by Gloeckler 1990). As
opposed to being a separate instrument, a mass spectrometer is typically incorporated
into an ESA as its detector system and is used to measure the speed of each particle.
The ESA ensures that only particles within a known, narrow range of energy per
charge pass through. As each particle enters the mass spectrometer, an electric field
accelerates it by a known amount. The particle then triggers a start signal by liberating
electrons from a thin foil,8 which are detected via an MCP. Next, the particle travels
a known distance Δs to another foil.9 The particle triggers a stop signal by passing
through this latter foil before finally reaching the detector. The time Δt between the
start and stop signals is the particle’s time of flight, a measurement of which allows
the particle’s speed v = Δs/Δt through the mass spectrometer to be inferred.
Several different designs have been developed for mass spectrometers for helio-
physics. In a time-of-flight versus energy (TOF/E) mass spectrometer, such as
Ulysses/SWICS (Gloeckler et al. 1992), ACE/SWICS (Gloeckler et al. 1998, Sect. 3.1),
and STEREO/IMPACT/PLASTIC (Galvin et al. 2008), solid-state detectors (SSDs)
are used to ultimately detect each ion. Unlike an electron multiplier, an SSD is able
to measure the energy of individual charged particles. Therefore, a TOF/E instrument
measures each ion’s initial energy per charge, speed through the instrument, and resid-
ual energy at the detector. Together, these quantities provide sufficient information to
determine the ion’s mass, charge, and initial speed. In contrast, a high-mass-resolution
spectrometer (HMRS) such as ACE/SWIMS (Gloeckler et al. 1998, Sect. 3.2) does
not need to measure the ions’ residual energy and can simply use MCP detectors.
An HMRS exploits the fact that passing through the start foil tends to decrease an
ion’s charge state to either 0 or + 1. The particle then passes through a known but
non-uniform electric field, which deflects the singly ionized particle to the detectors.

8 For example, a carbon foil supported by a nickel mesh was used on Ulysses/SWICS (Gloeckler et al.
1992), ACE/SWICS (Gloeckler et al. 1998), and STEREO/IMPACT/PLASTIC (Galvin et al. 2008).
9 For example, the SWICS instruments on both Ulysses and ACE (Gloeckler et al. 1992, 1998) use a gold
foil applied directly to the top of the detectors.

123
5 Page 38 of 136 D. Verscharen et al.

The electric field causes the time of flight to be mass dependent, so each particle’s
mass can be inferred.

2.3 Analyzing thermal-particle measurements

A particle spectrum, whether measured by a Faraday cup or an ESA, must be processed


in order to extract information about the observed particles. This involves inverting the
instrument’s response function—Eqs. (69) and (70) for a Faraday cup, and Eqs. (73)
and (77) for an ESA—so that particle moments or phase-space densities can be derived
from measured current or counts. This section briefly describes three methods for
achieving this: distribution-function imaging, moments analysis, and fitting of model
distribution functions.

2.3.1 Distribution-function imaging

Equation (75) suggests a very simple method for interpreting a particle spectrum from
an ESA. The number of counts ΔN j of j-particles is approximately proportional to
the value of the j-particles’ distribution function f j at some point in phase space. If
only j-particles are considered, then the set of measured ΔN -values (i.e., the particle
spectrum) can be used to give a set of values for f j across phase space. In this sense,
an ESA’s particle spectrum can be thought of as an image of a distribution function.
This is the method employed by Marsch et al. (1982a, b) in their well-known contour-
plots of proton and α-particle distribution functions from the Helios mission (see also
Figs. 5, 6 of this review). Since this technique is not focused on extracting the values
of particle moments, it is especially well suited to studying the three-dimensional
structure of distribution functions and non-Maxwellian features.
Nevertheless, distribution-function imaging carries significant limitations. First, in
the case of ion measurements, significant confusion can arise among the various ion
species in the plasma (see Sect. 2.2.3). If an ESA does not have a mass spectrometer,
it simply measures the total count of particles ΔN rather than each individual ΔN j .
Second, various assumptions are made in deriving Eq. (75). Notably, the field of view
and energy range were taken to be small relative to the scale of variations in the
distribution function. When these assumptions break down, this technique returns a
distorted image of f j . Third, this technique cannot be applied to observations from
a Faraday cup. Essentially, a Faraday cup’s large field of view means that each of its
ΔI -measurements samples a large region of phase space. The integrals in Eq. (69)
cannot be easily simplified to give an expression like Eq. (75).
Though ESA images of distribution functions can provide tremendous insight into
phase-space structure, care must be exercised to properly account for instrumental
effects. Any ESA has finite angular and energy resolutions, which must be considered
when interpreting their output. An irregularity in a distribution function may seem
significant in a contour plot but actually result from only a single datum with a low
number of particle counts. Such finite-resolution effects are often more pronounced in
proton versus electron data because protons, being supersonic, are concentrated into a
narrow beam of phase space. A related effect arises in both ion and electron data from

123
The multi-scale nature of the solar wind Page 39 of 136 5

the finite period of time required for an ESA to sweep through its angular and energy
ranges. Especially during periods of high variability in the solar wind, this may result
in distribution-function images that constitute “hybrids” of distinct plasma conditions.

2.3.2 Moments analysis

Moments analysis provides the most direct method for estimating particle moments
from a measured particle spectrum. Essentially, this technique relies on deriving rela-
tionships between the moments of a distribution function (see Sect. 1.4.1) and the
moments of the measured quantity: ΔI j for a Faraday cup or ΔN j for an ESA. For
the latter case, Eq. (75) shows that ΔN j is approximately proportional to f j . Thus,
each moment of f j can be approximated with a discrete integral of ΔN j : a sum over
all the measured ΔN -values. For a Faraday cup, the relationship between ΔI j and f j
in Eq. (69) is more complex, but similar expressions exist to relate the moments of
f j to sums of the measured ΔI -values (see, e.g., Kasper et al. 2006, Appendix A). In
either case, the calculations are relatively simple. For this reason, moments analyses
are commonly implemented in spacecraft flight computers, which often have limited
computational resources or limited down-link bandwidth for the transmission of full
particle spectra.
Moments analysis carries the significant limitation that it provides no mechanism
for easily distinguishing different components of a distribution function (e.g., its core
and beam), or, in the case of ions, for differentiating among species (see Sect. 2.2.3).
Additionally, the particle spectrum must provide excellent coverage of f j in phase
space so that the discrete integrals of the measured ΔI - or ΔN -values can reasonably
approximate the infinite integrals of f j that define its moments.

2.3.3 Fitting model distribution functions

In a fitting analysis of a particle spectrum, a model distribution (such as those defined in


Sect. 1.4.3) is chosen for each f j -component and particle species under consideration.
These model distributions are then substituted into the expression for ΔI for a Faraday
cup in Eq. (70) or ΔN for an ESA in Eq. (77). This substitution gives an expression
for the measured quantity, ΔI or ΔN , in terms of the fit parameters of the model
distributions: e.g., particle densities, velocities, and temperatures. This model can
then be fit to a measured spectrum to derive estimates of the particle moments.
Unlike moments analysis, fitting allows for the direct treatment of multiple f j -
components or ion species. It also allows data to be weighted based on the uncertainty
in each measurement and does not require that the particle spectrum cover almost all
of phase space. Indeed, Kasper et al. (2006) use the microkinetic limits on temperature
anisotropy to infer that fitting model distribution functions to ion measurements from
the Wind/SWE Faraday cups produces temperature values that are significantly more
accurate than those returned from a moments analysis.
The greatest disadvantage of fitting is the need to assume a model distribution. If
such a model does not capture all of the features of the actual distribution function, the
fitting results are unreliable. In addition, the complexity of the functions involved

123
5 Page 40 of 136 D. Verscharen et al.

usually necessitates the use of non-linear fitting algorithms (e.g., the Levenberg–
Marquardt algorithm; see Marquardt 1963), which are computationally intensive and
generally cannot be implemented on spacecraft computers.

2.4 Magnetometers

This section provides a brief overview of the three types of magnetometers most
commonly used on heliophysics missions: search-coil magnetometers, fluxgate mag-
netometers, and helium magnetometers. The reviews by Ness (1970), Acuña (1974,
2002), and Smith and Sonett (1976) provide much more detailed treatments of these
and other types of magnetometers.

2.4.1 Search-coil magnetometers

Though simpler in design than fluxgate and helium magnetometers, search-coil mag-
netometers have been less frequently flown on space-physics missions because of
their poor sensitivity to background magnetic fields and low-frequency magnetic fluc-
tuations. The search-coil magnetometer was first used in space on Pioneer 1 (Sonett
et al. 1960). Later, search coils were included in Wind/Waves (Bougeret et al. 1995),
Cluster/STAFF (Cornilleau-Wehrlin et al. 1997), and Themis/SCM (Roux et al. 2008).
Essentially, a search-coil magnetometer is a coil of wire that wraps around a portion
of a core made from a high-permeability material, which serves to amplify the magnetic
field. Let Bext denote the magnetic field external to the core, which is to be measured.
The magnetic field inside the core is

Bint = μc Bext , (78)

where μc is the effective relative permeability of the core. One complication is that
μc differs from μr , the relative permeability of the bulk material comprising the core.
In general,

μr
μc = , (79)
1 + Nd (μr − 1)

where Nd is the demagnetization factor, which reflects the core’s particular geometry
(see, e.g., Tumanski 2011, Sect. 2.4.3). For materials with relatively low permeability,
μc ≈ μr , but materials with high μr are usually favored for search coils as they
substantially boost sensitivity.
If the coil has N turns, then, by Faraday’s law according to Eq. (23), the voltage
induced in the coil is

N Aμc dBext,z
E =− , (80)
c dt

where A is the core’s cross-sectional area, and the core is oriented along the z-axis.
Thus, a measurement of E gives the rate of change in the axial component of Bext . If

123
The multi-scale nature of the solar wind Page 41 of 136 5

Bext,z (t) is sinusoidal,

Bext,z (t) = B0,z cos (2π νt + φ) , (81)

the coil voltage is

2π νN Aμc B0,z
E(t) = sin (2π νt + φ) . (82)
c

A single coil can only detect fluctuations in the Bext component parallel to the coil’s
axis. Thus, search-coil magnetometers often include three orthogonal coils to enable
measurements of the vector magnetic field.
The factor of ν in Eq. (82) indicates that a search coil’s sensitivity scales linearly
with frequency. Search-coil magnetometers are thus mostly used in the frequency
range from a few Hz to several kHz. A non-accelerating search coil is completely
insensitive to the background magnetic field. However, a search-coil magnetometer
on a spinning spacecraft can still measure a constant field since the field is non-constant
in the instrument’s frame of reference. This method was employed on Pioneer 1 to
make the first measurements of the interplanetary magnetic field (Sonett et al. 1960;
Rosenthal 1982).

2.4.2 Fluxgate magnetometers

The fluxgate magnetometer was first invented for terrestrial use by Aschenbrenner
and Goubau (1936), and since then, it has become the most widely used type of
magnetometer in heliophysics missions. Although the fluxgate magnetometer is more
complex than the search-coil magnetometer, it is much better suited to measuring the
background magnetic field and low-frequency ( 10 Hz) magnetic fluctuations.
A fluxgate magnetometer relies on the hysteresis of ferromagnetic materials. The
center-left plot in Fig. 11 shows an idealized representation of the hysteresis curve
for such a material. The magnetic field B inside the material depends not only on
the auxiliary field10 H applied to it but also on the history of the core’s magnetiza-
tion. Nevertheless, there exists a critical H -value, Hc , such that the magnetic field is
saturated at a strength Bs if |H| ≥ Hc .
In a typical design, a fluxgate magnetometer consists of a ferromagnetic core
wrapped by two coils of wire: a drive coil and a sense coil. A triangle-wave current is
applied to the drive coil to produce an auxiliary field Hd (t) that has an amplitude H0
and period Π (upper-left plot in Fig. 11). The core’s total auxiliary field is then

H (t) = Hd (t) + ΔHz , (83)

10 Unfortunately, no widely accepted term for H exists. Some authors (e.g., Jackson 1975) refer to it as the
“magnetic field” and use another term for B. Although there is some historical precedent for this naming
convention, Sommerfeld (1952) and Griffiths (2013) strongly criticize it and contend that B is the more
fundamental parameter. We follow the convention used widely in modern space physics of referring to B
as the “magnetic field.” For H, we choose the term “auxiliary field” from Griffiths (2013).

123
5 Page 42 of 136 D. Verscharen et al.

Fig. 11 The performance of an idealized, basic fluxgate magnetometer. The hysteresis plot of the fluxgate’s
ferromagnetic core is shown in the center left and indicates the magnetic field B in the core as a function
of the auxiliary field H applied to it. The value of H is the sum of the auxiliary field Hd from the fluxgate
magnetometer’s drive coil and the auxiliary field ΔHz associated with the magnetic field external to the
instrument. The upper-left plot shows Hd (t), and ΔHz is represented as a horizontal shift between the
two left plots. The value of ΔH has been greatly exaggerated for illustrative purposes. The H -values
for which the core is saturated are indicated by light-blue shading, and the times t when this occurs are
indicated by light-red shading. The center-right plot shows the core’s magnetic field B(t), which is limited
by the saturation value Bs . The lower-right plot shows the voltage Es (t) that B(t) induces in the fluxgate
magnetometer’s sense coil. After Ness (1970)

where the z-direction corresponds to the axis of the core, and ΔHz represents the
contribution of the external magnetic field, which is to be measured. The value of
H0 is chosen to be large enough that the core experiences both positive and negative
saturation during each cycle of Hd (t). As a result, the core’s magnetic field B(t) has
the form of a truncated triangle wave (center-right plot in Fig. 11). A non-zero value
of ΔHz produces a DC offset in B(t), which means that the core spends different
amounts of time in positive and negative saturation. By Faraday’s law according to
Eq. (23), the voltage induced in the fluxgate magnetometer’s sense coil is

123
The multi-scale nature of the solar wind Page 43 of 136 5

Ns A dB
Es = − , (84)
c dt
where Ns is the number of turns in the sense coil, and A is the core’s cross-sectional
area. Because of the offset and truncation in B(t), Es (t) has the form of an irregular
square wave (lower-right plot in Fig. 11). We denote the duration of a positive or
negative pulse as αΠ and the time from the start of a positive pulse to the start of the
next negative pulse as βΠ . Then,

Hc
α= (85)
4H0

and

1 ΔH
β= 1− . (86)
2 H0

Typically, the value of H0 is chosen so that it is substantially greater than ΔHz and
Hc , in which case both α and β are much less than one. The sense-coil voltage shown
in Fig. 11 (lower right) has the Fourier series expansion (Ness 1970)
∞ 
  sin (π αk) 2π kt
Es (t) = E0 1 − e−i2πβk cos , (87)
πk Π
k=1

where
2Ns ABs
E0 = − . (88)
cαΠ
In the absence of an external magnetic field, the values of ΔH and β would both be
zero, which would cause all even harmonics in the above series to vanish. Thus, the
second harmonic is typically measured in order to infer the value of ΔHz and thereby
the value of Bz .
A single fluxgate sensor, like a single search-coil, is only sensitive to one component
of the magnetic field. Consequently, fluxgate magnetometers often consist of three
orthogonal sensors so that the vector magnetic field can be measured.
A fluxgate magnetometer can be used to measure the background magnetic field and
low-frequency magnetic fluctuations up to a few 10’s of Hz (Ness 1970) but it has poor
sensitivity to fluctuations around or above the frequency of its drive coil. Consequently,
some missions carry not only fluxgate magnetometers but also search-coil magnetome-
ters, which are better suited to measuring high-frequency magnetic fluctuations. For
example, the Wind spacecraft includes both the MFI fluxgate magnetometers (Lepping
et al. 1995) and the Waves search-coil magnetometers (Bougeret et al. 1995). Likewise,
the four Cluster spacecraft include the FGM fluxgate magnetometers (Balogh et al.
1997) and the STAFF search-coil magnetometers (Cornilleau-Wehrlin et al. 1997).

123
5 Page 44 of 136 D. Verscharen et al.

More sophisticated designs for fluxgate magnetometers, which include additional


coils and more complex geometries for the core, have been developed to improve sensi-
tivity and to allow the instrument to be operated at higher frequencies. Notably, Geyger
(1962) introduced the use of toroidal cores, which were used, e.g., for the Pioneer 11
magnetometer (Acuña 1974), Voyager/MAG (Behannon et al. 1977), Wind/MFI (Lep-
ping et al. 1995), and STEREO/IMPACT/MAG (Acuña et al. 2008).

2.4.3 Helium magnetometers

Helium magnetometers belong to a large class of magnetometers known as optically


pumped magnetometers (Ness 1970; Acuña 2002). Though some optically pumped
magnetometers use the vapor of an alkali metal (e.g., sodium, cesium, or rubidium)
as their sensing medium, helium has been more widely used in space instruments.
The sensing element of a helium magnetometer is a cell containing helium gas
(Slocum and Reilly 1963). A radio-frequency oscillator is used to energize electrons
in the gas, which collisionally excite helium atoms from their ground state, 11 S0 , to
their first excited state, 23 S1 . Since 11 S0 is a singlet state, and 23 S1 is a triplet, the
transition between them via photon emission/absorption is doubly forbidden under
classical selection rules. As a result, the 23 S1 state is metastable.
Although collisional excitation produces equal populations for the three 23 S1 sub-
levels, optical pumping produces unequal populations for this triplet (Colegrove and
Franken 1960). A helium lamp serves a source of 1083 nm photons. This light is
then columnated into a beam, which passes through a circularly polarized filter before
reaching the cell. The 1083 nm wavelength corresponds to a helium atom’s transition
between the 23 S1 triplet state and the three closely-spaced 23 P states: 23 P0 , 23 P1 , 23 P2 .
A helium atom in the 23 S1 state can transition to a 23 P state by absorbing one of these
photons, after which it returns to 23 S1 via remission. However, since the photons are
circularly polarized, the atom, in the presence of a magnetic field, will preferentially
return to one of the 23 S1 sub-levels over the other two.
An infrared detector is used to measure how much of the helium lamp’s light
is able to pass through the cell. The transparency of helium to 1083 nm photons
depends directly on the pumping efficiency, which in turn varies with the strength
of the magnetic field and the field’s angle with respect to the beam path. Thus, the
magnetic field can be inferred from measurements of the intensity of transmitted light.
A vector helium magnetometer typically includes three orthogonal pairs of
Helmholtz coils so that an arbitrary magnetic field can be applied to the cell in addition
to the external magnetic field that is to be measured. In the usual operating mode, a
constant-magnitude magnetic field is rotated relative to the beam path at a frequency
of a few 100’s of Hz. This results in a periodic variation in the intensity of transmitted
light. For a full vector measurement of the external magnetic field, the applied mag-
netic field is rotated through two orthogonal planes, each of which has an axis parallel
to the beam path.
Vector helium magnetometers have been used on some heliophysics missions but
not as many as fluxgate magnetometers. In general, helium magnetometers are more
complex and often require more mass and power than fluxgate magnetometers (Acuña
2002). Nevertheless, helium magnetometers are effective for measuring strong mag-

123
The multi-scale nature of the solar wind Page 45 of 136 5

netic fields, which makes them useful for planetary missions such as Pioneers 10 and 11
(Smith et al. 1975). ISEE-3 (later renamed ICE; Frandsen et al. 1978) also carried a
vector helium magnetometer. Some missions, including Ulysses (Balogh et al. 1992)
and Cassini (Dunlop et al. 1999; Dougherty et al. 2004), carried both vector helium
and fluxgate magnetometers. The helium magnetometer on Cassini was unique in that
it could be operated in either a scalar or vector mode (i.e., measure either B or B). This
design was developed to improve measurements of Saturn’s strong magnetic field.

2.5 Electric-field measurements

Measurements of the vector electric field E in the solar wind are typically made over
a very wide range of frequencies from a few kHz to tens of MHz. The most common
probes of E are monopole and dipole antennas, the lengths of which can vary based
on scientific goals and practicalities. For example, the length (spacecraft to tip) of
each STEREO/Waves antenna is 6 m (Bale et al. 2008; Bougeret et al. 2008), while
Wind/Waves has antennas that are 7.5 m and 50 m long (Bougeret et al. 1995).
Electric-field instruments for heliophysics missions often utilize multiple receivers.
This not only helps to accommodate the wide range of frequencies but also allows
for different observation modes to be implemented. The simplest mode is waveform
capture, in which a time series of voltage measurements from each antenna is recorded.
This mode preserves the most information about E(t) but produces large amounts of
data and thus is generally used only as a burst mode. An alternative mode is spectrum
capture, in which only the power spectral density is recorded at a predetermined set
of frequencies. This significantly lowers the data volume while preserving frequency
information. As a matter of practice, this mode is often implemented with a narrow-
band receiver that is stepped through a series of discrete frequency ranges to measure
the total power in each.
Electric-field instruments also have uses beyond simply measuring E for its own
sake. Although these applications are beyond the scope of this review, two merit
brief mention here. The first is the measurement of the quasi-thermal noise spectrum,
which can be used to infer the properties of electrons (Meyer-Vernet and Perche
1989). When an antenna is surrounded by a plasma, the antenna’s frequency response
is altered in a predictable way at frequencies near the electron plasma frequency

ωpe . As shown in Eq. (7), ωpe is proportional to n e , so the determination of ωpe
from the quasi-thermal noise spectrum is a direct measure of the electron density
n e . In addition, the temperature and some non-thermal properties of electrons can be
extracted from the shape of the quasi-thermal noise spectrum. Second, antennas can
be used very effectively as dust detectors because of the large size of the antennas and
the distinctive electrical signal produced by a dust grain striking an antenna (Couturier
et al. 1981; Le Chat et al. 2009). The abundance and size-distribution of dust particles
have been studied using measurements from STEREO/Waves (Zaslavsky et al. 2012)
and Wind/Waves (Kellogg et al. 2016).

123
5 Page 46 of 136 D. Verscharen et al.

2.6 Multi-spacecraft techniques

Most of the observational results presented in this review are based on measurements
from individual spacecraft. Nevertheless, powerful techniques have been developed
to analyze simultaneous in-situ measurements from multiple spacecraft to distinguish
between spatial and temporal fluctuations in the plasma. This section offers a brief
description of the key concepts.
Spacecraft separated by relatively large distances ( 0.1 au) offer particular benefits
for observing remote or large-scale phenomena. For example, the primary motivation
of the aptly named STEREO mission (Kaiser et al. 2008) was to provide stereoscopic
observations of the Sun and the inner heliosphere. The in-situ particle instruments of
the PLASTIC suite were designed for studies of the temporal and spatial variations
of ICMEs (Galvin et al. 2008). Likewise, the Waves investigation allowed for the
triangulation (radiogoniometry) of radio-burst source regions (Bougeret et al. 2008,
Sect. 3.4), which has also been achieved using spacecraft from separate missions
(Steinberg et al. 1984; Hoang et al. 1998; Reiner et al. 1998).
Constellations of spacecraft with tighter spacings are used to observe local or small-
scale plasma phenomena, especially in Earth’s magnetosphere and magnetosheath.
This approach was largely pioneered with the Cluster mission (Escoubet et al. 1997)
and later employed and expanded upon for THEMIS/ARTEMIS (Angelopoulos 2008)
and MMS (Burch et al. 2016). In each of these missions, at least four spacecraft were
flown in a quasi-tetrahedral formation to utilize three basic techniques (Dunlop et al.
1988):

– In curlometry, a four-point measurement of the magnetic field B is used to estimate


∇ × B and thereby the current density j (Robert et al. 2000). This technique
relies on j being nearly uniform within the tetrahedron, so it is best suited to
study phenomena on spatial scales of order or larger than the dimension of the
constellation.
– For the wave-telescope technique, a Fourier analysis of B-measurements from the
four spacecraft is made to determine the frequency spectrum, directional distri-
bution, and mode of plasma fluctuations (Neubauer and Glassmeier 1990; Pinçon
and Motschmann 2000; Motschmann et al. 2000). Due to effects such as aliasing,
this method is most accurate in characterizing waves comparable in scale to the
spacecraft constellation (Sahraoui et al. 2010a).
– In a discontinuity analysis, the arrival times of a magnetic discontinuity (e.g., a
shock) at the spacecraft are compared so that the discontinuity’s orientation and
velocity can be inferred (Russell et al. 1983; Mottez and Chanteur 1994; Dunlop
and Woodward 2000). This method is most accurate for discontinuities whose
boundary regions are thin relative to the spacecraft separations.

3 Coulomb collisions

Collisions among particles provide the fundamental mechanism through which an


ionized or neutral gas increases its entropy and ultimately comes into thermal equi-

123
The multi-scale nature of the solar wind Page 47 of 136 5

librium. In a fully ionized plasma, hard scatterings rarely occur; instead, Coulomb
collisions, in which charged particles slightly deflect each other, are the primary colli-
sional means by which particles exchange momentum and energy. The solar wind’s low
density ensures that the rates of particle collisions remain relatively low. In contrast,
the denser plasma of the solar corona has a much higher collision rate, and collisional
processes are understood to be an important ingredient in the heating and acceleration
of coronal plasma (see Sect. 3.1). Unfortunately, this has led to the widespread mis-
conception that, beyond the solar corona, Coulomb collisions have no impact on the
evolution of solar-wind plasma. In reality, while collision rates in the solar wind can
be very low, the effects of collisions on the plasma never truly vanish.
This section overviews the effects that Coulomb collisions have on the microki-
netics and large-scale evolution of solar-wind plasma through interplanetary space.
Section 3.1 provides a simple dimensional analysis of Coulomb collisions, while
Sect. 3.2 overviews the more complete kinetic theory of particle collisions in plasmas.
Section 3.3 describes observations of solar-wind collisional relaxation.

3.1 Dimensional analysis of Coulomb collisions

Before addressing the detailed kinetic treatment of collisions, we use dimensional


analysis to derive a very rough expression for the rate of collisions in a plasma among
particles of the same species.
We consider a species whose particles have mass m j and charge q j . The j-particles
may be approximated as all traveling at the species’ thermal speed w j . When a pair
of j-particles collide, kinetic energy is temporarily converted into electric potential
energy. Assuming (very crudely) that this conversion is complete,

1 q 2j
2 m j w 2j = , (89)
2 xmin

where xmin is the particles’ distance of closest approach. Consequently,

πq 4j
σ ≡ π xmin
2
= (90)
m 2j w 4j

is the scattering cross-section for collisions among j-particles.


We now consider a volume V containing N j of the j-particles. The average time
t j that a j-particle goes between collisions is roughly equal to the time that it takes
to sweep out 1/N j of the total volume. Taking σ to be the particle’s effective cross-
sectional area,

1 V
= = σwjtj, (91)
nj Nj

123
5 Page 48 of 136 D. Verscharen et al.

where n j is the number density of j-particles. Thus,

1/2 3/2
1 m 2j w 3j 23/2 m j kB T j
tj = = = . (92)
n jwjσ πq 4j n j πq 4j n j

Though Eq. (92) was derived from a naïve treatment of Coulomb collisions, it can
be used to approximate the collisionality of a species such as protons. For example, at
r = 1 au from the Sun, n p ∼ 3 cm−3 and Tp ∼ 105 K. These correspond to a proton
collisional timescale of tp ∼ 108 s, which is substantially longer than the solar wind’s
typical expansion time to this distance; see Eq. (1). In contrast, in the middle corona
(see Fig. 2), n p ∼ 108 cm−3 and Tp ∼ 106 K, which give tp ∼ 350 s. These estimates,
though very rough, reveal that collisional effects have substantially more impact on
coronal versus solar-wind plasma.
The stark difference in collisionality between the solar corona and solar wind forms
the basis of exospheric models of the heliosphere. Although these models fall beyond
the scope of this review, they warrant some mention. Since the early work on exospheric
models by Jockers (1968, 1970) and Lemaire and Scherer (1971a, b), they have been
shown to account for some features of the interplanetary solar wind. For example, the
preferential heating of minor ions in a coronal exosphere can lead to the preferential
acceleration of these ions (Pierrard et al. 2004). Maksimovic et al. (2005) offer a
more complete overview of exospheric models, and the reviews by Marsch (1994) and
Echim et al. (2011) provide an even more detailed treatment of the subject.

3.2 Kinetic theory of collisions

A full treatment of the kinetic theory of collisions in plasmas is beyond the scope of
this review. Instead, this section serves as a brief description of how the collisional
term of the Boltzmann equation is used to derive collision rates for particle moments.
More complete presentations of the theory are given by Spitzer (1956), Longmire
(1963), Braginskii (1965), Wu (1966), Burgers (1969), Krall and Trivelpiece (1973,
Chapters 6 and 7), Schunk (1975, 1977), Lifshitz and Pitaevskii (1981, Chapter 4),
Klimontovich (1997), and Fitzpatrick (2015).

3.2.1 The collision term

Discussions of particle collisions in gases usually begin with the Boltzmann equation
(19) since the effects of collisions are neatly grouped into the collision term on the
right-hand side of the equation:

∂ fj ∂ fj ∂ fj δ fj
+v· +a· = , (93)
∂t ∂x ∂v δt c

where the derivative (δ/δt)c is known as the collision operator. The separation of
the collision term from the terms on the left-hand side becomes somewhat murky
for plasmas. Coulomb collisions occur through the interaction of the particle electric

123
The multi-scale nature of the solar wind Page 49 of 136 5

Fig. 12 Diagram of a j-particle scattering off of an i-particle via the electric force in the i-particle’s reference
frame, in which the j-particle has an initial velocity g ji and a final velocity gji ; see Eqs. (95) and (96)

fields, but the plasma’s background electric field contributes to the acceleration a. The
particle electric field is the field generated by a single particle, while the background
electric field is the collective result of all neighboring charged particles. Ultimately, the
distinction between collisions and the effects of the background fields is phenomeno-
logical. Under the molecular chaos hypothesis (or stoßzahlansatz), collisions among
particles are assumed to be uncorrelated and to occur randomly (Maxwell 1867).
To derive an expression for the collisional term, we consider the Coulomb scattering
of a j-particle off of an i-particle via the electric force. We define the particles’ initial
velocities as v j and vi , their final velocities as vj and vi , their masses as m j and m i ,
and their charges as q j and qi . We note that the j- and i-particles may be of the same
species. The center-of-mass velocity of the two particles is

m j v j + m i vi m j vj + m i vi
u ji ≡ = ≡ uji , (94)
m j + mi m j + mi

which is unchanged by the collision. Figure 12 depicts this scattering event in the
i-particle’s frame of reference, in which the j-particle has an initial velocity

g ji ≡ v j − vi (95)

and a final velocity

gji ≡ vj − vi . (96)

We denote the impact parameter as b and the scattering angle as θ . In a Coulomb


collision, these two quantities are related by

θ q j qi
tan = , (97)
2 m ji g 2ji b

123
5 Page 50 of 136 D. Verscharen et al.

where
m j mi
m ji ≡ (98)
m j + mi

is the reduced mass of the two particles (see, e.g., Thornton and Marion 2004; Fitz-
patrick 2015). We consider an infinitesimal portion of the impact-parameter plane (see
Fig. 12) as

dσ = b db dφ. (99)

All j-particles that originate from this region are scattered into an infinitesimal solid-
angle centered on θ :

dΩ = sin θ dθ dφ. (100)

To derive the differential cross-section for a Coulomb collision, we assume that the
colliding particles only interact electrostatically. Then, when we combine Eqs. (99)
and (100) with that for the Coulomb force, we arrive at the Rutherford cross-section
(Rutherford 1911; Geiger and Marsden 1913):

dσ q 2j qi2
= . (101)
dΩ 4m 2ji g 4ji sin4 (θ/2)

Now, we consider all i-particles in the infinitesimal volume of phase space d3 vi


that is centered on vi . The rate (i.e., the number of particles per unit time) at which
j-particles, originating from dσ , collide with i-particles in d3 vi is


f i (vi )g ji dσ d3 vi = f i (vi )g ji dΩ d3 vi . (102)

Thus, the rate of decrease in the value of f j (v j ) due to collisions with i-particles in
all regions of phase space is
 
δ fj dσ
=− 3
d vi dΩ f j (v j ) f i (vi )g ji . (103)
δt c,i,− dΩ

The above expression is negative because it only accounts for the decrease in f j (v j )
due to j-particles of velocity v j being scattered to other velocities by i-particles. The
value of f j (v j ) can also increase as collisions scatter j-particles of other velocities to
v j . Indeed, Coulomb collisions are symmetric: if j- and i-particles of initial velocities
vj and vi collide at an impact parameter b, their final velocities will be v j and vi .
Thus, the rate of increase in f j (v j ) due to collisions with i-particles is
 
δ fj dσ
= d3 vi dΩ f j (vj ) f i (vi )g ji . (104)
δt c,i,+ dΩ

123
The multi-scale nature of the solar wind Page 51 of 136 5

We note that, in the above equation, vj and vi are functions of v j , vi , and θ . The net
rate of change in f j (v j ) due to collisions with i-particles is

δ fj δ fj δ fj
= +
δt c,i δt c,i,+ δt c,i,−
    dσ
= d3 vi dΩ f j (vj ) f i (vi ) − f j (v j ) f i (vi ) g ji . (105)

Finally, the net rate of change in f j (v j ) due to collisions with all species (i.e., the full
collision term) is

δ fj  δ fj
=
δt c δt c,i
i
     dσ
= d3 vi dΩ f j (vj ) f i (vi ) − f j (v j ) f i (vi ) g ji . (106)

i

This includes Coulomb collisions of j-particles with other j-particles, so the above
sum must include i = j.

3.2.2 The Landau collision integral

Evaluating Eq. (106) is highly non-trivial but it is helped by the fact that the dominant
contribution comes from small-angle collisions: those that produce small θ -values.
Before invoking the small-θ limit, it is convenient to express the particles’ initial and
final velocities in terms of the center-of-mass velocity u ji = uji as

m ji
v j = u ji + g ji , (107)
mj
m ji 
vj = u ji + g , (108)
m j ji
m ji
vi = u ji − g ji , (109)
mi

and
m ji 
vi = u ji − g . (110)
m i ji

Thus,
m ji
vj = v j + Δg ji (111)
mj

and
m ji
vi = vi − Δg ji , (112)
mi

123
5 Page 52 of 136 D. Verscharen et al.

where

Δg ji ≡ gji − g ji . (113)
 
In the small-θ limit, Δg ji  is also small, so Eqs. (111) and (112) can be used as the
basis for a Taylor expansion of f j and f i about v = v j and v = vi , respectively.
Retaining terms through the second order gives

m ji ∂ fj m 2ji ∂2 f j
f j (vj ) ≈ f j (v j ) + Δg ji · + Δg ji Δg ji : (114)
mj ∂v j 2m 2j ∂v j ∂v j

and

m ji ∂ fi m 2ji ∂ 2 fi
f i (vi ) ≈ f i (vi ) − Δg ji · + Δg ji Δg ji : . (115)
mi ∂vi 2m i2 ∂vi ∂vi

These approximations can be substituted into Eq. (105), which, after considerable
simplification (see, e.g., Hellinger and Trávníček 2009; Fitzpatrick 2015), yields the
Landau collision integral/operator (Landau 1936, 1937):

δ fj 2πq 2j qi2
≈ ln Λ ji
δt c,i mj
 
∂ I3 g 2ji − g ji g ji f i (vi ) ∂ f j f j (v j ) ∂ f i
× · 3
d vi · − ,
∂v j g 3ji m j ∂v j m i ∂vi
(116)

where ln Λ ji is the Coulomb logarithm, which is the subject of Sect. 3.2.3 and is given
in Eq. (117).
Although Eq. (116) is an improvement over Eq. (105), actually calculating the
Landau collision integral remains a daunting task even for relatively simple sce-
narios. Often, additional approximations are introduced, and numerical methods are
employed. An alternative approach is the BGK operator, which explicitly models the
departure of a particle species’ distribution function from its equilibrium state (Bhatna-
gar et al. 1954). This method was later generalized for the case of magnetized plasmas
(Dougherty 1964, and references therein). Pezzi et al. (2015) present a numerical
comparison of the Landau and Dougherty collision operators.

3.2.3 The Coulomb logarithm

The factor ln Λ ji in Eq. (116) is known as the Coulomb logarithm:

b
ji,max
db b ji,max
ln Λ ji ≡ = ln . (117)
b b ji,min
b ji,min

123
The multi-scale nature of the solar wind Page 53 of 136 5

It arises from the Ω-integral in Eq. (105) via the relationship between b and θ according
to Eq. (97). Even though the derivation of Eq. (116) would seemingly imply that all
b from 0 to ∞ should be considered, the Coulomb logarithm diverges at both of
these limits. As a result, the integral in Eq. (117) has been given the more restrictive
limits b ji,min and b ji,max , which are discussed below. Though there is some degree of
arbitrariness in how these limits are defined, Eq. (117) is relatively insensitive to their
particular values. In practice, b ji,min  b ji,max , so the logarithm of their ratio only
changes appreciably when they are varied by orders of magnitude.
The integral in Eq. (117) diverges at small b due to the breakdown of the small-θ
limit used to derive Eq. (116): as the value of b decreases, the value of θ increases
until it can no longer be considered small. In reality, collisions with small b have a
minimal effect on the distribution function because of their relative rarity. As a result,
collisions with θ > θmax are negligible and may be safely disregarded. A typical
choice is θmax = 90◦ , which, by Eq. (97), corresponds to

q j qi
b ji,min = , (118)
m ji g 2ji

where g ji is the average speed of a j-particle relative to an i-particle. The quantity


m ji g 2ji roughly reflects the average kinetic energy of j- and i-particles in the plasma
frame. As a result,

q j qi
b ji,min = , (119)
kB T ji

where T ji is the average temperature of the j- and i-particles.


The divergent behavior of Eq. (117) at high b stems from a more subtle reason.
The analysis above begins by considering the scattering of a single particle by another.
Effectively, the motion of each particle is modeled as a series of hard scatters, between
which the particle’s velocity remains constant. In reality, Coulomb collisions are soft
scatters, and each plasma particle is simultaneously colliding with many other particles.
As a result, each particle is partially shielded from the influence of distant particles by
the particles closer to it. An appropriate choice, then, for b ji,max is the Debye length
λD (Cohen et al. 1950; Spitzer 1956) as defined in Eq. (11). Taking into account all
the particle species in the plasma,
 −1/2
4π  q2 n 
b ji,max = bmax ≡ , (120)
kB T


where q , n  , and T are the charge, number density, and temperature of each species
in the plasma. As a result of this choice, the value of b ji,max is the same for all pairs
of particle species.
This discussion of b ji,max raises some concern over the use of binary collisions at
all. In principle, a more accurate approach would be to use an analysis of Markovian
processes to derive the collision operator from the Fokker–Planck equation (Fokker

123
5 Page 54 of 136 D. Verscharen et al.

1914; Planck 1917). Nevertheless, Wu (1966, Sects. 2–6) notes that both analyses
produce the same result, Eq. (116), in the limit of small-angle scattering.

3.2.4 Rosenbluth potentials

An alternative expression for the Landau collision integral in Equation (116) can be
obtained by using the Rosenbluth potentials (Rosenbluth et al. 1957), which are defined
as

 
G i (v j ) ≡ g ji  f i (vi ) d3 vi (121)

and

1
Hi (v j ) ≡   f i (vi ) d3 vi . (122)
g ji 

Likewise, we define flux densities associated with friction

4πq 2j qi2 ∂ Hi
A ji ≡ ln Λ ji (123)
mi ∂v j

and with diffusion

2πq 2j qi2 ∂ 2Gi


D ji ≡ ln Λ ji . (124)
mj ∂v j ∂v j

With these quantities defined, we express the Landau collision operator as the velocity
divergence of the sum of these fluxes (see Montgomery and Tidman 1964; Marsch
2006; Fitzpatrick 2015), casting it in terms of a Fokker–Planck advection–diffusion
equation in velocity space:

δ fj 1 ∂ ∂
≈− · A ji − D ji · f j. (125)
δt c,i m j ∂v j ∂v j

3.2.5 Collisional timescales

Conceptually, a collisional timescale is the time required for collisions to significantly


reduce a non-equilibrium feature such as a drift or anisotropy (for examples of non-
equilibrium kinetic features in the solar wind, see Sects. 1.4.4, 1.4.5). Each specific
type of non-equilibrium feature has its own expression for its collisional timescale that
depends on the conditions in the plasma. These timescales are derived from moments
of the Boltzmann collision term, similar to the procedure described in Sect. 1.4.1.
This requires that assumptions be made about the particular form of the distribution
function of each particle species involved.

123
The multi-scale nature of the solar wind Page 55 of 136 5

As an example, we discuss the collisional slowing time for two particle species, j
and i.11 These species’ differential flow is

ΔU ji ≡ U j − Ui , (126)

where U j and Ui are the bulk velocities of species j and i, respectively. Then, the rate
of change in the differential flow due to collisions is
 
δ ΔU ji δU j δUi
= − . (127)
δt δt c δt c
c

We express the bulk velocities U j and Ui as moments of f j and f i , the distribution


functions of the j- and i-particles, according to Eq. (28) and find
       
δ ΔU ji δ 1 δ 1
= d3 v v f j (v) − d3 v v f i (v)
δt δt nj c δt ni c
c
  
1 δ fj 1 δ fi
= d3 v v − . (128)
nj δt c ni δt c

To continue this analysis, we must make a choice for the form of the collision terms
and for the distribution functions. Once these are set, the result, to first order, has the
form
 
δ ΔU ji
= − νs, ji ΔU ji , (129)
δt
c

where νs, ji is the collision frequency for the slowing of j particles by i particles. The
corresponding collisional timescale is defined to be

1
τs, ji ≡ . (130)
νs, ji

Collisional timescales are most commonly derived and used for the relaxation of
temperature anisotropy T⊥ j /T j , unequal temperatures T j /Ti , and differential flow
ΔU ji .
Specific expressions for these collisional timescales have been computed and/or
compiled by Spitzer (1956), Schunk (1975, 1977), Hernández and Marsch (1985),
Huba (2016), and Wilson et al. (2018). Typically, only one type of non-equilibrium
feature is considered in each collisional timescale but formulæ derived by Hellinger
and Trávníček (2009, 2010) consider all three of the features listed above. Hellinger
(2016) uses observations from the Wind spacecraft to demonstrate that they result in
substantially different collision and heating rates. Likewise, although most derivations

11 We note that j and i may refer to two different components of the same particle species (e.g., the proton
core and proton beam, or the electron core and the electron halo).

