Skip to main content
Philosophical transactions. Series A, Mathematical, physical, and engineering sciences logoLink to Philosophical transactions. Series A, Mathematical, physical, and engineering sciences
. 2015 May 13;373(2041):20140150. doi: 10.1098/rsta.2014.0150

Third-moment descriptions of the interplanetary turbulent cascade, intermittency and back transfer

Jesse T Coburn 1, Miriam A Forman 2, Charles W Smith 1,, Bernard J Vasquez 1, Julia E Stawarz 3
PMCID: PMC4394682  PMID: 25848079

Abstract

We review some aspects of solar wind turbulence with an emphasis on the ability of the turbulence to account for the observed heating of the solar wind. Particular attention is paid to the use of structure functions in computing energy cascade rates and their general agreement with the measured thermal proton heating. We then examine the use of 1 h data samples that are comparable in length to the correlation length for the fluctuations to obtain insights into local inertial range dynamics and find evidence for intermittency in the computed energy cascade rates. When the magnetic energy dominates the kinetic energy, there is evidence of anti-correlation in the cascade of energy associated with the outward- and inward-propagating components that we can only partially explain.

Keywords: turbulence, heating, cascade

1. Introduction

For most of the space age our view of solar wind fluctuations (magnetic, velocity, density, etc.) has been based on the theory of plasma waves. Attempts to incorporate turbulence concepts into this thinking have often been treated as little more than an afterthought that is either a secondary dynamic or a concept in direct conflict with the wave interpretation. The so-called weak turbulence theory illustrates this point wherein the primary dynamic is wave propagation with the secondary dynamic of waves exchanging energy on a time scale that is long compared with the wave period. Coleman illustrated this conflict with two papers: in the first [1], he argued that transverse magnetic fluctuations in the solar wind resemble low-frequency wave behaviour; whereas in the second [2], he argued that the reproducible spectrum of interplanetary fluctuations more nearly resembles the physics of hydrodynamic (HD) turbulence. If there was debate on these conflicting ideas, the wave viewpoint gained momentum with the observations that the fluctuations are transverse to the mean magnetic field, largely non-compressive, and with correlations suggestive of anti-Sunward propagation [3]. These are the properties of low-frequency Alfvén waves originating at the Alfvén radius.

The idea of the solar wind functioning as a giant wind tunnel for the purpose of studying magnetohydrodynamic (MHD) turbulence seems diminished, with many authors viewing turbulence as weakly interacting waves. There are problems with the wave interpretation and they are greatest when one considers the non-interacting wave viewpoint [4]. This view precludes the dynamics of many turbulent processes such as magnetic reconnection [5,6]. Within the context of solar wind heating, the greatest problem in the non-interacting wave viewpoint is that the wave energy available at scales that resonate with thermal particles is finite and insufficient to provide the observed heating rates [7]. Turbulence, and specifically the energy cascade that exists within well-developed turbulence, replenishes the energy at these scales and permits ongoing heating at a rate in general agreement with the observations [831].

Turbulence is the nonlinear evolution of a fluid generally considered to involve a wide range of spatial scales [32]. While it may arise from a single instability, it incorporates a broad range of dynamical processes acting in randomly distributed, coherent dynamics across the volume. It is not one nonlinear process—it is generally many dynamical processes. A fully developed turbulent spectrum is generally regarded to possess three subranges. The largest and most long-lived fluctuations constitute the ‘energy-containing range’. An example of these is the large-scale shear-flow regions in the solar wind. Once injected into the flow, these scales drive the turbulence and provide an energy reservoir for the excitation of smaller scales. The intermediate-scale fluctuations that are created by the slow destruction of the energy-containing range, but that are removed from the scales at which dissipation occurs, are called the ‘inertial range’. The inertial range conserves energy as various instabilities destroy and create coherent fluctuations in the process of moving energy to different spatial scales [3335]. This analysis focuses on these scales. The small-scale fluctuations form the ‘dissipation range’ where the spectrum steepens and turbulent energy is converted into heat [2,34,3643]. In the solar wind, there is likely to be a second inertial range at scales smaller than those where dissipation on thermal ions occurs and a second dissipation range where dissipation on thermal electrons occurs [44].

Statistical arguments derived from the fundamental equations without the approximations that lead to a finite set of dynamics or a basis set of oscillations have provided insights and reproducible results that can be confirmed by experiment and simulation. One of those reproducible results is the average transport of energy between scales. In three-dimensional HD and MHD turbulence which is statistically independent of location (homogeneous), energy is transported on average from large- to small-scale fluctuations. This is called a ‘cascade’. The cascade of energy is the central result that explains the heating of the solar wind: when small-scale fluctuations that permit one or another form of energy exchange between collective (fluid) motion are exhausted that energy is replaced by the turbulent cascade so that heating can continue. Its description does not rely on a selection of wave modes or a single dynamic describing their interaction. In this way, the large-scale fluctuations of the solar wind provide the energy that is transported to smaller scales where dissipation can occur. Small scales provide the heating dynamic, whereas the large scales provide the reservoir of energy.

There is a fundamental difficulty in assessing turbulent processes in the solar wind and it is best described by examination of the dynamical equations. The Navier–Stokes (N-S) equation that describes a traditional fluid is given by

1. 1.1

where ρ is the density, v is the velocity fluctuation, P is the pressure and ν is the viscosity. Three-dimensional gradients such as those seen here () cannot be evaluated using a single spacecraft. This prevents direct evaluation of the nonlinear terms.

This problem can be circumvented if one asks the correctly posed statistical question. If one assumes incompressibility, homogeneity, scale separation (there exists scales where energy is conserved that is separate from the scales where energy is dissipated) and addresses the statistics of the ensemble only [33,4547], then the energy cascade within the inertial range can be described as

1. 1.2

where Δv(L)≡v(x+L)−v(x) is the difference between the velocity measured at two points separated by the vector L and L is the gradient with respect to L. If one assumes that the statistics are stationary and isotropic and that the lag L resides within the inertial range, then they recover the familiar ‘Kolmogorov's 4/5 Law’ [48]

1. 1.3

where 〈…〉 denotes ensemble average and L is the separation vector. We will refer to this theory as K41a. The separation vector L can now be evaluated in any direction and in the solar wind we choose the anti-Sunward direction of the solar wind flow. Note that a simple scaling of equation (1.3) yields Inline graphic, where vk is called the ‘characteristic speed’ of an eddy of size L, where k≡2π/L.

The derivation of equations (1.2) and (1.3) is beyond the scope of this paper as it is somewhat lengthy. However, we note that K41a first assumes an expression for an isotropic cascade geometry derived by von Karman & Howarth [33] and derives equation (1.3) from that point. Frisch's textbook [47] also follows this approach. A more modern derivation can be performed that follows the Politano & Pouquet [49,50] derivation for MHD third moments wherein equation (1.2) is obtained without regard for geometry. This approach makes use of simple identities derived from homogeneity and the ability to move derivatives through ensemble averages.

Equations (1.1)–(1.3) require an ensemble average in their derivation. This permits the movement of derivatives from individual terms to the product of terms leading to cancellation and simplification supported by the assumption of homogeneity. As a practical matter, the average over the ensemble is normally taken to be an average over the data using as much data as is available. In order to facilitate error analysis, we normally take 1 h of data to represent a single sample within the ensemble and average over many hours to produce an ensemble average consistent with the above equations. However, it has been suggested that much smaller samples may produce averages for which the above equations are valid [51,52]. We will pursue this idea in §4a and following.

MHD also suffers from the same problem of multi-dimensional gradients in the primitive equations that are not readily analysable using single spacecraft. MHD turbulence can also be addressed in the above manner via third-moment expressions, but the derivation and the resulting expressions are somewhat more complicated. We will pursue them below. Direct assessment of the nonlinearity of solar wind fluctuations generally relies upon statistical arguments such as the average plasma properties at different heliocentric distances [8,9,1217,19,21,22,2428] as addressed by transport theory. The use of third-moment structure functions allow you to circumvent the multi-dimensional derivatives, but do require that some assumption regarding the underlying geometry of the turbulence be made. This permits us to replace the multi-dimensional derivatives with derivatives in a single measurable direction and still obtain a quantitative measure of the nonlinear dynamics. While these derivations are statistical in nature, they require only sufficient data to achieve convergence in order to provide a local measure of the nonlinear terms.

In this paper, we will review the use of third-moment expressions to compute estimates for the turbulent cascade and the general agreement of these results with local heating rates. Our recent work has shown that the natural variability of these computed cascade rates can be thought of as a direct measure of intermittency and we will present new results from this effort. Going beyond intermittency, we will show that there generally exists an anti-correlation between the energy cascade of the outward- and inward-propagating components of the solar wind fluctuations that is a fundamental aspect of the turbulence and we will attempt to explain this anti-correlation in terms of the large-scale energy-containing fluctuations. The reader is encouraged to read several particularly relevant papers in this same issue [43,5357].