123
5 Page 56 of 136 D. Verscharen et al.

assume Maxwellian or bi-Maxwellian distribution functions, Marsch and Livi (1985)


derive timescales for κ-distributions.

3.2.6 Coulomb number and collisional age

The majority of the heating and acceleration that gives rise to the solar wind’s non-
equilibrium properties occurs in and around the solar corona. Beyond that region, the
solar wind’s bulk velocity U remains approximately constant and radial (see, e.g.,
Hellinger et al. 2011, 2013). Thus, the time required for a parcel of plasma to travel
from the photosphere to a distance r is approximately the expansion time according
to Eq. (1):

r
τ= . (131)
Ur

The Coulomb number of the parcel of plasma is then defined as

τ r
Nc ≡ = , (132)
τc Ur τc

where τc is a collisional timescale. Notwithstanding the caveats noted below, the


Coulomb number essentially approximates the number of collisional timescales that
elapsed in a parcel of plasma during its journey from the Sun to an observer. In
collisionally old (Nc  1) plasma, collisional equilibration has proceeded much
farther than in collisionally young (Nc  1) plasma.
Although the Coulomb number has seen wide use in the analysis of solar-wind
observations (see Sect. 3.3), the concept carries significant limitations. The above def-
inition for Nc only allows for a single collision timescale τc . While the correct formula
for τc can be chosen for the non-equilibrium feature under consideration, accounting
for the interactions of multiple departures from equilibrium presents difficulties. More
fundamentally, the expression for Nc tacitly assumes that τc remains constant with dis-
tance r from the Sun. In reality, τc depends on density and temperature, both of which
have strong radial trends.
To address some of these issues, various studies (Hernández et al. 1987; Chhiber
et al. 2016; Kasper et al. 2017; Kasper and Klein 2019) employ an integrated Coulomb
number of the form
 
dt dr
Ac ≡ = . (133)
τc Ur (r )τc (r )

This formulation directly accounts for the radial dependences of densities, velocities,
and temperatures. These radial trends can either be derived from theoretical expec-
tations (e.g., for quasi-adiabatic expansion) or from empirical observations. Some
authors (e.g., Kasper et al. 2017) differentiate between the Coulomb number Nc and

123
The multi-scale nature of the solar wind Page 57 of 136 5

collisional age Ac , with the former defined by Eq. (132) and the latter defined by
Eq. (133).12
Maruca et al. (2013) introduce a close alternative to the Coulomb-number analysis,
retrograde collisional analysis, in which collisional timescales and radial trends are
used to “undo” the effects of collisions and estimate the state of the solar wind when
it was closer to the Sun.

3.3 Observations of collisional relaxation in the solar wind

This section summarizes observational studies of collisional relaxation’s effects on


solar-wind plasma as it expands through the heliosphere.

3.3.1 Ion collisions

Early observations of solar-wind ions indicate that α-particles tend to be significantly


faster and hotter than protons (see Sect. 1.4.4). Observations from IMP 6, IMP 7,
IMP 8, and OGO 5 (Feldman et al. 1974a;Neugebauer  1976; Neugebauer and Feld-
man 1979) demonstrate that the values of ΔUαp  and Tα /Tp decrease toward 0 and
1 with increasing Nc . This negative correlation indicates that α-particles are first
preferentially accelerated and heated in the corona and then partially equilibrate with
protons as the plasma expands through the inner heliosphere. Later studies using obser-
vations from Helios (Marsch et al. 1982a, 1983; Livi et al. 1986), ISEE 3 (Klein et al.
1985), Prognoz 7 (Yermolaev et al. 1989, 1991; Yermolaev and Stupin 1990), Ulysses
(Neugebauer et al. 1994), and Wind (Kasper et al. 2008, 2017; Maruca et al. 2013;
Hellinger 2016) confirm these early results. Interplanetary coronal mass ejections
(ICMEs) are a notable exception to this overall trend in that they exhibit enhance-
ments in Tα /Tp , which arise from ongoing heating during expansion (Liu et al. 2006).
Measurements of T⊥p and Tp from Wind reveal that the average value of the
anisotropy ratio T⊥p /Tp → 1 as the Coulomb number increases (Kasper et al. 2008,
2017). Further observations (Bale et al. 2009) show that both Coulomb collisions
and kinetic microinstabilities (see Sect. 6) have roles in limiting proton temperature
anisotropy. Numerical models confirm this interplay of collisional and wave–particle
effects (Tam and Chang 1999; Hellinger and Trávníček 2010; Matteini et al. 2012).
Figure 13 shows trends in four parameters with Coulomb number Nc in a dataset of
2.1-million data from the Wind/SWE Faraday cups compiled by Maruca et al. (2012,
2013). The values of Nc are calculated using the expression derived by Maruca et al.
(2013), which is based on the proton “self-collision time” described by Spitzer (1956).
For each parameter P, the (Nc , P)-plane is divided into 80 logarithmically spaced
Nc -bins and 40 linearly spaced P-bins. Once the data are binned, the grid is column-
normalized: the number of counts in each bin is divided by the number of counts in the
most-populated bin in its column. Thus, the color of each bin in Fig. 13 indicates the
relative likelihood of a P-value for a given Nc -value. Each of the four parameters in
Fig. 13 is an indicator of a departure from local thermal equilibrium. As Nc increases,
12 We adopt the new terminology of Kasper et al. (2017). We note, however, that some earlier publications
use the term “collisional age” for Nc (Kasper et al. 2008; Bale et al. 2009; Maruca et al. 2013).

123
5 Page 58 of 136 D. Verscharen et al.

Fig. 13 Trends in four parameters with Coulomb number Nc : a α–proton differential flow normalized to the
proton Alfvén speed, b α-to-proton relative temperature, c proton temperature anisotropy, and d α-particle
temperature anisotropy. The dataset, compiled by Maruca et al. (2012, 2013), consists of 2.1-million data
from the Wind/SWE Faraday cups. The color scale is linear, and red indicates the most-likely parameter
value for a given Nc -value. The probability densities of Coulomb number (top) and of each of the four
parameters (right) are also shown. After Kasper et al. (2008, 2017)

123
The multi-scale nature of the solar wind Page 59 of 136 5

 
the most-likely P-value approaches its equilibrium state: 0 for ΔUαp  /vAp and 1
for Tα /Tp , T⊥p /Tp , and T⊥α /Tα . Each parameter reaches equilibrium at a different
Nc -value because the formula for Nc uses the same self-collision time as a generic
collisional timescale rather than the specific collisional timescale for each parameter
P.
Column-normalizing plots (as has been done, e.g., for those in Figs. 13, 14) is
a powerful and well established technique for exploring collisional effects in solar-
wind plasma. It represents a refinement of the method used in some of the earliest
studies of collisional relaxation (e.g., Feldman et al. 1974a; Neugebauer 1976), in
which data were divided into logarithmically uniform Nc -intervals, and the average
Tα /Tp -value was plotted for each interval. Nevertheless, some caution is warranted in
producing and interpreting column-normalized plots in general. First, the procedure
of column-normalization modifies the weights of different data points and thus may
cause an overemphasis or underemphasis of bins in a statistical data set. Second, the
very act of column-normalization imposes causality: the parameter on the vertical axis
becomes a function of that on the horizontal axis. Though this is usually justified in
collisionalization studies because of the strong theoretical motivation for such a causal
relationship, column-normalization is not appropriate for all correlation studies. Third,
determining which parameters to plot is complicated by the many correlations that exist
among particle moments (e.g., the well established temperature–speed relationship  for
protons; Lopez and Freeman 1986). Even so, parameters such as Tα /Tp and ΔUαp 
have been qualitatively (Kasper et al. 2008) and quantitatively (Maruca et al. 2013)
demonstrated to be more strongly correlated with Nc than with n p , Upr , or Tp (all three
of which Nc depends on).
Observations also give insight into collisional effects on minor ions. ISEE 3 and
SOHO/CELIAS data show that, while mass-proportional temperatures are most com-
mon, the effects of collisional thermalization are apparent at low solar-wind speeds
(Bochsler et al. 1985; Hefti et al. 1998). Interestingly, von Steiger et al. (1995) and
von Steiger and Zurbuchen (2006) find no indications of a departure from mass-
proportional temperatures at any solar-wind speed. This may be due to the limited
number of data from very slow wind or from the ongoing heating of heavy ions.
Coulomb-number analyses of heavy-ion observations from ACE/SWICS show simi-
lar negative trends in the ion-to-proton temperature ratio with Coulomb number (Tracy
et al. 2015, 2016).
Although most observational studies of ion–ion collisions focus on the effects of
collisions on particle moments, some consider how collisions affect the structure of ion
distribution functions. Marsch and Goldstein (1983) note that the value of the collision
term in Eq. (106) varies across phase space and is highest for particles traveling at the
bulk speed of the plasma. This finding is consistent with proton distribution functions
observed by Helios, which show Maxwellian cores surrounded by non-Maxwellian
tails. A kinetic model of the collisional effects on proton distribution functions counter-
intuitively reveals that collisional isotropization can actually generate proton beams
(Livi and Marsch 1987), which themselves would then be ultimately eroded by colli-
sions.

123
5 Page 60 of 136 D. Verscharen et al.

3.3.2 Electron collisions

Collisions involving electrons, due to their higher rates (see, e.g., Wilson et al. 2018),
are thought to play an even more important role in solar-wind thermodynamics than
collisions involving only ions. As noted in Sect. 1.4.5, electron distribution functions
in the solar wind typically exhibit a three-component structure consisting of a core,
halo, and strahl. Many theories (e.g., Scudder and Olbert 1979a, b; Lie-Svendsen et al.
1997; Lie-Svendsen and Leer 2000) for the origin of these electron populations rely on
the transition from highly collisional plasma in the lower corona to weakly collisional
plasma in the upper corona.
Beyond the corona, numerous studies find that Coulomb collisions among electrons
continue to affect them in the interplanetary solar wind. An analysis of Mariner 10 data
(Ogilvie and Scudder 1978) reveals that collisions have the greatest influence on the
electron core while the electron halo remains weakly collisional. Electron distribution
functions observed by Helios show that Coulomb collisions have a significant impact
on the phase-space location of the core–halo boundary (Pilipp et al. 1987a, b, c). Kinetic
simulations suggest that the interplay of collisions and expansion in the solar wind
can give rise to the electron core, halo, and beam (Landi et al. 2010; Landi et al.
2012). Moreover, a kinetic model for the radial evolution of the strahl developed by
Horaites et al. (2018b) indicates that Coulomb collisions provide a significant source
of pitch-angle scattering for this population.
Solar-wind electrons typically exhibit less temperature anisotropy than ions (Chen
et al. 2016, Figure 1), which is at least partially ascribed to the higher rate of electron
versus ion collisions. Analytical models that account for electron expansion and colli-
sions in the interplanetary solar wind agree well with ISEE 3 and Ulysses observations
of electron temperature anisotropy (Phillips et al. 1989a; Phillips and Gosling 1990;
Phillips et al. 1993). A study of Wind observations by Salem et al. (2003) finds that
electron temperature anisotropy is strongly correlated with Coulomb number, with
collisionally old electrons being most likely to exhibit isotropy. As is the case for pro-
tons, data from Helios, Cluster, and Ulysses show that both Coulomb collisions and
kinetic microinstabilities play significant roles in isotropizing solar-wind electrons
(Štverák et al. 2008, 2015).
Collisions also significantly affect electron heat flux. According to Spitzer–Härm
theory (Spitzer and Härm 1953), the electron heat flux is proportional to the timescale
of electron–electron collisions. Statistical analyses of Wind electron measurements
show that this relationship holds true but only in highly collisional plasma (Salem
et al. 2003; Bale et al. 2013). Figure 14 shows the distribution of Wind/3DP electron
data in the plane of the normalized parallel heat flux versus the normalized electron
mean free path in the solar wind. We normalize qe to the free-streaming saturation heat
flux q0 ≡ 3n e kB Te we /2 and λmfp,e to the temperature gradient L T ≡ r /α, where r is
the heliocentric distance of the measurement and α describes the observed temperature
profile through Te ∝ r −α . The dimensionless quantity λmfp,e /L T is called the Knudsen
number. The black line shows the Spitzer–Härm prediction. The heat flux follows this
prediction at large collisionality but deviates in the collisionless limit.

123
The multi-scale nature of the solar wind Page 61 of 136 5

Fig. 14 Column-normalized distribution of Wind/3DP electron data as a function of the parallel heat flux
qe and the electron mean free path λmfp,e . The Spitzer–Härm prediction in this normalization is given
by qe /q0 = 1.07λmfp,e /L T and is shown as a black line. We use α = 2/7. The probability densities for
λmfp,e /L T (top) and qe /q0 (right) are also shown. After Salem et al. (2003) and Bale et al. (2013) and
using data provided by C. Salem

Spitzer–Härm theory is found to overestimate electron heat flux in moderately and


weakly collisional plasma, which is consistent with results from the kinetic simulations
of Landi et al. (2012) and Landi et al. (2014).
Occasionally, a parcel of solar-wind plasma is found to have an especially low or
high rate of Coulomb collisions, which offers insight into the most extreme effects of
collisions on electrons. In a study of several periods of very-low-density solar wind,
each period exhibits an unusually narrow electron strahl (Ogilvie et al. 2000). This
likely results from the combination of a low collision rate and the conservation of the
first adiabatic invariant, given in Eq. (44), to first order as suggested by Fairfield and
Scudder (1985). Conversely, data from ISEE 1 and ISEE 3 exhibit several heat-flux
dropouts (Fitzenreiter and Ogilvie 1992): periods of very low electron heat flux. The
weak electron halos observed during these dropouts likely result, at least in part, from
enhanced electron collisionality. Likewise, Larson et al. (2000) and Farrugia et al.
(2002), using the Wind and ACE spacecraft, identify weak halos in particularly dense
and cold magnetic clouds and find them to be consistent with collisional effects.

123
5 Page 62 of 136 D. Verscharen et al.

4 Plasma waves

Plasma waves are important processes for the transport and dissipation of energy in a
plasma. They can accelerate plasma flows and heat plasma by damping. Section 4.1
introduces basic concepts to describe plasma waves. Section 4.2 describes damping
and dissipation mechanisms, and Sect. 4.3 then presents types of plasma waves that
are relevant to the multi-scale evolution of the solar wind. For more details on the
broad topic of plasma waves, we refer to the excellent textbooks by Stix (1992) and
Swanson (2003).

4.1 Plasma waves as self-consistent electromagnetic and particle fluctuations

Waves are periodic or quasi-periodic spatio-temporal fluctuations which arise through


the action of a restoring force. The self-consistent electromagnetic interactions in a
plasma provide additional restoring forces that do not occur in a neutral gas. Therefore,
a plasma can exhibit many more types of wave modes than a neutral gas. In this section,
we introduce the linear theory of plasma waves. For further details on linear theory, we
refer the reader to the general review on solar-wind plasma waves by Ofman (2010)
and the textbooks by Stix (1992), Brambilla (1998), and Swanson (2003).
Linear wave theory considers a wave to be a fluctuating perturbation on an equi-
librium state. We assume that any physical quantity A of the system can be written
as

A(x, t) = A0 + δ A(x, t), (134)

where A0 is the constant background equilibrium, and δ A is the fluctuating perturbation


of A. Moreover, we assume that the fluctuating quantities in a wave behave like

δ A(x, t) = Re A(k, ω) exp (ik · x − iωt) , (135)

where A(k, ω) is the complex Fourier amplitude of A, the wavevector k is real, and
the frequency ω is complex. We define the real frequency as

ωr ≡ Re ω (136)

and the growth or damping rate as

γ ≡ Im ω. (137)

The linear dispersion relation is a mathematical expression based on a self-consistent


set of linearized equations for the plasma particles and the electromagnetic fields.
It connects the wavevector k with the frequency ω in such a way that its solutions
represent self-consistent waves in the plasma. If multiple solutions exist for a given
k, then each corresponds to a distinct mode. According to Eqs. (135) and (137), the
amplitude of the fluctuations decreases exponentially with time if γ < 0. As a solution

123
The multi-scale nature of the solar wind Page 63 of 136 5

to the linear dispersion relation, we describe such a wave as being linearly damped
(see Sect. 4.2.1). Likewise, if γ > 0, the wave amplitude increases exponentially with
time and the wave is linearly unstable (see Sect. 6).
Neglecting any background electric field E0 , we rewrite the electric and magnetic
fields according to Eq. (135) as

E(x, t) = δE(x, t) = Re E(k, ω) exp (ik · x − iωt) (138)

and

B(x, t) = B0 + δB(x, t) = B0 + Re B(k, ω) exp (ik · x − iωt) , (139)

using the complex Fourier amplitudes E(k, ω) and B(k, ω). In the following, we write
the Fourier amplitudes without their arguments (k, ω) and assume that |δB|  |B0 |.
Substituting Eqs. (138) and (139) into Maxwell’s equations (21) through (24), we
find in Fourier space

k · E = − 4πiρc , (140)
k · B = 0, (141)
ω
k × E − B = 0, (142)
c

and

ω 4πi
k×B+ E=− j, (143)
c c

where
 
ρc = ρc j = qjn j (144)
j j

is the charge density and



j= jj (145)
j

is the current density. In Eqs. (144) and (145), the sums are carried over all particle
species j in the plasma. The left-hand sides of Eqs. (140) through (143) represent
the interactions between the electric and magnetic fields, while the right-hand sides
represent the self-consistent effects of the particles on the fields.
We define the plasma susceptibility tensor χ j of species j through

4πi
χj ·E≡ jj (146)
ω

123
5 Page 64 of 136 D. Verscharen et al.

and the dielectric tensor  as



 ≡1+ χ j. (147)
j

The dielectric tensor is additive in the contributions from each plasma species j and
reflects the interaction between fields and particles. With these definitions, we find

4πi
·E=E+ j (148)
ω
and, by using Eq. (143),
ω
k×B+  · E = 0. (149)
c
Combining Eq. (142) with Eq. (149) leads to the wave equation:

n × (n × E) +  · E = D · E = 0, (150)

where n ≡ kc/ω is the refractive index and


⎛ ⎞
x x − n 2z x y x z + n x n z
D≡⎝  yx  yy − n 2x − n 2z  yz ⎠ (151)
zx + n z n x zy zz − n 2x

is the dispersion tensor. The phase velocity of a solution is given by ωk/k 2 . Non-trivial
solutions to the wave equation fulfill

det [D(k, ω)] = 0, (152)

which is the mathematical dispersion relation. The identification of plasma waves then
involves the calculation of a proper dielectric tensor for the plasma conditions at hand
as well as the derivation of the roots of Eq. (152).
If the calculation of  is based on the linearized Vlasov equation (Gary 1993),
Eq. (152) leads to the full hot-plasma dispersion relation, which is a standard-tool in
the calculation of plasma waves (Rönnmark 1982; Klein and Howes 2015; Verscharen
and Chandran 2018; Verscharen et al. 2018). In this model, Eq. (20) is linearized for
each plasma species j to first order in δ f j , under the assumption that f j = f 0 j + δ f j ,
as

∂δ f j ∂δ f j   ∂δ f qj 1 ∂ f0 j
j
+v· + Ω j v × b̂ · =− δE + v × δB · ,
∂t ∂x ∂v mj c ∂v
(153)

where the left-hand side describes the change of δ f j along the zeroth-order particle
trajectory, Ω j is calculated based on the background magnetic-field magnitude B0 ,

123
The multi-scale nature of the solar wind Page 65 of 136 5

and b̂ ≡ B0 /B0 . The resulting solutions for δ f j from integration along the particle
trajectories then define ρc and j according to Eqs. (25) and (26). We refer to the
textbooks by Melrose and McPhedran (1991), Stix (1992), and Gary (1993) for more
details on the calculation of .
In our discussion of wave modes in Sect. 4.3, we present analytical results for wave
dispersion and polarization relations based on different models and in different limits,
which we identify whenever necessary. Fluid models and kinetic models often lead
to different predictions in the dispersion relation and polarization properties of linear
waves (see, e.g., Verscharen et al. 2017; Wu et al. 2019). These differences result from
differences in the models’ underlying assumptions (e.g., the closure of the hierarchy
of moment equations; see Sect. 1.4.1). Furthermore, analytical calculations of the
dispersion relation often rely on mathematical approximations in certain limits (e.g.,
taking m e → 0 or T j → 0). Before we discuss the wave modes further, we describe
damping and dissipation mechanisms in the following section.

4.2 Damping and dissipation mechanisms

The damping and dissipation of plasma waves are important for the global behavior
of the plasma because these processes transfer energy between the electromagnetic
fields and the particles and are also candidates for the dissipation of turbulent plasma
fluctuations in the solar wind (see Sect. 5).
For our discussion, we distinguish between damping as a reduction in the amplitude
of field fluctuations (i.e., γ < 0) and dissipation as an irreversible increase in entropy
of a plasma species (i.e., dS j > 0, where S j is the entropy of species j). Lastly, we
define heating as an increase of the plasma’s thermal energy. In this section, we address
three important damping and dissipation mechanisms for plasma waves: (1) quasilin-
ear diffusion from Landau-resonant or cyclotron-resonant wave–particle interactions,
(2) nonlinear phase mixing, and (3) stochastic heating. So long as the Boltzmann
equation (19) is valid, dissipation in the sense of entropy generation can only occur
through particle–particle collisions. Even if collisions are not frequent enough to bring
the plasma distribution function into local thermodynamic equilibrium, phase-space
structures in the velocity distribution function can become small enough that colli-
sions lead to dissipation (cf Sect. 3.2). When we study the dissipation of “collisionless”
plasma waves, we, therefore, assume that collisions only affect small-scale structures
in the distribution function and investigate the processes that create these small-scale
structures, which in turn generate entropy through collisions. We note that deviations of
velocity distributions from local thermodynamic equilibrium (see Sects. 1.4.4, 1.4.5)
can affect the polarizations, transport ratios, and damping rates of the plasma normal
modes, as well as the heating mechanisms (Chandran et al. 2013; Kasper et al. 2013;
Klein and Howes 2015; Tong et al. 2015; Kunz et al. 2018).

4.2.1 Quasilinear diffusion

Quasilinear diffusion describes the evolution of the distribution function as velocity-


space diffusion that arises from the resonant interaction between waves and particles

123
5 Page 66 of 136 D. Verscharen et al.

(Marsch 2006). Quasilinear theory assumes the presence of a superposition of non-


interacting and randomly phased waves that are solutions to linear plasma-wave theory
as described in Sect. 4.1. The force term in the Vlasov equation is then averaged over
the gyro-phases of the unperturbed particle orbits so that a diffusion term for the
background distribution f 0 j in v⊥ and v results, independent of the gyro-phase of the
particles. This process is quasilinear in the sense that the fluctuations are solutions to the
linear dispersion relation (Sect. 4.1), which closes the system of equations, but the field
amplitudes enter the equations quadratically. In quasilinear theory, the background
distribution f 0 j evolves slowly compared to the timescale of the fluctuations 1/ωr .
Under the assumption of small wave amplitudes and |γ /ωr |  1, quasilinear diffusion
follows the equation (Shapiro and Shevchenko 1962; Kennel and Engelmann 1966;
Rowlands et al. 1966; Stix 1992)

+∞ 
∂ f0 j q 2j 1  1
= lim d3 k Ĝv⊥ δ ωr − k v − nΩ j |ψn |2 Ĝ f 0 j ,
∂t 2 2
8π m j V →∞ V n=− ∞ v⊥
(154)

where the pitch-angle operator is defined as

k v ∂ k v⊥ ∂
Ĝ ≡ 1 − + , (155)
ωr ∂v⊥ ωr ∂v

and
1   v

ψn ≡ √ E r eiφ Jn+1 (σ j ) + E l e−iφ Jn−1 (σ j ) + E z Jn (σ j ). (156)
2 v⊥

We define the wavevector components perpendicular and parallel to the background


magnetic field as k⊥ and k , respectively. The right-handed and left-handed √ com-
ponents of the Fourier-transformed electric field are E r ≡ E x − i E y / 2 and

E l ≡ E x + i E y / 2, respectively, Jn is the nth order Bessel function of the first
kind, σ j ≡ k⊥ v⊥ /Ω j , φ is the azimuthal angle of k, and V is the spatial volume
under consideration. Since Eq. (154) is a second-order differential equation in v⊥ and
v , it indeed corresponds to a diffusion in velocity space. The δ-function in Eq. (154)
guarantees that the only particles that participate in the resonant interactions are those
for which v is equal to the resonance speed:

ωr − nΩ j
vres ≡ . (157)
k

Due to the form of Ĝ, the diffusive flux of particles is tangent to semicircles in the
v −v⊥ plane defined by

2
ωr
v − + v⊥
2
= constant (158)
k

123
The multi-scale nature of the solar wind Page 67 of 136 5

Fig. 15 Quasilinear diffusion in the cyclotron-resonant damping of particles with v = vres < 0 (gray
shaded area) with waves of parallel phase speed ωr /k . The blue dotted circles represent isocontours of
the background distribution function f 0 j . The diffusion paths (blue arrows) are locally tangential to circles
around the point (v⊥ , v ) = (0, ωr /k ) (black circles). In this example, the resonant particles gain kinetic
energy, which corresponds to an increase in (v⊥ 2 + v 2 ). This energy is removed from the waves at ω and
 r
k , which are thus damped

and directed from larger to smaller values of f 0 j (Verscharen and Chandran 2013).
During the diffusion, the particles gain kinetic energy if (v⊥ 2 + v 2 ) increases and lose

it if this quantity decreases. The energy gained or lost by the particles is taken from
or given to the wave at the resonant k and ωr so that this wave’s amplitude changes.
The n = 0 term in the sum in Eq. (154) corresponds to Landau damping (1946) and
transit-time damping, and the n = 0 terms correspond to cyclotron damping.
We illustrate the quasilinear diffusion process for a cyclotron-damped wave in
Fig. 15. In this example, cyclotron-resonant particles with v = vres < 0 interact with
waves with ωr and k and diffuse in velocity space. The cyclotron-resonant damping
of left-handed waves propagating parallel to B0 exhibits these characteristics. We
illustrate the case of quasilinear diffusion for a cyclotron-resonant instability in Fig. 20
in Sect. 6.

4.2.2 Entropy cascade and nonlinear phase mixing

Since dissipation, by definition, is irreversible, all dissipation processes cause entropy


to increase. In a plasma with low collisionality, wave turbulence (see Sect. 5.2) is
associated with fluctuations in entropy13 that cascade to small scales, where collisions
have greater effects and ultimately dissipate these fluctuations. Applying Boltzmann’s
H -theorem to Eq. (19), we obtain the entropy relation

   3 
dS j d d3 r d r δ fj
= − 3
d v f j ln f j =− d3 v ln f j , (159)
dt dt V V δt c

13 These largely reversible fluctuations in entropy do not violate the second law of thermodynamics which
only applies to the total entropy of a closed system.

123
5 Page 68 of 136 D. Verscharen et al.

where S j is the entropy of species j, and V is the spatial volume under consideration.
Equation (159) shows that entropy only increases in the presence of particle–particle
collisions. We now separate f j into its equilibrium part f 0 j and its fluctuating part
δ f j as

f j (x, v, t) = f 0 j (v) + δ f j (x, v, t). (160)

We assume that the collision frequency is of order ωr ,14 and f 0 j is a Maxwellian as


in Eq. (59) with temperature T0 j . After averaging over the timescales greater than
the typical fluctuation time ∼ 1/ωr and summing over all species, we describe the
evolution of the generalized energy through the energy equation with the help of the
expression for the entropy from Eq. (159) as (Schekochihin et al. 2008)
⎛ ⎞
  kB T0 j δ f j2
dW d + d3 r ⎝ E2 B2 ⎠
= + d3 v
dt dt 8π V 2 f0 j
j
 3 
d r kB T0 j δ f j δ f j
=+ d3 v , (161)
V f0 j δt c
j

where W is the generalized energy and  is the externally supplied power (e.g., through
large-scale driving by shears or compressions).15
The entropy cascade constitutes the redistribution of generalized energy from elec-
tromagnetic fluctuations (E 2 + B 2 ) to entropy fluctuations (δ f j2 / f 0 j ) according to
Eq. (161). These fluctuations in entropy then cascade to smaller scales in velocity
space through a combination of linear and nonlinear phase mixing. Linear phase mix-
ing corresponds to Landau damping, which we describe in Sect. 4.2.1. The spread
in parallel velocity of the particle distribution leads to a dependency of the Landau–
resonant interactions between particles and the electric field on the particles’ parallel
velocity.
Nonlinear phase mixing often serves as a faster mechanism of entropy cascade. A
particle with a greater v⊥ has a greater ρ j and thus experiences a slower E × B drift
than a particle with smaller v⊥ (Dorland and Hammett 1993). Two particles of the
same species j but distinct perpendicular velocities v⊥ and v⊥  experience spatially

decorrelated fluctuations in the electric and magnetic fields if the difference between
the particles’ gyro-radii v⊥ /|Ω j | and v⊥ /|Ω | is greater than the perpendicular cor-
j
relation length 1/k⊥ of the field fluctuations (Schekochihin et al. 2008). In kinetic
theory, this process leads to spatial perpendicular mixing of ion distributions with dif-
ferent gyro-centers and hence to the creation of small-scale structure in the gyro-center
distribution. Small-scale structure in the fields in physical space thus leads to small-
scale structure in the distribution function in velocity space perpendicular to v⊥ as the

14 In gyrokinetic theory, the collision frequency and ω are both ordered to the intermediate timescale.
r
This ordering does not prevent us from considering the collisionless and collisional limits and justifies the
assumption of a Maxwellian f 0 j = f M (Howes et al. 2006; Schekochihin et al. 2008).
15 Although Eq. (161) was derived under the assumption of a Maxwellian background distribution, Kunz
et al. (2018) derive an expression for dW /dt assuming a drifting bi-Maxwellian f 0 j = f bM .

123
The multi-scale nature of the solar wind Page 69 of 136 5

Fig. 16 Trajectories of test particles in the plane perpendicular to B0 . We use a setup similar to the kinetic-
Alfvén-wave (KAW) simulations of stochastic heating described by Chandran et al. (2010). In the left
panel, we show solutions for a thermal-proton trajectory when the amplitude of the Alfvénic fluctuations
at k⊥ ρp ≈ 1 is small. The proton drifts due to the large-scale Alfvénic fluctuations, but its gyro-motion
is still circular to first order. In the right panel, we show the same solutions but with an amplitude of the
gyro-scale KAW fluctuations that is by a factor of five greater than in the left panel. The gyro-motion is
strongly perturbed and becomes stochastic, creating the conditions for stochastic heating

result of this nonlinear phase mixing (Tatsuno et al. 2009; Bañón Navarro et al. 2011;
Kawamori 2013; Navarro et al. 2016; Cerri et al. 2018). Once these velocity-space
structures are small enough, collisions can efficiently smooth them—see Eq. (106)
and the associated discussion—and thereby increase entropy and the perpendicular
temperature of the ions.

4.2.3 Stochastic heating

Stochastic heating is a non-resonant energy-diffusion process. It arises from field


fluctuations with spatial variations on the gyro-radius scale of the diffusing particles
(k⊥ ρ j ∼ 1) and frequencies that are small compared to the gyro-frequency (ωr 
|Ω j |) in a constant background magnetic field B0 (McChesney et al. 1987; Chen et al.
2001b; Johnson and Cheng 2001; Chaston et al. 2004; Fiksel et al. 2009).
If these fluctuations are low in amplitude, they induce only small perturbations
in the particles’ otherwise circular orbits. With increasing amplitude, however, the
fluctuations increasingly distort the gyro-orbits. If the amplitude of the gyro-scale
fluctuations is so large that the orbits become stochastic in the plane perpendicular
to B0 , particles experience stochastic increases and decreases in their kinetic energy
due to the fluctuations’ electric fields. Consequently, the particles diffuse in v⊥
2 , which

corresponds to perpendicular heating (Chandran et al. 2010; Klein and Chandran


2016). This process is consistent with observations of solar-wind protons (Bourouaine
and Chandran 2013; Martinović et al. 2019) and minor-ion temperatures and drifts
(Chandran 2010; Wang et al. 2011; Chandran et al. 2013).
Figure 16 shows the orbits of two thermal protons in test-particle simulations of
stochastic heating based on a superposition of randomly-phased kinetic Alfvén waves
(KAWs; see Sect. 4.3.2). If the amplitude of the gyro-scale fluctuations is small (left
panel), the magnetic moment is conserved and the particle trajectory corresponds to a
drifting quasi-circular motion. If the amplitude of the gyro-scale fluctuations is large

123
5 Page 70 of 136 D. Verscharen et al.

(right panel), the magnetic moment is no longer conserved. As a result, the particle’s
trajectory becomes stochastic, which corresponds to stochastic heating through the
waves’ electric fields.
The mechanisms of stochastic proton heating are different in the low-βp regime and
in the high-βp regime. In plasmas with low βp , the proton orbits become stochastic
mainly due to spatial variations in the electrostatic potential, and the protons primarily
gain energy from the slow temporal variations in the electrostatic potential associated
with the fluctuations (Chandran et al. 2010). In plasmas with high βp , the proton
orbits become stochastic mainly due to spatial variations in the magnetic field, and
the protons primarily gain energy from the solenoidal component of the electric field
(Hoppock et al. 2018). Despite these differences, stochastic heating remains a universal
candidate process to explain ion heating in the direction perpendicular to B0 in weakly
collisional plasmas.

4.3 Wave types in the solar wind

In this section, we discuss large-scale Alfvén waves, kinetic Alfvén waves, Alfvén/ion-
cyclotron waves, slow modes, and fast modes, which are the most important wave
types for the multi-scale dynamics of the solar wind. We note that the nomenclature of
wave types is not universal and that different names are commonly used for waves of
the same type depending on their location in wavevector space (e.g., TenBarge et al.
2012, Fig. 1).

4.3.1 Large-scale Alfvén waves

Alfvén waves are electromagnetic plasma waves for which magnetic tension serves
as the restoring force (Alfvén 1942; Alfvén 1943). To first order, these waves are
non-compressive. At large scales (i.e., kdp  1 and kρp  1), Alfvén waves obey
the linear dispersion relation

ω = ± |k |vA , (162)

where the upper√(lower) sign corresponds to propagation parallel (anti-parallel) to B0 ,


and vA ∗ ≡ B / 4πρ is the MHD Alfvén speed. The group-velocity vector is paral-
0
lel or anti-parallel to B0 , and large-scale Alfvén waves are only weakly damped in a
plasma with Maxwellian distribution functions. The fluctuating magnetic-field vector
δB is perpendicular to k and B0 . Alfvén waves are characterized by negligible fluc-
tuations in n j (i.e., they are non-compressive) and B ≡ |B|, but an (anti-)correlation
between velocity fluctuations δU j and magnetic-field fluctuations δB. In the MHD
approximation, this polarization property is given by

δU δB
∗ =∓ . (163)
vA B0

In the solar wind, the center-of-mass frame, in which we define ω and k, is dom-
inated by the proton flow so that U ≈ Up and ρ ≈ n p m p . Therefore, Eq. (163) is

123
The multi-scale nature of the solar wind Page 71 of 136 5

Fig. 17 Alfvénic correlations between δUp and δB. We show data from the Wind spacecraft’s SWE and
MFI instruments starting at 18:01:59 on 2018-05-06 for a total duration of 7 h. The top three panels show the
three components of the vector velocity (km/s; blue) and magnetic-field (nT; red) fluctuations. The vector
components are positively correlated in this example. The bottom panel shows that the density (cm−3 ;
green) and the absolute value of the magnetic field (nT; red) stay approximately constant

approximately δUp /vAp ≈ ∓ δB/B0 . Observations of the vector components of the


plasma velocity and the magnetic field in the solar wind often exhibit this polarization
(Unti and Neugebauer 1968; Belcher et al. 1969; Belcher and Davis 1971; Bruno et al.
1985; Velli and Pruneti 1997; Chandran et al. 2009; Boldyrev and Perez 2012; He
et al. 2012b, a; Podesta and TenBarge 2012), and we illustrate one such example in
Fig. 17.
In fact, since this polarization characterizes the majority of the solar wind’s
large-scale fluctuations, its large-scale turbulence is believed to be Alfvén-wave-like
turbulence (see Sect. 5.2). At large scales, the amplitudes of the Alfvénic fluctuations
in the solar wind are often so large that their behavior becomes nonlinear. Their polar-
ization fulfills B = constant, while the magnetic-field and velocity vectors often show
a spherical or arc-like polarization (Tsurutani et al. 1994; Riley et al. 1996; Vasquez
and Hollweg 1996). Although Alfvén waves predominantly occur in the fast solar
wind, D’Amicis and Bruno (2015) identify a type of slow wind that also carries large-
amplitude Alfvén waves and shows many other characteristics usually associated with
fast wind (D’Amicis et al. 2019).
We note that left-circularly polarized and parallel-propagating Alfvén waves are a
solution of the full nonlinear MHD and multi-fluid equations (Marsch and Verscharen
2011). At large scales, these waves follow a polarization relation that follows directly
from the multi-fluid equations:

δU j U j δB
∗ =∓ 1∓ ∗ , (164)
vA vA B0

123
5 Page 72 of 136 D. Verscharen et al.

where the upper and lower signs describe the propagation direction as in Eq. (162).
Equation (164) shows that a particle species with U j ≈ vA ∗ does not participate

in the bulk-velocity polarization motion associated with parallel-propagating large-


scale Alfvén waves: in the reference frame of these particles, the wave has no electric
field. Observations confirm that α-particles (see Sect. 1.4.4) with Uα ≈ vA ∗ exhibit

δUα ≈ 0, which is an effect known as surfing α-particles (Marsch et al. 1982a;


Goldstein et al. 1995; Matteini et al. 2015b).
There are two extensions of the Alfvén wave to smaller scales: the kinetic Alfvén
wave (KAW) at k⊥ ρp  1 and k⊥  k , and the Alfvén/ion-cyclotron (A/IC) wave at
k dp  1 and k⊥  k . Although KAWs and A/IC waves belong to the Alfvén-wave
family (Andre 1985; Yoon and Fang 2008; Klein and Howes 2015), we discuss them
separately in the following two sections due to their great importance for the physics
of the solar wind.

4.3.2 Kinetic Alfvén waves

Kinetic Alfvén waves (KAWs) are the short-wavelength extension of the Alfvén-wave
branch for k⊥  k . This type of wave has received much attention since large-scale
turbulence in the solar wind is Alfvén-wave-like and supports a cascade with increasing
anisotropy toward k⊥  k (see Sect. 5.2). Thus, KAWs are the prime candidate for
extending the Alfvénic cascade to small scales.
When k⊥ ρp  1, finite-Larmor-radius effects modify the properties of the Alfvén
wave. The linear KAW dispersion relation in the gyrokinetic limit with isotropic tem-
peratures is given by (Howes et al. 2006)

|k |vAp k⊥ ρp
ω = ± . (165)
2
βp +
1 + Te /Tp

KAWs are electromagnetic, are elliptically right-hand polarized, and have a frequency
 Ωp in this limit. While large-scale Alfvén waves are non-compressive, KAWs
exhibit fluctuations in the particle density n j and the magnetic-field strength B. Obser-
vations of polarization properties of proton-scale and sub-proton-scale fluctuations in
the solar wind and other space plasmas often find an agreement with the predicted
KAW polarization (Bale et al. 2005; Salem et al. 2012; Chen et al. 2013; Podesta
2013; Roberts et al. 2013; Klein et al. 2014b; Šafránková et al. 2019; Zhu et al. 2019).
The compressive behavior of KAWs introduces fluctuations in the parallel electric
field, allowing KAWs to experience Landau damping (see Sect. 4.2.1). Hybrid fluid-
gyrokinetic simulations suggest that KAW turbulence leads to preferential electron
heating at low βp and to preferential ion heating at high βp (Kawazura et al. 2019). At
low βp , thermal protons do not satisfy the Landau-resonance condition according to
Eq. (157) with n = 0. In this case, the KAW turbulence cascades to even smaller scales,
ultimately leading to preferential electron heating through electron Landau damping
and subsequent collisions. At the same time, nonlinear phase mixing of the ions (see
Sect. 4.2.2) creates smaller structures in the ions’ v⊥ distribution, which eventually

123
The multi-scale nature of the solar wind Page 73 of 136 5

dissipate via collisions and perpendicularly heat the ions. At high βp , KAWs efficiently
dissipate through proton Landau damping and subsequent collisions, which result in
preferential parallel proton heating (Quataert 1998; Leamon et al. 1999; Howes 2010;
Plunk 2013; TenBarge et al. 2013; He et al. 2015; Told et al. 2015; Hughes et al.
2017; Howes et al. 2018). Under certain conditions, KAW turbulence approaches the
local ion-cyclotron frequency in the plasma frame, at which point perpendicular ion
heating through cyclotron-resonant processes (see Sect. 4.2.1) occurs (Arzamasskiy
et al. 2019).
In their stochastic-heating model (see Sect. 4.2.3), Chandran et al. (2010) determine
the proton heating rate for stochastic heating by KAWs in low-βp plasma to be

δvρ
3  c 
2
Q ⊥ = c1 exp − , (166)
ρp ¯

where the empirical factors c1 and c2 are constants, δvρ is the amplitude of the gyro-
scale fluctuations in the E × B velocity, and ¯ ≡ δvρ /w⊥p . Test-particle simulations
using plasma parameters consistent with low-βp solar-wind streams suggest that c1 ≈
0.75 and c2 ≈ 0.34 (Chandran et al. 2010), while reduced MHD simulations suggest
larger values for c1 and smaller values for c2 (Xia et al. 2013).
In intermediate- to high-βp plasma (1  βp  30), the stochastic KAW proton
heating rate is given by (Hoppock et al. 2018)

3
δvρ  σ2
Q ⊥ = σ1 βp exp − , (167)
ρp δ̄

where σ1 and σ2 are constants, δ̄ ≡ δ Bρ /B0 , and δ Bρ is the amplitude of gyro-scale


fluctuations in the magnetic field. Test-particle simulations suggest that σ1 = 5 and
σ2 = 0.21.16

4.3.3 Alfvén/ion-cyclotron waves

Alfvén/ion-cyclotron (A/IC) waves are the short-wavelength extension of the Alfvén-


wave branch for k  k⊥ . The anisotropic Alfvénic turbulent cascade on its own
cannot generate A/IC waves. However, A/IC waves have received considerable atten-
tion due to their ability to heat ions preferentially in the direction perpendicular to B0
through cyclotron resonance (see Sect. 4.2.1; Dusenbery and Hollweg 1981; Isenberg
and Hollweg 1983; Gomberoff and Elgueta 1991; Hollweg 1999; Araneda et al. 2009;
Rudakov et al. 2012).

16 The use of ¯ in Eq. (166) and δ̄ in Eq. (167) reflects the importance of the two different stochastization
mechanisms discussed in Sect. 4.2.3: the electrostatic potential in low-βp plasmas and the magnetic field
in high-βp plasmas.

123
5 Page 74 of 136 D. Verscharen et al.

The linear dispersion relation for quasi-parallel A/IC waves in the cold-plasma limit
(i.e., β j → 0) is given by (Verscharen 2012)
 
ωr k 2 dp2 4
=± 1+ 2 2 −1 . (168)
Ωp 2 k dp

In this regime, the A/IC wave is also known as the L-mode. The frequency is always less
than Ωp , and the quasi-parallel A/IC wave is almost fully left-circularly polarized—
the same sense of rotation as the cyclotron motion of positively charged particles.
This polarization accounts for the frequency cutoff at the proton cyclotron frequency,
above which plasmas are opaque to A/IC waves. For finite-temperature plasmas, ωr
asymptotes to an even smaller value than Ωp since, with increasing temperature, an
increasing number of particles resonate with the Doppler-shifted wave frequency in
their reference frame.
The amplitudes of the perpendicular components of the fluctuating proton and elec-
tron bulk velocities are equal in the limit of k → 0. The amplitude of the perpendicular
proton bulk velocity then increases as ωr → Ωp , while the amplitude of the perpen-
dicular electron bulk velocity remains approximately constant. Therefore, the proton
contribution to the polarization current increases with ωr , until the protons carry most
of the current.
The inherent ambiguities of single-spacecraft measurements (see Sect. 2.6) com-
plicate the identification of A/IC waves within background solar-wind turbulence.
However, A/IC-storms have been observed as enhancements in the magnetic-field
power spectrum at ωr  Ωp with predominantly left-handed polarization (Jian et al.
2009, 2010; He et al. 2011; Jian et al. 2014; Boardsen et al. 2015; Wicks et al. 2016).
A/IC waves damp on particles that fulfill the cyclotron-resonance condition accord-
ing to Eq. (157) in Sect. 4.2.1 with n = + 1,

ωr = k v + Ωp . (169)

This effect heats ions very efficiently in the perpendicular direction. More specifically,
the quasilinear pitch-angle diffusion through the n = + 1 resonance creates a char-
acteristic plateau along pitch-angle gradients, which has often been observed in the
fast solar wind (Cranmer 2001; Isenberg 2001; Marsch and Tu 2001; Tu and Marsch
2001; Hollweg and Isenberg 2002; Gary et al. 2005; Kasper et al. 2013; Cranmer
2014; Woodham et al. 2018). These observations strongly support the A/IC-heating
scenario, but difficulties remain in explaining the origin of these waves in the solar
wind. Microinstabilities may play an important role in the generation of A/IC waves
as we discuss in Sect. 6.

4.3.4 Slow modes

Although most solar-wind fluctuations are non-compressive, about 2% of the fluctu-


ating power is in compressive modes in the inertial range (Chen 2016; Šafránková

123
The multi-scale nature of the solar wind Page 75 of 136 5

et al. 2019). Due to its polarization properties, the slow mode is a major candidate to
explain these compressive fluctuations.
The linear dispersion relation of slow modes in the MHD limit is given by

ωr = ± kC− , (170)

where
 ! 1/2
∗ 1 κ  1  κ 2
C± ≡ vA 1 + βp ± 1 + βp − 2κβp cos θ
2 (171)
2 2 2 2

is the fast (upper sign; see Sect. 4.3.5) and slow (lower sign) magnetosonic speed,
κ is the polytropic index, and θ is the angle between k and B0 . Oblique MHD slow
modes at βp < 2/κ are characterized by an anti-correlation between fluctuations in
density δn j and magnetic-field strength δ|B|. In this limit, the mode is largely acoustic
in nature, and the mode’s velocity perturbation is closely aligned with B0 . In the high-
βp limit, the MHD slow mode is largely tensional in nature, and the mode’s velocity
perturbation δU is predominantly (anti-)parallel to B0 . In both of these limits of the
MHD slow wave, the vector δB lies in the k−B0 plane. In the limit of θ = 0◦ , the MHD
slow wave is either a pure acoustic wave with δB = 0 when βp < 2κ or degenerate
with the Alfvén wave when βp > 2κ. In the limit of θ = 90◦ , the slow mode does not
propagate.
Polarization properties are often more useful than phase speeds in defining the type
of plasma wave. Therefore, we more generally define slow modes as the solutions
to the dispersion relation that exhibit the anti-correlation between δn j and δ|B| that
characterizes the MHD slow mode’s low-βp limit. In kinetic theory, two solutions
exhibit this anti-correlation.17 We consequently identify both of them with the kinetic
slow mode (Verscharen et al. 2017).
The first solution is the ion-acoustic wave (Narita and Marsch 2015), which obeys
the linear dispersion relation

3kB Tp + kB Te
ωr = ± |k | (172)
mp

which can be obtained in the gyrokinetic limit (Verscharen et al. 2017). The phase
speed of this wave is the ion-acoustic speed, which indicates that the parallel pressures
of protons and electrons provide this mode’s restoring force, while the proton mass
provides its inertial force. The protons behave like a one-dimensional adiabatic fluid
since κp = 3, while the electrons behave like an isothermal fluid since κe = 1, where
κ j is the polytropic index of species j.

17 In fact, kinetic linear theory has an infinite number of solutions with this anti-correlation. However,
almost all of them are so heavily damped with |γ |  |ωr | that they are irrelevant for all practical purposes
to the solar wind.

123
5 Page 76 of 136 D. Verscharen et al.

The second type of kinetic slow mode is the non-propagating mode,18 which obeys
the linear dispersion relation

ωr = 0. (173)

If any plasma species has a sufficiently strong temperature anisotropy with T⊥ j > T j ,
the non-propagating mode can become unstable and then gives rise to the mirror-mode
instability (see Sect. 6.1.1).
The anti-correlation of δn j and δ|B|, which defines slow modes, is frequently
observed in the solar wind (Yao et al. 2011; Kellogg and Horbury 2005; Chen et al.
2012b; Howes et al. 2012; Klein et al. 2012; Roberts et al. 2017; Yang et al. 2017a;
Roberts et al. 2018). Figure 18 shows a period of solar-wind measurements that exem-
plify this anti-correlation over a wide range of scales.
Ion-acoustic waves mainly damp through Landau damping (Barnes 1966). Since
the mode’s phase speed is of order the proton thermal speed (unless Te  Tp ),
the ion-acoustic mode predominantly heats ions in the field-parallel direction. We
note that the damping rate of slow modes is significant even at scales  dp . On
this basis, slow modes have at times been rejected as candidates for the compressive
fluctuations in the solar wind. Nevertheless, at very large angles between k and B0 ,
the damping rate decreases significantly, and the ion-acoustic wave and the MHD
slow wave no longer propagate. Instead, they become non-propagating structures that
exhibit pressure balance,

B2
Ptot ≡ P + = constant. (174)

These pressure-balanced structures have been observed often and across many scales
both in the solar wind and in plasma simulations (Burlaga and Ogilvie 1970; Marsch
and Tu 1990b, 1993; Tu and Marsch 1994; Bavassano et al. 2004; Verscharen et al.
2012a; Yao et al. 2013a, b). A recent study suggests that slow modes also play an
important role in how low-frequency, low-β j plasma turbulence partitions heating
between ions and electrons (Schekochihin et al. 2019).

4.3.5 Fast modes

Fast modes are another type of compressive fluctuation, although they are non-
compressive in parallel propagation. Their linear dispersion relation in the MHD
approximation is given by

ωr = ± kC+ , (175)

where C+ is the fast magnetosonic speed according to Eq. (171). Oblique MHD fast
modes at βp < 2/κ are characterized by a positive correlation between fluctuations in
18 The non-propagating kinetic slow mode is sometimes called the kinetic entropy mode in reference to
the non-propagating MHD entropy mode. Although both modes share this non-propagating behavior, the
MHD entropy mode is different from the kinetic slow mode in the sense that it does not exhibit variations
in δ|B|.

123
The multi-scale nature of the solar wind Page 77 of 136 5

Fig. 18 Time series of n e (cm−3 ; green) and |B| (nT; red) in the solar wind on multiple scales, each of which
has fluctuations that clearly exhibit the anti-correlation between δn e and δ|B| that characterizes slow waves.
These panels show data from the Cluster EFW and FGM instruments measured for 1 h starting at 22:30:00
on 2001-04-05. Following the technique by Yao et al. (2011), we show from top to bottom decreasing
interval lengths. The gray lines in each plot indicate the start and end points of the interval shown in the
plot immediately below it. We use a running average to filter the spacecraft spin tones from the data

density δn j and magnetic-field strength δ|B|. In this limit, the mode’s restoring force is
a combination of the total-pressure-gradient force and the magnetic-tension force, and
its velocity perturbation δU lies in the k−B0 plane. In the high-βp limit, the MHD fast
mode is largely acoustic in nature, and the mode’s velocity perturbation δU is mainly
parallel to k. In the limit of θ = 0◦ , the MHD fast wave is either degenerate with the
Alfvén wave when βp < 2κ or a purely acoustic wave with its velocity perturbation
δU parallel to k when βp > 2κ. In the limit of θ = 90◦ , the MHD fast mode is
a magnetoacoustic pressure wave. In the MHD fast wave, the vector δB lies in the
k−B0 plane. Analogous to the case of generalized slow modes, we define fast modes
as the solutions to the linear dispersion relation that exhibit a characteristic positive
correlation between δn j and δ|B| known from the low-βp limit of the MHD fast mode.
On smaller scales, the fast-mode family includes the whistler mode, the lower-
hybrid mode, and the kinetic magnetosonic mode. We refer to all modes of this family
as fast-magnetosonic/whistler (FM/W) waves. In the limit kde  1 in a cold plasma
with quasi-parallel direction of propagation, the linear FM/W-wave dispersion relation

123
5 Page 78 of 136 D. Verscharen et al.

is approximately given by
 
ωr k 2 dp2 4
=± 1+ 2 2 +1 , (176)
Ωp 2 k dp

which connects to the Alfvén-wave branch at small k as in Eq. (168). The quasi-parallel
FM/W wave is also known as the R-mode. In the limit kdp  1 and allowing for oblique
propagation with cos2 θ  m e /m p , the cold-plasma FM/W-wave dispersion relation
can be approximated by

ωr k|k |de2
≈± . (177)
|Ωe | 1 + k 2 de2

In the limit k → ∞, this dispersion relation asymptotes toward ∼ |Ωe | cos θ . In this
regime, the FM/W wave is known as the whistler wave. The amplitudes of the perpen-
dicular components of the fluctuating proton and electron bulk velocities are equal in
the limit of k → 0. The amplitude of the fluctuations in the perpendicular electron bulk
velocity then increases as ωr → |Ωe | while the amplitude of the fluctuations in the
perpendicular proton bulk velocity decreases until the proton bulk velocity is almost
zero. Therefore, the electron contribution to the polarization current increases with ωr
until the electrons carry most of the current. The electrons remain magnetized at these
frequencies, while the protons are unmagnetized. The phase speed of whistler waves
is proportional to k, so waves with a higher frequency travel faster than waves with a
lower frequency. This strongly dispersive behavior of whistler waves is responsible for
their name since they were first discovered as whistling sounds with decreasing pitch
in radio measurements of ionospheric disturbances caused by lightning (Barkhausen
1919; Storey 1953).
In the highly-oblique limit (cos2 θ  m e /m p ), the FM/W wave corresponds to the
lower-hybrid wave. A useful approximation for its linear dispersion relation in the
cold-plasma limit is (Verdon et al. 2009)
 
ωr2 1 mp cos2 θ
≈ 1+ , (178)
ωLH
2 1 + ωe2 /k 2 c2 m e 1 + ωpe
2 /k 2 c2

where
ωpp
ωLH ≡  (179)
ωpe
2
1+
Ωe2

is the lower-hybrid frequency. Under typical solar-wind conditions, βp  10−3 , and


the lower-hybrid wave is very strongly Landau-damped. However, this mode may be
driven unstable by certain electron configurations and thus account for some of the
electrostatic noise observed in the solar wind (Marsch and Chang 1982; Lakhina 1985;
Migliuolo 1985; McMillan and Cairns 2006).

123
The multi-scale nature of the solar wind Page 79 of 136 5

Quasi-parallel FM/W waves are right-hand polarized—the same sense of rotation as


the cyclotron motion of electrons. This polarization results in a frequency cutoff at the
electron gyro-frequency. FM/W waves are almost undamped at ion scales (kde  1).
When they reach the electron scales, they cyclotron-resonate with thermal electrons
very efficiently through the n = − 1 resonance (see Sect. 4.2.1). This leads to efficient
perpendicular electron heating. Oblique FM/W modes can resonate with ions through
other resonances, including the Landau resonance with n = 0.
Quasi-perpendicular FM/W waves have been an alternative candidate to KAWs for
explaining the observed solar-wind fluctuations at k⊥ ρp  1 (Coroniti et al. 1982;
He et al. 2012a; Sahraoui et al. 2012; Narita et al. 2016). However, their existence
is unlikely to result from the large-scale Alfvénic cascade since this scenario would
necessitate a transition from Alfvénic modes to fast modes at some point in the cascade.
The solar wind only rarely exhibits pronounced time intervals with a positive corre-
lation between δn j and δ|B| at large scales (Klein et al. 2012). However, a number of
observations of polarization properties of fluctuations reveal occasional consistency
with the predictions for FM/W waves (Beinroth and Neubauer 1981; Marsch and
Bourouaine 2011; Chang et al. 2014; Gary et al. 2016a; Narita et al. 2016). FM/W
modes may be the result of a class of microinstabilities (see Sects. 6.1.1, 6.1.2) and
thus may be important for the thermodynamics of the solar wind beyond the turbulent
cascade.

5 Plasma turbulence

After a brief introduction to the phenomenology of plasma turbulence in Sect. 5.1,


we discuss the important concepts of wave turbulence in Sect. 5.2 and critical balance
in Sect. 5.3. Section 5.4 closes our description of turbulence with a brief discussion
of more advanced topics. There are many excellent textbooks and review articles on
plasma turbulence (e.g., Tu and Marsch 1995; Bavassano 1996; Petrosyan et al. 2010;
Bruno and Carbone 2013). We refer the reader to this literature for a deeper discussion
of the topic.

5.1 Phenomenology of plasma turbulence in the solar wind

Turbulence is a state of fluids in which their characteristic quantities such as their


velocity or density fluctuate in an effectively unpredictable way.19 Fluids with low
viscosity transition easily into a turbulent flow pattern. Turbulence is inherently a multi-
scale phenomenon. Energy enters the system at large scales. Nonlinear interactions
between fluctuations on comparable scales then transfer the energy to fluctuations on
different scales with a net transfer of energy to smaller and smaller scales. This cascade
of energy occurs through the interaction of neighboring eddies in the fluid that break
up into smaller eddies. At the smallest scales, the fluctuations eventually dissipate into

19 We use the term “unpredictable” here to refer to the statistic nature of turbulence and the notion of
randomness (Leslie 1973). The fluctuations in these quantities are still bound within certain limits and
exhibit correlations.

123
5 Page 80 of 136 D. Verscharen et al.

heat through collisions and raise the medium’s entropy. In a neutral fluid, the injection
at large scales may represent a slow (compared to the characteristic time associated
with the turbulent cascade) stirring mechanism. The dissipation is a consequence
of the viscous interaction, which strengthens with decreasing scale. Turbulence in a
plasma, however, is different from turbulence in a neutral fluid due to the additional,
electromagnetic interactions and the presence of additional, non-viscous dissipation
channels at the characteristic plasma scales (ρ j , d j , λ j , etc.). The solar wind, due to
its low collisionality, exemplifies such a turbulent plasma.
The multi-scale nature of turbulence leads to a broad power-law in the power spectral
density of the fluctuating quantities. For fluid turbulence, a dimensional scale analysis
shows that the power spectral density in the inertial range, which is the range of scales
between the large injection scales and the small dissipation scales, follows a power
law in wavenumber k (see also Fig. 19). Kolmogorov (1941a, b) estimates the power
index of the power spectral density of the fluid velocity fluctuations by employing the
following dimensional analysis. He identifies the dissipation rate with the constant rate
of energy transfer  in the inertial range under steady-state conditions. For an eddy of
size  and velocity difference δU across its extent, the characteristic time to turn over
is approximately τnl ∼ /δU . The transfer rate of energy density for this eddy, on the
other hand, is related to the energy density E through  ∼ E/τnl = constant, where
E ∼ (δU )2 . Combining these relations, we find E ∼ ()2/3 . Relating scale and
wavenumber through  ∼ 1/k and defining the power spectral density as E(k) ∼ E/k
then leads to

E(k) ∼  2/3 k −5/3 . (180)

Such a power law in k is characteristic of turbulent fluids. Indeed, spectra of the


solar wind’s magnetic field, which have been measured in progressively greater detail
for decades, often exhibit this power law (Coleman 1968; Kiyani et al. 2015). We
show an exemplar power spectrum of solar-wind magnetic fluctuations in frequency
in Fig. 19, which spans almost eight orders of magnitude in frequency (for other
examples, see Leamon et al. 1998; Alexandrova et al. 2009; Sahraoui et al. 2010b;
Bruno et al. 2017). We use the same instruments and data intervals in January and
February of 2007 as Kiyani et al. (2015) and compose a spectrum based on a direct
fast Fourier analysis of a 58-day interval from ACE MFI, a 51-h interval from ACE
MFI, a 1-h interval from Cluster 4 FGM, and the same 1-h interval from Cluster 4
STAFF-SC. These time intervals are nested: each interval lies within the next longer
time interval.
When a single spacecraft measures a time series of a fluctuating quantity, it cannot
distinguish between local temporal variations and variations due to the convection of
spatial structures over the spacecraft with the solar-wind speed. Even purely spatial
variations appear as temporal variations, so a power spectrum in frequency reflects
the combined effects of temporal and spatial variations (Taylor 1938). More precisely,
the Doppler shift connects the observed frequency f sc of fluctuations in the spacecraft
frame to the wavevector k and the frequency f 0 of the fluctuations in the plasma frame
through

123
The multi-scale nature of the solar wind Page 81 of 136 5

Fig. 19 Power spectral density of magnetic-field fluctuations in the solar wind during a time interval
with βp ∼ 1. The black lines show power laws with the power indices − 1, − 5/3, and − 2.8, which are
characteristic of the injection, inertial, and dissipation ranges, respectively. The frequency is measured
in the spacecraft reference frame. The average plasma parameters are B = 4.528 nT, n p = 1.02 cm−3 ,
n e = 1.12 cm−3 , Tp = 1.26 MK, Te = 0.138 MK, and Up = 658 km/s. After Kiyani et al. (2015)

1
f sc = f 0 + k · ΔU, (181)

where ΔU is the velocity difference between the spacecraft frame and the plasma
frame. For low-frequency fluctuations (i.e., f 0  k · ΔU), Taylor’s hypothesis sim-
plifies the Doppler-shift relationship in Eq. (181) to

1
f sc ≈ k · ΔU, (182)

which is often used in the analysis of solar-wind fluctuations (for a more detailed
discussion of its applicability, see Howes et al. 2014b; Klein et al. 2014a, 2015;
Bourouaine and Perez 2018). In Fig. 19, we use Taylor’s hypothesis to convert the
convected frequencies associated with the scales d j and ρ j as f d j ≡ Up /2π d j and
f ρ j ≡ Up /2πρ j , respectively, based on the average Cluster 4 FGM, CIS, and PEACE
measurements during the 1-h time interval used in this analysis.
Figure 19 shows all three of the typical ranges observed in the solar wind. At the
lowest frequencies ( f sc  10−4 Hz), is the injection range, which follows a power law
with f sc−1 . For comparison, we note that the expansion time of τ = 2.4 d corresponds
to a frequency of about 5 × 10−6 Hz, while the solar rotation period τrot = 25 d
corresponds to a frequency of about 5 × 10−7 Hz (see Sect. 1.1). The nature and
origin of fluctuations in the injection range are not well understood (Matthaeus and
Goldstein 1986; Verdini et al. 2012; Consolini et al. 2015). The fluctuations exhibit
Alfvénic polarization properties (see Sect. 4.3.1) and B ≈ constant (Matteini et al.
2018; Bruno et al. 2019).

123
5 Page 82 of 136 D. Verscharen et al.

At intermediate frequencies (10−4 Hz  f sc  1 Hz), the inertial range of magnetic


−5/3
fluctuations approximately follows a power law with fsc , which roughly agrees with
Kolmogorov’s theory according to Eq. (180). Fluctuations in other quantities, such as
bulk velocity (Boldyrev et al. 2011) and density (Kellogg and Horbury 2005), have
similar but not identical spectral indices compared to the magnetic fluctuations. The
differences between the magnetic-field and velocity spectra are interpreted as resulting
from significant residual energy being generated at large scales. At high frequencies
( f sc ∼ 1 Hz), the magnetic-field spectrum steepens again toward a power law approx-
imately following f sc−2.8 , which may indicate the beginning of the dissipation range.
The power index at small scales varies, however, and the origin of this break is still
unclear. Recent work suggests that there is a further transition at the electron scales
toward an even steeper slope of the spectrum (Alexandrova et al. 2009; Sahraoui et al.
2009). The e-folding de-correlation time of the 51-h time interval is τc = 18.3 min,
and we define f τc ≡ 1/2π τc as the spacecraft frequency associated with the e-folding
de-correlation length. Like most properties of the solar wind, the fluctuations change
with distance from the Sun. For instance, solar-wind expansion causes the overall level
of fluctuation amplitudes to decrease with distance (Bavassano et al. 1982; Burlaga
and Goldstein 1984). The power of the large-scale magnetic-field fluctuations beyond
a few tens of R decreases approximately ∝ r −3 as predicted by WKB theory (Belcher
and Burchsted 1974; Hollweg 1974). Moreover, the positions of the spectral break-
points vary with distance (Matthaeus and Goldstein 1982; Bavassano and Smith 1986;
Roberts et al. 1987). The spacecraft-frame frequency f b1 of the breakpoint between
the injection range and the inertial range decreases with distance r from the Sun as
f b1 ∝ r −1.5 (Bruno et al. 2009), while the frequency f b2 of the breakpoint between the
inertial range and the dissipation range decreases as f b2 ∝ r −1.09 (Bruno and Trenchi
2014).
The importance of damping and dissipation of plasma turbulence in the solar wind
is underlined by the finding that the energy cascade rate through the inertial range in
solar-wind turbulence (e.g., MacBride et al. 2008) is typically sufficient to explain the
observed heating of the solar wind (see Sect. 1.4.6). These studies are based on the
relationship found by Politano and Pouquet (1998), which estimates the energy transfer
rate assuming isotropy, incompressibility, homogeneity, and equipartition between
magnetic and kinetic energies. However, it is as yet unclear what underlying physics
mechanisms heat the plasma through the damping and dissipation of the turbulent
fluctuations.

5.2 Wave turbulence and its composition

In order to understand the effects of solar-wind turbulence on the multi-scale evolution


of the plasma, we must determine the nature of the fluctuations. Iroshnikov (1963)
and Kraichnan (1965) suggest that MHD turbulence in a strongly magnetized medium
is a manifestation of nonlinear collisions between counter-propagating Alfvén-wave
packets. According to their statistically isotropic theory, the Alfvén-wave-collision
mechanism leads to a power law of the magnetic-field spectrum with

123
The multi-scale nature of the solar wind Page 83 of 136 5

E(k) ∼ k −3/2 (183)

in the inertial range. This work introduced the framework of wave turbulence (see also
Howes et al. 2014a) into plasma-turbulence research. Wave turbulence accounts for
the fact that a plasma, unlike a neutral fluid, carries plasma waves as linear normal
modes for the system (see Sect. 4.1). The linear response of the system still plays a
role in the dynamics of the turbulence, even though the evolution of the turbulence is
nonlinear. Therefore, fluctuations in wave turbulence retain certain characteristics of
the plasma’s linear normal modes such as propagation and polarization properties. In
the wave-turbulence framework, the identification of the nature of plasma turbulence is
thus informed by the identification of the dominant wave modes of the turbulence. As
a caveat to this picture, we note that nonlinear interactions may generate fluctuations
that are not (linear) normal modes of the system as those described in Sect. 4.3.
These driven modes may behave unexpectedly, and linear theory does not predict
their properties.
There are two important timescales associated with fluctuations in wave turbulence:
the linear time τlin and the nonlinear time τnl . The linear time is associated with the
evolution of the plasma’s dominant wave modes due to propagation along B0 . It is
related to the wave frequency through

1
τlin ∼ . (184)
ωr

The nonlinear time is associated with the nonlinear interaction between the modes
perpendicular to the field direction, which leads to the nonlinear cascade process. It
is related to the perpendicular wavenumber k⊥ and the perpendicular fluctuations in
velocity δU⊥ through

1
τnl ∼ . (185)
k⊥ δU⊥

Turbulence is called strong when τlin  τnl and weak when τlin  τnl . Wave turbulence
can exist in the strong and in the weak regime, and we emphasize that the terms wave
turbulence and weak turbulence are not interchangeable.
In the weak-turbulence paradigm, the collision of two waves with frequencies ω1
and ω2 and with wavevectors k1 and k2 most efficiently leads to a resultant wave
with frequency (Montgomery and Turner 1981; Shebalin et al. 1983; Montgomery
and Matthaeus 1995)

ω3 = ω1 + ω2 (186)

and wavevector

k3 = k1 + k2 . (187)

Assuming Alfvén waves with ω = ± k vA∗ (see Sect. 4.3.1), where k ≡ k · B /B ,


 0 0
these wave–wave resonances cannot feed an MHD Alfvén-wave triad with ω3 = 0.

123
5 Page 84 of 136 D. Verscharen et al.

Although k⊥ can increase, these triads lead to a situation with k → 0, where k⊥ ≡


|k − k B0 /B0 |. This weak-turbulence process plays an important role in the onset of
plasma turbulence because it creates increasingly perpendicular wavevectors. Indeed,
spacecraft observations show a strong wavevector anisotropy with k⊥  k in the
solar wind for the majority of turbulent fluctuations (Dasso et al. 2005; Hamilton et al.
2008; Tessein et al. 2009; MacBride et al. 2010; Wicks et al. 2010; Chen et al. 2011a;
Ruiz et al. 2011; Chen et al. 2012a; Horbury et al. 2012; Oughton et al. 2015; Lacombe
et al. 2017).
Indirect measurements of the two-point correlation function

R(r) ≡ B(x) · B(x + r) (188)

and the magnetic helicity



H≡ A · B d3 x, (189)

where · · ·  indicates the average over many positions x, and A is the magnetic vector
potential, independently reveal the existence of two highly-anisotropic components of
turbulence (Matthaeus et al. 1990; Tu and Marsch 1993; Bieber et al. 1996; Podesta
and Gary 2011b; He et al. 2012b). The first component consists of highly-oblique fluc-
tuations with k⊥  k . The second component consists of fluctuations that are more
field-aligned (k⊥  k ) and have lower amplitudes. This discovery led to the notion
of the simultaneous existence of two-dimensional (k  0) turbulent fluctuations
and slab (k⊥  0) wave-like fluctuations. Although this slab+2D model successfully
reproduces the bimodal nature of the fluctuations in the solar wind, it does not account
for a broader distribution of power in three-dimensional wavevector space.
Since waves and turbulence are interlinked through the concept of wave turbulence,
a good understanding of the linear properties of plasma waves (Sect. 4.3) is important
to understand the nature of the fluctuations and their dissipation mechanisms. By
combining these concepts, we achieve a deeper insight into the dissipation mechanisms
of turbulence. Working in the framework of wave turbulence, however, we emphasize
again that we refer to waves as both the classical linear wave modes and the carriers
of the turbulent fluctuations in wave turbulence.

5.3 The concept of critical balance

Critical balance describes the state of strong wave turbulence in which the linear and
the nonlinear timescales from Eqs. (184) and (185) are of the same order (Sridhar and
Goldreich 1994; Goldreich and Sridhar 1995; Lithwick et al. 2007):

ωr (k , k⊥ ) ∼ k⊥ δU⊥ . (190)

The physics justification for critical balance is based on a causality argument (Howes
2015). Initially, a weak-turbulence interaction of two counter-propagating plasma
waves as quantified in Eqs. (186) and (187) generates a pseudo-wave packet with

123
The multi-scale nature of the solar wind Page 85 of 136 5

k  0 and with k⊥ greater than that of either of the first two waves. However, causal-
ity forbids the final state of the turbulence from being completely two-dimensional. If
it were, two planes at different locations along the background magnetic field would
have to be identical if truly k = 0, which precludes any structure along B0 (Mont-
gomery and Turner 1982). These two arbitrary planes, though, can only be identical if
they are able to causally communicate with each other, which occurs via the exchange
of Alfvén waves between them. This interplay between the generation of smaller k
through weak-turbulence interactions and the requirement of causal connection along
B0 creates a situation in which the timescale of the nonlinear interactions in one plane
(i.e., τnl ) is of order the timescale of the communication between the two planes (i.e.,
τlin ). This describes the critical-balance condition in Eq. (190). In this model, the wave
collision creates a pseudo-wave packet with k  0, which then interacts with another
propagating wave from the pool of fluctuations. This results in a new propagating wave
with an even higher k⊥ . This multi-wave process, mediated by pseudo-wave packets
and propagating wave packets, generates anisotropy while still satisfying causality
through the field-parallel propagating waves. This process fills the critical-balance
cone, which is the wavevector space satisfying Eq. (190), as it distributes power in
three-dimensional wavevector space at increasing wavenumbers. Turbulence in the
critical-balance state is still strong turbulence (rather than weak), notwithstanding that
it retains properties of the associated plasma normal modes according to the wave-
turbulence paradigm.
Although the justification of critical balance is still under debate (Matthaeus et al.
2014; Zank et al. 2017), there is a growing body of evidence from spacecraft mea-
surements for the existence of conditions consistent with critical balance and wave
turbulence in the solar wind (for a summary, see Chen 2016). We note, however,
that the fluctuations in the solar wind do not consist of only one prescribed type of
fluctuations (quasi-parallel waves, non-propagating structures and vortices, critically
balanced wave turbulence, etc.) but rather a combination of these.
The concept of critical balance can be further illustrated in the MHD approxima-
tion (see Sect. 1.4.2), which has a long and successful history in plasma-turbulence
research. For incompressible MHD turbulence (∇ · U = 0) consisting of transverse
(δB ⊥ B0 and δU ⊥ B0 ) fluctuations, the Elsasser (1950) formulation of the MHD
equations is a useful parameterization, which has been applied successfully to solar-
wind measurements (Grappin et al. 1990; Marsch and Tu 1990a). We define the
Elsasser variables

δB
z± ≡ δU ∓ √ (191)
4πρ

for forward (upper sign) and backward (lower sign) propagating Alfvén waves with
respect to the background field B0 . Using these variables, we rewrite the MHD momen-
tum equation (51) and Faraday’s law (52) as

∂z± ∗ 1
± vA · ∇ z± = − z∓ · ∇ z± − ∇ Ptot , (192)
∂t ρ

123
5 Page 86 of 136 D. Verscharen et al.

∗ ≡ B /√4πρ is the MHD Alfvén speed and P


tot ≡ P + B /8π . The
where vA 2
0
terms on the left-hand side of Eq. (192) represent the linear behavior of z± , while
the terms on the right-hand side represent their nonlinear behavior. The linear terms
are responsible for propagation effects, while the nonlinear terms are responsible for
the cross-scale interactions, which are the building blocks of Alfvén-wave turbulence.
Using Eqs. (184) and (185), we estimate the frequencies associated with the linear
timescale τlin and the nonlinear timescale τnl from the spatial operators on z± in
Eq. (192) as

1 ∗ v∗
∼ vA ·∇ ∼ A (193)
τlin 

and
1 δU
∼ z∓ · ∇ ∼ , (194)
τnl ⊥

where we define the characteristic scales  and ⊥ parallel and perpendicular with
respect to B0 . In critical balance, τlin ∼ τnl so that

δU v∗
∼ A, (195)
⊥ 

which corresponds to k⊥ δU ∼ k vA ∗ as in Eq. (190). Critical balance predicts that the

inertial-range power spectrum of magnetic-field fluctuations in the direction perpen-


dicular to B0 follows the Kolmogorov slope given by Eq. (180), where k is replaced
by k⊥ . The inertial-range power spectrum of magnetic fluctuations in the direction
parallel to B0 then follows E(k ) ∼ k−2 .
The phenomenological model of dynamic alignment describes an extension of
critical balance (Boldyrev 2005, 2006; Mallet et al. 2015). In this model, the turbulent
velocity fluctuations δU increasingly align their directions with the directions of the
mangetic-field fluctuations δB as the energy cascades toward smaller scales. This
framework predicts two limits depending on the strength of the background magnetic
field. If the background field is strong, the turbulent spectrum follows the Iroshnikov–
Kraichnan slope given by Eq. (183), where k is replaced by k⊥ , in the perpendicular
direction. Conversely, if the background field is weak, the perpendicular spectrum
follows the Kolmogorov slope given by Eq. (180), where k is replaced by k⊥ . This
prediction is consistent with MHD simulations of driven turbulence (Müller et al.
2003). In the fully aligned state, either z+ or z− is exactly zero, so nonlinear interactions
cease.

5.4 Advanced topics

We briefly address three topics of great importance for solar-wind turbulence research
that go beyond the direct focus of our review on the multi-scale nature of the solar
wind: intermittency, reconnection, and anti-phase-mixing.

123
The multi-scale nature of the solar wind Page 87 of 136 5

5.4.1 Intermittency

The two-point speed increment is defined as δu(r ) ≡ U (x + r ) − U (x), where


x is the distance along a straight path through a volume of plasma and · · ·  is the
average over many x. Though the probability distribution of δu(r ) in the solar wind
has a Gaussian distribution at larger scales r , it exhibits non-Gaussian features at
smaller r (Marsch and Tu 1994; Sorriso-Valvo et al. 1999, 2001; Osman et al. 2014a).
Specifically, the distribution develops enhanced tails, which indicate that sharp changes
in velocity occur more frequently than predicted by Gaussian statistics. The increments
in the magnetic field also exhibit this statistical property. These findings suggest that
the solar-wind turbulence is intermittent (i.e., exhibiting bursty patches of increased
turbulence) and forms localized regions of enhanced fluctuations.
The diagnostic called Partial Variance of Increments (PVI) is defined as (Greco
et al. 2008)

|δB(t, τ )|
PVI ≡ "# $, (196)
|δB(t, τ )|2

where δB(t, τ ) ≡ B(t + τ ) − B(t) is the magnetic-field increment in a time-series


measurement of B(t) (Greco et al. 2018). PVI enables the identification of intermit-
tency and allows for the statistical comparison of intermittency in plasma simulations
and solar-wind observations (Wang et al. 2013; Greco et al. 2016). Large PVI values
indicate coherent structures, which are organized and persistent turbulent flow patterns
and are believed to be the building blocks of intermittency. Because non-linearities
are locally quenched inside these coherent structures, they survive longer than the sur-
rounding turbulence. The slow solar wind exhibits greater enhancements in PVI values
than the fast solar wind (Servidio et al. 2011; Greco et al. 2012), which demonstrates
that the slow solar wind contains a greater density of coherent structures than the fast
solar wind (see also Bruno et al. 2003). Regions of increased plasma heating and non-
Maxwellian features in the particle distribution functions tend to occur in and around
coherent structures (Osman et al. 2011; Wan et al. 2012; Karimabadi et al. 2013; Wu
et al. 2013; Wan et al. 2015; Parashar and Matthaeus 2016; Yang et al. 2017b).
Intermittency is a general feature known from fluid turbulence (McComb 1990).
However, it remains unclear how intermittency and wave turbulence interact in the
solar wind and what role intermittency plays in the dissipation of turbulence (Wang
et al. 2014; Wan et al. 2015, 2016; Zhdankin et al. 2016; Perrone et al. 2017; Howes
et al. 2018; Mallet et al. 2019).

5.4.2 Magnetic reconnection

Magnetic reconnection refers to the rearrangement of the magnetic field in a highly-


conducting fluid through resistive diffusion, which leads to a conversion of magnetic-
field energy into particle energy. In regard to plasma turbulence, magnetic reconnection
is a process that is closely related to intermittency. Intermittency is associated with
localized large gradients in the magnetic field, which, according to Ampère’s law in

123
5 Page 88 of 136 D. Verscharen et al.

Eq. (23), corresponds to current sheets: localized regions of enhanced current j, which
are a type of coherent structure as introduced in Sect. 5.4.1 (Karimabadi et al. 2013;
TenBarge and Howes 2013; Howes 2016). Current sheets are candidate regions for
magnetic reconnection, which demonstrates the direct link between turbulence and
reconnection (Matthaeus et al. 1984; Servidio et al. 2009, 2010; Osman et al. 2014b),
and reconnection acts as a dissipation channel for the turbulent fluctuations (Retinò
et al. 2007; Sundkvist et al. 2007; Cerri and Califano 2017; Shay et al. 2018). On the
other hand, reconnection sites are inherently unstable to the tearing instability, which
progressively fragments them into smaller and smaller current sheets (Loureiro et al.
2007; Lapenta 2008; Loureiro and Uzdensky 2016; Tenerani et al. 2016). In this way,
reconnection sites generate a cascade to smaller scales by themselves and thus drive
turbulence. In these progressively fragmented current sheets, the reconnection time
gradually becomes faster than any other timescale, including the nonlinear time (Pucci
and Velli 2014). When this condition is established, reconnection is able to interrupt
the cascade of Alfvén-wave turbulence (Boldyrev and Loureiro 2017; Loureiro and
Boldyrev 2017; Mallet et al. 2017). Therefore, reconnection must be considered when
studying turbulence dynamics at small scales.
For further information on the connection between turbulence, coherent structures,
and reconnection, we recommend the review article by Matthaeus and Velli (2011)
and the comprehensive textbook by Frisch (1995).

5.4.3 Anti-phase-mixing

In Sects. 4.2.1 and 4.2.2, we discuss the formation of smaller velocity-space struc-
ture in the particle distribution function through linear and nonlinear phase mixing.
Anti-phase-mixing, which is a stochastic variant of the plasma echo effect (Gould
et al. 1967), is a process by which small-scale structure is removed from the distribu-
tion function in a turbulent plasma. For electrostatic turbulence, Parker et al. (2016)
and Schekochihin et al. (2016) describe phase mixing and anti-phase-mixing in terms
of the flux of energy in Hermite space of the particle distribution function. Phase
mixing creates a transfer of energy from small to large Hermite moments. In a turbu-
lent plasma with a low collision rate, a stochastic plasma echo creates a transfer of
energy from large to small Hermite moments: effectively from small-scale structure to
large-scale structure in velocity space. It therefore suppresses small-scale structure in
the distribution function and thus non-Maxwellian features that may have otherwise
led to collisional damping after ongoing phase mixing as described in Sect. 4.2.2.
Anti-phase-mixing not only counteracts collisionless damping mechanisms but also
leads to a fluid-like behavior of fluctuations even at low collisionality because higher-
order-moment closures become unnecessary (Meyrand et al. 2019). This process is
potentially responsible for the observed fluid-like behavior of compressive and KAW-
like fluctuations in space plasmas (Verscharen et al. 2017; Wu et al. 2019).

123
The multi-scale nature of the solar wind Page 89 of 136 5

6 Kinetic microinstabilities

Instabilities are mechanisms that transfer energy from free-energy sources, such as
the non-equilibrium particle distributions described in Sects. 1.4.4 and 1.4.5 or large-
amplitude waves, to plasma normal modes that initially have amplitudes at the thermal-
noise level (Rosenbluth 1965). The amplitude of these normal modes then grows
exponentially with time as shown in Eq. (138),

A(x, t) ∝ eγ t , (197)

where γ > 0 is the growth rate of the instability, out of the thermal noise during the
linear phase of the instability, while it extracts energy from its free-energy source. After
the linear phase, the normal-mode amplitude reaches some saturation level, at which
point nonlinear behavior occurs that limits the exponential growth of the instability.
In this section, we focus on small-scale instabilities that have characteristic wave-
lengths of order the particle kinetic scales d j and ρ j and that affect the large-scale
dynamic evolution of the solar wind. We divide these instabilities into two categories.
First, we discuss those associated with non-thermal structure in the particle velocity
distributions, including temperature anisotropies and beams. These instabilities lead
to wave–particle interactions that drive unstable growth. Second, we discuss those
instabilities caused by large-amplitude fluctuations, producing wave–wave interac-
tions that drive unstable growth. This taxonomy provides the organizational structure
for this section.
Generically, both types of instabilities generate small-scale fluctuations in the elec-
tric and/or magnetic field. While the turbulent cascade is dominated by interactions
that are local in wavevector space (see Sect. 5.1), instabilities directly inject energy
into the fluctuation spectrum at small scales. The scattering of particles on these small-
scale field structures acts as an effective viscosity for the large-scale plasma behavior
and thereby influences the thermodynamic evolution of the solar wind (Kunz et al.
2011, 2014; Rincon et al. 2015; Riquelme et al. 2015, 2016, 2017, 2018). As we focus
on the effects of small-scale structure on larger-scale behavior, we point the interested
reader to the complementary review by Matteini et al. (2012) on the complementary
effects of large-scale solar-wind behavior on kinetic-scale phenomena. In particular,
the discussion of the effects of background inhomogeneities at larger scales are left
for later editions of this review.

6.1 Wave–particle instabilities

Wave–particle instabilities are driven by departures of velocity distribution functions


from the Maxwellian equilibrium given in Eq. (59). Such departures are frequently
observed in the solar wind (see Sect. 1.4.4, 1.4.5), but not all of the associated energy is
available to drive the system unstable. For instance, unequal temperatures between dif-
ferent plasma species are not known by themselves to drive wave–particle instabilities,
which has major implications for accretion-disk dynamics in astrophysics (Begelman
and Chiueh 1988; Narayan and McClintock 2008; Sironi and Narayan 2015). A non-

123
5 Page 90 of 136 D. Verscharen et al.