2. Heating

The in situ heating of the solar wind is now well established [5862]. Turbulence, both as a source and a means, is cited from the acceleration [53,63] to the termination region [8,9,1217,19,21,22,2428]. If we consider the turbulent cascade as a mechanism for transporting energy from the reservoir of the large-scale, energy-containing range to the small scales where dissipation can occur, we will need a rate. Kolmogorov argued that the central dynamic in the transport of energy within the N-S equation is the interaction of two circulating eddies [34]. He asserts that the lifetime of an eddy will be the time it takes to revolve one time (the ‘turnover time’ L/vk, where vk is the characteristic speed of the eddy and L is the size). This sets the rate for energy transport to be the energy contained within the eddy divided by its lifetime, Inline graphic. Dimensional analysis produces a prediction for the energy spectrum in the inertial range (k−5/3 where k is the wavenumber). The resulting prediction for the power spectrum is Pk=CKϵ2/3k−5/3, where CK is determined by experiment to be CK≃1.6 [64]. We will refer to this theory as K41b.

Iroshnikov [65] and Kraichnan [66] (IK) argue that it is the properties of propagating waves that define the interaction time in MHD turbulence so that they predict a different form for the power spectrum (Pk=AIK(ϵVA)1/2k−3/2 where VA is the Alfvén speed) and a different rate of energy cascade in the process. Numerous other predictions have followed [8,6770]. It is important to realize that the spectrum of magnetic and velocity fluctuations in the solar wind do not agree with either the K41b or IK prediction.

The more modern view of the K41b versus IK controversy is based upon the orientation of the wave vectors called ‘geometry’ or ‘spectral anisotropy’ [57]. If the wave vectors are quasi-parallel to the background mean magnetic field, it may be that the IK formalism based on wave dynamics has merit. If the wave vectors are quasi-perpendicular to the mean magnetic field, the wave frequencies for incompressible fluctuations approach zero and the K41b formalism is likely to be more appropriate. Recent results suggest that the energy-containing scales, or at least the correlation scale, is dominated by wave vectors aligned with the mean magnetic field [71]. Earlier studies at still greater spatial scales deep within the energy-containing range suggest wave vectors that are radially aligned and not related to the mean field direction [72]. There is strong evidence that inertial-range fluctuations are dominated by perpendicular wave vectors [39,42,7377]. This includes observations associated with solar energetic particle (SEP) events [74]. Theory and simulation support this claim [7881] and several theories of inertial range dynamics build on the idea [55,68,82]. We should note that when the analysis of Bieber et al. [74] is applied to the observations of Belcher & Davis [3] the conclusion is that those fluctuations are also dominated by an abundance of perpendicular wave vectors.

The apparent dominance of perpendicular wave vectors within the inertial range may change in the dissipation scales [39,42,76]. Some authors argue that it does not [8389] and argue that kinetic Alfvén waves provide the observed steepening, but some of these observations have been brought into question [90]. Dispersion is sometimes invoked as an explanation for the steepening of the power spectrum at ion dissipation scales [55,9195] and whistler waves are also discussed in the context of forming the dissipation range at ion scales [92,9698]. The central issue to that question rests with the dissipation processes that remove energy from the turbulence and produce heat. What processes are available depends to a great degree on the underlying geometry, or spectral anisotropy, at the smallest scales of the inertial range. The resultant change in the spectral anisotropy at dissipation scales is a diagnostic of the dissipation processes.

One demonstration of the role of turbulence in the solar wind is the ability of turbulence transport theory to reproduce the solar wind proton temperature as a function of heliocentric distance and time [8,9,1217,19,21,22,2428]. Most published applications of this theory have used K41a scaling laws under the tacit assumption that the turbulence is dominated by two-dimensional dynamics, but one analysis used IK theory and reproduced the observations about as well [25]. There is a degree of ambiguity in the theory and this points to the usefulness of a more exact and demanding analysis that may eventually resolve the dilemma.

Turbulence transport theory describes the cascade of energy and assumes that however dissipation occurs it will adapt to the changing cascade rate [41,48,99]. It does not specify the means by which dissipation occurs. In HD, the dissipation scale resides within the fluid approximation. When the energy cascade is greater dissipation moves to smaller scales where the dissipation term ν2v is stronger [48]. This does not happen in the solar wind [41]. MHD is a fluid approximation to plasma dynamics. Dissipation sets in at the ion inertial scale [43,54,55,57,100], where the fluid approximation breaks down and kinetic physics takes over. However, there is new evidence to suggest that the spectral break may occur at the proton Larmor radius [101,102] under some conditions. Either way, this results in a spectrum at smaller scales that steepens with increased cascade rate [41] which does not happen in HD.