Maxwellian velocity-space structure must conform to specific conditions in order to


drive an instability: i.e., to transfer energy from the particles to the electric and mag-
netic fields. This process simultaneously leads to an exponentially growing mode and
drives the system closer to local thermodynamic equilibrium. Once the system no
longer meets the conditions for instability, the march toward equilibrium halts, and
the system lingers in a state of marginal stability; i.e., the conditions for which γ = 0.
This effect has been identified in numerical simulations (Matteini et al. 2006; Hellinger
and Trávníček 2008), but recent work suggests that dynamic interactions between the
ions and electrons may modify the stability threshold conditions (Yoon and Sarfraz
2017). Gary (1993) and Yoon (2017) offer more details into the theory of unstable
wave–particle interactions in the solar wind.
A variety of different schemes are used to classify wave–particle instabilities (Krall
and Trivelpiece 1973; Treumann and Baumjohann 1997; Schekochihin et al. 2010;
Klein and Howes 2015). Most focus on the spatial scales at which unstable modes are
driven: macroinstabilities and microinstabilities respectively drive unstable modes
with wavelengths much greater than and comparable to kinetic scales. Other classifi-
cations focus on the mechanisms that drive the unstable modes: configuration-space
instabilities are driven by the departure of macroscopic quantities from thermodynamic
equilibrium and thus can be modeled by fluid equations, and kinetic or velocity-space
instabilities are driven by resonant interactions with structures in the particle velocity
distributions.
A prototypical macroscopic configuration-space instability is the Chew–
Goldberger–Low (CGL) firehose instability (Chew et al. 1956), in which the pressure
p⊥ perpendicular to the magnetic field becomes insufficient to counteract the centrifu-
gal force experienced by the particles along a bend in the magnetic field. Without a
sufficiently robust restoring force, initial magnetic perturbations are not damped but
in fact amplified, leading to the growth of a large-scale unstable Alfvén mode.20
A typical microscopic kinetic instability is the ion-cyclotron instability, which is
physically very similar to the cyclotron-resonant damping of A/IC waves discussed in
Sect. 4.2.1 but with γ > 0. A left-hand circularly polarized wave with finite k may
resonantly interact with particles from a narrow range of parallel velocities ≈ vres that
satisfy the resonance condition in Eq. (157) for n = + 1. These resonant particles
diffuse according to the quasilinear diffusion relation in Eq. (154) along trajectories
tangent to semi-circles defined by Eq. (158) around the point (v⊥ , v ) = (0, ωr /k )
in velocity space. At the same time, quasilinear diffusion demands that the particles
diffuse from higher f 0 j toward lower f 0 j . We discuss the differences between the
damped and the unstable cases with the help of Fig. 20, which shows the same situation
as Fig. 15 but a different shape of f 0 j (blue dashed lines). This new shape of f 0 j now
exhibits a temperature anisotropy with T⊥p > Tp , which causes particles to diffuse
toward smaller v⊥ in Fig. 20 rather than toward larger v⊥ as in Fig. 15. This change in
behavior is a direct consequence of the altered alignment between the diffusion paths
(black semi-circles) and the contours of f 0 j (blue dashed lines). The diffusive particle
motion now causes the resonant particles to lose kinetic energy (i.e., a decrease in
20 The CGL marginal stability threshold arises at larger pressure anisotropies than those derived from kinetic
theory (Klein and Howes 2015; Hunana and Zank 2017), which, combined with the limited relevance of a
fluid theory to a weakly collisionless system, limits this instability’s relevance to the solar wind.

123
The multi-scale nature of the solar wind Page 91 of 136 5

Fig. 20 Quasilinear diffusion for an anisotropic particle distribution f 0 j (isocontours shown as blue
dashed lines) unstable to left-hand circularly polarized ion-cyclotron waves with frequency ωr and par-
allel wavenumber k . Unlike the cyclotron-resonant damping case (Fig. 15), the velocity-space diffusion
along the pitch-angle gradients of f 0 j (black semi-circles) at v = vres (gray shaded area) causes res-
2 + v 2 ), which is transferred to the growing
onant particles to lose kinetic energy (i.e., to decrease in v⊥ 
electromagnetic wave. This mechanism drives the kinetic ion-cyclotron instability

v⊥
2 + v 2 ), which is transferred to growing field fluctuations. Importantly, the direction

of the energy flow between the fields and the particle distribution depends on the local
sign of the pitch-angle gradient of f 0 j at the resonance speed according to Eq. (155). In
addition to temperature anisotropies, drifting populations and other non-Maxwellian
features can lead to pitch-angle gradients that drive resonant instabilities.
Despite their apparent similarity, the macro/micro and configuration/kinetic
schemes are not synonymous. Some instabilities occur at large spatial scales but are
driven by velocity-space effects. For example, the mirror-mode instability (South-
wood and Kivelson 1993) is driven by the interaction between the slow-mode-like
anti-phase response of bulk thermal and magnetic fluctuations, δ p and δ|B|, and the
in-phase response felt by particles with v ∼ 0. This latter population is approxi-
mately stationary along the background magnetic field and gains or loses energy with
changes in the magnetic-field strength. On the other hand, the bulk population, which
does move parallel to the magnetic field in a slow-mode-like polarized wave (see
Sect. 4.3.4), is able to effectively conserve energy via transfer between parallel and
perpendicular degrees of freedom.
The numerical evaluation of linear instabilities in kinetic theory follows the same
procedure as the numerical evaluation of wave dispersion relations described in
Sect. 4.1: the linearized Vlasov equation is used to calculate the dielectric tensor .
Solutions to the dispersion relation in Eq. (152) with γ > 0 for a particular wavevector
k represent linear kinetic instabilities, which grow with time according to Eq. (197).
Following from the linear set of Vlasov–Maxwell equations, these solutions are inde-
pendent of the fluctuation amplitude. In contrast, the wave–wave instabilities discussed
in Sect. 6.2 depend on fluctuation amplitude.

123
5 Page 92 of 136 D. Verscharen et al.

The behavior of instabilities in the inhomogeneous and turbulent solar wind as


well as the nonlinear evolution of plasma instabilities are important matters of ongo-
ing research. Most numerical evaluations of linear instabilities assume homogeneous
plasma conditions, which are not fulfilled in the solar wind in general. For instance, the
expansion of the plasma, the interaction of different plasma streams, and the ubiquitous
turbulence create inhomogeneities and temporal variability that call into question the
assumption of homogeneity. Nevertheless, the solar wind’s parameter space is often
observed to be restricted by the linear-instability thresholds, which suggests that linear
theory bears some applicability to the solar wind.
We define the marginal stability threshold as a contour of constant maximum growth
rate γm at any k through parameter space for a given instability. The choice of the
relevant γm is somewhat arbitrary. Assuming that only a couple of parameters (e.g.,
β j and T⊥ j /T j ) have a significant impact on the growth rate of a specific instability,
it is possible to construct a parametric model for the instability threshold. The inverse
relation between a species’ temperature anisotropy and β j serves as the prototypical
example of such a threshold model, given for instance by Gary et al. (1994a, b), Gary
and Lee (1994), and Hellinger et al. (2006):

T⊥ j a
=1+ b
, (198)
T j β j − c

where a, b, and c are constant parameters calculated from fits to solutions of the hot-
plasma dispersion relation. This form for the inverse relation is introduced by Hellinger
et al. (2006) for a bi-Maxwellian proton background distribution function according
to Eq. (61) and an isotropic Maxwellian electron distribution. The values of a, b, and
c are different for the four unstable modes that can be driven by proton temperature
anisotropies (i.e., the ion-cyclotron, parallel firehose, mirror-mode, or oblique firehose
instability), as well as the desired maximum growth rates. Verscharen et al. (2016)
compare the parameters a, b, and c for thresholds depending on maximum growth
rates. Table 3 lists best-fit values for these parameters for three different γm /Ωp -values
for each of the four instabilities driven by proton temperature anisotropy. The growth
rates have been calculated for a quasi-neutral plasma consisting of bi-Maxwellian
protons and Maxwellian electrons with Te = Tp and vAp /c = 10−4 . The values of
a, b, and c change in the presence of other plasma components, including beams and
minor ion components, which may act as additional sources of free energy or may
stabilize unstable growth (Price et al. 1986; Podesta and Gary 2011a; Maruca et al.
2012; Matteini et al. 2015a). If the underlying distribution has a shape other than
bi-Maxwellian—e.g., if the particles have a κ-distribution according to Eq. (62) or a
bi-κ-distribution according to Eq. (63)—these threshold curves can be significantly
different (Summers and Thorne 1991; Xue et al. 1993; Summers et al. 1994; Xue et al.
1996; Astfalk et al. 2015; Astfalk and Jenko 2016). The exploration of more general
phase-space densities requires direct numerical integration of the dispersion relation
(Dum et al. 1980; Matsuda and Smith 1992; Astfalk and Jenko 2017; Horaites et al.
2018a; Verscharen et al. 2018). Such general distributions produce instabilities that are
either enhanced or suppressed relative to those associated with bi-Maxwellian particle
distributions.

123
The multi-scale nature of the solar wind Page 93 of 136 5

Table 3 Fit parameters for


Instability a b c
isocontours of constant
maximum growth rate γm = 10−2 Ωp
γm = 10−2 Ωp , γm = 10−3 Ωp ,
Ion-cyclotron 0.649 0.400 0.000
and γm = 10−4 Ωp in the
βp −T⊥p /Tp plane for use in Mirror-mode 1.040 0.633 − 0.012
Eq. (198) Parallel firehose − 0.647 0.583 0.713
Oblique firehose − 1.447 1.000 − 0.148

γm = 10−3 Ωp
Ion-cyclotron 0.437 0.428 − 0.003
Mirror-mode 0.801 0.763 − 0.063
Parallel firehose − 0.497 0.566 0.543
Oblique firehose − 1.390 1.005 − 0.111

γm = 10−4 Ωp
Ion-cyclotron 0.367 0.364 0.011
Mirror-mode 0.702 0.674 − 0.009
Parallel firehose − 0.408 0.529 0.410
Oblique firehose − 1.454 1.023 − 0.178
Calculated with the NHDS code (Verscharen and Chandran 2018) and
adapted from Verscharen et al. (2016)

Table 4 lists the wave–particle instabilities that are most important in regulating the
large-scale dynamics of the solar wind. Many foundational publications (e.g., Hollweg
1975; Schwartz and Roxburgh 1980; Gary 1993) provide more complete catalogues.
Two of the most common free-energy sources are distinct temperatures or pressures
perpendicular and parallel to the background magnetic field and the presence of faster
populations that form a shoulder on or a beam distinct from the core population (Fig. 4).
These two specific cases are considered in Sects. 6.1.1 and 6.1.2, with particular
emphasis on their impact on the macroscale behavior of the solar wind. Significant
work has been done on the effects of instabilities in other space environments such as
the magnetosphere and magnetosheath (Maruca et al. 2018, and references therein),
but these results lie beyond the scope of this work.

6.1.1 Temperature anisotropy

Wave–particle instabilities associated with temperature anisotropies serve as a canoni-


cal example for the effects of wave–particle instabilities on the solar wind’s large-scale
evolution. Initial investigations of instability limits on solar-wind proton temperature
anisotropy address either the T⊥p > Tp limit or the T⊥p < Tp limit separately. For
the former, Gary et al. (2001) find that the ion-cyclotron stability threshold limits the
maximum anisotropy of observations from the ACE spacecraft. For the latter limit,
Kasper et al. (2002) find that the Wind spacecraft’s temperature-anisotropy values
are mostly bounded by the parallel firehose instability threshold. Subsequent work
(Hellinger et al. 2006) shows that, for the slow solar wind, the distribution of tem-
perature anisotropies is well constrained for T⊥p /Tp > 1 and T⊥p /Tp < 1 by the

123
5

Table 4 Wave–particle instabilities relevant to the solar wind organized by free-energy source

Instability Classification Unstable normal mode References

123
T⊥i /Ti > 1a
Ion-cyclotron Micro/resonant Parallel A/IC Kennel and Petschek (1966) and Davidson and Ogden
Page 94 of 136

(1975)
Mirror-mode Macro/resonant Non-propagating oblique kinetic Tajiri (1967), Southwood and Kivelson (1993) and Kivelson
slow mode and Southwood (1996)

T⊥i /Ti < 1


Parallel firehose Micro/resonant Parallel FM/W Quest and Shapiro (1996) and Gary et al. (1998)
Oblique firehose Micro/resonant Non-propagating oblique Alfvén Hellinger and Matsumoto (2000)

T⊥e /Te < 1


Parallel electron firehose Micro/resonant Parallel FM/W Hollweg and Völk (1970) and Gary and Madland (1985)
Oblique electron firehose Micro/configuration Oblique non-propagating Alfvén Li and Habbal (2000) and Kunz et al. (2018)

T⊥e /Te > 1


Whistler anisotropy Micro/resonant Parallel FM/W Kennel and Petschek (1966) and Scharer and Trivelpiece
(1967)

P⊥ /P < 1b
CGL firehose Macro/configuration Non-propagating oblique Alfvén Chew et al. (1956)

Electromagnetic beam
Ion/ion RH resonant Micro/resonant Parallel FM/W Barnes (1970)
Ion/ion nonresonant Macro/configuration Backward propagating firehose-like Sentman et al. (1981) and Winske and Gary (1986)
Ion/ion LH resonant Micro/resonant Parallel A/IC Sentman et al. (1981)
Electron/ion Micro/resonant FM/W and A/IC modes Akimoto et al. (1987)
Electron heat flux Micro/resonant Parallel FM/W Gary et al. (1975, 1994c, 1999), Gary and Li (2000),
Horaites et al. (2018a) and Tong et al. (2018)
D. Verscharen et al.
Table 4 continued

Instability Classification Unstable normal mode References

Ion drift Micro/resonant Parallel and oblique FM/W Verscharen and Chandran (2013)
Ion drift Micro/resonant Parallel and oblique A/IC Verscharen and Chandran (2013)

Ion drift and anisotropy Micro/resonant Parallel FM/W and A/IC Verscharen et al. (2013a) and Bourouaine et al. (2013)
For each instability, we list its name, classification, name of the unstable normal mode, and further references
a Resonant instabilities due to temperature anisotropies can arise for each ion species (index i; see Maruca et al. 2012)
b Configuration-space instabilities are triggered by contributions to the total excess pressure from each plasma species (Kunz et al. 2015; Chen et al. 2016)
The multi-scale nature of the solar wind
Page 95 of 136

123
5
5 Page 96 of 136 D. Verscharen et al.

Fig. 21 Probability distribution of the pristine solar wind in the βp –T⊥p /Tp plane. The instability thresh-
olds for the four instabilities associated with proton temperature anisotropy according to Eq. (198) and
Table 3 with γm = 10−2 Ωp are plotted for comparison. We only plot bins containing at least 25 counts.
A significant fraction of the distribution exceeds the two resonant thresholds (ion-cyclotron and parallel
firehose), while the non-resonant mirror-mode and oblique-firehose thresholds set more precise boundaries
to the data distribution

threshold of each of the configuration-space instabilities: i.e., the mirror-mode and


oblique firehose instabilities. The probability distribution of data in the βp −T⊥p /Tp
plane using measurements from the Wind spacecraft is illustrated in Fig. 21.21 We use
the same dataset as described by Maruca and Kasper (2013).
Interestingly, as seen in Fig. 21, the solar wind is not constrained by all possible
temperature-anisotropy thresholds: a significant portion of the βp −T⊥p /Tp distribu-
tion extends beyond the ion-cyclotron threshold, which, for βp  1, sets a stricter
limit on the departure from isotropy than the mirror-mode instability threshold, as is
pointed out by Hellinger et al. (2006). Several justifications for this apparent inactivity
of the ion-cyclotron instability have been proposed: low efficiency of energy extrac-
tion (Shoji et al. 2009), stabilizing effects of minor ions and/or drifts (Maruca 2012;
Maruca et al. 2012), or quasilinear flattening of the resonant region (Isenberg et al.
2013).

21 Plots of the data distribution in the β −T /T plane have become colloquially known as “Brazil
p ⊥p p
plots” due to the characteristic shape of the data distribution for near-Earth solar wind.

123
The multi-scale nature of the solar wind Page 97 of 136 5

A naïve model for the expanding solar wind would have T⊥ j and T j follow the
double-adiabatic prediction [see Eqs. (44) and (45) in Sect. 1.4.1]. Using data from
Helios and Ulysses at different heliocentric distances, Matteini et al. (2007) show
that the distribution in βp −T⊥p /Tp space follows a radial trend, albeit one with a
smaller radial gradient than that predicted by double-adiabatic expansion, until the
system encounters the instability thresholds. Then, the distribution’s anisotropy is
constrained by the parametric thresholds to the stable parameter space.
Identifying polarization and other linear quantities associated with the predicted
instabilities allows us to infer the presence of modes driven by temperature-anisotropy
instabilities. For instance, the signal of strongly peaked magnetic helicity near par-
allel ion-kinetic scales (He et al. 2011; Podesta and Gary 2011b; Klein et al. 2014b)
indicates the presence of parallel-propagating FM/W or A/IC waves associated with
proton temperature-anisotropy instabilities. Wind observations provide evidence for
enhanced magnetic fluctuations near threshold boundaries (Bale et al. 2009), suggest-
ing that instabilities are active near these thresholds in generating unstable modes
which are associated with such fluctuations. Ion temperature (Maruca et al. 2011;
Bourouaine et al. 2013) and intermittency (Osman et al. 2012; Servidio et al. 2014)
are also found to be enhanced in marginally unstable parameter regions. Calculating
polarization as a function of T⊥p /Tp and βp reveals the presence of a population
of A/IC waves in the region in which they are expected to become unstable (Telloni
and Bruno 2016). The identification of parallel-propagating A/IC waves (e.g., Jian
et al. 2009, 2010, 2014; Gary et al. 2016b) that do not naturally arise from critically
balanced turbulence (see Sect. 5.3) serves as further, indirect evidence for the action
of these instabilities.
We emphasize that caution must be exercised in the analysis of β j −T⊥ j /T j
plots. Hellinger and Trávníček (2014) raise concerns about the effects of projecting
the distribution of quantities onto any reduced parameter space. By partitioning the
data into different temperature quartiles and studying the temperature-anisotropy dis-
tribution of each, they find that enhanced quantities near the instability thresholds
may primarily result from underlying correlations between solar-wind tempera-
tures and speeds. Moreover, it is important to carefully account for the blurring of
temperature-anisotropy observations due to the finite time required to construct a
velocity distribution measurement (Verscharen and Marsch 2011; Maruca and Kasper
2013).
In addition to instabilities triggered by the temperature anisotropy of the core pro-
ton velocity distribution, anisotropic distributions of the other plasma components,
including the electrons (Hollweg and Völk 1970; Gary and Madland 1985; Li and
Habbal 2000; Kunz et al. 2018) and heavy ions (Ofman et al. 2001; Maruca et al.
2012; Bourouaine et al. 2013) can lead to resonant instabilities. We discuss the com-
bined effect of these sources of free energy in Sect. 6.1.3.

6.1.2 Beams and heat flux

The relative drift between plasma components is another common source of free
energy that can drive wave–particle instabilities. The velocity difference between the
two components (of the same or different species) can contribute to excess parallel

123
5 Page 98 of 136 D. Verscharen et al.

pressure or induce non-zero currents, and the drifting distributions themselves may
resonate with unstable waves (e.g., the parallel propagating beam instability described
by Verscharen et al. 2013b). As with temperature anisotropies, some thresholds asso-
ciated with drifts and beams constrain the observed data distributions in parameter
space.
Beam and heat-flux instabilities regulate non-thermal features in the electron dis-
tribution function. For instance, Tong et al. (2018) find compelling evidence that the
heat-flux-driven Alfvén-wave instability limits the electron core drift with respect to
the halo and the protons. To some degree, this result contradicts the earlier work of
Bale et al. (2013), who find that the collisional transport rather than a heat-flux insta-
bility is more active in limiting the electron-core drift (see also Sect. 3.3.2). However,
collisions and kinetic instabilities can co-exist in the solar wind and simultaneously
regulate the heat flux. The electron-strahl heat flux can drive oblique instabilities of
the lower-hybrid and the oblique FM/W wave (Omelchenko et al. 1994; Shevchenko
and Galinsky 2010; Vasko et al. 2019; Verscharen et al. 2019a).
Likewise, ion beams can drive plasma instabilities. Bourouaine et al. (2013) report
constraints on the drift of α-particles relative to protons through parallel-propagating
A/IC and FM/W instabilities. These ion-beam instabilities result in a quasi-continuous
deceleration of the α-particles, which leads to a quasi-continuous release of energy
from the α-particle kinetic energy into field fluctuations (Verscharen et al. 2015). Fig-
ure 22 shows, as functions of distance from the Sun, the rate of energy-density release
Q flow derived from energy conservation as well as the empirical perpendicular heating
rates Q ⊥p for protons and Q ⊥α for α-particles. Q flow > Q ⊥α at distances between
0.3 and 1 au, and Q flow > Q ⊥p at distances between 0.3 and 0.4 au. This finding
suggests that the energy release through α-particle instabilities comprises a signifi-
cant fraction of the solar wind’s overall energy, and that large-scale solar-wind models
must account for α-particle thermodynamics. Due to the lack of in-situ measurements
at smaller heliocentric distances, we are unable to compare Q flow with Q ⊥p or Q ⊥α
closer to the Sun yet; however, we expect this trend to continue toward the acceleration
region of the solar wind.

6.1.3 Multiple sources of free energy

Under typical solar-wind conditions, multiple sources of free energy are simultane-
ously available to drive distinct unstable modes. For example, beams, temperature
anisotropies, and anisothermal temperatures between species are all frequently and
simultaneously present in solar-wind plasma (Kasper et al. 2008, 2017). The intro-
duction of an additional source of free energy can act either to enhance an instability’s
growth rate or act to stabilize the system.
The thresholds of configuration-space instabilities (i.e., the mirror-mode and the
oblique firehose instabilities) depend on the total free energy in the system (Chen et al.
2016). The threshold of the oblique firehose instability limits the observed plasma to
the stable parameter space, when the combined effects of ion and electron anisotropies
as well as relative drifts between the plasma species are considered. Less than 1% of
the observations exceed this threshold, and, for these intervals, the proton, electron,
and α-particle components all significantly contribute to the system’s unstable growth.

123
The multi-scale nature of the solar wind Page 99 of 136 5

Fig. 22 Rate of energy release Q flow from the deceleration of α-particles through kinetic microinstabilities
as a function of distance in the inner heliosphere. We assume that the α-particle drift speed is always fixed to
the local threshold for the FM/W instability based on average fast-solar-wind measurements from Helios.
Q flow then follows from energy conservation. Q ⊥p and Q ⊥α are calculated based on Eq. (44), setting
q⊥ j = 0 and the right-hand side to Q ⊥ j . Using empirical profiles for B, p⊥ j , n j , and U j for j = p and
j = α then gives the empirical heating rates Q ⊥p and Q ⊥α . Adapted from Verscharen et al. (2015)

According to an analytical model of the coupling between the effects of temperature


anisotropy and drifts (Ibscher and Schlickeiser 2014), the combined effects of these
free-energy sources yield a threshold in the region of parameter space with βp < 1
and T⊥p < Tp . This is consistent with the lack of solar-wind observations in this
region of parameter space (see Fig. 21). However, Bale et al. (2009) do not find
enhanced fluctuations or other indications of unstable-mode generation in this region,
and Vafin et al. (2019) explain the lack of data in this region through collisional effects.
The coupling of temperature anisotropy and beams has been incorporated into an
improved threshold model for limiting proton-temperature-anisotropy observations
(Vafin et al. 2018), which may be tested in future in-situ observations of low-βp
systems such as the near-Sun solar wind. Verscharen et al. (2013a) provide testable
limits on temperature anisotropy and α-particle drifts, which Bourouaine et al. (2013)
find to largely agree with solar-wind observations. Numerical simulations (e.g., by
Maneva and Poedts 2018) are also used to study the simultaneous impact of drifts
and temperature anisotropies. The coupling between electrons and ions modifies the
solar-wind expansion, preventing a uniform progression of the bulk thermodynamic
properties toward the firehose threshold (Yoon and Sarfraz 2017). This effect occurs
in addition to the effects of collisions on drawing the solar wind toward isotropy (see
Sect. 3.3), which is found to be important but insufficient for a complete description
of the solar wind’s observed state (Yoon 2016).
Instead of relying solely on analytical threshold models, which are formally valid
for low-dimensional sub-spaces (e.g., βp and T⊥p /Tp only) of the full parameter
space that characterizes the solar wind, the Nyquist instability criterion accounts for
the simultaneous effects of all wave–particle free-energy sources (Nyquist 1932).

123
5 Page 100 of 136 D. Verscharen et al.

Fig.  23 Illustration of the Nyquist instability criterion. Black lines indicate isocontours of
det D(k, ωr + iγ ) for a stable (left) and unstable (right) system, with the normal-mode solutions indi-
cated with red dots. The contour integral is performed over the entire upper half plane, symbolized by
the blue curve (which would formally extend out to ωr → ±∞). Applying the residue theorem yields a
non-negative integer Wn equal to the number of unstable modes supported by the system

This method determines whether a system supports any growing modes at a particular
given wavevector k by performing a complex contour integration, which is illustrated
in Fig. 23. The normal modes of a system are the solutions to det [D(k, ω)] = 0
according to Eq. (152), where D is the system’s dispersion tensor. As described in
Sect. 4.1, the form of D depends on the set of system parameters such as temperature,
density, and drift of each plasma component. The number of modes satisfying γ > 0
can be ascertained by applying the residue theorem to the integral

%
1 dω
Wn = , (199)
2πi det [D(k, ω)]

where the contour is taken over the upper half plane of complex frequency space ω =
ωr +iγ . The integration in Eq. (199) is much easier to compute than the determination
of the dispersion relation for all individual potentially unstable modes. This method
has more than half a century of productive use in the study of plasma stability (Jackson
1958; Buneman 1959; Penrose 1960; Gardner 1963).
Klein et al. (2017) present a modern automatic implementation of the Nyquist
instability criterion for the case of an arbitrary number of drifting bi-Maxwellian com-
ponents. The application of this criterion to a statistically random set of solar-wind
observations modeled as a collection of proton core, proton beam, and α-particle com-
ponents (each with distinct anisotropies, densities, and drifts) finds that a majority of
intervals are unstable (Klein et al. 2018). Most of the unstable modes are resonant insta-
bilities at ion-kinetic scales and with growth rates less than the instrument integration
time and convected kinetic scales. About 10% of the intervals have instabilities with
growth rates of order the nonlinear turbulent cascade rate 1/τnl at proton-kinetic scales,
which indicates that they may grow quickly enough to compete with the background
turbulence.

123
The multi-scale nature of the solar wind Page 101 of 136 5

6.2 Wave–wave instabilities

Wave–wave instabilities, in contrast to wave–particle instabilities, depend sensitively


on the amplitudes of the plasma fluctuations. The finite amplitudes of fluctuating waves
lead to violations of the linearization used to derive the wave–particle instabilities
discussed in Sect. 6.1. Instead, nonlinear effects allow for wave–wave coupling to
lead to unstable wave growth, which places limits on the amplitudes of magnetic and
velocity fluctuations.

6.2.1 Parametric-decay instability

The parametric-decay instability (PDI) is a classic wave–wave instability first


described by Galeev and Oraevskii (1963) and Sagdeev and Galeev (1969) for a three-
wave interaction. It belongs to a broader class of parametric instabilities that also
includes beat and modulational instabilities (Hollweg 1994). In the low-βp limit, the
PDI causes a finite-amplitude forward-propagating Alfvén wave, known as the pump
mode, to decay into a backward-propagating Alfvén wave and a forward-propagating
acoustic wave. Goldstein (1978) provides a generalization of this instability for
circularly-polarized Alfvén waves in finite-βp plasmas. The dynamics of such instabil-
ities are important for the evolution of the solar wind. As described in Sect. 4.3.4, the
compressive acoustic mode can efficiently dissipate and thus heat the plasma (Barnes
1966). Furthermore, the generation of counter-propagating Alfvén waves is essential
for driving the turbulent cascade (see Sect. 5.2). Malara and Velli (1996) show that,
even in the large-amplitude limit and when the pump mode is non-monochromatic,
the PDI continues to operate without a significant reduction in its growth rate. Theo-
retical work suggests that the PDI may develop an inverse cascade near the Sun and,
therefore, be essential in driving solar-wind turbulence (Chandran 2018).
A number of numerical simulations investigate the presence and effects of decay
instabilities under conditions approximating the solar wind (Matteini et al. 2010; Ver-
scharen et al. 2012b; Tenerani and Velli 2013, 2017; Shoda and Yokoyama 2018;
Shoda et al. 2018). A recent analysis of solar-wind observations at 1 au (Bowen et al.
2018) indicates a strong correlation between observed compressive fluctuations and
higher estimated PDI growth rates, which is consistent with the parametric decay of
Alfvén modes. Parametric instabilities are also observed in laboratory plasma experi-
ments (Dorfman and Carter 2016).

6.2.2 Limits on large-amplitude magnetic fluctuations

In addition to decay instabilities, finite-amplitude waves are capable of self-


destabilization. Linearly polarized, large-amplitude Alfvén waves drive compressions
in the plasma, which reduce the amplitude of the Alfvénic fluctuations if δ|B| = 0 (see
also Sect. 4.3.1 of this review; Hollweg 1971). This effect may lead to the observed
preference for Alfvénic fluctuations with B = constant. A related example of such
behavior occurs if the amplitude δ B⊥ /B0 of the perpendicular magnetic fluctuations
−1/2
exceeds the threshold ∼ βp (Squire et al. 2016). Beyond this limit, the pressure

123
5 Page 102 of 136 D. Verscharen et al.

anisotropy associated with the wave fluctuations exceeds the parallel-firehose limit
and destroys the restoring force associated with the magnetic tension, which desta-
bilizes the wave. Numerical simulations confirm signatures of this instability, which
are currently also being sought in solar-wind observations under high-βp conditions
(Squire et al. 2017a, b; Tenerani and Velli 2018).

6.3 The fluctuating-anisotropy effect

Large-scale compressive fluctuations with finite amplitudes and ωr  Ωp modify


the plasma moments, including β j and T⊥ j /T j according to Eqs. (44) and (45).
These and potentially other plasma moments (like the relative drifts between species)
fluctuate with the large-scale compressive fluctuations (Squire et al. 2017a, b; Ten-
erani and Velli 2018). If the amplitude of these fluctuations is sufficiently large, these
modifications can move the system from a stable to an unstable configuration with
respect to anisotropy-driven kinetic microinstabilities (Verscharen et al. 2016). The
instability then acts to modify the velocity distribution, e.g., by pitch-angle scattering
particles. It suppresses further growth of the anisotropy, which leads to a reduction in
the amplitude of the large-scale compressive fluctuations and an isotropization of the
particles. Whether this process occurs depends on the polarization and amplitude of the
large-scale compressive mode. Compressive ion-acoustic modes (see Sect. 4.3.4) with
reasonable magnetic fluctuation amplitudes (δ|B|/B0  0.04) can trigger this effect
with temperature-anisotropy-driven instabilities under typical solar-wind conditions
at 1 au. This fluctuating-anisotropy effect can be generalized to a fluctuating-moment
effect, which includes, for instance, variations in relative drift speeds that may trigger
additional instabilities.

7 Conclusions

We briefly summarize our discussion of the multi-scale nature of the solar wind, give
an outlook on future developments in the field, and outline the broader impact of this
research topic.

7.1 Summary

As we summarize in Fig. 24, the solar wind’s dynamics and thermodynamics result
from an intricate multi-scale coupling between global expansion effects and local
kinetic processes. The global expansion shapes particle distribution functions slowly
compared to most of the collective plasma timescales and creates the ubiquitous non-
equilibrium features of solar-wind particles. It also generates gradients in the plasma
bulk parameters that drive Sunward-propagating waves, which subsequently interact
with anti-Sunward-propagating waves to generate turbulence. By creating microphys-
ical features and turbulence, the expansion couples to small scales and sets the stage
for collisional relaxation, the dissipation of waves and turbulence, and kinetic microin-
stabilities to act locally. On the other hand, these local processes couple to the global

123
The multi-scale nature of the solar wind Page 103 of 136 5

Fig. 24 Summary of the multi-scale couplings in the solar wind. We describe the effects of collisions in
Sect. 3, the effects of waves in Sect. 4, the effects of turbulence in Sect. 5, and the effects of microinstabilities
in Sect. 6. The arrows illustrate the connections and interactions discussed in this review article

scales and modify the large-scale plasma flow by, for example, accelerating the plasma,
changing the plasma temperatures, introducing temperature anisotropies, regulating
heat flux, or generating electromagnetic structures for particles to scatter on. These
effects then modify the expansion. Figure 24 includes some processes (e.g., reflection-
driven waves) that we will discuss in the next major update of this Living Review.
We derive our understanding of the solar wind’s multi-scale evolution from detailed
measurements of its particles and fields. In-situ observations provide perspective on
small-scale processes, while remote observations provide perspective on large-scale
processes. Therefore, we rely on the combination of in-situ and remote observations,
in concert with theoretical modeling efforts and numerical simulations to elucidate the
multi-scale evolution of the solar wind. This review describes the current state of the
art of the field based on a combination of observational discoveries and fundamental
plasma physics.

7.2 Future outlook

Major new space missions such as Parker Solar Probe (PSP; Fox et al. 2016) and Solar
Orbiter (SO; Müller et al. 2013) are dedicated to the study of the processes at the heart
of this review.
PSP, which launched in August 2018 and achieved its first perihelion in November
2018, is beginning to measure in-situ plasma properties with unprecedented energy
and temporal resolution and at unexplored heliocentric distances (see Fig. 8). New
findings derived from PSP will transform our understanding of plasma processes near
the Sun. PSP is expected to provide our first in-situ observations of the corona, which

123
5 Page 104 of 136 D. Verscharen et al.

are anticipated to draw together the heliospheric and solar communities and to enable
novel combinations of in-situ and remote observations.
SO will measure the solar-wind properties through both in-situ measurements of
the local plasma conditions and remote observations of the Sun’s surface. A major
goal for SO is linkage science: connecting processes in and near the Sun with the
behavior of solar-wind plasma across all relevant scales. SO’s inclined orbit will carry
it out of the ecliptic plane and enable it to sample solar wind from polar coronal holes
with its more extensive instrumentation package compared to PSP. Both PSP and SO
will drive research into the multi-scale nature of the solar wind for decades.
Other heliospheric missions that are currently being developed and proposed will
directly address the topics of this review. These include mission concepts to investigate
the nature of waves and turbulence through multi-point and multi-scale measurements
as well as mission concepts to resolve the smallest natural plasma scales in the solar
wind (e.g., National Academy of Sciences, Engineering, and Medicine 2016; Klein
et al. 2019; Matthaeus et al. 2019; TenBarge et al. 2019; Verscharen et al. 2019b).
These efforts demonstrate that the heliophysics community understands the need to
investigate the multi-scale couplings of plasma processes and their impact on the
dynamics and thermodynamics of the solar wind.
We also anticipate major advances in modeling in the near future. Previously,
numerical simulations of processes that connect over large scale separations required
computational resources too great for them to be practical. Therefore, most models
either focused on global expansion dynamics (e.g., global MHD simulations) or on
local plasma processes (e.g., homogeneous-box particle-in-cell simulations).22 How-
ever, our increasing numerical capabilities will allow us to simulate self-consistently
the coupling across scales of global and local processes in the near future. Even though
a full particle-in-cell model of the heliosphere with realistic properties may still lie
decades in the future, the ongoing improvement in our modeling capabilities will
advance our understanding of the multi-scale nature of the solar wind.

7.3 Broader impact

All magnetized plasmas exhibit a broad range of characteristic length scales and
timescales. These span from the largest scales of the system to its microscopic scales:
those of plasma oscillations, particle gyration, and electrostatic and electromagnetic
shielding. The vast system sizes of space and astrophysical plasmas lead to especially
large separations among these characteristic plasma scales. The solar wind exem-
plifies such a multi-scale astrophysical plasma, and the combination of solar-wind
observations with fundamental plasma physics has improved our understanding of
astrophysical plasma throughout the Universe. The solar wind’s expansion through
the heliosphere introduces additional global scales that couple to the small-scale
plasma processes. We anticipate that, in the coming years, the connection of small-
scale kinetic processes with the large-scale thermodynamics of astrophysical plasmas

22 Notable exceptions to this dichotomy in global and local scales include expanding-box models and
ad-hoc inclusions of kinetic processes through effective transport coefficients in global models.

123
The multi-scale nature of the solar wind Page 105 of 136 5

will be a major research focus not only in heliophysics but throughout the astrophysics
community.
The solar wind is the ideal place to study the multi-scale nature of astrophysical
plasmas. The conditions of space and astrophysical plasmas cannot be reproduced and
sampled with comparable accuracy in laboratories. With the notable exception of the
very local interstellar medium, the only astrophysical plasmas that have been observed
in situ are in the heliosphere.
Research into this topic serves a broader impact beyond the purely academic under-
standing of space and astrophysical plasmas. The study of the solar wind’s multi-scale
nature enables a better understanding of its dynamics and thermodynamics based on
first principles. This knowledge will be invaluable to the design of physics-based mod-
els for space weather and to guiding our efforts toward the successful prediction of
space hazards for our increasingly technological and spacefaring society.

Acknowledgements This work was supported by the STFC Ernest Rutherford Fellowship ST/P003826/1,
the STFC Consolidated Grant ST/S000240/1, as well as NASA grants NNX16AM23G and NNX16AG81G.
We acknowledge the use of data from Wind’s SWE instrument (PIs K. W. Ogilvie and A. F. Viñas), from
Wind’s MFI instrument (PI A. Szabo), and from Ulysses’ SWOOPS instrument (PI D. J. McComas).
Lynn Wilson is Wind’s Project Scientist. We extend our gratitude to Chadi Salem for providing processed
Wind/3DP (PI S. D. Bale) data to create Fig. 14. We also acknowledge the use of data from ESA’s Cluster
Science Archive (Laakso et al. 2010) from the EFW instrument (PI M. André), the FGM instrument (PI
C. Carr), the STAFF instrument (PI P. Canu), the CIS instrument (PI I. Dandouras), and the PEACE
instrument (PI A. Fazakerley). We acknowledge vigorous feedback on Fig. 8 from D. J. McComas at the
2019 SHINE Meeting. Data access was provided by the National Space Science Data Center (NSSDC) Space
Physics Data Facility (SPDF) and NASA/GSFC’s Space Physics Data Facility’s CDAWeb and COHOWeb
services. This review has made use of the SAO/NASA Astrophysics Data System (ADS).

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 Interna-
tional License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution,
and reproduction in any medium, provided you give appropriate credit to the original author(s) and the
source, provide a link to the Creative Commons license, and indicate if changes were made.

References
Abramowitz M, Stegun IA (1972) Handbook of mathematical functions. Dover, New York
Acuña MH (1974) Fluxgate magnetometers for outer planets exploration. IEEE Trans Magn 10(3):519–523.
https://doi.org/10.1109/TMAG.1974.1058457
Acuña MH (2002) Space-based magnetometers. Rev Sci Instrum 73(11):3717–3736. https://doi.org/10.
1063/1.1510570
Acuña MH, Curtis D, Scheifele JL, Russell CT, Schroeder P, Szabo A, Luhmann JG (2008) The
STEREO/IMPACT magnetic field experiment. Space Sci Rev 136:203–226. https://doi.org/10.1007/
s11214-007-9259-2
Aellig MR, Lazarus AJ, Kasper JC, Ogilvie KW (2001a) Rapid measurements of solar wind ions with
the Triana PlasMag faraday cup. Astrophys Space Sci 277:305–307. https://doi.org/10.1023/A:
1012229729242
Aellig MR, Lazarus AJ, Steinberg JT (2001b) The solar wind helium abundance: variation with wind speed
and the solar cycle. Geophys Res Lett 28:2767–2770. https://doi.org/10.1029/2000GL012771
Ahnert P (1943) Der Komet 1942 g (Whipple-Fedtke). Z Astrophys 22:288
Akimoto K, Gary SP, Omidi N (1987) Electron/ion whistler instabilities and magnetic noise bursts. J
Geophys Res 92:11209–11214. https://doi.org/10.1029/JA092iA10p11209

123
5 Page 106 of 136 D. Verscharen et al.