While transport theory allows us to reproduce the temperatures observed by Helios, Ulysses, Voyager 1 & 2 and Pioneer 10 & 11, long-term measurements at 1 AU permit us to obtain a more exacting comparison. Using the previously published analyses of Helios data, together with Advanced Composition Explorer (ACE) data, Vasquez et al. [61] obtained an expression describing proton heating at 1 AU that varies with wind speed and observed temperature. More recently, these results have been extended to include high latitudes [103,104] and non-zero cross helicity (called ‘imbalanced turbulence’ by some) [62].

3. Third moment theory

The incompressible MHD equations can be written as

3. 3.1

where Z±v±b are the Elsässer variables [105] and we define the magnetic field B in units of the Alfvén speed, Inline graphic. As noted below, equation (1.1) the multidimensional gradients in equation (3.1) prevent direct evaluation of the nonlinear terms when only a single spacecraft is used. The use of multiple spacecraft also poses limitations.

The Elsässer variables Z± are general representations of MHD dynamics. However, low-frequency waves can also be conveniently described in these terms so that Z (Z+) corresponds to waves propagating parallel (anti-parallel) to the background mean magnetic field. This is often the language of solar wind observations, so we will adopt it as a convenient way to distinguish the two components of the turbulence. However, the reader should note that wave vectors perpendicular to the mean magnetic field are a common point of discussion in solar wind turbulence. Any energy associated with these wave vectors can still be described in terms of Z± even though some of the associated wave forms are non-propagating. Unless otherwise stated, we do not mean to imply a turbulence theory based on wave interactions alone. We simply use the language as a convenient way of connecting with the observations that whichever of Z± that is associated with ‘outward’, or anti-Sunward, propagation is generally seen to be energetically dominant.

In the analysis that follows, we employ the term ‘pseudo-energy’ to represent the quantities |Z±|2 so that the total energy is Inline graphic. The energy cascade associated with each pseudo-energy can be computed using these techniques and compared.

If we apply the same considerations used to derive equation (1.2), we can derive an expression governing the cascade of energy in incompressible MHD [49,50]

3. 3.2

where ϵ± refers to the energy cascade of the pseudo-energies |Z±|2. The expression can be evaluated under various assumed geometries, but the simplest is isotropic which is not especially relevant to the solar wind. However, other attempts at anisotropic geometries we have tried have given similar results. If we assume isotropy, invoke the Taylor frozen-in-flow assumption [106], and employ separation along the solar wind flow, equation (3.2) can be written as

3. 3.3

where Vsw is the solar wind speed, R denotes the radial component away from the Sun, L=−Vswτ and τ is the time separation in the measured data. The total energy cascade is given by ϵT=(ϵ++ϵ)/2.

The fluctuations in the solar wind are typically indicative of dominance by the outwardly propagating component. This diagnostic is accomplished via the correlation of the velocity and magnetic field fluctuations. One quantification of this is the cross-helicity [107110] defined to be Inline graphic, where v and b are the fluctuations of the velocity and magnetic field relative to their means, and N is the number of measurements in the sample. Likewise, the bulk energy of the fluctuations can be defined as Inline graphic. A measure of the significance of the cross-helicity correlation is given by the normalized cross helicity −1≤σC≤1, where σCHC/E. The cascade of cross helicity can also be expressed in terms of equation (3.3) as ϵC≡(ϵ+ϵ)/2.

It is important to stress that the third-moment formalism described above does not rely on a single view of turbulence, a prescription of wave modes or a selection of a specific dynamic. It is entirely general subject to the assumptions listed. It can as accurately describe K41a dynamics as it does IK turbulence, or other theories [8,6770]. This makes it an objective metric against which diverse theories can be tested.

We have tested the third-moment expressions against the thermal proton heating results [61] under the assumption that whatever energy cascades to small scales is dissipated primarily by heating protons [23,30]. The comparison is strongly favourable in both solar maximum and solar minimum. At the present time, 20–40% of the energy cascade is available to heat heavy ions and electrons when the cascade is strong (wind speed and proton temperature are both high) and less when the cascade is weak [23], which is in agreement with empirical constraints [111].