Alexandrova O, Saur J, Lacombe C, Mangeney A, Mitchell J, Schwartz SJ, Robert P (2009) Universality of
solar-wind turbulent spectrum from MHD to electron scales. Phys Rev Lett 103(16):165003. https://
doi.org/10.1103/PhysRevLett.103.165003. arXiv:0906.3236
Alfvén H (1942) Existence of electromagnetic–hydrodynamic waves. Nature 150:405–406. https://doi.org/
10.1038/150405d0
Alfvén H (1943) On the existence of electromagnetic–hydrodynamic waves. Arkv Mat Astron Fys 29B:1–7
Alterman BL, Kasper JC, Stevens ML, Koval A (2018) A comparison of alpha particle and proton beam
differential flows in collisionally young solar wind. Astrophys J 864:112. https://doi.org/10.3847/
1538-4357/aad23f. arXiv:1809.01693
Anderson BR, Skoug RM, Steinberg JT, McComas DJ (2012) Variability of the solar wind suprathermal
electron strahl. J Geophys Res 117:A04107. https://doi.org/10.1029/2011JA017269
Andre M (1985) Dispersion surfaces. J. Plasma Phys 33:1–19. https://doi.org/10.1017/S0022377800002270
Andrews GB, Zurbuchen TH, Mauk BH, Malcom H, Fisk LA, Gloeckler G, Ho GC, Kelley JS, Koehn PL,
Lefevere TW, Livi SS, Lundgren RA, Raines JM (2007) The energetic particle and plasma spectrometer
instrument on the MESSENGER spacecraft. Space Sci Rev 131:523–556. https://doi.org/10.1007/
s11214-007-9272-5
Angelopoulos V (2008) The THEMIS mission. Space Sci Rev 141:5–34. https://doi.org/10.1007/s11214-
008-9336-1
Araneda JA, Maneva Y, Marsch E (2009) Preferential heating and acceleration of α particles by Alfvén-
cyclotron waves. Phys Rev Lett 102(17):175001. https://doi.org/10.1103/PhysRevLett.102.175001
Arzamasskiy L, Kunz MW, Chandran BDG, Quataert E (2019) Hybrid-kinetic simulations of ion
heating in Alfvénic turbulence. Astrophys J 879(1):53. https://doi.org/10.3847/1538-4357/ab20cc.
arXiv:1901.11028
Asbridge JR, Bame SJ, Feldman WC (1974) Abundance differences in solar wind double streams. Sol Phys
37:451–467. https://doi.org/10.1007/BF00152503
Asbridge JR, Bame SJ, Feldman WC, Montgomery MD (1976) Helium and hydrogen velocity differences
in the solar wind. J Geophys Res 81:2719–2727. https://doi.org/10.1029/JA081i016p02719
Aschenbrenner H, Goubau G (1936) Eine Anordnung zur Registrierung rascher magnetischer Störungen.
Hochfrequenztech Elektroakust 47:177–181
Astfalk P, Jenko F (2016) Parallel and oblique firehose instability thresholds for bi-kappa distributed protons.
J Geophys Res 121:2842–2852. https://doi.org/10.1002/2015JA022267
Astfalk P, Jenko F (2017) LEOPARD: a grid-based dispersion relation solver for arbitrary gyrotropic dis-
tributions. J Geophys Res 122:89–101. https://doi.org/10.1002/2016JA023522
Astfalk P, Görler T, Jenko F (2015) DSHARK: a dispersion relation solver for obliquely propagating waves in
bi-kappa-distributed plasmas. J Geophys Res 120:7107–7120. https://doi.org/10.1002/2015JA021507
Bagenal F (2013) Planetary magnetospheres. In: Oswalt TD, French LM, Kalas P (eds) Planets, stars and
stellar systems. Springer, Dordrecht, p 251. https://doi.org/10.1007/978-94-007-5606-9_6
Bale SD, Kellogg PJ, Mozer FS, Horbury TS, Reme H (2005) Measurement of the electric fluctuation
spectrum of magnetohydrodynamic turbulence. Phys Rev Lett 94(21):215002. https://doi.org/10.1103/
PhysRevLett.94.215002
Bale SD, Ullrich R, Goetz K, Alster N, Cecconi B, Dekkali M, Lingner NR, Macher W, Manning RE,
McCauley J, Monson SJ, Oswald TH, Pulupa M (2008) The electric antennas for the STEREO/WAVES
experiment. Space Sci Rev 136:529. https://doi.org/10.1007/s11214-007-9251-x
Bale SD, Kasper JC, Howes GG, Quataert E, Salem C, Sundkvist D (2009) Magnetic fluctuation power near
proton temperature anisotropy instability thresholds in the solar wind. Phys Rev Lett 103(21):211101.
https://doi.org/10.1103/PhysRevLett.103.211101. arXiv:0908.1274
Bale SD, Pulupa M, Salem C, Chen CHK, Quataert E (2013) Electron heat conduction in the solar wind:
transition from Spitzer–Härm to the collisionless limit. Astrophys J Lett 769:L22. https://doi.org/10.
1088/2041-8205/769/2/L22. arXiv:1303.0932
Balogh A, Beek TJ, Forsyth RJ, Hedgecock PC, Marquedant RJ, Smith EJ, Southwood DJ, Tsurutani BT
(1992) The magnetic field investigation on the Ulysses mission—instrumentation and preliminary
scientific results. Astron Astrophys Suppl Ser 92(2):221–236
Balogh A, Gonzalez-Esparza JA, Forsyth RJ, Burton ME, Goldstein BE, Smith EJ, Bame SJ (1995)
Interplanetary shock waves: Ulysses observations in and out of the ecliptic plane. Space Sci Rev
72:171–180. https://doi.org/10.1007/BF00768774
Balogh A, Dunlop MW, Cowley SWH, Southwood DJ, Thomlinson JG, Glassmeier KH, Musmann G, Luhr
H, Buchert S, Acuna MH, Fairfield DH, Slavin JA, Riedler W, Schwingenschuh K, Kivelson MG

123
The multi-scale nature of the solar wind Page 107 of 136 5

(1997) The cluster magnetic field investigation. Space Sci Rev 79:65–91. https://doi.org/10.1023/A:
1004970907748
Bañón Navarro A, Morel P, Albrecht-Marc M, Carati D, Merz F, Görler T, Jenko F (2011) Free energy cascade
in gyrokinetic turbulence. Phys Rev Lett 106(5):055001. https://doi.org/10.1103/PhysRevLett.106.
055001. arXiv:1008.3974
Bame SJ, Asbridge JR, Feldman WC, Montgomery MD, Kearney PD (1975) Solar wind heavy ion abun-
dances. Sol Phys 43:463–473. https://doi.org/10.1007/BF00152368
Barakat AR, Schunk RW (1982) Transport equations for multicomponent anisotropic space plasmas—a
review. Plasma Phys 24:389–418. https://doi.org/10.1088/0032-1028/24/4/004
Barkhausen H (1919) Zwei mit Hilfe der neuen Verstärker entdeckter Erscheinungen. Phys Z 20:401–403
Barnes A (1966) Collisionless damping of hydromagnetic waves. Phys Fluids 9:1483–1495. https://doi.
org/10.1063/1.1761882
Barnes A (1970) Theory of generation of bow-shock-associated hydromagnetic waves in the upstream
interplanetary medium. Cosmic Electrodyn 1:90–114
Bavassano B (1996) The solar wind: a turbulent magnetohydrodynamic medium. Space Sci Rev 78:29–32.
https://doi.org/10.1007/BF00170789
Bavassano B, Smith EJ (1986) Radial variation of interplanetary Alfvénic fluctuations Pioneer 10 and
11 observations between 1 and 5 AU. J Geophys Res 91:1706–1710. https://doi.org/10.1029/
JA091iA02p01706
Bavassano B, Dobrowolny M, Mariani F, Ness NF (1982) Radial evolution of power spectra of interplanetary
Alfvénic turbulence. J Geophys Res 87:3617–3622. https://doi.org/10.1029/JA087iA05p03617
Bavassano B, Pietropaolo E, Bruno R (2004) Compressive fluctuations in high-latitude solar wind. Ann
Geophys 22:689–696. https://doi.org/10.5194/angeo-22-689-2004
Begelman MC, Chiueh T (1988) Thermal coupling of ions and electrons by collective effects in two-
temperature accretion flows. Astrophys J 332:872. https://doi.org/10.1086/166698
Behannon KW (1978) Heliocentric distance dependence of the interplanetary magnetic field. Rev Geophys
Space Phys 16:125–145. https://doi.org/10.1029/RG016i001p00125
Beinroth HJ, Neubauer FM (1981) Properties of whistler mode waves between 0.3 and 1.0 AU from Helios
observations. J Geophys Res 86:7755–7760. https://doi.org/10.1029/JA086iA09p07755
Behannon KW, Acuna MH, Burlaga LF, Lepping RP, Ness NF, Neubauer FM (1977) Magnetic field exper-
iment for Voyagers 1 and 2. Space Sci Rev 21(3):235–257. https://doi.org/10.1007/BF00211541
Belcher JW, Burchsted R (1974) Energy densities of Alfvén waves between 0.7 and 1.6 AU. J Geophys Res
79(31):4765. https://doi.org/10.1029/JA079i031p04765
Belcher JW, Davis L Jr (1971) Large-amplitude Alfvén waves in the interplanetary medium, 2. J Geophys
Res 76:3534. https://doi.org/10.1029/JA076i016p03534
Belcher JW, Davis L Jr, Smith EJ (1969) Large-amplitude Alfvén waves in the interplanetary medium:
mariner 5. J Geophys Res 74:2302. https://doi.org/10.1029/JA074i009p02302
Berger L, Wimmer-Schweingruber RF, Gloeckler G (2011) Systematic measurements of ion-proton
differential streaming in the solar wind. Phys Rev Lett 106(15):151103. https://doi.org/10.1103/
PhysRevLett.106.151103
Bhatnagar PL, Gross EP, Krook M (1954) A model for collision processes in gases. I. Small amplitude
processes in charged and neutral one-component systems. Phys Rev 94(3):511–525. https://doi.org/
10.1103/PhysRev.94.511
Bieber JW, Wanner W, Matthaeus WH (1996) Dominant two-dimensional solar wind turbulence with impli-
cations for cosmic ray transport. J Geophys Res 101:2511–2522. https://doi.org/10.1029/95JA02588
Biermann L (1951) Kometenschweife und solare Korpuskularstrahlung. Z Astrophys 29:274
Birkeland K (1914) A possible connection between magnetic and meteorologic phenomena. Mon Weather
Rev 42:211
Boardsen SA, Jian LK, Raines JL, Gershman DJ, Zurbuchen TH, Roberts DA, Korth H (2015) MESSENGER
survey of in situ low frequency wave storms between 0.3 and 0.7 AU. J Geophys Res 120(A9):10.
https://doi.org/10.1002/2015JA021506
Bochsler P (2000) Abundances and charge states of particles in the solar wind. Rev Geophys 38:247–266.
https://doi.org/10.1029/1999RG000063
Bochsler P (2007) Minor ions in the solar wind. Astron Astrophys Rev 14:1–40. https://doi.org/10.1007/
s00159-006-0002-x
Bochsler P, Geis J, Joos R (1985) Kinetic temperatures of heavy ions in the solar wind. J Geophys Res
90:10. https://doi.org/10.1029/JA090iA11p10779

123
5 Page 108 of 136 D. Verscharen et al.

Bogoliubov NN (1946) Kinetic equations. J Phys USSR 10:265–274


Boldyrev S (2005) On the spectrum of magnetohydrodynamic turbulence. Astrophys J 626:L37–L40. https://
doi.org/10.1086/431649. arXiv:astro-ph/0503053
Boldyrev S (2006) Spectrum of magnetohydrodynamic turbulence. Phys Rev Lett 96:115002. https://doi.
org/10.1103/PhysRevLett.96.115002. arXiv:astro-ph/0511290
Boldyrev S, Loureiro NF (2017) Magnetohydrodynamic turbulence mediated by reconnection. Astrophys
J 844:125. https://doi.org/10.3847/1538-4357/aa7d02. arXiv:1706.07139
Boldyrev S, Perez JC (2012) Spectrum of kinetic-Alfvén turbulence. Astrophys J Lett 758:L44. https://doi.
org/10.1088/2041-8205/758/2/L44. arXiv:1204.5809
Boldyrev S, Perez JC, Borovsky JE, Podesta JJ (2011) Spectral scaling laws in magnetohydrodynamic
turbulence simulations and in the solar wind. Astrophys J Lett 741(1):L19. https://doi.org/10.1088/
2041-8205/741/1/L19. arXiv:1106.0700
Bougeret JL, Kaiser ML, Kellogg PJ, Manning R, Goetz K, Monson SJ, Monge N, Friel L, Meetre CA,
Perche C, Sitruk L, Hoang S (1995) Waves: the radio and plasma wave investigation on the Wind
spacecraft. Space Sci Rev 71:231–263. https://doi.org/10.1007/BF00751331
Bougeret JL, Goetz K, Kaiser ML, Bale SD, Kellogg PJ, Maksimovic M, Monge N, Monson SJ, Astier
PL, Davy S, Dekkali M, Hinze JJ, Manning RE, Aguilar-Rodriguez E, Bonnin X, Briand C, Cairns
IH, Cattell CA, Cecconi B, Eastwood J, Ergun RE, Fainberg J, Hoang S, Huttunen KEJ, Krucker S,
Lecacheux A, MacDowall RJ, Macher W, Mangeney A, Meetre CA, Moussas X, Nguyen QN, Oswald
TH, Pulupa M, Reiner MJ, Robinson PA, Rucker H, Salem C, Santolik O, Silvis JM, Ullrich R, Zarka
P, Zouganelis I (2008) S/WAVES: the radio and plasma wave investigation on the STEREO mission.
Space Sci Rev 136:487–528. https://doi.org/10.1007/s11214-007-9298-8
Bourouaine S, Chandran BDG (2013) Observational test of stochastic heating in low-β fast-solar-wind
streams. Astrophys J 774:96. https://doi.org/10.1088/0004-637X/774/2/96. arXiv:1307.3789
Bourouaine S, Perez JC (2018) On the limitations of Taylor’s hypothesis in Parker Solar Probe’s measure-
ments near the Alfvén critical point. Astrophys J Lett 858(2):L20. https://doi.org/10.3847/2041-8213/
aabccf
Bourouaine S, Marsch E, Neubauer FM (2011) On the relative speed and temperature ratio of solar wind
alpha particles and protons: collisions versus wave effects. Astrophys J Lett 728:L3. https://doi.org/
10.1088/2041-8205/728/1/L3
Bourouaine S, Verscharen D, Chandran BDG, Maruca BA, Kasper JC (2013) Limits on alpha particle tem-
perature anisotropy and differential flow from kinetic instabilities: solar wind observations. Astrophys
J Lett 777:L3. https://doi.org/10.1088/2041-8205/777/1/L3. arXiv:1309.4010
Bowen TA, Badman S, Hellinger P, Bale SD (2018) Density fluctuations in the solar wind driven by
Alfvén wave parametric decay. Astrophys J Lett 854:L33. https://doi.org/10.3847/2041-8213/aaabbe.
arXiv:1712.09336
Braginskii SI (1965) Transport processes in a plasma. Rev Plasma Phys 1:205–311
Brambilla M (1998) Kinetic theory of plasma waves, homogeneous plasmas. Oxford University Press,
Oxford
Bridge HS, Dilworth C, Rossi B, Scherb F, Lyon EF (1960) An instrument for the investigation of inter-
planetary plasma. J Geophys Res 65(10):3053–3055. https://doi.org/10.1029/JZ065i010p03053
Bridge HS, Belcher JW, Butler RJ, Lazarus AJ, Mavretic AM, Sullivan JD, Siscoe GL, Vasyliunas VM
(1977) The plasma experiment on the 1977 Voyager mission. Space Sci Rev 21:259–287. https://doi.
org/10.1007/BF00211542
Bruno R, Bavassano B (1997) On the winding of the IMF spiral for slow and fast wind within the inner
heliosphere. Geophys Res Lett 24:2267. https://doi.org/10.1029/97GL02183
Bruno R, Carbone V (2013) The solar wind as a turbulence laboratory. Living Rev Sol Phys 10:2. https://
doi.org/10.12942/lrsp-2013-2
Bruno R, Trenchi L (2014) Radial dependence of the frequency break between fluid and kinetic scales in
the solar wind fluctuations. Astrophys J Lett 787:L24. https://doi.org/10.1088/2041-8205/787/2/L24
Bruno R, Bavassano B, Villante U (1985) Evidence for long period Alfvén waves in the inner solar system.
J Geophys Res 90:4373–4377. https://doi.org/10.1029/JA090iA05p04373
Bruno R, Carbone V, Sorriso-Valvo L, Bavassano B (2003) Radial evolution of solar wind intermit-
tency in the inner heliosphere. J Geophys Res 108:1130. https://doi.org/10.1029/2002JA009615.
arXiv:astro-ph/0303578
Bruno R, Carbone V, Vörös Z, D’Amicis R, Bavassano B, Cattaneo MB, Mura A, Milillo A, Orsini S, Veltri
P, Sorriso-Valvo L, Zhang T, Biernat H, Rucker H, Baumjohann W, Jankovičová D, Kovács P (2009)

123
The multi-scale nature of the solar wind Page 109 of 136 5

Coordinated study on solar wind turbulence during the venus-express, ACE and Ulysses alignment of
August 2007. Earth Moon Planets 104:101–104. https://doi.org/10.1007/s11038-008-9272-9
Bruno R, Telloni D, DeIure D, Pietropaolo E (2017) Solar wind magnetic field background spectrum from
fluid to kinetic scales. Mon Not R Astron Soc 472:1052–1059. https://doi.org/10.1093/mnras/stx2008
Bruno R, Telloni D, Sorriso-Valvo L, Marino R, De Marco R, D’Amicis R (2019) The low-frequency break
observed in the slow solar wind magnetic spectra. Astron Astrophys 627:A96. https://doi.org/10.1051/
0004-6361/201935841. arXiv:1906.11767
Buneman O (1959) Dissipation of currents in ionized media. Phys Rev 115:503–517. https://doi.org/10.
1103/PhysRev.115.503
Burch JL, Moore TE, Torbert RB, Giles BL (2016) Magnetospheric multiscale overview and science objec-
tives. Space Sci Rev 199:5–21. https://doi.org/10.1007/s11214-015-0164-9
Burgers JM (1969) Flow equations for composite gases. Academic Press, New York
Burlaga LF, Goldstein ML (1984) Radial variations of large-scale magnetohydrodynamic fluctuations in
the solar wind. J Geophys Res 89:6813–6817. https://doi.org/10.1029/JA089iA08p06813
Burlaga LF, Ogilvie KW (1970) Magnetic and thermal pressures in the solar wind. Sol Phys 15:61–71.
https://doi.org/10.1007/BF00149472
Burlaga LF, Ness NF, Acuña MH, Lepping RP, Connerney JEP, Richardson JD (2008) Magnetic fields at
the solar wind termination shock. Nature 454:75–77. https://doi.org/10.1038/nature07029
Camporeale E, Carè A, Borovsky JE (2017) Classification of solar wind with machine learning. J Geophys
Res 122:10910–10920. https://doi.org/10.1002/2017JA024383. arXiv:1710.02313
Carlson CW, Curtis DW, Paschmann G, Michel W (1983) An instrument for rapidly measuring plasma
distribution functions with high resolution. Adv Space Res 2(7):67–70. https://doi.org/10.1016/0273-
1177(82)90151-X
Cerri SS, Califano F (2017) Reconnection and small-scale fields in 2D–3V hybrid-kinetic driven turbulence
simulations. New J Phys 19(2):025007. https://doi.org/10.1088/1367-2630/aa5c4a
Cerri SS, Kunz MW, Califano F (2018) Dual phase-space cascades in 3D hybrid-Vlasov–Maxwell turbu-
lence. Astrophys J Lett 856:L13. https://doi.org/10.3847/2041-8213/aab557. arXiv:1802.06133
Chamberlain JW (1961) Interplanetary gas. III. A hydrodynamic model of the corona. Astrophys J 133:675.
https://doi.org/10.1086/147070
Chandran BDG (2010) Alfvén-wave turbulence and perpendicular ion temperatures in coronal holes. Astro-
phys J 720:548–554. https://doi.org/10.1088/0004-637X/720/1/548
Chandran BDG (2018) Parametric instability, inverse cascade and the range of solar-wind turbulence. J
Plasma Phys 84(1):905840106. https://doi.org/10.1017/S0022377818000016. arXiv:1712.09357
Chandran BDG, Quataert E, Howes GG, Xia Q, Pongkitiwanichakul P (2009) Constraining low-frequency
Alfvénic turbulence in the solar wind using density-fluctuation measurements. Astrophys J 707:1668–
1675. https://doi.org/10.1088/0004-637X/707/2/1668. arXiv:0908.0757
Chandran BDG, Li B, Rogers BN, Quataert E, Germaschewski K (2010) Perpendicular ion heating by
low-frequency Alfvén-wave turbulence in the solar wind. Astrophys J 720:503–515. https://doi.org/
10.1088/0004-637X/720/1/503. arXiv:1001.2069
Chandran BDG, Dennis TJ, Quataert E, Bale SD (2011) Incorporating kinetic physics into a two-fluid
solar-wind model with temperature anisotropy and low-frequency Alfvén-wave turbulence. Astrophys
J 743:197. https://doi.org/10.1088/0004-637X/743/2/197. arXiv:1110.3029
Chandran BDG, Verscharen D, Quataert E, Kasper JC, Isenberg PA, Bourouaine S (2013) Stochastic heating,
differential flow, and the alpha-to-proton temperature ratio in the solar wind. Astrophys J 776:45.
https://doi.org/10.1088/0004-637X/776/1/45
Chang O, Peter Gary S, Wang J (2014) Energy dissipation by whistler turbulence: three-dimensional particle-
in-cell simulations. Phys Plasmas 21(5):052305. https://doi.org/10.1063/1.4875728
Chapman S (1917) On the times of sudden commencement of magnetic storms. Proc Phys Soc London
30:205–214. https://doi.org/10.1088/1478-7814/30/1/317
Chashei IV, Shishov VI (1997) Shock waves propagation in the turbulent interplanetary plasma. Adv Space
Res 20:75–78. https://doi.org/10.1016/S0273-1177(97)00484-5
Chaston CC, Bonnell JW, Carlson CW, McFadden JP, Ergun RE, Strangeway RJ, Lund EJ (2004) Auroral
ion acceleration in dispersive Alfvén waves. J Geophys Res 109:A04205. https://doi.org/10.1029/
2003JA010053
Chen CHK (2016) Recent progress in astrophysical plasma turbulence from solar wind observations. J
Plasma Phys 82(6):535820602. https://doi.org/10.1017/S0022377816001124. arXiv:1611.03386

123
5 Page 110 of 136 D. Verscharen et al.

Chen CHK, Mallet A, Yousef TA, Schekochihin AA, Horbury TS (2011) Anisotropy of Alfvénic turbulence
in the solar wind and numerical simulations. Mon Not R Astron Soc 415:3219–3226. https://doi.org/
10.1111/j.1365-2966.2011.18933.x. arXiv:1009.0662
Chen L, Lin Z, White R (2001b) On resonant heating below the cyclotron frequency. Phys Plasmas 8:4713–
4716. https://doi.org/10.1063/1.1406939
Chen CHK, Mallet A, Schekochihin AA, Horbury TS, Wicks RT, Bale SD (2012a) Three-dimensional
structure of solar wind turbulence. Astrophys J 758:120. https://doi.org/10.1088/0004-637X/758/2/
120. arXiv:1109.2558
Chen CHK, Salem CS, Bonnell JW, Mozer FS, Bale SD (2012b) Density fluctuation spectrum of solar wind
turbulence between ion and electron scales. Phys Rev Lett 109(3):035001. https://doi.org/10.1103/
PhysRevLett.109.035001. arXiv:1205.5063
Chen CHK, Boldyrev S, Xia Q, Perez JC (2013) Nature of subproton scale turbulence in the solar wind.
Phys Rev Lett 110(22):225002. https://doi.org/10.1103/PhysRevLett.110.225002
Chen CHK, Matteini L, Schekochihin AA, Stevens ML, Salem CS, Maruca BA, Kunz MW, Bale SD
(2016) Multi-species measurements of the firehose and mirror instability thresholds in the solar wind.
Astrophys J Lett 825:L26. https://doi.org/10.3847/2041-8205/825/2/L26. arXiv:1606.02624
Chew GF, Goldberger ML, Low FE (1956) The Boltzmann equation and the one-fluid hydromagnetic
equations in the absence of particle collisions. Proc R Soc Lonon A 236:112–118. https://doi.org/10.
1098/rspa.1956.0116
Chhiber R, Usmanov AV, Matthaeus WH, Goldstein ML (2016) Solar wind collisional age from a global
magnetohydrodynamics simulation. Astrophys J 821:34. https://doi.org/10.3847/0004-637X/821/1/
34
Chiuderi C, Velli M (2015) Basics of plasma astrophysics. Springer, Milano. https://doi.org/10.1007/978-
88-470-5280-2
Cohen RS, Spitzer L Jr, Routly PM (1950) The electrical conductivity of an ionized gas. Phys Rev 80(2):230–
238. https://doi.org/10.1103/PhysRev.80.230
Cohen CMS, Collier MR, Hamilton DC, Gloeckler G, Sheldon RB, von Steiger R, Wilken B (1996) Kinetic
temperature ratios of O6+ and He2+ : observations from Wind/MASS and Ulysses/SWICS. Geophys
Res Lett 23:1187–1190. https://doi.org/10.1029/96GL00587
Colegrove FD, Franken PA (1960) Optical pumping of helium in the 3 S1 metastable state. Phys Rev
119(2):680–690. https://doi.org/10.1103/PhysRev.119.680
Coleman PJ Jr (1968) Turbulence, viscosity, and dissipation in the solar-wind plasma. Astrophys J 153:371.
https://doi.org/10.1086/149674
Consolini G, De Marco R, Carbone V (2015) On the emergence of a 1/k spectrum in the sub-inertial domains
of turbulent media. Astrophys J 809(1):21. https://doi.org/10.1088/0004-637X/809/1/21
Cornilleau-Wehrlin N, Chauveau P, Louis S, Meyer A, Nappa JM, Perraut S, Rezeau L, Robert P, Roux
A, de Villedary C, de Conchy Y, Friel L, Harvey CC, Hubert D, Lacombe C, Manning R, Wouters F,
Lefeuvre F, Parrot M, Pincon JL, Poirier B, Kofman W, Louarn P (1997) The Cluster Spatio-Temporal
Analysis of Field Fluctuations (STAFF) experiment. Space Sci Rev 79:107–136. https://doi.org/10.
1023/A:1004979209565
Coroniti FV, Kennel CF, Scarf FL, Smith EJ (1982) Whistler mode turbulence in the disturbed solar wind.
J Geophys Res 87:6029–6044. https://doi.org/10.1029/JA087iA08p06029
Couturier P, Hoang S, Meyer-Vernet N, Steinberg JL (1981) Quasi-thermal noise in a stable plasma at
rest: theory and observations from ISEE 3. J Geophys Res 86:11127–11138. https://doi.org/10.1029/
JA086iA13p11127
Cranmer SR (2001) Ion cyclotron diffusion of velocity distributions in the extended solar corona. J Geophys
Res 106:24937–24954. https://doi.org/10.1029/2001JA000012
Cranmer SR (2014) Ensemble simulations of proton heating in the solar wind via turbulence and ion
cyclotron resonance. Astrophys J Supp 213:16. https://doi.org/10.1088/0067-0049/213/1/16
Cranmer SR, van Ballegooijen AA (2005) On the generation, propagation, and reflection of Alfvén waves
from the solar photosphere to the distant heliosphere. Astrophys J Supp 156:265–293. https://doi.org/
10.1086/426507. arXiv:astro-ph/0410639
Cranmer SR, Field GB, Kohl JL (1999) Spectroscopic constraints on models of ion cyclotron resonance
heating in the polar solar corona and high-speed solar wind. Astrophys J 518:937–947. https://doi.
org/10.1086/307330

123
The multi-scale nature of the solar wind Page 111 of 136 5

Cranmer SR, Matthaeus WH, Breech BA, Kasper JC (2009) Empirical constraints on proton and electron
heating in the fast solar wind. Astrophys J 702:1604–1614. https://doi.org/10.1088/0004-637X/702/
2/1604. arXiv:0907.2650
D’Amicis R, Bruno R (2015) On the origin of highly Alfvénic slow solar wind. Astrophys J 805:84. https://
doi.org/10.1088/0004-637X/805/1/84
D’Amicis R, Matteini L, Bruno R (2019) On the slow solar wind with high Alfvénicity: from composition
and microphysics to spectral properties. Mon Not R Astron Soc 483(4):4665–4677. https://doi.org/
10.1093/mnras/sty3329. arXiv:1812.01899
Dasso S, Milano LJ, Matthaeus WH, Smith CW (2005) Anisotropy in fast and slow solar wind fluctuations.
Astrophys J Lett 635:L181–L184. https://doi.org/10.1086/499559
Davidson RC, Ogden JM (1975) Electromagnetic ion cyclotron instability driven by ion energy anisotropy
in high-beta plasmas. Phys Fluids 18:1045–1050. https://doi.org/10.1063/1.861253
Dorfman S, Carter TA (2016) Observation of an Alfvén wave parametric instability in a laboratory plasma.
Phys Rev Lett 116:195002. https://doi.org/10.1103/PhysRevLett.116.195002
Dorland W, Hammett GW (1993) Gyrofluid turbulence models with kinetic effects. Phys Fluids B 5:812–
835. https://doi.org/10.1063/1.860934
Dougherty JP (1964) Model Fokker–Planck equation for a plasma and its solution. Phys Fluids 7(11):1788–
1799. https://doi.org/10.1063/1.2746779
Dougherty MK, Kellock S, Southwood DJ, Balogh A, Smith EJ, Tsurutani BT, Gerlach B, Glassmeier
KH, Gleim F, Russell CT, Erdos G, Neubauer FM, Cowley SWH (2004) The Cassini magnetic field
investigation. Space Sci Rev 114:331–383. https://doi.org/10.1007/s11214-004-1432-2
Dum CT, Marsch E, Pilipp W (1980) Determination of wave growth from measured distribution functions
and transport theory. J Plasma Phys 23:91–113. https://doi.org/10.1017/S0022377800022170
Dunlop MW, Woodward TI (2000) Multi-spacecraft discontinuity analysis: orientation and motion. In:
Paschmann G, Daly PW (eds) Analysis methods for multi-spacecraft data, 1st edn, no. SR-001 in ISSI
Scientific Report, International Space Science Institute (ISSI), Bern, Chap 11, pp 271–306. http://
www.issibern.ch/forads/sr-001-11.pdf
Dunlop MW, Southwood DJ, Glassmeier KH, Neubauer FM (1988) Analysis of multipoint magnetometer
data. Adv Space Res 8:273–277. https://doi.org/10.1016/0273-1177(88)90141-X
Dunlop MW, Dougherty MK, Kellock S, Southwood DJ (1999) Operation of the dual magnetometer
on Cassini: science performance. Planet Space Sci 47:1389–1405. https://doi.org/10.1016/S0032-
0633(99)00060-4
Dupree TH (1961) Dynamics of ionized gases. Phys Fluids 4:696–702. https://doi.org/10.1063/1.1706386
Durovcová T, Šafránková J, Němeček Z, Richardson JD (2017) Evolution of proton and alpha particle
velocities through the solar cycle. Astrophys J 850:164. https://doi.org/10.3847/1538-4357/aa9618
Dusenbery PB, Hollweg JV (1981) Ion-cyclotron heating and acceleration of solar wind minor ions. J
Geophys Res 86:153–164. https://doi.org/10.1029/JA086iA01p00153
Ebert RW, McComas DJ, Elliott HA, Forsyth RJ, Gosling JT (2009) Bulk properties of the slow and
fast solar wind and interplanetary coronal mass ejections measured by Ulysses: three polar orbits of
observations. J Geophys Res 114:A01109. https://doi.org/10.1029/2008JA013631
Echim MM, Lemaire J, Lie-Svendsen Ø (2011) A review on solar wind modeling: kinetic and fluid aspects.
Surv Geophys 32:1–70. https://doi.org/10.1007/s10712-010-9106-y
Eddington AS (1910) c 1908 (Morehouse), the envelopes of. Mon Not R Astron Soc 70:442–458. https://
doi.org/10.1093/mnras/70.5.442
Edlén B (1943) Die Deutung der Emissionslinien im Spektrum der Sonnenkorona. Z Astrophys 22:30
Elsasser WM (1950) The hydromagnetic equations. Phys Rev 79:183–183. https://doi.org/10.1103/
PhysRev.79.183
Escoubet CP, Schmidt R, Goldstein ML (1997) Cluster-science and mission overview. Space Sci Rev
79:11–32. https://doi.org/10.1023/A:1004923124586
Fairfield DH, Scudder JD (1985) Polar rain-solar coronal electrons in the Earth’s magnetosphere. J Geophys
Res 90:4055–4068. https://doi.org/10.1029/JA090iA05p04055
Farrugia CJ, Popecki M, Möbius E, Jordanova VK, Desai MI, Fitzenreiter RJ, Ogilvie KW, Matsui H,
Lepri S, Zurbuchen T, Mason GM, Lawrence GR, Burlaga LF, Lepping RP, Dwyer JR, McComas D
(2002) Wind and ACE observations during the great flow of 1–4 May 1998: relation to solar activity and
implications for the magnetosphere. J Geophys Res 107:1240. https://doi.org/10.1029/2001JA000188
Feldman WC, Asbridge JR, Bame SJ (1974a) The solar wind He2+ to H+ temperature ratio. J Geophys
Res 79:2319. https://doi.org/10.1029/JA079i016p02319

123
5 Page 112 of 136 D. Verscharen et al.

Feldman WC, Asbridge JR, Bame SJ, Montgomery MD (1974b) Interpenetrating solar wind streams. Rev
Geophys Space Phys 12:715–723. https://doi.org/10.1029/RG012i004p00715
Feldman WC, Asbridge JR, Bame SJ, Montgomery MD, Gary SP (1975) Solar wind electrons. J Geophys
Res 80:4181–4196. https://doi.org/10.1029/JA080i031p04181
Feldman WC, Asbridge JR, Bame SJ, Gary SP, Montgomery MD, Zink SM (1976) Evidence for the
regulation of solar wind heat flux at 1 AU. J Geophys Res 81:5207–5211. https://doi.org/10.1029/
JA081i028p05207
Feldman WC, Asbridge JR, Bame SJ, Gosling JT, Lemons DS (1979) The core electron temperature profile
between 0.5 and 1.0 AU in the steady-state high speed solar wind. J Geophys Res 84:4463–4467.
https://doi.org/10.1029/JA084iA08p04463
Fiksel G, Almagri AF, Chapman BE, Mirnov VV, Ren Y, Sarff JS, Terry PW (2009) Mass-dependent ion
heating during magnetic reconnection in a laboratory plasma. Phys Rev Lett 103(14):145002. https://
doi.org/10.1103/PhysRevLett.103.145002
Fitzenreiter RJ, Ogilvie KW (1992) Heat flux dropouts in the solar wind and Coulomb scattering effects. J
Geophys Res 97(A12):19213–19219. https://doi.org/10.1029/92JA00432
Fitzenreiter RJ, Ogilvie KW, Chornay DJ, Keller J (1998) Observations of electron velocity distribution
functions in the solar wind by the WIND spacecraft: high angular resolution strahl measurements.
Geophys Res Lett 25:249–252. https://doi.org/10.1029/97GL03703
Fitzpatrick R (2015) Plasma physics: an introduction. CRC Press, Boca Raton
Fokker AD (1914) Die mittlere Energie rotierender elektrischer Dipole im Strahlungsfeld. Ann Phys
348(5):810–820. https://doi.org/10.1002/andp.19143480507
Formisano V, Palmiotto F, Moreno G (1970) α-Particle observations in the solar wind. Sol Phys 15:479–498.
https://doi.org/10.1007/BF00151853
Forsyth RJ, Balogh A, Smith EJ (2002) The underlying direction of the heliospheric magnetic field through
the Ulysses first orbit. J Geophys Res 107:1405. https://doi.org/10.1029/2001JA005056
Fox NJ, Velli MC, Bale SD, Decker R, Driesman A, Howard RA, Kasper JC, Kinnison J, Kusterer M,
Lario D, Lockwood MK, McComas DJ, Raouafi NE, Szabo A (2016) The Solar Probe Plus mission:
humanity’s first visit to our star. Space Sci Rev 204:7–48. https://doi.org/10.1007/s11214-015-0211-
6
Frandsen AMA, Connor BV, van Amersfoort J, Smith EJ (1978) The ISEE-C vector helium magnetometer.
IEEE Trans Geosci Electron 16(3):195–198. https://doi.org/10.1109/TGE.1978.294545
Frisch U (1995) Turbulence. Cambridge University Press, Cambridge
Galeev AA, Oraevskii VN (1963) The stability of Alfvén waves. Sov Phys Dokl 7:988
Galvin AB, Kistler LM, Popecki MA, Farrugia CJ, Simunac KDC, Ellis L, Möbius E, Lee MA, Boehm M,
Carroll J, Crawshaw A, Conti M, Demaine P, Ellis S, Gaidos JA, Googins J, Granoff M, Gustafson A,
Heirtzler D, King B, Knauss U, Levasseur J, Longworth S, Singer K, Turco S, Vachon P, Vosbury M,
Widholm M, Blush LM, Karrer R, Bochsler P, Daoudi H, Etter A, Fischer J, Jost J, Opitz A, Sigrist M,
Wurz P, Klecker B, Ertl M, Seidenschwang E, Wimmer-Schweingruber RF, Koeten M, Thompson B,
Steinfeld D (2008) The Plasma and Suprathermal Ion Composition (PLASTIC) investigation on the
STEREO observatories. Space Sci Rev 136:437–486. https://doi.org/10.1007/s11214-007-9296-x
Gardini A, Laurenza M, Storini M (2011) SEP events and multi-spacecraft observations: constraints on
theory. Adv Space Res 47:2127–2139. https://doi.org/10.1016/j.asr.2011.01.025
Gardner CS (1963) Bound on the energy available from a plasma. Phys Fluids 6:839–840. https://doi.org/
10.1063/1.1706823
Gary SP (1993) Theory of space plasma microinstabilities. Cambridge University Press, Cambridge
Gary SP, Lee MA (1994) The ion cyclotron anisotropy instability and the inverse correlation between proton
anisotropy and proton beta. J Geophys Res 99:11297–11302. https://doi.org/10.1029/94JA00253
Gary SP, Li H (2000) Whistler heat flux instability at high beta. Astrophys J 529:1131–1135. https://doi.
org/10.1086/308294
Gary SP, Madland CD (1985) Electromagnetic electron temperature anisotropy instabilities. J Geophys Res
90:7607–7610. https://doi.org/10.1029/JA090iA08p07607
Gary SP, Saito S (2007) Broadening of solar wind strahl pitch-angles by the electron/electron instability:
particle-in-cell simulations. Geophys Res Lett 34:L14111. https://doi.org/10.1029/2007GL030039
Gary SP, Feldman WC, Forslund DW, Montgomery MD (1975) Heat flux instabilities in the solar wind. J
Geophys Res 80:4197. https://doi.org/10.1029/JA080i031p04197

123
The multi-scale nature of the solar wind Page 113 of 136 5

Gary SP, Anderson BJ, Denton RE, Fuselier SA, McKean ME (1994a) A limited closure relation for
anisotropic plasmas from the Earth’s magnetosheath. Phys Plasmas 1:1676–1683. https://doi.org/10.
1063/1.870670
Gary SP, McKean ME, Winske D, Anderson BJ, Denton RE, Fuselier SA (1994b) The proton cyclotron
instability and the anisotropy/β inverse correlation. J Geophys Res 99:5903–5914. https://doi.org/10.
1029/93JA03583
Gary SP, Scime EE, Phillips JL, Feldman WC (1994c) The whistler heat flux instability: threshold conditions
in the solar wind. J Geophys Res 99:23391–23400. https://doi.org/10.1029/94JA02067
Gary SP, Li H, O’Rourke S, Winske D (1998) Proton resonant firehose instability: temperature anisotropy and
fluctuating field constraints. J Geophys Res 103:14567–14574. https://doi.org/10.1029/98JA01174
Gary SP, Skoug RM, Daughton W (1999) Electron heat flux constraints in the solar wind. Phys Plasmas
6:2607–2612. https://doi.org/10.1063/1.873532
Gary SP, Skoug RM, Steinberg JT, Smith CW (2001) Proton temperature anisotropy constraint in the solar
wind: ACE observations. Geophys Res Lett 28:2759–2762. https://doi.org/10.1029/2001GL013165
Gary SP, Smith CW, Skoug RM (2005) Signatures of Alfvén-cyclotron wave-ion scattering: Advanced
Composition Explorer (ACE) solar wind observations. J Geophys Res 110:A07108. https://doi.org/
10.1029/2004JA010569
Gary SP, Hughes RS, Wang J (2016a) Whistler turbulence heating of electrons and ions: three-dimensional
particle-in-cell simulations. Astrophys J 816:102. https://doi.org/10.3847/0004-637X/816/2/102
Gary SP, Jian LK, Broiles TW, Stevens ML, Podesta JJ, Kasper JC (2016b) Ion-driven instabilities in the
solar wind: wind observations of 19 March 2005. J Geophys Res 121:30–41. https://doi.org/10.1002/
2015JA021935
Gazis PR, Lazarus AJ (1982) Voyager observations of solar wind proton temperature: 1–10 AU. Geophys
Res Lett 9:431–434. https://doi.org/10.1029/GL009i004p00431
Gazis PR, Barnes A, Mihalov JD, Lazarus AJ (1994) Solar wind velocity and temperature in the outer
heliosphere. J Geophys Res 99:6561–6573. https://doi.org/10.1029/93JA03144
Geiger H, Marsden E (1913) The laws of deflexion of α particles through large angles. Philos Mag
25(148):604–623. https://doi.org/10.1080/14786440408634197
Gershman DJ, Zurbuchen TH, Fisk LA, Gilbert JA, Raines JM, Anderson BJ, Smith CW, Korth H, Solomon
SC (2012) Solar wind alpha particles and heavy ions in the inner heliosphere observed with MES-
SENGER. J Geophys Res 117:A00M02. https://doi.org/10.1029/2012JA017829
Geyger WA (1962) The ring-core magnetometer—a new type of second-harmonic flux-gate magnetometer.
Trans Am Inst Electr Eng Part I Commun Electron 81(1):65–73. https://doi.org/10.1109/TCE.1962.
6373206
Giacalone J, Drake JF, Jokipii JR (2012) The acceleration mechanism of anomalous cosmic rays. Space Sci
Rev 173:283–307. https://doi.org/10.1007/s11214-012-9915-z
Gloeckler G (1990) Ion composition measurement techniques for space plasmas. Rev Sci Instrum 61:3613–
3620. https://doi.org/10.1063/1.1141581
Gloeckler G, Geiss J (1989) The abundances of elements and isotopes in the solar wind. In: Waddington
CJ (ed) Cosmic abundances of matter. ASP conference series, vol 183. American Institute of Physics,
New York, pp 49–71. https://doi.org/10.1063/1.37985
Gloeckler G, Geiss J (1998) Interstellar and inner source pickup ions observed with SWICS on Ulysses.
Space Sci Rev 86:127–159. https://doi.org/10.1023/A:1005019628054
Gloeckler G, Geiss J, Balsiger H, Bedini P, Cain JC, Fischer J, Fisk LA, Galvin AB, Gliem F, Hamilton DC,
Hollweg JV, Ipavich FM, Joos R, Livi S, Lundgren RA, Mall U, McKenzie JF, Ogilvie KW, Ottens
F, Rieck W, Tums EO, von Steiger R, Weiss W, Wilken B (1992) The Solar Wind Ion Composition
Spectrometer. Astron Astrophys Suppl Ser 92(2):267–289
Gloeckler G, Cain J, Ipavich FM, Tums EO, Bedini P, Fisk LA, Zurbuchen TH, Bochsler P, Fischer J,
Wimmer-Schweingruber RF, Geiss J, Kallenbach R (1998) Investigation of the composition of solar
and interstellar matter using solar wind and pickup ion measurements with SWICS and SWIMS on
the ACE spacecraft. Space Sci Rev 86:497–539. https://doi.org/10.1023/A:1005036131689
Goldreich P, Sridhar S (1995) Toward a theory of interstellar turbulence. 2: Strong Alfvénic turbulence.
Astrophys J 438:763–775. https://doi.org/10.1086/175121
Goldstein ML (1978) An instability of finite amplitude circularly polarized Alfvén waves. Astrophys J
219:700–704. https://doi.org/10.1086/155829
Goldstein BE, Neugebauer M, Smith EJ (1995) Alfvén waves, alpha particles, and pickup ions in the solar
wind. Geophys Res Lett 22:3389–3392. https://doi.org/10.1029/95GL03182

123
5 Page 114 of 136 D. Verscharen et al.

Goldstein BE, Neugebauer M, Zhang LD, Gary SP (2000) Observed constraint on proton–proton relative
velocities in the solar wind. Geophys Res Lett 27:53–56. https://doi.org/10.1029/1999GL003637
Gomberoff L, Elgueta R (1991) Resonant acceleration of alpha particles by ion cyclotron waves in the solar
wind. J Geophys Res 96:9801–9804. https://doi.org/10.1029/91JA00613
Gosling JT (2012) Magnetic reconnection in the solar wind. Space Sci Rev 172:187–200. https://doi.org/
10.1007/s11214-011-9747-2
Gosling JT, Baker DN, Bame SJ, Feldman WC, Zwickl RD, Smith EJ (1987) Bidirectional solar wind
electron heat flux events. J Geophys Res 92:8519–8535. https://doi.org/10.1029/JA092iA08p08519
Gould RW, O’Neil TM, Malmberg JH (1967) Plasma wave echo. Phys Rev Lett 19:219–222. https://doi.
org/10.1103/PhysRevLett.19.219
Graham GA, Rae IJ, Owen CJ, Walsh AP, Arridge CS, Gilbert L, Lewis GR, Jones GH, Forsyth C, Coates
AJ, Waite JH (2017) The evolution of solar wind strahl with heliospheric distance. J Geophys Res
122:3858–3874. https://doi.org/10.1002/2016JA023656
Grappin R, Mangeney A, Marsch E (1990) On the origin of solar wind MHD turbulence: Helios data
revisited. J Geophys Res 95(A6):8197–8209. https://doi.org/10.1029/JA095iA06p08197
Grard RJL, Buechner J, Scholer M, Burgess D (1991) Planetary magnetospheric physics I. In: Proceedings.
Symposium 6 and topical meeting of the COSPAR interdisciplinary scientific commission D (Meetings
D2, D4) of the COSPAR 28. Plenary meeting, The Hague (Netherlands), 25 June–6 July 1990. Adv
Space Res 11
Greco A, Chuychai P, Matthaeus WH, Servidio S, Dmitruk P (2008) Intermittent MHD structures and
classical discontinuities. Geophys Res Lett 35:L19111. https://doi.org/10.1029/2008GL035454
Greco A, Matthaeus WH, D’Amicis R, Servidio S, Dmitruk P (2012) Evidence for nonlinear development
of magnetohydrodynamic scale intermittency in the inner heliosphere. Astrophys J 749:105. https://
doi.org/10.1088/0004-637X/749/2/105
Greco A, Perri S, Servidio S, Yordanova E, Veltri P (2016) The complex structure of magnetic field discon-
tinuities in the turbulent solar wind. Astrophys J Lett 823:L39. https://doi.org/10.3847/2041-8205/
823/2/L39. arXiv:1511.03084
Greco A, Matthaeus WH, Perri S, Osman KT, Servidio S, Wan M, Dmitruk P (2018) Partial variance of
increments method in solar wind observations and plasma simulations. Space Sci Rev 214:1. https://
doi.org/10.1007/s11214-017-0435-8
Griffiths DJ (2013) Introduction to electrodynamics, 4th edn. Pearson Education, Boston
Gringauz KI, Bezrokikh VV, Ozerov VD, Rybchinskii RE (1960) A study of the interplanetary ionized gas,
high-energy electrons and corpuscular radiation from the Sun by means of the three-electrode trap for
charged particles on the second Soviet Cosmic Rocket. Sov Phys Dokl 5:361
Grotrian W (1939) Zur Frage der Deutung der Linien im Spektrum der Sonnenkorona. Naturwissenschaften
27:214–214. https://doi.org/10.1007/BF01488890
Gurgiolo C, Goldstein ML, Viñas AF, Fazakerley AN (2012) Direct observations of the formation of the
solar wind halo from the strahl. Ann Geophys 30:163–175. https://doi.org/10.5194/angeo-30-163-
2012
Gurnett DA, Kurth WS, Kirchner DL, Hospodarsky GB, Averkamp TF, Zarka P, Lecacheux A, Manning R,
Roux A, Canu P, Cornilleau-Wehrlin N, Galopeau P, Meyer A, Boström R, Gustafsson G, Wahlund JE,
Åhlen L, Rucker HO, Ladreiter HP, Macher W, Woolliscroft LJC, Alleyne H, Kaiser ML, Desch MD,
Farrell WM, Harvey CC, Louarn P, Kellogg PJ, Goetz K, Pedersen A (2004) The Cassini radio and
plasma wave investigation. Space Sci Rev 114:395–463. https://doi.org/10.1007/s11214-004-1434-0
Halekas JS, Taylor ER, Dalton G, Johnson G, Curtis DW, McFadden JP, Mitchell DL, Lin RP, Jakosky
BM (2015) The solar wind ion analyzer for MAVEN. Space Sci Rev 195:125–151. https://doi.org/10.
1007/s11214-013-0029-z
Hamilton K, Smith CW, Vasquez BJ, Leamon RJ (2008) Anisotropies and helicities in the solar wind inertial
and dissipation ranges at 1 AU. J Geophys Res 113:A01106. https://doi.org/10.1029/2007JA012559
Hammond CM, Feldman WC, McComas DJ, Phillips JL, Forsyth RJ (1996) Variation of electron-strahl
width in the high-speed solar wind: Ulysses observations. Astron Astrophys 316:350–354
He J, Marsch E, Tu C, Yao S, Tian H (2011) Possible evidence of Alfvén-cyclotron waves in the angle
distribution of magnetic helicity of solar wind turbulence. Astrophys J 731:85. https://doi.org/10.
1088/0004-637X/731/2/85
He J, Tu C, Marsch E, Yao S (2012a) Do oblique Alfvén/ion-cyclotron or fast-mode/whistler waves dominate
the dissipation of solar wind turbulence near the proton inertial length? Astrophys J Lett 745:L8. https://
doi.org/10.1088/2041-8205/745/1/L8

123
The multi-scale nature of the solar wind Page 115 of 136 5

He J, Tu C, Marsch E, Yao S (2012b) Reproduction of the observed two-component magnetic helicity in


solar wind turbulence by a superposition of parallel and oblique Alfvén waves. Astrophys J 749:86.
https://doi.org/10.1088/0004-637X/749/1/86
He J, Wang L, Tu C, Marsch E, Zong Q (2015) Evidence of landau and cyclotron resonance between protons
and kinetic waves in solar wind turbulence. Astrophys J Lett 800:L31. https://doi.org/10.1088/2041-
8205/800/2/L31
Heber B, Fichtner H, Scherer K (2006) Solar and heliospheric modulation of galactic cosmic rays. Space
Sci Rev 125:81–93. https://doi.org/10.1007/s11214-006-9048-3
Hefti S, Grünwaldt H, Ipavich FM, Bochsler P, Hovestadt D, Aellig MR, Hilchenbach M, Kallenbach R,
Galvin AB, Geiss J, Gliem F, Gloeckler G, Klecker B, Marsch E, Möbius E, Neugebauer M, Wurz
P (1998) Kinetic properties of solar wind minor ions and protons measured with SOHO/CELIAS. J
Geophys Res 103:29697–29704. https://doi.org/10.1029/1998JA900022
Heidrich-Meisner V, Peleikis T, Kruse M, Berger L, Wimmer-Schweingruber R (2016) Observations of high
and low Fe charge states in individual solar wind streams with coronal-hole origin. Astron Astrophys
593:A70. https://doi.org/10.1051/0004-6361/201527998
Hellinger P (2016) Ion collisional transport coefficients in the solar wind at 1 AU. Astrophys J 825(2):120.
https://doi.org/10.3847/0004-637X/825/2/120
Hellinger P, Matsumoto H (2000) New kinetic instability: oblique Alfvén fire hose. J Geophys Res
105:10519–10526. https://doi.org/10.1029/1999JA000297
Hellinger P, Trávníček PM (2008) Oblique proton fire hose instability in the expanding solar wind: hybrid
simulations. J Geophys Res 113:A10109. https://doi.org/10.1029/2008JA013416
Hellinger P, Trávníček PM (2009) On Coulomb collisions in bi-Maxwellian plasmas. Phys Plasmas
16(5):054501. https://doi.org/10.1063/1.3139253
Hellinger P, Trávníček PM (2010) Langevin representation of Coulomb collisions for bi-Maxwellian plas-
mas. J Comp Phys 229:5432–5439. https://doi.org/10.1016/j.jcp.2010.04.009
Hellinger P, Trávníček PM (2014) Solar wind protons at 1 AU: trends and bounds. Constraints and corre-
lations. Astrophys J Lett 784:L15. https://doi.org/10.1088/2041-8205/784/1/L15. arXiv:1402.4611
Hellinger P, Trávníček P, Kasper JC, Lazarus AJ (2006) Solar wind proton temperature anisotropy: lin-
ear theory and WIND/SWE observations. Geophys Res Lett 33:L09101. https://doi.org/10.1029/
2006GL025925
Hellinger P, Matteini L, Štverák Š, Trávníček PM, Marsch E (2011) Heating and cooling of protons in the
fast solar wind between 0.3 and 1 AU: Helios revisited. J Geophys Res 116:A09105. https://doi.org/
10.1029/2011JA016674
Hellinger P, Trávníček PM, Štverák Š, Matteini L, Velli M (2013) Proton thermal energetics in the solar
wind: Helios reloaded. J Geophys Res 118:1351–1365. https://doi.org/10.1002/jgra.50107
Hernández R, Marsch E (1985) Collisional time scales for temperature and velocity exchange between drift-
ing Maxwellians. J Geophys Res 90(A11):11062–11066. https://doi.org/10.1029/JA090iA11p11062
Hernández R, Livi S, Marsch E (1987) On the He2+ to H+ temperature ratio in slow solar wind. J Geophys
Res 92:7723–7727. https://doi.org/10.1029/JA092iA07p07723
Hoang S, Maksimovic M, Bougeret JL, Reiner MJ, Kaiser ML (1998) Wind-Ulysses source location of radio
emissions associated with the January 1997 coronal mass ejection. Geophys Res Lett 25(14):2497–
2500. https://doi.org/10.1029/98GL00571
Hodgson R (1859) On a curious appearance seen in the Sun. Mon Not R Astron Soc 20:15–16. https://doi.
org/10.1093/mnras/20.1.15
Hoffmeister C (1943) Physikalische Untersuchungen an Kometen. I. Die Beziehungen des primären
Schweifstrahls zum Radiusvektor. Z Astrophys 22:265
Hollweg JV (1971) Density fluctuations driven by Alfvén waves. J Geophys Res 76(22):5155. https://doi.
org/10.1029/JA076i022p05155
Hollweg JV (1974) Transverse Alfvén waves in the solar wind: arbitrary k, v0 , B0 , and |δ B|. J Geophys
Res 79(10):1539. https://doi.org/10.1029/JA079i010p01539
Hollweg JV (1975) Waves and instabilities in the solar wind. Rev Geophys Space Phys 13:263–289. https://
doi.org/10.1029/RG013i001p00263
Hollweg JV (1994) Beat, modulational, and decay instabilities of a circularly polarized Alfvén wave. J
Geophys Res 99:23. https://doi.org/10.1029/94JA02185
Hollweg JV (1999) Cyclotron resonance in coronal holes: 1. Heating and acceleration of protons, O5+ , and
Mg9+ . J Geophys Res 104:24781–24792. https://doi.org/10.1029/1999JA900300

123
5 Page 116 of 136 D. Verscharen et al.

Hollweg JV, Isenberg PA (2002) Generation of the fast solar wind: a review with emphasis on the resonant
cyclotron interaction. J Geophys Res 107:1147. https://doi.org/10.1029/2001JA000270
Hollweg JV, Lee MA (1989) Slow twists of solar magnetic flux tubes and the polar magnetic field of the
Sun. Geophys Res Lett 16:919–922. https://doi.org/10.1029/GL016i008p00919
Hollweg JV, Völk HJ (1970) New plasma instabilities in the solar wind. J Geophys Res 75:5297. https://
doi.org/10.1029/JA075i028p05297
Hoppock IW, Chandran BDG, Klein KG, Mallet A, Verscharen D (2018) Stochastic proton heating by
kinetic-Alfvén-wave turbulence in moderately high-β plasmas. J Plasma Phys 84(6):905840615.
https://doi.org/10.1017/S0022377818001277. arXiv:1811.08873
Horaites K, Astfalk P, Boldyrev S, Jenko F (2018a) Stability analysis of core-strahl electron distributions
in the solar wind. Mon Not R Astron Soc 480:1499. https://doi.org/10.1093/mnras/sty1808
Horaites K, Boldyrev S, Wilson LB III, Viñas AF, Merka J (2018b) Kinetic theory and fast wind observations
of the electron strahl. Mon Not R Astron Soc 474:115–127. https://doi.org/10.1093/mnras/stx2555
Horbury TS, Wicks RT, Chen CHK (2012) Anisotropy in space plasma turbulence: solar wind observations.
Space Sci Rev 172:325–342. https://doi.org/10.1007/s11214-011-9821-9
Howard TA, Tappin SJ (2009) Interplanetary coronal mass ejections observed in the heliosphere: 1. Review
of theory. Space Sci Rev 147:31–54. https://doi.org/10.1007/s11214-009-9542-5
Howes GG (2010) A prescription for the turbulent heating of astrophysical plasmas. Mon Not R Astron
Soc 409:L104–L108. https://doi.org/10.1111/j.1745-3933.2010.00958.x. arXiv:1009.4212
Howes GG (2015) A dynamical model of plasma turbulence in the solar wind. Philos Trans R Soc A
373:20140145. https://doi.org/10.1098/rsta.2014.0145. arXiv:1502.04109
Howes GG (2016) The dynamical generation of current sheets in astrophysical plasma turbulence. Astrophys
J Lett 827:L28. https://doi.org/10.3847/2041-8205/827/2/L28. arXiv:1607.07465
Howes GG, Cowley SC, Dorland W, Hammett GW, Quataert E, Schekochihin AA (2006) Astrophysical
gyrokinetics: basic equations and linear theory. Astrophys J 651:590–614. https://doi.org/10.1086/
506172. arXiv:astro-ph/0511812
Howes GG, Bale SD, Klein KG, Chen CHK, Salem CS, TenBarge JM (2012) The slow-mode nature of
compressible wave power in solar wind turbulence. Astrophys J Lett 753:L19. https://doi.org/10.1088/
2041-8205/753/1/L19. arXiv:1106.4327
Howes GG, Klein KG, TenBarge JM (2014a) The quasilinear premise for the modeling of plasma turbulence.
ArXiv e-prints arXiv:1404.2913
Howes GG, Klein KG, TenBarge JM (2014b) Validity of the Taylor hypothesis for linear kinetic waves in
the weakly collisional solar wind. Astrophys J 789(2):106. https://doi.org/10.1088/0004-637X/789/
2/106. arXiv:1405.5460
Howes GG, McCubbin AJ, Klein KG (2018) Spatially localized particle energization by Landau damping
in current sheets produced by strong Alfvén wave collisions. J Plasma Phys 84(1):905840105. https://
doi.org/10.1017/S0022377818000053. arXiv:1708.00757
Huba JD (2016) NRL plasma formulary. Techncial report, Naval Research Laboratory, Washington, DC.
http://www.nrl.navy.mil/ppd/content/nrl-plasma-formulary
Hughes RS, Gary SP, Wang J, Parashar TN (2017) Kinetic Alfvén turbulence: electron and ion heating by
particle-in-cell simulations. Astrophys J Lett 847:L14. https://doi.org/10.3847/2041-8213/aa8b13
Hunana P, Zank GP (2017) On the parallel and oblique firehose instability in fluid models. Astrophys J
839(1):13. https://doi.org/10.3847/1538-4357/aa64e3. arXiv:1703.06221
Hundhausen AJ (1970) Composition and dynamics of the solar wind plasma. Rev Geophys Space Phys
8:729–811. https://doi.org/10.1029/RG008i004p00729
Hundhausen AJ, Asbridge JR, Bame SJ, Gilbert HE, Strong IB (1967a) Vela 3 satellite observations of solar
wind ions: a preliminary report. J Geophys Res 72:87. https://doi.org/10.1029/JZ072i001p00087
Hundhausen AJ, Bame SJ, Ness NF (1967b) Solar wind thermal anisotropies: Vela 3 and IMP 3. J Geophys
Res 72:5265. https://doi.org/10.1029/JZ072i021p05265
Ibscher D, Schlickeiser R (2014) Solar wind kinetic instabilities at small plasma betas. Phys Plasmas
21:022110. https://doi.org/10.1063/1.4863497
Iroshnikov PS (1963) Turbulence of a conducting fluid in a strong magnetic field. Astron Zh 40:742
Isenberg PA (2001) Heating of coronal holes and generation of the solar wind by ion-cyclotron resonance.
Space Sci Rev 95:119–131
Isenberg PA, Hollweg JV (1983) On the preferential acceleration and heating of solar wind heavy ions. J
Geophys Res 88:3923–3935. https://doi.org/10.1029/JA088iA05p03923

123
The multi-scale nature of the solar wind Page 117 of 136 5

Isenberg PA, Maruca BA, Kasper JC (2013) Self-consistent ion cyclotron anisotropy-beta relation for solar
wind protons. Astrophys J 773:164. https://doi.org/10.1088/0004-637X/773/2/164. arXiv:1307.1059
Issautier K, Meyer-Vernet N, Moncuquet M, Hoang S (1998) Solar wind radial and latitudinal structure–
electron density and core temperature from ULYSSES thermal noise spectroscopy. J Geophys Res
103:1969. https://doi.org/10.1029/97JA02661
Jackson JD (1958) Plasma oscillations. Techncial report NP-7977; GM-TR-0165-00535, Space Technology
Labs
Jackson JD (1975) Classical electrodynamics, 2nd edn. Wiley, New York
Jian LK, Russell CT, Luhmann JG, Strangeway RJ, Leisner JS, Galvin AB (2009) Ion cyclotron waves in
the solar wind observed by STEREO near 1 AU. Astrophys J Lett 701:L105–L109. https://doi.org/
10.1088/0004-637X/701/2/L105
Jian LK, Russell CT, Luhmann JG, Anderson BJ, Boardsen SA, Strangeway RJ, Cowee MM, Wenn-
macher A (2010) Observations of ion cyclotron waves in the solar wind near 0.3 AU. J Geophys Res
115(A14):A12115. https://doi.org/10.1029/2010JA015737
Jian LK, Wei HY, Russell CT, Luhmann JG, Klecker B, Omidi N, Isenberg PA, Goldstein ML, Figueroa-
Viñas A, Blanco-Cano X (2014) Electromagnetic waves near the proton cyclotron frequency: STEREO
observations. Astrophys J 786:123. https://doi.org/10.1088/0004-637X/786/2/123
Jockers K (1968) On the stability of the solar wind. Solar Phys 3(4):603–610. https://doi.org/10.1007/
BF00151941
Jockers K (1970) Solar wind models based on exospheric theory. Astron Astrophys 6:219–239
Johnson NL (1979) Handbook of Soviet lunar and planetary exploration, science and technology series, vol
47. American Astronautical Society, San Diego
Johnson JR, Cheng CZ (2001) Stochastic ion heating at the magnetopause due to kinetic Alfvén waves.
Geophys Res Lett 28:4421–4424. https://doi.org/10.1029/2001GL013509
Kaiser ML, Kucera TA, Davila JM, St Cyr OC, Guhathakurta M, Christian E (2008) The STEREO mission:
an introduction. Space Sci Rev 136:5–16. https://doi.org/10.1007/s11214-007-9277-0
Karimabadi H, Roytershteyn V, Wan M, Matthaeus WH, Daughton W, Wu P, Shay M, Loring B, Borovsky
J, Leonardis E, Chapman SC, Nakamura TKM (2013) Coherent structures, intermittent turbulence,
and dissipation in high-temperature plasmas. Phys Plasmas 20(1):012303. https://doi.org/10.1063/1.
4773205
Kasper JC (2002) Solar wind plasma: kinetic properties and micro-instabilities. PhD thesis, Massachusetts
Institute of Technology. http://hdl.handle.net/1721.1/29937
Kasper JC, Klein KG (2019) Strong preferential ion heating is limited to within the solar Alfvén surface.
Astrophys J Lett 877(2):L35. https://doi.org/10.3847/2041-8213/ab1de5. arXiv:1906.02763
Kasper JC, Lazarus AJ, Gary SP (2002) Wind/SWE observations of firehose constraint on solar wind proton
temperature anisotropy. Geophys Res Lett 29:1839. https://doi.org/10.1029/2002GL015128
Kasper JC, Lazarus AJ, Steinberg JT, Ogilvie KW, Szabo A (2006) Physics-based tests to identify the
accuracy of solar wind ion measurements: a case study with the Wind faraday cups. J Geophys Res
111(A11):3105. https://doi.org/10.1029/2005JA011442
Kasper JC, Stevens ML, Lazarus AJ, Steinberg JT, Ogilvie KW (2007) Solar wind helium abundance as a
function of speed and heliographic latitude: variation through a solar cycle. Astrophys J 660:901–910.
https://doi.org/10.1086/510842
Kasper JC, Lazarus AJ, Gary SP (2008) Hot solar-wind helium: direct evidence for local heating by
Alfvén-cyclotron dissipation. Phys Rev Lett 101(26):261103. https://doi.org/10.1103/PhysRevLett.
101.261103
Kasper JC, Stevens ML, Korreck KE, Maruca BA, Kiefer KK, Schwadron NA, Lepri ST (2012) Evolution
of the relationships between helium abundance, minor ion charge state, and solar wind speed over the
solar cycle. Astrophys J 745:162. https://doi.org/10.1088/0004-637X/745/2/162
Kasper JC, Maruca BA, Stevens ML, Zaslavsky A (2013) Sensitive test for ion-cyclotron resonant heating
in the solar wind. Phys Rev Lett 110(9):091102. https://doi.org/10.1103/PhysRevLett.110.091102
Kasper JC, Klein KG, Weber T, Maksimovic M, Zaslavsky A, Bale SD, Maruca BA, Stevens ML, Case
AW (2017) A zone of preferential ion heating extends tens of solar radii from the sun. Astrophys J
849:126. https://doi.org/10.3847/1538-4357/aa84b1
Kawamori E (2013) Experimental verification of entropy cascade in two-dimensional electrostatic turbu-
lence in magnetized plasma. Phys Rev Lett 110(9):095001. https://doi.org/10.1103/PhysRevLett.110.
095001

123
5 Page 118 of 136 D. Verscharen et al.

Kawazura Y, Barnes M, Schekochihin AA (2019) Thermal disequilibration of ions and electrons by


collisionless plasma turbulence. Proc Natl Acad Sci 116(3):771–776. https://doi.org/10.1073/pnas.
1812491116. arXiv:1807.07702
Kellogg PJ, Horbury TS (2005) Rapid density fluctuations in the solar wind. Ann Geophys 23:3765–3773.
https://doi.org/10.5194/angeo-23-3765-2005
Kellogg PJ, Goetz K, Monson SJ (2016) Dust impact signals on the Wind spacecraft. J Geophys Res
121(2):966–991. https://doi.org/10.1002/2015JA021124
Kennel CF, Engelmann F (1966) Velocity space diffusion from weak plasma turbulence in a magnetic field.
Phys Fluids 9:2377–2388. https://doi.org/10.1063/1.1761629
Kennel CF, Petschek HE (1966) Limit on stably trapped particle fluxes. J Geophys Res 71:1
Kivelson MG, Bagenal F (2007) Planetary magnetospheres. In: McFadden LA, Weissman PR, Johnson TV
(eds) Encyclopedia of the solar system. Elsevier, Berlin, pp 519–540. https://doi.org/10.1016/B978-
012088589-3/50032-3
Kivelson MG, Southwood DJ (1996) Mirror instability II: the mechanism of nonlinear saturation. J Geophys
Res 101:17365–17372. https://doi.org/10.1029/96JA01407
Kiyani KH, Osman KT, Chapman SC (2015) Dissipation and heating in solar wind turbulence: from the
macro to the micro and back again. Philos Trans R Soc A. https://doi.org/10.1098/rsta.2014.0155
Klein KG, Chandran BDG (2016) Evolution of the proton velocity distribution due to stochastic heat-
ing in the near-Sun solar wind. Astrophys J 820:47. https://doi.org/10.3847/0004-637X/820/1/47.
arXiv:1602.05114
Klein KL, Dalla S (2017) Acceleration and propagation of solar energetic particles. Space Sci Rev 212:1107–
1136. https://doi.org/10.1007/s11214-017-0382-4. arXiv:1705.07274
Klein KG, Howes GG (2015) Predicted impacts of proton temperature anisotropy on solar wind turbulence.
Phys Plasmas 22(3):032903. https://doi.org/10.1063/1.4914933. arXiv:1503.00695
Klein LW, Ogilvie KW, Burlaga LF (1985) Coulomb collisions in the solar wind. J Geophys Res
90(A8):7389–7396. https://doi.org/10.1029/JA090iA08p07389
Klein KG, Howes GG, TenBarge JM, Bale SD, Chen CHK, Salem CS (2012) Using synthetic spacecraft
data to interpret compressible fluctuations in solar wind turbulence. Astrophys J 755:159. https://doi.
org/10.1088/0004-637X/755/2/159. arXiv:1206.6564
Klein KG, Howes GG, TenBarge JM (2014a) The violation of the Taylor hypothesis in measurements of
solar wind turbulence. Astrophys J Lett 790(2):L20. https://doi.org/10.1088/2041-8205/790/2/L20.
arXiv:1406.5470
Klein KG, Howes GG, TenBarge JM, Podesta JJ (2014b) Physical interpretation of the angle-dependent
magnetic helicity spectrum in the solar wind: the nature of turbulent fluctuations near the proton gyro-
radius scale. Astrophys J 785:138. https://doi.org/10.1088/0004-637X/785/2/138. arXiv:1403.2306
Klein KG, Perez JC, Verscharen D, Mallet A, Chandran BDG (2015) A modified version of Taylor’s
hypothesis for Solar Probe Plus observations. Astrophys J Lett 801(1):L18. https://doi.org/10.1088/
2041-8205/801/1/L18. arXiv:1412.3786
Klein KG, Kasper JC, Korreck KE, Stevens ML (2017) Applying Nyquist’s method for stability
determination to solar wind observations. J Geophys Res 122:9815–9823. https://doi.org/10.1002/
2017JA024486
Klein KG, Alterman BL, Stevens ML, Vech D, Kasper JC (2018) Majority of solar wind intervals support
ion-driven instabilities. Phys Rev Lett 120:205102. https://doi.org/10.1103/PhysRevLett.120.205102
Klein KG, Alexandrova O, Bookbinder J, Caprioli D, Case AW, Chandran BDG, Chen LJ, Horbury T, Jian
L, Kasper JC, Le Contel O, Maruca BA, Matthaeus W, Retino A, Roberts O, Schekochihin A, Skoug
R, Smith C, Steinberg J, Spence H, Vasquez B, TenBarge JM, Verscharen D, Whittlesey P (2019)
[Plasma 2020 Decadal] Multipoint measurements of the solar wind: a proposed advance for studying
magnetized turbulence. ArXiv e-prints arXiv:1903.05740
Klimontovich YL (1967) The statistical theory of non-equilibrium processes in a plasma. Pergamon Press,
Oxford
Klimontovich YL (1997) Physics of collisionless plasma. Phys Usp 40:21–51. https://doi.org/10.1070/
PU1997v040n01ABEH000200
Kohl JL, Noci G, Cranmer SR, Raymond JC (2006) Ultraviolet spectroscopy of the extended solar corona.
Astron Astrophys Rev 13:31–157. https://doi.org/10.1007/s00159-005-0026-7
Kohlhase CE, Penzo PA (1977) Voyager mission description. Space Sci Rev 21(2):77–101. https://doi.org/
10.1007/BF00200846

123
The multi-scale nature of the solar wind Page 119 of 136 5

Kolmogorov A (1941a) The local structure of turbulence in incompressible viscous fluid for very large
Reynolds numbers. Dokl Akad Nauk SSSR 30:301–305
Kolmogorov AN (1941b) Dissipation of energy in locally isotropic turbulence. Dokl Akad Nauk SSSR
32:16
Kraichnan RH (1965) Inertial-range spectrum of hydromagnetic turbulence. Phys Fluids 8:1385–1387.
https://doi.org/10.1063/1.1761412
Krall NA, Trivelpiece AW (1973) Principles of plasma physics. Series in pure and applied physics. McGraw-
Hill Kogakusha, Tokyo
Krüger H, Landgraf M, Altobelli N, Grün E (2007) Interstellar dust in the solar system. Space Sci Rev
130:401–408. https://doi.org/10.1007/s11214-007-9181-7. arXiv:0706.3110
Kunz MW, Schekochihin AA, Cowley SC, Binney JJ, Sanders JS (2011) A thermally stable heating mech-
anism for the intracluster medium: turbulence, magnetic fields and plasma instabilities. Mon Not R
Astron Soc 410:2446–2457. https://doi.org/10.1111/j.1365-2966.2010.17621.x. arXiv:1003.2719
Kunz MW, Schekochihin AA, Stone JM (2014) Firehose and mirror instabilities in a collisionless
shearing plasma. Phys Rev Lett 112(20):205003. https://doi.org/10.1103/PhysRevLett.112.205003.
arXiv:1402.0010
Kunz MW, Schekochihin AA, Chen CHK, Abel IG, Cowley SC (2015) Inertial-range kinetic turbulence
in pressure-anisotropic astrophysical plasmas. J Plasma Phys 81:325810501. https://doi.org/10.1017/
S0022377815000811
Kunz MW, Abel IG, Klein KG, Schekochihin AA (2018) Astrophysical gyrokinetics: turbulence in pressure-
anisotropic plasmas at ion scales and beyond. J Plasma Phys 84(2):715840201. https://doi.org/10.1017/
S0022377818000296. arXiv:1712.02269
Laakso H, Perry C, McCaffrey S, Herment D, Allen AJ, Harvey CC, Escoubet CP, Gruenberger C, Taylor
MGGT, Turner R (2010) Cluster active archive: overview. In: Laakso H, Taylor M, Escoubet C (eds)
The cluster active archive, Astrophysics and Space Science Proceedings, vol 11. Springer, Dordrecht,
pp 3–37. https://doi.org/10.1007/978-90-481-3499-1_1
Lacombe C, Alexandrova O, Matteini L, Santolík O, Cornilleau-Wehrlin N, Mangeney A, de Conchy Y,
Maksimovic M (2014) Whistler mode waves and the electron heat flux in the solar wind: Cluster
observations. Astrophys J 796:5. https://doi.org/10.1088/0004-637X/796/1/5. arXiv:1410.6187
Lacombe C, Alexandrova O, Matteini L (2017) Anisotropies of the magnetic field fluctuations at kinetic
scales in the solar wind: Cluster observations. Astrophys J 848:45. https://doi.org/10.3847/1538-4357/
aa8c06. arXiv:1710.02341
Lakhina GS (1985) Electromagnetic lower hybrid instability in the solar wind. Astrophys Space Sci 111:325–
334. https://doi.org/10.1007/BF00649972
Laming JM (2015) The FIP and inverse FIP effects in solar and stellar coronae. Living Rev Sol Phys 12:2.
https://doi.org/10.1007/lrsp-2015-2. arXiv:1504.08325
Landau LD (1936) Kinetic equation for the Coulomb effect. Phys Z Sowjetunion 10:154
Landau LD (1937) Kinetic equation for the Coulomb effect. Zh Eksp Teor Fiz 7:203
Landau LD (1946) On the vibrations of the electronic plasma. Zh Eksp Teor Fiz 10:25–34
Landau LD, Lifshitz EM (1969) Statistical physics. Part 1. Pergamon Press, Oxford
Landi S, Pantellini F, Matteini L (2010) Radial evolution of the electron velocity distribution in the helio-
sphere: role of collisions. In: 12th International Solar Wind Conference, vol 1216, pp 218–222. https://
doi.org/10.1063/1.3395841
Landi S, Matteini L, Pantellini F (2012) On the competition between radial expansion and Coulomb colli-
sions in shaping the electron velocity distribution function: kinetic simulations. Astrophys J 760:143.
https://doi.org/10.1088/0004-637X/760/2/143
Landi S, Matteini L, Pantellini F (2014) Electron heat flux in the solar wind: are we observing the collisional
limit in the 1 AU data? Astrophys J Lett 790:L12. https://doi.org/10.1088/2041-8205/790/1/L12
Lapenta G (2008) Self-feeding turbulent magnetic reconnection on macroscopic scales. Phys Rev Lett
100(23):235001. https://doi.org/10.1103/PhysRevLett.100.235001. arXiv:0805.0426
Larson DE, Lin RP, Steinberg J (2000) Extremely cold electrons in the January 1997 magnetic cloud.
Geophys Res Lett 27:157–160. https://doi.org/10.1029/1999GL003632
Lazarus AJ, Bridge HS, Davis J (1966) Preliminary results from the Pioneer 6 M. I. T. plasma experiment.
J Geophys Res 71:3787–3790. https://doi.org/10.1029/JZ071i015p03787
Lazarus AJ, Siscoe GL, Ness NF (1968) Plasma and magnetic field observations during the magnetosphere
passage of Pioneer 7. J Geophys Res 73:2399–2409. https://doi.org/10.1029/JA073i007p02399

123
5 Page 120 of 136 D. Verscharen et al.

Le Chat G, Issautier K, Meyer-Vernet N, Zouganelis I, Maksimovic M, Moncuquet M (2009) Quasi-thermal


noise in space plasma: “kappa” distributions. Phys Plasmas 16:102903–102903. https://doi.org/10.
1063/1.3243495
Le Chat G, Issautier K, Meyer-Vernet N, Hoang S (2011) Large-scale variation of solar wind electron
properties from quasi-thermal noise spectroscopy: Ulysses measurements. Sol Phys 271:141–148.
https://doi.org/10.1007/s11207-011-9797-3
Leamon RJ, Smith CW, Ness NF, Matthaeus WH, Wong HK (1998) Observational constraints on the
dynamics of the interplanetary magnetic field dissipation range. J Geophys Res 103:4775. https://doi.
org/10.1029/97JA03394
Leamon RJ, Smith CW, Ness NF, Wong HK (1999) Dissipation range dynamics: kinetic Alfvén waves and
the importance of βe . J Geophys Res 104:22331–22344. https://doi.org/10.1029/1999JA900158
Lemaire J, Scherer M (1971a) Kinetic models of the solar wind. J Geophys Res 76(31):7479–7490. https://
doi.org/10.1029/JA076i031p07479
Lemaire J, Scherer M (1971b) Simple model for an ion-exosphere in an open magnetic field. Phys Fluids
14(8):1683–1694. https://doi.org/10.1063/1.1693664
Lemaire J, Scherer M (1973) Kinetic models of the solar and polar winds. Rev Geophys Space Phys
11:427–468. https://doi.org/10.1029/RG011i002p00427
Lepping R (2000) Solar wind shock waves and discontinuities. In: Murdin P (ed) Encyclopedia of astronomy
and astrophysics. Institute of Physics Publishing, Bristol, p 2307. https://doi.org/10.1888/0333750888/
2307
Lepping RP, Acũna MH, Burlaga LF, Farrell WM, Slavin JA, Schatten KH, Mariani F, Ness NF, Neubauer
FM, Whang YC, Byrnes JB, Kennon RS, Panetta PV, Scheifele J, Worley EM (1995) The WIND
magnetic field investigation. Space Sci Rev 71:207–229. https://doi.org/10.1007/BF00751330
Lepri ST, Landi E, Zurbuchen TH (2013) Solar wind heavy ions over solar cycle 23: ACE/SWICS mea-
surements. Astrophys J 768:94. https://doi.org/10.1088/0004-637X/768/1/94
Leslie DC (1973) Developments in the theory of turbulence. Clarendon Press, Oxford
Levy EH (1976) The interplanetary magnetic field structure. Nature 261:394. https://doi.org/10.1038/
261394a0
Li X, Habbal SR (2000) Electron kinetic firehose instability. J Geophys Res 105:27377–27386. https://doi.
org/10.1029/2000JA000063
Lie-Svendsen Ø, Leer E (2000) The electron velocity distribution in the high-speed solar wind: modeling
the effects of protons. J Geophys Res 105:35–46. https://doi.org/10.1029/1999JA900438
Lie-Svendsen Ø, Hansteen VH, Leer E (1997) Kinetic electrons in high-speed solar wind streams: formation
of high-energy tails. J Geophys Res 102(A3):4701–4718. https://doi.org/10.1029/96JA03632
Lifshitz EM, Pitaevskii LP (1981) Physical kinetics. Pergamon Press, Oxford
Lin RP (1998) WIND observations of suprathermal electrons in the interplanetary medium. Space Sci Rev
86:61–78. https://doi.org/10.1023/A:1005048428480
Lin RP, Anderson KA, Ashford S, Carlson C, Curtis D, Ergun R, Larson D, McFadden J, McCarthy M,
Parks GK, Rème H, Bosqued JM, Coutelier J, Cotin F, D’Uston C, Wenzel KP, Sanderson TR, Henrion
J, Ronnet JC, Paschmann G (1995) A three-dimensional plasma and energetic particle investigation
for the WIND spacecraft. Space Sci Rev 71:125–153. https://doi.org/10.1007/BF00751328
Lithwick Y, Goldreich P, Sridhar S (2007) Imbalanced strong MHD turbulence. Astrophys J 655:269–274.
https://doi.org/10.1086/509884
Liu Y, Richardson JD, Belcher JW, Kasper JC, Elliott HA (2006) Thermodynamic structure of collision-
dominated expanding plasma: heating of interplanetary coronal mass ejections. J Geophys Res
111:A01102. https://doi.org/10.1029/2005JA011329
Livadiotis G (2017) Statistical origin and properties of kappa distributions. J Phys: Conf Ser 900:012014.
https://doi.org/10.1088/1742-6596/900/1/012014
Livadiotis G, McComas DJ (2013) Understanding kappa distributions: a toolbox for space science and
astrophysics. Space Sci Rev 175:183–214. https://doi.org/10.1007/s11214-013-9982-9
Livi S, Marsch E (1987) Generation of solar wind proton tails and double beams by Coulomb collisions. J
Geophys Res 92:7255–7261. https://doi.org/10.1029/JA092iA07p07255
Livi S, Marsch E, Rosenbauer H (1986) Coulomb collisional domains in the solar wind. J Geophys Res
91:8045–8050. https://doi.org/10.1029/JA091iA07p08045
Longmire CL (1963) Elementary plasma physics. Interscience Publishers, New York
Lopez RE, Freeman JW (1986) Solar wind proton temperature–velocity relationship. J Geophys Res
91:1701–1705. https://doi.org/10.1029/JA091iA02p01701

123
The multi-scale nature of the solar wind Page 121 of 136 5

Loureiro NF, Boldyrev S (2017) Role of magnetic reconnection in magnetohydrodynamic turbulence. Phys
Rev Lett 118(24):245101. https://doi.org/10.1103/PhysRevLett.118.245101
Loureiro NF, Uzdensky DA (2016) Magnetic reconnection: from the Sweet–Parker model to stochastic
plasmoid chains. Plasma Phys Control Fusion 58(1):014021. https://doi.org/10.1088/0741-3335/58/
1/014021. arXiv:1507.07756
Loureiro NF, Schekochihin AA, Cowley SC (2007) Instability of current sheets and formation
of plasmoid chains. Phys Plasmas 14(10):100703–100703. https://doi.org/10.1063/1.2783986.
arXiv:astro-ph/0703631
MacBride BT, Smith CW, Forman MA (2008) The turbulent cascade at 1 AU: energy transfer and the
third-order scaling for MHD. Astrophys J 679(2):1644–1660. https://doi.org/10.1086/529575
MacBride BT, Smith CW, Vasquez BJ (2010) Inertial-range anisotropies in the solar wind from 0.3 to 1
AU: Helios 1 observations. J Geophys Res 115:A07105. https://doi.org/10.1029/2009JA014939
Maksimovic M, Pierrard V, Riley P (1997) Ulysses electron distributions fitted with kappa functions.
Geophys Res Lett 24:1151–1154. https://doi.org/10.1029/97GL00992
Maksimovic M, Gary SP, Skoug RM (2000) Solar wind electron suprathermal strength and temper-
ature gradients: Ulysses observations. J Geophys Res 105:18337–18350. https://doi.org/10.1029/
2000JA900039
Maksimovic M, Pierrard V, Lemaire J (2001) On the exospheric approach for the solar wind acceleration.
Astrophys Space Sci 277:181–187. https://doi.org/10.1023/A:1012250027289
Maksimovic M, Zouganelis I, Chaufray JY, Issautier K, Scime EE, Littleton JE, Marsch E, McComas DJ,
Salem C, Lin RP, Elliott H (2005) Radial evolution of the electron distribution functions in the fast solar
wind between 0.3 and 1.5 AU. J Geophys Res 110:A09104. https://doi.org/10.1029/2005JA011119
Malara F, Velli M (1996) Parametric instability of a large-amplitude nonmonochromatic Alfvén wave. Phys
Plasmas 3(12):4427–4433. https://doi.org/10.1063/1.872043
Mallet A, Schekochihin AA, Chandran BDG (2015) Refined critical balance in strong Alfvénic turbulence.
Mon Not R Astron Soc 449:L77–L81. https://doi.org/10.1093/mnrasl/slv021. arXiv:1406.5658
Mallet A, Schekochihin AA, Chandran BDG (2017) Disruption of sheet-like structures in Alfvénic tur-
bulence by magnetic reconnection. Mon Not R Astron Soc 468:4862–4871. https://doi.org/10.1093/
mnras/stx670. arXiv:1612.07604
Mallet A, Klein KG, Chandran BDG, Grošelj D, Hoppock IW, Bowen TA, Salem CS, Bale SD (2019)
Interplay between intermittency and dissipation in collisionless plasma turbulence. J Plasma Phys
85(3):175850302. https://doi.org/10.1017/S0022377819000357. arXiv:1807.09301
Maneva YG, Poedts S (2018) Generation and evolution of anisotropic turbulence and related energy trans-
fer in drifting proton-alpha plasmas. Astron Astrophys 613:A10. https://doi.org/10.1051/0004-6361/
201731204
Mann I, Czechowski A, Meyer-Vernet N, Zaslavsky A, Lamy H (2010) Dust in the interplanetary
medium. Plasma Phys Control Fusion 52:124012. https://doi.org/10.1088/0741-3335/52/12/124012.
arXiv:1008.1742
Mariani F, Ness NF, Burlaga LF, Bavassano B, Villante U (1978) The large-scale structure of the interplan-
etary magnetic field between 1 and 0.3 AU during the primary mission of Helios 1. J Geophys Res
83:5161–5166. https://doi.org/10.1029/JA083iA11p05161
Mariani F, Villante U, Bruno R, Bavassano B, Ness NF (1979) An extended investigation of Helios 1 and 2
observations—the interplanetary magnetic field between 0.3 and 1 AU. Sol Phys 63:411–421. https://
doi.org/10.1007/BF00174545
Marquardt D (1963) An algorithm for least-squares estimation of nonlinear parameters. J Soc Ind Appl
Math 11(2):431–441. https://doi.org/10.1137/0111030
Marsch E (1994) Theoretical models for the solar wind. Adv Space Res 14(4):103–121. https://doi.org/10.
1016/0273-1177(94)90170-8
Marsch E (2006) Kinetic physics of the solar corona and solar wind. Living Rev Sol Phys 3:1. https://doi.
org/10.12942/lrsp-2006-1
Marsch E (2012) Helios: evolution of distribution functions 0.3–1 AU. Space Sci Rev 172:23–39. https://
doi.org/10.1007/s11214-010-9734-z
Marsch E (2018) Solar wind and kinetic heliophysics. Ann Geophys Discuss 2018:1–41. https://doi.org/
10.5194/angeo-2018-36. https://www.ann-geophys-discuss.net/angeo-2018-36/
Marsch E, Bourouaine S (2011) Velocity-space diffusion of solar wind protons in oblique waves and weak
turbulence. Ann Geophys 29:2089–2099. https://doi.org/10.5194/angeo-29-2089-2011

123
5 Page 122 of 136 D. Verscharen et al.

Marsch E, Chang T (1982) Lower hybrid waves in the solar wind. Geophys Res Lett 9:1155–1158. https://
doi.org/10.1029/GL009i010p01155
Marsch E, Goldstein H (1983) The effects of Coulomb collisions on solar wind ion velocity distributions.
J Geophys Res 88:9933–9940. https://doi.org/10.1029/JA088iA12p09933
Marsch E, Livi S (1985) Coulomb collision rates for self-similar and kappa distributions. Phys Fluids
28(5):1379–1386. https://doi.org/10.1063/1.864971
Marsch E, Tu CY (1990a) On the radial evolution of MHD turbulence in the inner heliosphere. J Geophys
Res 95(A6):8211–8229. https://doi.org/10.1029/JA095iA06p08211
Marsch E, Tu CY (1990b) Spectral and spatial evolution of compressible turbulence in the inner solar wind.
J Geophys Res 95:11945–11956. https://doi.org/10.1029/JA095iA08p11945
Marsch E, Tu CY (1993) Correlations between the fluctuations of pressure, density, temperature and mag-
netic field in the solar wind. Ann Geophys 11:659–677
Marsch E, Tu CY (1994) Non-Gaussian probability distributions of solar wind fluctuations. Ann Geophys
12:1127–1138. https://doi.org/10.1007/s00585-994-1127-8
Marsch E, Tu CY (2001) Heating and acceleration of coronal ions interacting with plasma waves through
cyclotron and Landau resonance. J Geophys Res 106:227–238. https://doi.org/10.1029/2000JA000042
Marsch E, Verscharen D (2011) On nonlinear Alfvén-cyclotron waves in multi-species plasma. J Plasma
Phys 77:385–403. https://doi.org/10.1017/S0022377810000541. arXiv:1101.1060
Marsch E, Rosenbauer H, Schwenn R, Mühlhäuser KH, Denskat KU (1981) Pronounced proton core
temperature anisotropy, ion differential speed, and simultaneous Alfven wave activity in slow solar
wind at 0.3 AU. J Geophys Res 86:9199–9203. https://doi.org/10.1029/JA086iA11p09199
Marsch E, Rosenbauer H, Schwenn R, Mühlhäuser KH, Neubauer FM (1982a) Solar wind helium ions:
observations of the Helios solar probes between 0.3 and 1 AU. J Geophys Res 87:35–51. https://doi.
org/10.1029/JA087iA01p00035
Marsch E, Schwenn R, Rosenbauer H, Mühlhäuser KH, Pilipp W, Neubauer FM (1982b) Solar wind
protons—three-dimensional velocity distributions and derived plasma parameters measured between
0.3 and 1 AU. J Geophys Res 87:52–72. https://doi.org/10.1029/JA087iA01p00052
Marsch E, Mühlhäuser KH, Rosenbauer H, Schwenn R (1983) On the equation of state of solar wind
ions derived from Helios measurements. J Geophys Res 88:2982–2992. https://doi.org/10.1029/
JA088iA04p02982
Marsch E, Pilipp WG, Thieme KM, Rosenbauer H (1989) Cooling of solar wind electrons inside 0.3 AU.
J Geophys Res 94:6893–6898. https://doi.org/10.1029/JA094iA06p06893
Marsch E, Ao XZ, Tu CY (2004) On the temperature anisotropy of the core part of the proton velocity distri-
bution function in the solar wind. J Geophys Res 109:A04102. https://doi.org/10.1029/2003JA010330
Martinović MM, Klein KG, Bourouaine S (2019) Radial evolution of stochastic heating in low-β solar
wind. Astrophys J 879(1):43. https://doi.org/10.3847/1538-4357/ab23f4. arXiv:1905.13355
Maruca BA (2012) Instability-driven limits on ion temperature anisotropy in the solar wind: observations and
linear Vlasov theory. PhD thesis, Harvard University. http://nrs.harvard.edu/urn-3:HUL.InstRepos:
9547903
Maruca BA, Kasper JC (2013) Improved interpretation of solar wind ion measurements via high-resolution
magnetic field data. Adv Space Res 52:723–731. https://doi.org/10.1016/j.asr.2013.04.006
Maruca BA, Kasper JC, Bale SD (2011) What are the relative roles of heating and cooling in generating solar
wind temperature anisotropies? Phys Rev Lett 107(20):201101. https://doi.org/10.1103/PhysRevLett.
107.201101
Maruca BA, Kasper JC, Gary SP (2012) Instability-driven limits on helium temperature anisotropy in the
solar wind: observations and linear Vlasov analysis. Astrophys J 748:137. https://doi.org/10.1088/
0004-637X/748/2/137
Maruca BA, Bale SD, Sorriso-Valvo L, Kasper JC, Stevens ML (2013) Collisional thermalization of
hydrogen and helium in solar-wind plasma. Phys Rev Lett 111(24):241101. https://doi.org/10.1103/
PhysRevLett.111.241101. arXiv:1311.5473
Maruca BA, Chasapis A, Gary SP, Bandyopadhyay R, Chhiber R, Parashar TN, Matthaeus WH, Shay
MA, Burch JL, Moore TE, Pollock CJ, Giles BJ, Paterson WR, Dorelli J, Gershman DJ, Torbert RB,
Russell CT, Strangeway RJ (2018) MMS observations of beta-dependent constraints on ion temperature
anisotropy in Earth’s magnetosheath. Astrophys J 866:25. https://doi.org/10.3847/1538-4357/aaddfb.
arXiv:1806.08886
Matsuda Y, Smith GR (1992) A microinstability code for a uniform magnetized plasma with an arbitrary
distribution function. J Comput Phys 100:229–235. https://doi.org/10.1016/0021-9991(92)90230-V

123
The multi-scale nature of the solar wind Page 123 of 136 5

Matteini L, Landi S, Hellinger P, Velli M (2006) Parallel proton fire hose instability in the expanding solar
wind: hybrid simulations. J Geophys Res 111:A10101. https://doi.org/10.1029/2006JA011667
Matteini L, Landi S, Hellinger P, Pantellini F, Maksimovic M, Velli M, Goldstein BE, Marsch E (2007)
Evolution of the solar wind proton temperature anisotropy from 0.3 to 2.5 AU. Geophys Res Lett
34:L20105. https://doi.org/10.1029/2007GL030920
Matteini L, Landi S, Velli M, Hellinger P (2010) Kinetics of parametric instabilities of Alfvén waves: evolu-
tion of ion distribution functions. J Geophys Res 115:A09106. https://doi.org/10.1029/2009JA014987
Matteini L, Hellinger P, Landi S, Trávníček PM, Velli M (2012) Ion kinetics in the solar wind: coupling
global expansion to local microphysics. Space Sci Rev 172:373–396. https://doi.org/10.1007/s11214-
011-9774-z
Matteini L, Hellinger P, Schwartz SJ, Landi S (2015a) Fire hose instability driven by alpha particle temper-
ature anisotropy. Astrophys J 812:13. https://doi.org/10.1088/0004-637X/812/1/13
Matteini L, Horbury TS, Pantellini F, Velli M, Schwartz SJ (2015b) Ion kinetic energy conservation and
magnetic field strength constancy in multi-fluid solar wind Alfvénic turbulence. Astrophys J 802:11.
https://doi.org/10.1088/0004-637X/802/1/11
Matteini L, Stansby D, Horbury TS, Chen CHK (2018) On the 1/f spectrum in the solar wind and its
connection with magnetic compressibility. Astrophys J Lett 869(2):L32. https://doi.org/10.3847/2041-
8213/aaf573. arXiv:1812.05716
Matthaeus WH, Goldstein ML (1982) Measurement of the rugged invariants of magnetohydrodynamic tur-
bulence in the solar wind. J Geophys Res 87:6011–6028. https://doi.org/10.1029/JA087iA08p06011
Matthaeus WH, Goldstein ML (1986) Low-frequency 1/f noise in the interplanetary magnetic field. Phys
Rev Lett 57(4):495–498. https://doi.org/10.1103/PhysRevLett.57.495
Matthaeus WH, Velli M (2011) Who needs turbulence? A review of turbulence effects in the heliosphere
and on the fundamental process of reconnection. Space Sci Rev 160:145–168. https://doi.org/10.1007/
s11214-011-9793-9
Matthaeus WH, Ambrosiano JJ, Goldstein ML (1984) Particle-acceleration by turbulent magnetohydrody-
namic reconnection. Phys Rev Lett 53:1449–1452. https://doi.org/10.1103/PhysRevLett.53.1449
Matthaeus WH, Goldstein ML, Roberts DA (1990) Evidence for the presence of quasi-two-dimensional
nearly incompressible fluctuations in the solar wind. J Geophys Res 95:20673–20683. https://doi.org/
10.1029/JA095iA12p20673
Matthaeus WH, Oughton S, Osman KT, Servidio S, Wan M, Gary SP, Shay MA, Valentini F, Roytershteyn V,
Karimabadi H, Chapman SC (2014) Nonlinear and linear timescales near kinetic scales in solar wind
turbulence. Astrophys J 790:155. https://doi.org/10.1088/0004-637X/790/2/155. arXiv:1404.6569
Matthaeus WH, Bandyopadhyay R, Brown MR, Borovsky J, Carbone V, Caprioli D, Chasapis A, Chhiber
R, Dasso S, Dmitruk P, Del Zanna L, Dmitruk PA, Franci L, Gary SP, Goldstein ML, Gomez D, Greco
A, Horbury TS, Ji H, Kasper JC, Klein KG, Landi S, Li H, Malara F, Maruca BA, Mininni P, Oughton
S, Papini E, Parashar TN, Petrosyan A, Pouquet A, Retino A, Roberts O, Ruffolo D, Servidio S,
Spence H, Smith CW, Stawarz JE, TenBarge J, Vasquez1 BJ, Vaivads A, Valentini F, Velli M, Verdini
A, Verscharen D, Whittlesey P, Wicks R, Bruno R, Zimbardo G (2019) [Plasma 2020 Decadal] The
essential role of multi-point measurements in turbulence investigations: the solar wind beyond single
scale and beyond the Taylor hypothesis. ArXiv e-prints arXiv:1903.06890
Maxwell JC (1867) On the dynamical theory of gases. Philos Trans Roy Soc London 157:49–88. https://
doi.org/10.1098/rstl.1867.0004
McChesney JM, Stern RA, Bellan PM (1987) Observation of fast stochastic ion heating by drift waves.
Phys Rev Lett 59:1436–1439. https://doi.org/10.1103/PhysRevLett.59.1436
McComas DJ, Bame SJ, Feldman WC, Gosling JT, Phillips JL (1992) Solar wind halo electrons from 1–4
AU. Geophys Res Lett 19:1291–1294. https://doi.org/10.1029/92GL00631
McComas DJ, Bame SJ, Barker P, Feldman WC, Phillips JL, Riley P, Griffee JW (1998a) Solar wind
electron proton alpha monitor (SWEPAM) for the Advanced Composition Explorer. Space Sci Rev
86:563–612. https://doi.org/10.1023/A:1005040232597
McComas DJ, Bame SJ, Barraclough BL, Feldman WC, Funsten HO, Gosling JT, Riley P, Skoug R, Balogh
A, Forsyth R, Goldstein BE, Neugebauer M (1998b) Ulysses’ return to the slow solar wind. Geophys
Res Lett 25:1–4. https://doi.org/10.1029/97GL03444
McComas DJ, Barraclough BL, Funsten HO, Gosling JT, Santiago-Muñoz E, Skoug RM, Goldstein BE,
Neugebauer M, Riley P, Balogh A (2000) Solar wind observations over Ulysses’ first full polar orbit.
J Geophys Res 105:10419–10434. https://doi.org/10.1029/1999JA000383

123
5 Page 124 of 136 D. Verscharen et al.

McComas DJ, Elliott HA, Schwadron NA, Gosling JT, Skoug RM, Goldstein BE (2003) The three-
dimensional solar wind around solar maximum. Geophys Res Lett 30:1517. https://doi.org/10.1029/
2003GL017136
McComas DJ, Ebert RW, Elliott HA, Goldstein BE, Gosling JT, Schwadron NA, Skoug RM (2008) Weaker
solar wind from the polar coronal holes and the whole Sun. Geophys Res Lett 35:L18103. https://doi.
org/10.1029/2008GL034896
McComas DJ, Dayeh MA, Allegrini F, Bzowski M, DeMajistre R, Fujiki K, Funsten HO, Fuselier SA,
Gruntman M, Janzen PH, Kubiak MA, Kucharek H, Livadiotis G, Möbius E, Reisenfeld DB, Reno
M, Schwadron NA, Sokół JM, Tokumaru M (2012) The first three years of IBEX observations and
our evolving heliosphere. Astrophys J Suppl Ser 203:1. https://doi.org/10.1088/0067-0049/203/1/1
McComb WD (1990) The physics of fluid turbulence. Oxford University Press, Oxford
McMillan BF, Cairns IH (2006) Lower hybrid turbulence driven by parallel currents and associated electron
energization. Phys Plasmas 13(5):052104. https://doi.org/10.1063/1.2198212
Melrose DB, McPhedran RC (1991) Electromagnetic processes in dispersive media: a treatment based on
the dielectric tensor. Cambridge University Press, Cambridge
Meyer-Vernet N, Perche C (1989) Tool kit for antennae and thermal noise near the plasma frequency. J
Geophys Res 94:2405–2415. https://doi.org/10.1029/JA094iA03p02405
Meyrand R, Kanekar A, Dorland W, Schekochihin AA (2019) Fluidization of collisionless plasma
turbulence. Proc Natl Acad Sci 116(4):1185–1194. https://doi.org/10.1073/pnas.1813913116.
arXiv:1808.04284
Migliuolo S (1985) Lower hybrid waves in finite-beta plasmas, destabilized by electron beams. J Geophys
Res 90:377–385. https://doi.org/10.1029/JA090iA01p00377
Mihalov JD, Wolfe JH (1978) Pioneer-10 observation of the solar wind proton temperature heliocentric
gradient. Sol Phys 60:399–406. https://doi.org/10.1007/BF00156539
Mikić Z, Lee MA (2006) An introduction to theory and models of CMEs, shocks, and solar energetic
particles. Space Sci Rev 123:57–80. https://doi.org/10.1007/s11214-006-9012-2
Milne EA (1926) On the possibility of the emission of high-speed atoms from the Sun and stars. Mon Not
R Astron Soc 86:459–473. https://doi.org/10.1093/mnras/86.7.459
Montgomery MD (1972) Average thermal characteristics of solar wind electrons. NASA Spec Publ 308:208
Montgomery D, Matthaeus WH (1995) Anisotropic modal energy transfer in interstellar turbulence. Astro-
phys J 447:706. https://doi.org/10.1086/175910
Montgomery D, Turner L (1981) Anisotropic magnetohydrodynamic turbulence in a strong external mag-
netic field. Phys Fluids 24:825–831. https://doi.org/10.1063/1.863455
Montgomery DC, Tidman DA (1964) Plasma kinetic theory. McGraw-Hill, New York
Montgomery D, Turner L (1982) Two-and-a-half-dimensional magnetohydrodynamic turbulence. Phys
Fluids 25:345–349. https://doi.org/10.1063/1.863741
Montgomery MD, Bame SJ, Hundhausen AJ (1968) Solar wind electrons: Vela 4 measurements. J Geophys
Res 73:4999. https://doi.org/10.1029/JA073i015p04999
Motschmann U, Glassmeier KH, Pinçon JL (2000) Multi-spacecraft filtering: plasma mode recognition.
In: Paschmann G, Daly PW (eds) Analysis methods for multi-spacecraft data, ISSI Scientific Report
SR-001 (Electronic edition 1.1), International Space Science Institute (ISSI), Bern, Chap 4, pp 79–90.
http://www.issibern.ch/forads/sr-001-04.pdf
Mott-Smith HM, Langmuir I (1926) The theory of collectors in gaseous discharges. Phys Rev 28:727–763.
https://doi.org/10.1103/PhysRev.28.727
Mottez F, Chanteur G (1994) Surface crossing by a group of satellites: a theoretical study. J Geophys Res
99(A7):13499–13507. https://doi.org/10.1029/93JA03326
Müller WC, Biskamp D, Grappin R (2003) Statistical anisotropy of magnetohydrodynamic turbulence.
Phys Rev E 67:066302. https://doi.org/10.1103/PhysRevE.67.066302. arXiv:physics/0306045
Müller D, Marsden RG, St Cyr OC, Gilbert HR (2013) Solar Orbiter: exploring the Sun-heliosphere con-
nection. Sol Phys 285:25–70. https://doi.org/10.1007/s11207-012-0085-7
Narayan R, McClintock JE (2008) Advection-dominated accretion and the black hole event horizon. New
Astron Rev 51:733–751. https://doi.org/10.1016/j.newar.2008.03.002. arXiv:0803.0322
Narita Y, Marsch E (2015) Kinetic slow mode in the solar wind and its possible role in turbulence dissipation
and ion heating. Astrophys J 805:24. https://doi.org/10.1088/0004-637X/805/1/24
Narita Y, Nakamura R, Baumjohann W, Glassmeier KH, Motschmann U, Giles B, Magnes W, Fischer D,
Torbert RB, Russell CT, Strangeway RJ, Burch JL, Nariyuki Y, Saito S, Gary SP (2016) On electron-

123
The multi-scale nature of the solar wind Page 125 of 136 5

scale whistler turbulence in the solar wind. Astrophys J Lett 827:L8. https://doi.org/10.3847/2041-
8205/827/1/L8
National Academy of Sciences, Engineering, and Medicine (2016) Achieving science with CubeSats: think-
ing inside the box. The National Academies Press, Washington, DC. https://doi.org/10.17226/23503
Navarro AB, Teaca B, Told D, Groselj D, Crandall P, Jenko F (2016) Structure of plasma heating in gyroki-
netic Alfvénic turbulence. Phys Rev Lett 117(24):245101. https://doi.org/10.1103/PhysRevLett.117.
245101. arXiv:1607.07480
Ness NF (1970) Magnetometers for space research. Space Sci Rev 11(4):459–554. https://doi.org/10.1007/
BF00183028
Neubauer FM, Glassmeier KH (1990) Use of an array of satellites as a wave telescope. J Geophys Res
95(A11):19115–19122. https://doi.org/10.1029/JA095iA11p19115
Neugebauer M (1976) The role of Coulomb collisions in limiting differential flow and temperature differ-
ences in the solar wind. J Geophys Res 81:78–82. https://doi.org/10.1029/JA081i001p00078
Neugebauer MM, Feldman WC (1979) Relation between superheating and superacceleration of helium in
the solar wind. Solar Phys 63:201–205. https://doi.org/10.1007/BF00155710
Neugebauer M, Snyder CW (1962) Solar plasma experiment. Science 138:1095–1097. https://doi.org/10.
1126/science.138.3545.1095-a
Neugebauer M, Goldstein BE, Bame SJ, Feldman WC (1994) Ulysses near-ecliptic observations of differ-
ential flow between protons and alphas in the solar wind. J Geophys Res 99:2505–2511. https://doi.
org/10.1029/93JA02615
Neugebauer M, Goldstein BE, Smith EJ, Feldman WC (1996) Ulysses observations of differential alpha-
proton streaming in the solar wind. J Geophys Res 101:17047–17056. https://doi.org/10.1029/
96JA01406
Newbury JA, Russell CT, Phillips JL, Gary SP (1998) Electron temperature in the ambient solar wind:
typical properties and a lower bound at 1 AU. J Geophys Res 103:9553–9566. https://doi.org/10.
1029/98JA00067
Nyquist H (1932) Regeneration theory. Bell Syst Tech J 11(1):126–147
Ofman L (2010) Wave modeling of the solar wind. Living Rev Sol Phys 7:4. https://doi.org/10.12942/lrsp-
2010-4
Ofman L, Viñas A, Gary SP (2001) Constraints on the O+5 anisotropy in the solar corona. Astrophys J Lett
547:L175–L178. https://doi.org/10.1086/318900
Ogilvie KW (1975) Differences between the bulk speeds of hydrogen and helium in the solar wind. J
Geophys Res 80:1335–1338. https://doi.org/10.1029/JA080i010p01335
Ogilvie KW, Coplan MA (1995) Solar wind composition. Rev Geophys 33:615–622. https://doi.org/10.
1029/95RG00122
Ogilvie KW, Scudder JD (1978) The radial gradients and collisional properties of solar wind electrons. J
Geophys Res 83(A8):3776–3782. https://doi.org/10.1029/JA083iA08p03776
Ogilvie KW, Chornay DJ, Fritzenreiter RJ, Hunsaker F, Keller J, Lobell J, Miller G, Scudder JD, Sittler
EC Jr, Torbert RB, Bodet D, Needell G, Lazarus AJ, Steinberg JT, Tappan JH, Mavretic A, Gergin E
(1995) SWE, a comprehensive plasma instrument for the Wind spacecraft. Space Sci Rev 71:55–77.
https://doi.org/10.1007/BF00751326
Ogilvie KW, Fitzenreiter R, Desch M (2000) Electrons in the low-density solar wind. J Geophys Res
105:27277–27288. https://doi.org/10.1029/2000JA000131
Omelchenko YA, Shapiro VD, Shevchenko VI, Ashour-Abdalla M, Schriver D (1994) Modified lower
hybrid fan instability excited by precipitating auroral electrons. J Geophys Res 99:5965–5976. https://
doi.org/10.1029/93JA01323
Osman KT, Matthaeus WH, Greco A, Servidio S (2011) Evidence for inhomogeneous heating in the solar
wind. Astrophys J Lett 727:L11. https://doi.org/10.1088/2041-8205/727/1/L11
Osman KT, Matthaeus WH, Hnat B, Chapman SC (2012) Kinetic signatures and intermittent turbulence
in the solar wind plasma. Phys Rev Lett 108(26):261103. https://doi.org/10.1103/PhysRevLett.108.
261103. arXiv:1203.6596
Osman KT, Kiyani KH, Chapman SC, Hnat B (2014a) Anisotropic intermittency of magnetohy-
drodynamic turbulence. Astrophys J Lett 783:L27. https://doi.org/10.1088/2041-8205/783/2/L27.
arXiv:1311.5938
Osman KT, Matthaeus WH, Gosling JT, Greco A, Servidio S, Hnat B, Chapman SC, Phan TD (2014b)
Magnetic reconnection and intermittent turbulence in the solar wind. Phys Rev Lett 112(21):215002.
https://doi.org/10.1103/PhysRevLett.112.215002. arXiv:1403.4590

123
5 Page 126 of 136 D. Verscharen et al.

Oughton S, Matthaeus WH, Wan M, Osman KT (2015) Anisotropy in solar wind plasma turbulence. Philos
Trans R Soc London A 373:20140152–20140152. https://doi.org/10.1098/rsta.2014.0152
Owens MJ, Forsyth RJ (2013) The heliospheric magnetic field. Living Rev Sol Phys 10:5. https://doi.org/
10.12942/lrsp-2013-5
Owens MJ, Crooker NU, Schwadron NA (2008) Suprathermal electron evolution in a Parker spiral magnetic
field. J Geophys Res 113(A12):A11104. https://doi.org/10.1029/2008JA013294
Owens MJ, Lockwood M, Riley P, Linker J (2017) Sunward strahl: a method to unambiguously determine
open solar flux from in situ spacecraft measurements using suprathermal electron data. J Geophys Res
122(A11):10. https://doi.org/10.1002/2017JA024631
Pagel C, Gary SP, de Koning CA, Skoug RM, Steinberg JT (2007) Scattering of suprathermal electrons in the
solar wind: ACE observations. J Geophys Res 112:A04103. https://doi.org/10.1029/2006JA011967
Parashar TN, Matthaeus WH (2016) Propinquity of current and vortex structures: effects on collisionless
plasma heating. Astrophys J 832:57. https://doi.org/10.3847/0004-637X/832/1/57. arXiv:1610.02912
Parker EN (1958) Dynamics of the interplanetary gas and magnetic fields. Astrophys J 128:664. https://
doi.org/10.1086/146579
Parker JT, Highcock EG, Schekochihin AA, Dellar PJ (2016) Suppression of phase mixing in
drift-kinetic plasma turbulence. Phys Plasmas 23(7):070703. https://doi.org/10.1063/1.4958954.
arXiv:1603.06968
Paschmann G, Øieroset M, Phan T (2013) In-situ observations of reconnection in space. Space Sci Rev
178:385–417. https://doi.org/10.1007/s11214-012-9957-2
Penrose O (1960) Electrostatic instabilities of a uniform non-Maxwellian plasma. Phys Fluids 3:258–265.
https://doi.org/10.1063/1.1706024
Perrone D, Alexandrova O, Roberts OW, Lion S, Lacombe C, Walsh A, Maksimovic M, Zouganelis I (2017)
Coherent structures at ion scales in fast solar wind: Cluster observations. Astrophys J 849:49. https://
doi.org/10.3847/1538-4357/aa9022. arXiv:1709.09644
Petrosyan A, Balogh A, Goldstein ML, Léorat J, Marsch E, Petrovay K, Roberts B, von Steiger R, Vial JC
(2010) Turbulence in the solar atmosphere and solar wind. Space Sci Rev 156:135–238. https://doi.
org/10.1007/s11214-010-9694-3
Pezzi O, Valentini F, Veltri P (2015) Collisional relaxation: Landau versus Dougherty operator. J Plasma
Phys 81:305810107. https://doi.org/10.1017/S0022377814000877
Phillips JL, Gosling JT (1990) Radial evolution of solar wind thermal electron distributions due to expansion
and collisions. J Geophys Res 95:4217–4228. https://doi.org/10.1029/JA095iA04p04217
Phillips JL, Gosling JT, McComas DJ, Bame SJ, Gary SP, Smith EJ (1989a) Anisotropic thermal elec-
tron distributions in the solar wind. J Geophys Res 94(A6):6563–6579. https://doi.org/10.1029/
JA094iA06p06563
Phillips JL, Gosling JT, McComas DJ, Bame SJ, Smith EJ (1989b) ISEE 3 observations of solar wind thermal
electrons with T-perpendicular greater than T-parallel. J Geophys Res 94:13377–13386. https://doi.
org/10.1029/JA094iA10p13377
Phillips JL, Bame SJ, Gosling JT, McComas DJ, Goldstein BE, Balogh A (1993) Solar wind thermal
electrons from 1.15 to 5.34 AU: Ulysses observations. Adv Space Res 13(6):47–50. https://doi.org/
10.1016/0273-1177(93)90389-S
Phillips JL, Bame SJ, Barnes A, Barraclough BL, Feldman WC, Goldstein BE, Gosling JT, Hoogeveen
GW, McComas DJ, Neugebauer M, Suess ST (1995) Ulysses solar wind plasma observations from
pole to pole. Geophys Res Lett 22:3301–3304. https://doi.org/10.1029/95GL03094
Pierrard V, Lamy H, Lemaire J (2004) Exospheric distributions of minor ions in the solar wind. J Geophys
Res 109:A02118. https://doi.org/10.1029/2003JA010069
Pilipp WG, Miggenrieder H, Mühlhäuser KH, Rosenbauer H, Schwenn R, Neubauer FM (1987a) Variations
of electron distribution functions in the solar wind. J Geophys Res 92:1103–1118. https://doi.org/10.
1029/JA092iA02p01103
Pilipp WG, Miggenrieder H, Montgomery MD, Mühlhäuser KH, Rosenbauer H, Schwenn R (1987b)
Characteristics of electron velocity distribution functions in the solar wind derived from the Helios
plasma experiment. J Geophys Res 92:1075–1092. https://doi.org/10.1029/JA092iA02p01075
Pilipp WG, Miggenrieder H, Montgomery MD, Mühlhäuser KH, Rosenbauer H, Schwenn R (1987c)
Unusual electron distribution functions in the solar wind derived from the Helios plasma experi-
ment: double-strahl distributions and distributions with an extremely anisotropic core. J Geophys Res
92(A2):1093–1102. https://doi.org/10.1029/JA092iA02p01093

123
The multi-scale nature of the solar wind Page 127 of 136 5

Pilipp WG, Mühlhäuser KH, Miggenrieder H, Rosenbauer H, Schwenn R (1990) Large-scale variations of
thermal electron parameters in the solar wind between 0.3 and 1 AU. J Geophys Res 95:6305–6329.
https://doi.org/10.1029/JA095iA05p06305
Pinçon JL, Motschmann U (2000) Multi-spacecraft filtering: general framework. In: Paschmann G, Daly PW
(eds) Analysis methods for multi-spacecraft data, ISSI Scientific Report SR-001 (Electronic edition
1.1), International Space Science Institute (ISSI), Bern, Chap 3, pp 65–78. http://www.issibern.ch/
forads/sr-001-03.pdf
Planck M (1917) Über einen Satz der statistischen Dynamik und seine Erweiterung in der Quantentheorie.
Sitzungsber Preuss Akad 24:324–341. https://biodiversitylibrary.org/page/29213319
Plunk GG (2013) Landau damping in a turbulent setting. Phys Plasmas 20(3):032304. https://doi.org/10.
1063/1.4794851. arXiv:1206.3415
Podesta JJ (2013) Evidence of kinetic Alfvén waves in the solar wind at 1 AU. Sol Phys 286:529–548.
https://doi.org/10.1007/s11207-013-0258-z
Podesta JJ, Gary SP (2011a) Effect of differential flow of alpha particles on proton pressure anisotropy
instabilities in the solar wind. Astrophys J 742:41. https://doi.org/10.1088/0004-637X/742/1/41
Podesta JJ, Gary SP (2011b) Magnetic helicity spectrum of solar wind fluctuations as a function of the angle
with respect to the local mean magnetic field. Astrophys J 734:15. https://doi.org/10.1088/0004-637X/
734/1/15
Podesta JJ, TenBarge JM (2012) Scale dependence of the variance anisotropy near the proton gyrora-
dius scale: additional evidence for kinetic Alfvén waves in the solar wind at 1 AU. J Geophys Res
117(A16):A10106. https://doi.org/10.1029/2012JA017724
Politano H, Pouquet A (1998) von Kármán–Howarth equation for magnetohydrodynamics and its conse-
quences on third-order longitudinal structure and correlation functions. Phys Rev E 57(1):R21–R24.
https://doi.org/10.1103/PhysRevE.57.R21
Pontin DI (2011) Three-dimensional magnetic reconnection regimes: a review. Adv Space Res 47:1508–
1522. https://doi.org/10.1016/j.asr.2010.12.022. arXiv:1101.0924
Porsche H (1977) General aspects of the mission Helios 1 and 2. J Geophys Res Z Geophys 42(6):551–559
Potgieter MS (2008) Challenges to cosmic ray modeling: from beyond the solar wind termination shock.
Adv Space Res 41:245–258. https://doi.org/10.1016/j.asr.2007.01.051
Potgieter MS (2013) Solar modulation of cosmic rays. Living Rev Sol Phys 10:3. https://doi.org/10.12942/
lrsp-2013-3. arXiv:1306.4421
Price CP, Swift DW, Lee LC (1986) Numerical simulation of nonoscillatory mirror waves at the Earth’s
magnetosheath. J Geophys Res 91:101–112. https://doi.org/10.1029/JA091iA01p00101
Pucci F, Velli M (2014) Reconnection of quasi-singular current sheets: the “ideal” tearing mode. Astrophys
J Lett 780:L19. https://doi.org/10.1088/2041-8205/780/2/L19
Quataert E (1998) Particle heating by Alfvénic turbulence in hot accretion flows. Astrophys J 500:978–991.
https://doi.org/10.1086/305770. arXiv:astro-ph/9710127
Quest KB, Shapiro VD (1996) Evolution of the fire-hose instability: linear theory and wave-wave coupling.
J Geophys Res 101:24457–24470. https://doi.org/10.1029/96JA01534
Raymond JC (1999) Composition variations in the solar corona and solar wind. Space Sci Rev 87:55–66.
https://doi.org/10.1023/A:1005157914229
Reiner MJ, Fainberg J, Kaiser ML, Stone RG (1998) Type III radio source located by Ulysses/Wind trian-
gulation. J Geophys Res 103(A2):1923. https://doi.org/10.1029/97JA02646
Reisenfeld DB, Gary SP, Gosling JT, Steinberg JT, McComas DJ, Goldstein BE, Neugebauer M (2001)
Helium energetics in the high-latitude solar wind: Ulysses observations. J Geophys Res 106:5693–
5708. https://doi.org/10.1029/2000JA000317
Retinò A, Sundkvist D, Vaivads A, Mozer F, André M, Owen CJ (2007) In situ evidence of magnetic
reconnection in turbulent plasma. Nature Phys 3:236–238. https://doi.org/10.1038/nphys574
Rice WKM, Zank GP (2003) Particle acceleration at CME driven shock waves. Adv Space Res 31:901–906.
https://doi.org/10.1016/S0273-1177(02)00797-4
Richardson JD, Wang C, Burlaga LF (2004) The solar wind in the outer heliosphere. Adv Space Res
34:150–156. https://doi.org/10.1016/j.asr.2003.03.066
Riley P, Sonett CP, Tsurutani BT, Balogh A, Forsyth RJ, Hoogeveen GW (1996) Properties of arc-polarized
Alfvén waves in the ecliptic plane: Ulysses observations. J Geophys Res 101:19987–19994. https://
doi.org/10.1029/96JA01743
Rincon F, Schekochihin AA, Cowley SC (2015) Non-linear mirror instability. Mon Not R Astron Soc
447:L45–L49. https://doi.org/10.1093/mnrasl/slu179. arXiv:1407.4707

123
5 Page 128 of 136 D. Verscharen et al.

Riquelme MA, Quataert E, Verscharen D (2015) Particle-in-cell simulations of continuously driven mirror
and ion cyclotron instabilities in high beta astrophysical and heliospheric plasmas. Astrophys J 800:27.
https://doi.org/10.1088/0004-637X/800/1/27. arXiv:1402.0014
Riquelme MA, Quataert E, Verscharen D (2016) PIC simulations of the effect of velocity space instabilities
on electron viscosity and thermal conduction. Astrophys J 824:123. https://doi.org/10.3847/0004-
637X/824/2/123. arXiv:1602.03126
Riquelme M, Osorio A, Quataert E (2017) Stochastic electron acceleration by the whistler instability in a
growing magnetic field. Astrophys J 850:113. https://doi.org/10.3847/1538-4357/aa95ba
Riquelme M, Quataert E, Verscharen D (2018) PIC simulations of velocity-space instabilities in a decreasing
magnetic field: viscosity and thermal conduction. Astrophys J 854:132. https://doi.org/10.3847/1538-
4357/aaa6d1. arXiv:1708.03926
Robert P, Dunlop MW, Roux A, Chanteur G (2000) Accuracy of current density determination. In:
Paschmann G, Daly PW (eds) Analysis methods for multi-spacecraft data, ISSI Scientific Report
SR-001 (Electronic edition 1.1), International Space Science Institute (ISSI), Bern, Chap 16, pp 395–
418. http://www.issibern.ch/forads/sr-001-16.pdf
Roberts DA, Goldstein ML, Klein LW, Matthaeus WH (1987) Origin and evolution of fluctuations in the
solar wind: Helios observations and Helios-Voyager comparisons. J Geophys Res 92:12023–12035.
https://doi.org/10.1029/JA092iA11p12023
Roberts OW, Li X, Li B (2013) Kinetic plasma turbulence in the fast solar wind measured by Cluster.
Astrophys J 769:58. https://doi.org/10.1088/0004-637X/769/1/58
Roberts OW, Narita Y, Li X, Escoubet CP, Laakso H (2017) Multipoint analysis of compressive fluctuations
in the fast and slow solar wind. J Geophys Res 122:6940–6963. https://doi.org/10.1002/2016JA023552
Roberts OW, Narita Y, Escoubet CP (2018) Multi-scale analysis of compressible fluctuations in the solar
wind. Ann Geophys 36:47–52. https://doi.org/10.5194/angeo-36-47-2018
Rönnmark K (1982) Waves in homogeneous, anisotropic multicomponent plasmas (WHAMP). Techni-
cal report KGI–179, Kiruna Geophysical Institute. http://inis.iaea.org/search/search.aspx?orig_q=RN:
14744092
Rosenbauer H, Schwenn R, Marsch E, Meyer B, Miggenrieder H, Montgomery MD, Mühlhäuser KH,
Pilipp W, Voges W, Zink SM (1977) A survey on initial results of the Helios plasma experiment. J
Geophys Res Z Geophys 42(6):561–580
Rosenbluth MN (1965) Microinstabilities. In: Lectures presented at the Trieste Seminar on Plasma Physics,
p 485
Rosenbluth MN, MacDonald WM, Judd DL (1957) Fokker–Planck equation for an inverse-square force.
Phys Rev 107(1):1–6. https://doi.org/10.1103/PhysRev.107.1
Rosenthal A (1982) A record of NASA space missions since 1958. Technical report TM-109260, NASA.
https://ntrs.nasa.gov/search.jsp?R=19940003358
Roux A, Le Contel O, Coillot C, Bouabdellah A, de la Porte B, Alison D, Ruocco S, Vassal MC (2008) The
search coil magnetometer for THEMIS. Space Sci Rev 141:265–275. https://doi.org/10.1007/s11214-
008-9455-8
Rowlands J, Shapiro VD, Shevchenko VI (1966) Quasilinear theory of plasma cyclotron instability. Sov
Phys JETP 23:651–660
Rudakov L, Crabtree C, Ganguli G, Mithaiwala M (2012) Quasilinear evolution of plasma distribution
functions and consequences on wave spectrum and perpendicular ion heating in the turbulent solar
wind. Phys Plasmas 19(4):042704–042704. https://doi.org/10.1063/1.3698407
Ruiz ME, Dasso S, Matthaeus WH, Marsch E, Weygand JM (2011) Aging of anisotropy of solar wind
magnetic fluctuations in the inner heliosphere. J Geophys Res 116(A15):A10102. https://doi.org/10.
1029/2011JA016697. arXiv:1110.4012
Russell CT, Mellott MM, Smith EJ, King JH (1983) Multiple spacecraft observations of interplanetary
shocks: four spacecraft determination of shock normals. J Geophys Res 88(A6):4739–4748. https://
doi.org/10.1029/JA088iA06p04739
Rutherford E (1911) The scattering of α and β particles by matter and the structure of the atom. Philos Mag
21(125):669–688. https://doi.org/10.1080/14786440508637080
Ryan JM, Lockwood JA, Debrunner H (2000) Solar energetic particles. Space Sci Rev 93:35–53. https://
doi.org/10.1023/A:1026580008909
Sabine E (1851) V. On periodical laws discoverable in the mean effects of the larger magnetic distur-
bances. Philos Trans R Soc London 141:123–139. https://doi.org/10.1098/rstl.1851.0007. http://rstl.
royalsocietypublishing.org/content/141/123.short

123
The multi-scale nature of the solar wind Page 129 of 136 5

Sabine E (1852) VIII. On periodical laws discoverable in the mean effects of the larger magnetic
disturbance.—No. II. Philos Trans R Soc London 142:103–124. https://doi.org/10.1098/rstl.1852.
0009
Šafránková J, Němeček Z, Přech L, Koval A, Čermák I, Beránek M, Zastenker G, Shevyrev N, Chesalin
L (2008) A new approach to solar wind monitoring. Adv Space Res 41:153–159. https://doi.org/10.
1016/j.asr.2007.08.034
Šafránková J, Němeček Z, Němec F, Verscharen D, Chen CHK, Durovcová T, Riazantseva MO (2019)
Scale-dependent polarization of solar wind velocity fluctuations at the inertial and kinetic scales.
Astrophys J 870:40. https://doi.org/10.3847/1538-4357/aaf239
Sagdeev RZ, Galeev AA (1969) Nonlinear plasma theory. Benjamin, New York
Sahraoui F, Goldstein ML, Robert P, Khotyaintsev YV (2009) Evidence of a cascade and dissipation of
solar-wind turbulence at the electron gyroscale. Phys Rev Lett 102(23):231102. https://doi.org/10.
1103/PhysRevLett.102.231102
Sahraoui F, Belmont G, Goldstein ML, Rezeau L (2010a) Limitations of multispacecraft data techniques
in measuring wave number spectra of space plasma turbulence. J Geophys Res 115:A04206. https://
doi.org/10.1029/2009JA014724
Sahraoui F, Goldstein ML, Belmont G, Canu P, Rezeau L (2010b) Three dimensional anisotropic k spectra
of turbulence at subproton scales in the solar wind. Phys Rev Lett 105(13):131101. https://doi.org/10.
1103/PhysRevLett.105.131101
Sahraoui F, Belmont G, Goldstein ML (2012) New insight into short-wavelength solar wind fluctuations from
vlasov theory. Astrophys J 748:100. https://doi.org/10.1088/0004-637X/748/2/100. arXiv:1109.1484
Saito S, Gary SP (2007) All whistlers are not created equally: scattering of strahl electrons in the solar wind
via particle-in-cell simulations. Geophys Res Lett 34:L01102. https://doi.org/10.1029/2006GL028173
Salem C, Hubert D, Lacombe C, Bale SD, Mangeney A, Larson DE, Lin RP (2003) Electron properties
and Coulomb collisions in the solar wind at 1 AU: Wind observations. Astrophys J 585:1147–1157.
https://doi.org/10.1086/346185
Salem CS, Howes GG, Sundkvist D, Bale SD, Chaston CC, Chen CHK, Mozer FS (2012) Identification of
kinetic Alfvén wave turbulence in the solar wind. Astrophys J Lett 745:L9. https://doi.org/10.1088/
2041-8205/745/1/L9
Sauvaud JA, Larson D, Aoustin C, Curtis D, Médale JL, Fedorov A, Rouzaud J, Luhmann J, Moreau T,
Schröder P, Louarn P, Dandouras I, Penou E (2008) The IMPACT Solar Wind Electron Analyzer
(SWEA). Space Sci Rev 136:227–239. https://doi.org/10.1007/s11214-007-9174-6
Scharer JE, Trivelpiece AW (1967) Cyclotron wave instabilities in a plasma. Phys Fluids 10:591–595.
https://doi.org/10.1063/1.1762153
Schekochihin AA, Cowley SC, Dorland W, Hammett GW, Howes GG, Plunk GG, Quataert E, Tatsuno
T (2008) Gyrokinetic turbulence: a nonlinear route to dissipation through phase space. Plasma Phys
Control Fusion 50(12):124024. https://doi.org/10.1088/0741-3335/50/12/124024. arXiv:0806.1069
Schekochihin AA, Cowley SC, Rincon F, Rosin MS (2010) Magnetofluid dynamics of magnetized cosmic
plasma: firehose and gyrothermal instabilities. Mon Not R Astron Soc 405:291–300. https://doi.org/
10.1111/j.1365-2966.2010.16493.x. arXiv:0912.1359
Schekochihin AA, Parker JT, Highcock EG, Dellar PJ, Dorland W, Hammett GW (2016) Phase mixing versus
nonlinear advection in drift-kinetic plasma turbulence. J Plasma Phys 82(2):905820212. https://doi.
org/10.1017/S0022377816000374. arXiv:1508.05988
Schekochihin AA, Kawazura Y, Barnes MA (2019) Constraints on ion versus electron heating by plasma
turbulence at low beta. J Plasma Phys 85(3):905850303. https://doi.org/10.1017/S0022377819000345.
arXiv:1812.09792
Schunk RW (1975) Transport equations for aeronomy. Planet Space Sci 23:437–485. https://doi.org/10.
1016/0032-0633(75)90118-X
Schunk RW (1977) Mathematical structure of transport equations for multispecies flows. Rev Geophys
Space Phys 15:429–445. https://doi.org/10.1029/RG015i004p00429
Schwartz SJ, Roxburgh IW (1980) Instabilities in the solar wind. Philos Trans R Soc London A 297:555–563.
https://doi.org/10.1098/rsta.1980.0231
Schwenn R, Rosenbauer H, Miggenrieder H (1975) The plasma experiment on board Helios. Raumfahrt-
forschung 19:226–232
Scime EE, Bame SJ, Feldman WC, Gary SP, Phillips JL, Balogh A (1994) Regulation of the solar wind
electron heat flux from 1 to 5 AU: Ulysses observations. J Geophys Res 99:23. https://doi.org/10.
1029/94JA02068

123
5 Page 130 of 136 D. Verscharen et al.

Scime EE, Bame SJ, Phillips JL, Balogh A (1995) Latitudinal variations in the solar wind electron heat
flux. Space Sci Rev 72:105–108. https://doi.org/10.1007/BF00768762
Scime EE, Badeau AE Jr, Littleton JE (1999) The electron heat flux in the polar solar wind: Ulysses
observations. Geophys Res Lett 26:2129–2132. https://doi.org/10.1029/1999GL900503
Scime EE, Littleton JE, Gary SP, Skoug R, Lin N (2001) Solar cycle variations in the electron heat flux:
Ulysses observations. Geophys Res Lett 28:2169–2172. https://doi.org/10.1029/2001GL012925
Scudder JD, Olbert S (1979a) A theory of local and global processes which affect solar wind electrons:
1. The origin of typical 1 AU velocity distribution functions-steady state theory. J Geophys Res
84(A6):2755–2772. https://doi.org/10.1029/JA084iA06p02755
Scudder JD, Olbert S (1979b) A theory of local and global processes which affect solar wind
electrons: 2. Experimental support. J Geophys Res 84(A11):6603–6620. https://doi.org/10.1029/
JA084iA11p06603
Sentman DD, Edmiston JP, Frank LA (1981) Instabilities of low frequency, parallel propagating electro-
magnetic waves in the Earth’s foreshock region. J Geophys Res 86:7487–7497. https://doi.org/10.
1029/JA086iA09p07487
Serbu GP (1972) Explorer 35 observations of solar-wind electron density, temperature, and anisotropy. J
Geophys Res 77:1703. https://doi.org/10.1029/JA077i010p01703
Servidio S, Matthaeus WH, Shay MA, Cassak PA, Dmitruk P (2009) Magnetic reconnection in two-
dimensional magnetohydrodynamic turbulence. Phys Rev Lett 102(11):115003. https://doi.org/10.
1103/PhysRevLett.102.115003
Servidio S, Matthaeus WH, Shay MA, Dmitruk P, Cassak PA, Wan M (2010) Statistics of magnetic recon-
nection in two-dimensional magnetohydrodynamic turbulence. Phys Plasmas 17(3):032315. https://
doi.org/10.1063/1.3368798
Servidio S, Greco A, Matthaeus WH, Osman KT, Dmitruk P (2011) Statistical association of discontinuities
and reconnection in magnetohydrodynamic turbulence. J Geophys Res 116:A09102. https://doi.org/
10.1029/2011JA016569
Servidio S, Osman KT, Valentini F, Perrone D, Califano F, Chapman S, Matthaeus WH, Veltri P (2014)
Proton kinetic effects in Vlasov and solar wind turbulence. Astrophys J Lett 781:L27. https://doi.org/
10.1088/2041-8205/781/2/L27. arXiv:1306.6455
Shapiro VD, Shevchenko VI (1962) The nonlinear theory of intercation between charged particle beams
and a plasma in a magnetic field. Sov Phys JETP 15:1053–1061
Sharma P, Hammett GW, Quataert E, Stone JM (2006) Shearing box simulations of the MRI in a collisionless
plasma. Astrophys J 637:952–967. https://doi.org/10.1086/498405. arXiv:astro-ph/0508502
Shay MA, Haggerty CC, Matthaeus WH, Parashar TN, Wan M, Wu P (2018) Turbulent heating due to
magnetic reconnection. Phys Plasmas 25(1):012304. https://doi.org/10.1063/1.4993423
Shebalin JV, Matthaeus WH, Montgomery D (1983) Anisotropy in MHD turbulence due to a mean magnetic
field. J Plasma Phys 29:525–547. https://doi.org/10.1017/S0022377800000933
Shevchenko VI, Galinsky VL (2010) Stability of the strahl electron distribution function and its dynamics.
Nonlinear Proc Geophys 17:593–597. https://doi.org/10.5194/npg-17-593-2010
Shoda M, Yokoyama T (2018) Anisotropic magnetohydrodynamic turbulence driven by parametric decay
instability: the onset of phase mixing and Alfvén wave turbulence. Astrophys J Lett 859:L17. https://
doi.org/10.3847/2041-8213/aac50c
Shoda M, Yokoyama T, Suzuki TK (2018) Frequency-dependent Alfvén-wave propagation in the solar
wind: onset and suppression of parametric decay instability. Astrophys J 860:17. https://doi.org/10.
3847/1538-4357/aac218
Shoji M, Omura Y, Tsurutani BT, Verkhoglyadova OP, Lembege B (2009) Mirror instability and L-mode
electromagnetic ion cyclotron instability: competition in the Earth’s magnetosheath. J Geophys Res.
https://doi.org/10.1029/2008JA014038
Sironi L, Narayan R (2015) Electron heating by the ion cyclotron instability in collisionless accretion flows.
I. Compression-driven instabilities and the electron heating mechanism. Astrophys J 800:88. https://
doi.org/10.1088/0004-637X/800/2/88. arXiv:1411.5685
Slocum RE, Reilly FN (1963) Low field helium magnetometer for space applications. IEEE Trans Nucl Sci
10(1):165–171. https://doi.org/10.1109/TNS.1963.4323257
Smith EJ, Sonett CP (1976) Extraterrestrial magnetic fields: achievements and opportunities. IEEE Trans
Geosci Electron 14(3):154–171. https://doi.org/10.1109/TGE.1976.294447
Smith EJ, Connor BV, Foster GT Jr (1975) Measuring the magnetic fields of Jupiter and the outer solar
system. IEEE Trans Magn 11(4):962–980. https://doi.org/10.1109/TMAG.1975.1058779

123
The multi-scale nature of the solar wind Page 131 of 136 5

Smith HM, Marsch E, Helander P (2012) Electron transport in the fast solar wind. Astrophys J 753:31.
https://doi.org/10.1088/0004-637X/753/1/31
Sommerfeld A (1952) Electrodynamics. Academic Press, New York
Sonett CP, Judge DL, Sims AR, Kelso JM (1960) A radial rocket survey of the distant geomagnetic field.
Geophys Res Lett 65(1):55–68. https://doi.org/10.1029/JZ065i001p00055
Sorriso-Valvo L, Carbone V, Veltri P, Consolini G, Bruno R (1999) Intermittency in the solar wind turbulence
through probability distribution functions of fluctuations. Geophys Res Lett 26:1801–1804. https://
doi.org/10.1029/1999GL900270. arXiv:physics/9903043
Sorriso-Valvo L, Carbone V, Giuliani P, Veltri P, Bruno R, Antoni V, Martines E (2001) Intermittency in
plasma turbulence. Planet Space Sci 49:1193–1200. https://doi.org/10.1016/S0032-0633(01)00060-
5
Southwood DJ, Kivelson MG (1993) Mirror instability. I. Physical mechanism of linear instability. J Geophys
Res 98:9181–9187. https://doi.org/10.1029/92JA02837
Spitzer L Jr (1956) Physics of fully ionized gases, Interscience Tracts on Physics and Astronomy, vol 3.
Interscience, New York
Spitzer L Jr, Härm R (1953) Transport phenomena in a completely ionized gas. Phys Rev 89(5):977–981.
https://doi.org/10.1103/PhysRev.89.977
Squire J, Quataert E, Schekochihin AA (2016) A stringent limit on the amplitude of Alfvénic perturbations in
high-beta low-collisionality plasmas. Astrophys J Lett 830:L25. https://doi.org/10.3847/2041-8205/
830/2/L25. arXiv:1605.02759
Squire J, Kunz MW, Quataert E, Schekochihin AA (2017a) Kinetic simulations of the interruption of large-
amplitude shear-Alfvén waves in a high-β plasma. Phys Rev Lett 119:155101. https://doi.org/10.1103/
PhysRevLett.119.155101
Squire J, Schekochihin AA, Quataert E (2017b) Amplitude limits and nonlinear damping of shear-Alfvén
waves in high-beta low-collisionality plasmas. New J Phys 19:055005. https://doi.org/10.1088/1367-
2630/aa6bb1
Sridhar S, Goldreich P (1994) Toward a theory of interstellar turbulence. 1. Weak Alfvénic turbulence.
Astrophys J 432:612–621. https://doi.org/10.1086/174600
Srivastava N, Schwenn R (2000) The origin of the solar wind: an overview. In: Scherer K, Fichtner H,
Marsch E (eds) The outer heliosphere: beyond the planets. Copernicus, Katlenburg-Lindau, pp 12–40
Steinberg JL, Hoang S, Lecacheux A, Aubier MG, Dulk GA (1984) Type III radio bursts in the interplanetary
medium—the role of propagation. Astron Astrophys 140:39–48
Steinberg JT, Lazarus AJ, Ogilvie KW, Lepping R, Byrnes J (1996) Differential flow between solar wind
protons and alpha particles: first WIND observations. Geophys Res Lett 23:1183–1186. https://doi.
org/10.1029/96GL00628
Stewart B (1861) On the great magnetic disturbance which extended from August 28 to September 7, 1859,
as recorded by photography at the Kew Observatory. Philos Trans R Soc London Ser I 151:423–430
Stix TH (1992) Waves in plasmas. American Institute of Physics, New York
Stone EC, Cummings AC, McDonald FB, Heikkila BC, Lal N, Webber WR (2005) Voyager 1 explores
the termination shock region and the heliosheath beyond. Science 309:2017–2020. https://doi.org/10.
1126/science.1117684
Storey LRO (1953) An investigation of whistling atmospherics. Philos Trans R Soc London A 246:113–141.
https://doi.org/10.1098/rsta.1953.0011
Štverák Š, Trávníček P, Maksimovic M, Marsch E, Fazakerley AN, Scime EE (2008) Electron tempera-
ture anisotropy constraints in the solar wind. J Geophys Res 113:A03103. https://doi.org/10.1029/
2007JA012733
Štverák Š, Maksimovic M, Trávníček PM, Marsch E, Fazakerley AN, Scime EE (2009) Radial evolu-
tion of nonthermal electron populations in the low-latitude solar wind: Helios, Cluster, and Ulysses
observations. J Geophys Res 114:A05104. https://doi.org/10.1029/2008JA013883
Štverák Š, Trávníček PM, Hellinger P (2015) Electron energetics in the expanding solar wind via Helios
observations. J Geophys Res 120:8177–8193. https://doi.org/10.1002/2015JA021368
Summers D, Thorne RM (1991) The modified plasma dispersion function. Phys Fluids B 3:1835–1847.
https://doi.org/10.1063/1.859653
Summers D, Xue S, Thorne RM (1994) Calculation of the dielectric tensor for a generalized Lorentzian
(kappa) distribution function. Phys Plasmas 1:2012–2025. https://doi.org/10.1063/1.870656
Sundkvist D, Retinò A, Vaivads A, Bale SD (2007) Dissipation in turbulent plasma due to reconnection in
thin current sheets. Phys Rev Lett 99(2):025004. https://doi.org/10.1103/PhysRevLett.99.025004

123
5 Page 132 of 136 D. Verscharen et al.

Swanson DG (2003) Plasma waves, 2nd edn. Institute of Physics Publishing, Bristol
Tajiri M (1967) Propagation of hydromagnetic waves in collisionless plasma. II. Kinetic approach. J Phys
Soc Jpn 22:1482
Tam SWY, Chang T (1999) Kinetic evolution and acceleration of the solar wind. Geophys Res Lett 26:3189–
3192. https://doi.org/10.1029/1999GL010689
Tao J, Wang L, Zong Q, Li G, Salem CS, Wimmer-Schweingruber RF, He J, Tu C, Bale SD (2016) Quiet-
time suprathermal (∼ 0.1–1.5 keV) electrons in the solar wind. Astrophys J 820:22. https://doi.org/
10.3847/0004-637X/820/1/22
Tatsuno T, Dorland W, Schekochihin AA, Plunk GG, Barnes M, Cowley SC, Howes GG (2009) Nonlinear
phase mixing and phase-space cascade of entropy in gyrokinetic plasma turbulence. Phys Rev Lett
103(1):015003. https://doi.org/10.1103/PhysRevLett.103.015003. arXiv:0811.2538
Taylor GI (1938) The spectrum of turbulence. Proc R Soc London Ser A 164:476–490. https://doi.org/10.
1098/rspa.1938.0032
Telloni D, Bruno R (2016) Linking fluid and kinetic scales in solar wind turbulence. Mon Not R Astron
Soc 463:L79–L83. https://doi.org/10.1093/mnrasl/slw135
TenBarge JM, Howes GG (2013) Current sheets and collisionless damping in kinetic plasma turbulence.
Astrophys J Lett 771:L27. https://doi.org/10.1088/2041-8205/771/2/L27. arXiv:1304.2958
TenBarge JM, Podesta JJ, Klein KG, Howes GG (2012) Interpreting magnetic variance anisotropy mea-
surements in the solar wind. Astrophys J 753:107. https://doi.org/10.1088/0004-637X/753/2/107.
arXiv:1205.0749
TenBarge JM, Howes GG, Dorland W (2013) Collisionless damping at electron scales in solar wind turbu-
lence. Astrophys J 774:139. https://doi.org/10.1088/0004-637X/774/2/139
TenBarge JM, Alexandrova O, Boldyrev S, Califano F, Cerri SS, Chen CHK, Howes GG, Horbury T, Isenberg
PA, Ji H, Klein KG, Krafft C, Kunz M, Loureiro NF, Mallet A, Maruca BA, Matthaeus WH, Meyrand
R, Quataert E, Perez JC, Roberts OW, Sahraoui F, Salem CS, Schekochihin AA, Spence H, Squire
J, Told D, Verscharen D, Wicks RT (2019) [Plasma 2020 Decadal] Disentangling the spatiotemporal
structure of turbulence using multi-spacecraft data. ArXiv e-prints arXiv:1903.05710
Tenerani A, Velli M (2013) Parametric decay of radial Alfvén waves in the expanding accelerating solar
wind. J Geophys Res 118:7507–7516. https://doi.org/10.1002/2013JA019293
Tenerani A, Velli M (2017) Evolving waves and turbulence in the outer corona and inner heliosphere: the
accelerating expanding box. Astrophys J 843:26. https://doi.org/10.3847/1538-4357/aa71b9
Tenerani A, Velli M (2018) Nonlinear firehose relaxation and constant-B field fluctuations. Astrophys J
Lett 867(2):L26. https://doi.org/10.3847/2041-8213/aaec01. arXiv:1808.04453
Tenerani A, Velli M, Pucci F, Landi S, Rappazzo AF (2016) ‘Ideally’ unstable current sheets and the
triggering of fast magnetic reconnection. J Plasma Phys 82(5):535820501. https://doi.org/10.1017/
S002237781600088X. arXiv:1608.05066
Tessein JA, Smith CW, MacBride BT, Matthaeus WH, Forman MA, Borovsky JE (2009) Spectral indices
for multi-dimensional interplanetary turbulence at 1 AU. Astrophys J 692:684–693. https://doi.org/
10.1088/0004-637X/692/1/684
Thomas BT, Smith EJ (1980) The Parker spiral configuration of the interplanetary magnetic field between
1 and 8.5 AU. J Geophys Res 85:6861–6867. https://doi.org/10.1029/JA085iA12p06861
Thornton ST, Marion JB (2004) Classical dynamics of particles and systems, 5th edn. Brooks/Cole, Belmont
Told D, Jenko F, TenBarge JM, Howes GG, Hammett GW (2015) Multiscale nature of the dissipation range
in gyrokinetic simulations of Alfvénic turbulence. Phys Rev Lett 115(2):025003. https://doi.org/10.
1103/PhysRevLett.115.025003. arXiv:1505.02204
Tong Y, Bale SD, Chen CHK, Salem CS, Verscharen D (2015) Effects of electron drifts on the collisionless
damping of kinetic Alfvén waves in the solar wind. Astrophys J Lett 804:L36. https://doi.org/10.1088/
2041-8205/804/2/L36
Tong Y, Bale SD, Salem C, Pulupa M (2018) Observed instability constraints on electron heat flux in the
solar wind. ArXiv e-prints arXiv:1801.07694
Tracy PJ, Kasper JC, Zurbuchen TH, Raines JM, Shearer P, Gilbert J (2015) Thermalization of heavy ions
in the solar wind. Astrophys J 812:170. https://doi.org/10.1088/0004-637X/812/2/170
Tracy PJ, Kasper JC, Raines JM, Shearer P, Gilbert JA, Zurbuchen TH (2016) Constraining solar wind
heating processes by kinetic properties of heavy ions. Phys Rev Lett 116(25):255101. https://doi.org/
10.1103/PhysRevLett.116.255101
Treumann RA, Baumjohann W (1997) Advanced space plasma physics. Imperial College Press, London

123
The multi-scale nature of the solar wind Page 133 of 136 5

Tsallis C (1988) Possible generalization of Boltzmann–Gibbs statistics. J Stat Phys 52:479–487. https://
doi.org/10.1007/BF01016429
Tsurutani BT, Ho CM, Smith EJ, Neugebauer M, Goldstein BE, Mok JS, Arballo JK, Balogh A, Southwood
DJ, Feldman WC (1994) The relationship between interplanetary discontinuities and Alfvén waves:
Ulysses observations. Geophys Res Lett 21:2267–2270. https://doi.org/10.1029/94GL02194
Tu CY, Marsch E (1993) A model of solar wind fluctuations with two components—Alfven waves and
convective structures. J Geophys Res 98:1257–1276. https://doi.org/10.1029/92JA01947
Tu CY, Marsch E (1994) On the nature of compressive fluctuations in the solar wind. J Geophys Res 99:21.
https://doi.org/10.1029/94JA00843
Tu CY, Marsch E (1995) MHD structures, waves and turbulence in the solar wind: observations and theories.
Space Sci Rev 73:1–210. https://doi.org/10.1007/BF00748891
Tu CY, Marsch E (2001) On cyclotron wave heating and acceleration of solar wind ions in the outer corona.
J Geophys Res 106:8233–8252. https://doi.org/10.1029/2000JA000024
Tu CY, Marsch E, Qin ZR (2004) Dependence of the proton beam drift velocity on the proton core plasma
beta in the solar wind. J Geophys Res 109:A05101. https://doi.org/10.1029/2004JA010391
Tumanski S (2011) Handbook of magnetic measurements. CRC Press, Boca Raton
Unti TWJ, Neugebauer M (1968) Alfvén waves in the solar wind. Phys Fluids 11:563–568. https://doi.org/
10.1063/1.1691953
Vafin S, Lazar M, Fichtner H, Schlickeiser R, Drillisch M (2018) Solar wind temperature anisotropy con-
straints from streaming instabilities. Astron Astrophys 613:A23. https://doi.org/10.1051/0004-6361/
201731852
Vafin S, Riazantseva M, Pohl M (2019) Coulomb collisions as a candidate for temperature anisotropy
constraints in the solar wind. Astrophys J 871:L11. https://doi.org/10.3847/2041-8213/aafb11
Vasko IY, Krasnoselskikh V, Tong Y, Bale SD, Bonnell JW, Mozer FS (2019) Whistler fan instability driven
by strahl electrons in the solar wind. Astrophys J 871:L29. https://doi.org/10.3847/2041-8213/ab01bd
Vasquez BJ, Hollweg JV (1996) Formation of arc-shaped Alfvén waves and rotational discontinuities from
oblique linearly polarized wave trains. J Geophys Res 101(A6):13527–13540. https://doi.org/10.1029/
96JA00612
Velli M (1994) From supersonic winds to accretion: comments on the stability of stellar winds and related
flows. Astrophys J Lett 432:L55. https://doi.org/10.1086/187510
Velli M (2001) Hydrodynamics of the solar wind expansion. Astrophys Space Sci 277:157–167. https://
doi.org/10.1023/A:1012237708634
Velli M, Pruneti F (1997) Alfvén waves in the solar corona and solar wind. Plasma Phys Control Fusion
39:B317–B324. https://doi.org/10.1088/0741-3335/39/12B/024
Verdini A, Grappin R, Pinto R, Velli M (2012) On the origin of the 1/f spectrum in the solar wind magnetic
field. Astrophys J Lett 750(2):L33. https://doi.org/10.1088/2041-8205/750/2/L33. arXiv:1203.6219
Verdon AL, Cairns IH, Melrose DB, Robinson PA (2009) Warm electromagnetic lower hybrid wave dis-
persion relation. Phys Plasmas 16(5):052105. https://doi.org/10.1063/1.3132628
Verscharen D (2012) On convected wave structures and spectral transfer in space plasmas: applications to
solar corona and solar wind. PhD thesis, Max Planck Institute for Solar System Research, Lindau, Ger-
many; Technical University Braunschweig, Braunschweig, Germany. https://doi.org/10.5281/zenodo.
50886
Verscharen D, Chandran BDG (2013) The dispersion relations and instability thresholds of oblique plasma
modes in the presence of an ion beam. Astrophys J 764:88. https://doi.org/10.1088/0004-637X/764/
1/88. arXiv:1212.5192
Verscharen D, Chandran BDG (2018) NHDS: the New Hampshire Dispersion relation Solver. Res Not Am
Astron Soc 2(2):13. https://doi.org/10.3847/2515-5172/aabfe3. arXiv:1804.10096
Verscharen D, Marsch E (2011) Apparent temperature anisotropies due to wave activity in the solar wind.
Ann Geophys 29:909–917. https://doi.org/10.5194/angeo-29-909-2011. arXiv:1106.5878
Verscharen D, Marsch E, Motschmann U, Müller J (2012a) Kinetic cascade beyond magnetohydrodynamics
of solar wind turbulence in two-dimensional hybrid simulations. Phys Plasmas 19(2):022305–022305.
https://doi.org/10.1063/1.3682960. arXiv:1201.2784
Verscharen D, Marsch E, Motschmann U, Müller J (2012b) Parametric decay of oblique Alfvén waves in
two-dimensional hybrid simulations. Phys Rev E 86(2):027401. https://doi.org/10.1103/PhysRevE.
86.027401. arXiv:1207.6144

123
5 Page 134 of 136 D. Verscharen et al.

Verscharen D, Bourouaine S, Chandran BDG (2013a) Instabilities driven by the drift and temperature
anisotropy of alpha particles in the solar wind. Astrophys J 773:163. https://doi.org/10.1088/0004-
637X/773/2/163. arXiv:1307.1823
Verscharen D, Bourouaine S, Chandran BDG, Maruca BA (2013b) A parallel-propagating Alfvénic ion-
beam instability in the high-beta solar wind. Astrophys J 773:8. https://doi.org/10.1088/0004-637X/
773/1/8. arXiv:1306.2531
Verscharen D, Chandran BDG, Bourouaine S, Hollweg JV (2015) Deceleration of alpha particles in the
solar wind by instabilities and the rotational force: implications for heating, azimuthal flow, and
the Parker spiral magnetic field. Astrophys J 806:157. https://doi.org/10.1088/0004-637X/806/2/157.
arXiv:1411.4570
Verscharen D, Chandran BDG, Klein KG, Quataert E (2016) Collisionless isotropization of the solar-wind
protons by compressive fluctuations and plasma instabilities. Astrophys J 831:128. https://doi.org/10.
3847/0004-637X/831/2/128. arXiv:1605.07143
Verscharen D, Chen CHK, Wicks RT (2017) On kinetic slow modes, fluid slow modes, and pressure-
balanced structures in the solar wind. Astrophys J 840:106. https://doi.org/10.3847/1538-4357/
aa6a56. arXiv:1703.03040
Verscharen D, Klein KG, Chandran BDG, Stevens ML, Salem CS, Bale SD (2018) ALPS: the Arbitrary
Linear Plasma Solver. J Plasma Phys 84(4):905840403. https://doi.org/10.1017/S0022377818000739.
arXiv:1803.04697
Verscharen D, Chandran BDG, Jeong SY, Salem CS, Pulupa MP, Bale SD (2019a) Self-induced scattering of
strahl electrons in the solar wind. Astrophys. J. 886:136. https://doi.org/10.3847/1538-4357/ab4c30
Verscharen D, Wicks RT, Alexandrova O, Bruno R, Burgess D, Chen CHK, D’Amicis R, De Keyser J,
Dudok de Wit T, Franci L, He J, Henri P, Kasahara S, Khotyaintsev Y, Klein KG, Lavraud B, Maruca
BA, Maksimovic M, Plaschke F, Poedts S, Reynolds CS, Roberts O, Sahraoui F, Saito S, Salem CS,
Saur J, Servidio S, Stawarz JE, Stverak S, Told D (2019b) A case for electron-astrophysics. ArXiv
e-prints arXiv:1908.02206
Vocks C, Salem C, Lin RP, Mann G (2005) Electron halo and strahl formation in the solar wind by resonant
interaction with whistler waves. Astrophys J 627:540–549. https://doi.org/10.1086/430119
von Steiger R, Zurbuchen TH (2002) Kinetic properties of heavy solar wind ions from Ulysses-SWICS.
Adv Space Res 30:73–78. https://doi.org/10.1016/S0273-1177(02)00174-6
von Steiger R, Zurbuchen TH (2006) Kinetic properties of heavy solar wind ions from Ulysses-SWICS.
Geophys Res Lett 33:L09103. https://doi.org/10.1029/2005GL024998
von Steiger R, Geiss J, Gloeckler G, Galvin AB (1995) Kinetic properties of heavy ions in the solar wind
from SWICS/Ulysses. Space Sci Rev 72:71–76. https://doi.org/10.1007/BF00768756
von Steiger R, Schwadron NA, Fisk LA, Geiss J, Gloeckler G, Hefti S, Wilken B, Wimmer-Schweingruber
RF, Zurbuchen TH (2000) Composition of quasi-stationary solar wind flows from Ulysses/Solar
Wind Ion Composition Spectrometer. J Geophys Res 105:27217–27238. https://doi.org/10.1029/
1999JA000358
Wan M, Matthaeus WH, Karimabadi H, Roytershteyn V, Shay M, Wu P, Daughton W, Loring B, Chapman
SC (2012) Intermittent dissipation at kinetic scales in collisionless plasma turbulence. Phys Rev Lett
109(19):195001. https://doi.org/10.1103/PhysRevLett.109.195001
Wan M, Matthaeus WH, Roytershteyn V, Karimabadi H, Parashar T, Wu P, Shay M (2015) Intermittent
dissipation and heating in 3D kinetic plasma turbulence. Phys Rev Lett 114(17):175002. https://doi.
org/10.1103/PhysRevLett.114.175002
Wan M, Matthaeus WH, Roytershteyn V, Parashar TN, Wu P, Karimabadi H (2016) Intermittency, coherent
structures and dissipation in plasma turbulence. Phys Plasmas 23(4):042307. https://doi.org/10.1063/
1.4945631
Wang B, Wang CB, Yoon PH, Wu CS (2011) Stochastic heating and acceleration of minor ions by Alfvén
waves. Geophys Res Lett 38(10):L10103. https://doi.org/10.1029/2011GL047729
Wang L, Lin RP, Salem C, Pulupa M, Larson DE, Yoon PH, Luhmann JG (2012) Quiet-time interplanetary
∼ 2–20 keV superhalo electrons at solar minimum. Astrophys J Lett 753:L23. https://doi.org/10.1088/
2041-8205/753/1/L23
Wang X, Tu C, He J, Marsch E, Wang L (2013) On Intermittent turbulence heating of the solar wind:
differences between tangential and rotational discontinuities. Astrophys J Lett 772:L14. https://doi.
org/10.1088/2041-8205/772/2/L14
Wang X, Tu C, He J, Marsch E, Wang L (2014) The influence of intermittency on the spectral anisotropy
of solar wind turbulence. Astrophys J Lett 783:L9. https://doi.org/10.1088/2041-8205/783/1/L9

123
The multi-scale nature of the solar wind Page 135 of 136 5

Webb DF, Howard TA (2012) Coronal mass ejections: observations. Living Rev Sol Phys 9:3. https://doi.
org/10.12942/lrsp-2012-3
Weber EJ, Davis L Jr (1967) The angular momentum of the solar wind. Astrophys J 148:217–227. https://
doi.org/10.1086/149138
Whang YC (1971) Higher moment equations and the distribution function of the solar-wind plasma. J
Geophys Res 76:7503. https://doi.org/10.1029/JA076i031p07503
Wicks RT, Horbury TS, Chen CHK, Schekochihin AA (2010) Power and spectral index anisotropy of the
entire inertial range of turbulence in the fast solar wind. Mon Not R Astron Soc 407:L31–L35. https://
doi.org/10.1111/j.1745-3933.2010.00898.x. arXiv:1002.2096
Wicks RT, Alexander RL, Stevens M, Wilson LB III, Moya PS, Viñas A, Jian LK, Roberts DA, O’Modhrain
S, Gilbert JA, Zurbuchen TH (2016) A proton-cyclotron wave storm generated by unstable proton
distribution functions in the solar wind. Astrophys J 819:6. https://doi.org/10.3847/0004-637X/819/
1/6
Wilson LB III, Stevens ML, Kasper JC, Klein KG, Maruca BA, Bale SD, Bowen TA, Pulupa MP, Salem CS
(2018) The statistical properties of solar wind temperature parameters near 1 AU. Astrophys J Supp
236:41. https://doi.org/10.3847/1538-4365/aab71c. arXiv:1802.08585
Wilson LB III, Chen LJ, Wang S, Schwartz SJ, Turner DL, Stevens ML, Kasper JC, Osmane A, Caprioli D,
Bale SD, Pulupa MP, Salem CS, Goodrich KA (2019) Electron energy partition across interplanetary
shocks. I. Methodology and data product. Astrophys J Supp 243(1):8. https://doi.org/10.3847/1538-
4365/ab22bd. arXiv:1902.01476
Winske D, Gary SP (1986) Electromagnetic instabilities driven by cool heavy ion beams. J Geophys Res
91:6825–6832. https://doi.org/10.1029/JA091iA06p06825
Woodham LD, Wicks RT, Verscharen D, Owen CJ (2018) The role of proton cyclotron resonance as a
dissipation mechanism in solar wind turbulence: a statistical study at ion-kinetic scales. Astrophys J
856:49. https://doi.org/10.3847/1538-4357/aab03d. arXiv:1801.07344
Wu TY (1966) Kinetic equations of gases and plasmas. Addison-Wesley, Reading
Wu P, Perri S, Osman K, Wan M, Matthaeus WH, Shay MA, Goldstein ML, Karimabadi H, Chapman S
(2013) Intermittent heating in solar wind and kinetic simulations. Astrophys J Lett 763:L30. https://
doi.org/10.1088/2041-8205/763/2/L30
Wu H, Verscharen D, Wicks RT, Chen CH, He J, Nicolaou G (2019) The fluid-like and kinetic behavior of
kinetic Alfvén turbulence in space plasma. Astrophys J 870:106. https://doi.org/10.3847/1538-4357/
aaef77. arXiv:1808.09763
Xia Q, Perez JC, Chandran BDG, Quataert E (2013) Perpendicular ion heating by reduced mag-
netohydrodynamic turbulence. Astrophys J 776:90. https://doi.org/10.1088/0004-637X/776/2/90.
arXiv:1309.0742
Xu F, Borovsky JE (2015) A new four-plasma categorization scheme for the solar wind. J Geophys Res
120:70–100. https://doi.org/10.1002/2014JA020412
Xue S, Thorne RM, Summers D (1993) Electromagnetic ion-cyclotron instability in space plasmas. J
Geophys Res 98:17475–17484. https://doi.org/10.1029/93JA00790
Xue S, Thorne RM, Summers D (1996) Excitation of magnetosonic waves in the undisturbed solar wind.
Geophys Res Lett 23:2557–2560. https://doi.org/10.1029/96GL02202
Yang L, Wang L, Li G, He J, Salem CS, Tu C, Wimmer-Schweingruber RF, Bale SD (2015) The angular
distribution of solar wind superhalo electrons at quiet times. Astrophys J Lett 811:L8. https://doi.org/
10.1088/2041-8205/811/1/L8
Yang L, He J, Tu C, Li S, Zhang L, Marsch E, Wang L, Wang X, Feng X (2017a) Multiscale pressure-balanced
structures in three-dimensional magnetohydrodynamic turbulence. Astrophys J 836:69. https://doi.org/
10.3847/1538-4357/836/1/69. arXiv:1612.01496
Yang Y, Matthaeus WH, Parashar TN, Wu P, Wan M, Shi Y, Chen S, Roytershteyn V, Daughton W
(2017b) Energy transfer channels and turbulence cascade in Vlasov–Maxwell turbulence. Phys Rev E
95(6):061201. https://doi.org/10.1103/PhysRevE.95.061201
Yao S, He JS, Marsch E, Tu CY, Pedersen A, Rème H, Trotignon JG (2011) Multi-scale anti-correlation
between electron density and magnetic field strength in the solar wind. Astrophys J 728:146. https://
doi.org/10.1088/0004-637X/728/2/146
Yao S, He JS, Tu CY, Wang LH, Marsch E (2013a) Small-scale pressure-balanced structures driven by
mirror-mode waves in the solar wind. Astrophys J 776:94. https://doi.org/10.1088/0004-637X/776/
2/94

123
5 Page 136 of 136 D. Verscharen et al.

Yao S, He JS, Tu CY, Wang LH, Marsch E (2013b) Small-scale pressure-balanced structures driven by
oblique slow mode waves measured in the solar wind. Astrophys J 774:59. https://doi.org/10.1088/
0004-637X/774/1/59
Yermolaev YI, Stupin VV (1990) Some alpha-particle heating and acceleration mechanisms in the solar
wind: Prognoz 7 measurements. Planet Space Sci 38(10):1305–1313. https://doi.org/10.1016/0032-
0633(90)90133-B
Yermolaev YI, Stupin VV, Zastenker GN, Khamitov GP, Kozak I (1989) Variations of solar wind proton
and alpha-particle hydrodynamic parameters: Prognoz 7 observations. Adv Space Res 9(4):123–126.
https://doi.org/10.1016/0273-1177(89)90104-X
Yermolaev YI, Stupin VV, Kozak I (1991) Dynamics of proton and alpha-particle velocities and temperatures
in the solar wind: Prognoz 7 observations. Adv Space Res 11(1):79–82. https://doi.org/10.1016/0273-
1177(91)90095-2
Yoon PH (2016) Proton temperature relaxation in the solar wind by combined collective and collisional
processes. J Geophys Res 121:10665–10676. https://doi.org/10.1002/2016JA023044
Yoon PH (2017) Kinetic instabilities in the solar wind driven by temperature anisotropies. Rev Mod Plasma
Phys 1(1):4. https://doi.org/10.1007/s41614-017-0006-1
Yoon PH, Fang TM (2008) Dispersion surfaces for low-frequency modes. Plasma Phys Control Fusion
50:125002. https://doi.org/10.1088/0741-3335/50/12/125002
Yoon PH, Sarfraz M (2017) Interplay of electron and proton instabilities in expanding solar wind. Astrophys
J 835:246. https://doi.org/10.3847/1538-4357/835/2/246
Young DT, Berthelier JJ, Blanc M, Burch JL, Coates AJ, Goldstein R, Grande M, Hill TW, Johnson RE,
Kelha V, McComas DJ, Sittler EC, Svenes KR, Szegö K, Tanskanen P, Ahola K, Anderson D, Bakshi
S, Baragiola RA, Barraclough BL, Black RK, Bolton S, Booker T, Bowman R, Casey P, Crary FJ,
Delapp D, Dirks G, Eaker N, Funsten H, Furman JD, Gosling JT, Hannula H, Holmlund C, Huomo
H, Illiano JM, Jensen P, Johnson MA, Linder DR, Luntama T, Maurice S, McCabe KP, Mursula
K, Narheim BT, Nordholt JE, Preece A, Rudzki J, Ruitberg A, Smith K, Szalai S, Thomsen MF,
Viherkanto K, Vilppola J, Vollmer T, Wahl TE, Wüest M, Ylikorpi T, Zinsmeyer C (2004) Cassini
Plasma Spectrometer Investigation. Space Sci Rev 114:1–112. https://doi.org/10.1007/s11214-004-
1406-4
Zank GP (1999) Interaction of the solar wind with the local interstellar medium: a theoretical perspective.
Space Sci Rev 89:413–688. https://doi.org/10.1023/A:1005155601277
Zank GP, Cairns IH, Webb GM (1995) The termination shock: physical processes. Adv Space Res 15:453–
462. https://doi.org/10.1016/0273-1177(94)00129-O
Zank GP, Adhikari L, Hunana P, Shiota D, Bruno R, Telloni D (2017) Theory and transport of nearly
incompressible magnetohydrodynamic turbulence. Astrophys J 835:147. https://doi.org/10.3847/
1538-4357/835/2/147
Zank GP, Adhikari L, Zhao LL, Mostafavi P, Zirnstein EJ, McComas DJ (2018) The pickup ion-mediated
solar wind. Astrophys J 869:23. https://doi.org/10.3847/1538-4357/aaebfe
Zaslavsky A, Meyer-Vernet N, Mann I, Czechowski A, Issautier K, Le Chat G, Pantellini F, Goetz K,
Maksimovic M, Bale SD, Kasper JC (2012) Interplanetary dust detection by radio antennas: mass
calibration and fluxes measured by STEREO/WAVES. J Geophys Res 117:A05102. https://doi.org/
10.1029/2011JA017480
Zhdankin V, Boldyrev S, Chen CHK (2016) Intermittency of energy dissipation in Alfvénic turbulence.
Mon Not R Astron Soc 457:L69–L73. https://doi.org/10.1093/mnrasl/slv208. arXiv:1512.07355
Zhu X, He J, Verscharen D, Zhao J (2019) Composition of wave modes in magnetosheath turbulence from
sub-ion to sub-electron scales. Astrophys J 878(1):48. https://doi.org/10.3847/1538-4357/ab1be7
Zurbuchen TH, Richardson IG (2006) In-situ solar wind and magnetic field signatures of interplanetary
coronal mass ejections. Space Sci Rev 123:31–43. https://doi.org/10.1007/s11214-006-9010-4
Zurbuchen TH, Hefti S, Fisk LA, Gloeckler G, von Steiger R (1999) The transition between fast and
slow solar wind from composition data. Space Sci Rev 87:353–356. https://doi.org/10.1023/A:
1005126718714
Zurbuchen TH, Fisk LA, Gloeckler G, von Steiger R (2002) The solar wind composition throughout
the solar cycle: a continuum of dynamic states. Geophys Res Lett 29:1352. https://doi.org/10.1029/
2001GL013946

Publisher’s Note Springer Nature remains neutral with regard to jurisdictional claims in published maps
and institutional affiliations.

123

You might also like