A simple scaling law can be inferred from equation (3.3) that resembles K41a. Where the HD cascade scales as Inline graphic, MHD exhibits a different dynamic wherein the Z+ field cascades the Z field and vice versa. This leads to a simple scaling of equation (3.3) taking the form [112]

3. 3.4

and

3. 3.5

See related discussions [8,9,113]. We will apply that scaling below.

4. Analysis

In this study, we use magnetic field, velocity and proton density data from the Advanced Composition Explorer (ACE) spacecraft [114116]. Magnetic field data have been averaged to the same 64 s cadence as the thermal proton data. We use 12 years of measurements from early 1998 through mid-2010. Interplanetary shocks and their driver gases have been removed.

The nominal correlation length for fluctuations in the solar wind is 1 h. Therefore, the statistical means and uncertainties derived from many samples when each is 1 h in duration or longer is determined by simple Gaussian statistics. In this study, we use 1 h samples of 64 s resolution data to compute individual realizations from which statistical quantities are obtained. Measurement errors are insignificant in comparison to statistical variations and are not considered here. In order to obtain meaningful estimates for each 1 h sample, we require that greater than 70% of the data have usable magnetic field, velocity and density measurements. We further require, as an estimate of stationarity within the sample, that the mean magnetic field and average wind speed of the first and last half of every hour of data differ from the average for the hour by no more than 10%. This further removes transients, strong gradients including shear, etc., and thereby imposes homogeneity upon each sample. This avoids the complications associated with third-moment expressions that are applicable to shear regions [117119]. It also fails to address questions of in situ generation of turbulence in these regions [120122]. As with our earlier papers, we interpolate the results of each 1 h sample onto a common spatial grid that is equivalent to measurements with a 400 km s−1 wind speed. This is done under the assumption that the in situ dynamics are defined by spatial scale rather than the associated Doppler shifted signal that is dependent on the wind speed. This analysis is based upon data selection for the average energy and cross helicity of a 1 h interval. We will avoid examples of large |σC| for reasons described in [123].

We compute third-order structure functions as defined in equation (3.3) and recognize the strong tendency for outward propagating fluctuations to possess greater energy than inward propagating fluctuations as determined from the Z±. Using the computed mean field for the sample, we associate Z± with the inward- and outward-propagating components Zin,out and gather our statistics for Inline graphic accordingly. This permits us to write third-moment forms equivalent to equation (3.3):

4. 4.1
4. 4.2
4. 4.3

The total energy cascade ϵT can then be computed from Inline graphic whether we employ Z± or Zin,out.

The data spacing is 64 s and the data products derived from equation (4.1)–(4.3) are averages of lagged products. For simplicity, we refer to lags not by the time spacing, but by the data spacing. Therefore, a calculation based on lag=5 means using measurements that are five data points (5×64=320 s) apart.

(a). Intermittency

In most of our past publications on this subject [18,20,23,29,30,119,124,125], we have used relatively long data intervals (typically about 12 h in duration) and many such samples to compute average properties of the third moment and implied cascade. The exception to this is a study performed on 3 h samples [126] that revealed the underlying dynamics of intermittency. Here, we repeat that analysis using a different energy range for selection of our intervals and the shorter 1 h sample size.

We limit the average bulk energy of the fluctuations to 300<E<700 km2 s−2 and the average normalized cross helicity to |σC|≤0.75. The results that we show here have been seen in every energy and helicity range we have examined to date. This selection results in a sufficient number of samples to produce a clean distribution function and we choose these conditions for this reason only. It does eliminate times of high |σC|, which may prove interesting in themselves, but this is where we have chosen to begin this investigation. There are 7075 1 h samples analysed in this study. Functions Inline graphic, Inline graphic and Inline graphic are computed for each 1 h sample. We perform a linear fit to the functions Inline graphic over lags 2–20 (128–1280 s separation) and evaluate the goodness of the fit according to Inline graphic where xi is the time into the hourly sample (64, 128, 192 s, etc.), and yi is the associated value of D3. Figure 1 shows the resulting distribution of R2 values. A total of 4787 samples (approx. 68%) had R2≥0.7. We consider these 1 h samples to be well-fit by a line passing through the origin as is required of equation (4.3).

Figure 1.

Figure 1.

Analysis of 1 h samples using 12 year of ACE MAG and SWEPAM data showing the distribution of R2, the square of the correlation coefficient for Inline graphic versus L, for lags 2–20 corresponding to 128–1280 s. See equations (4.1)–(4.3) for definition of Inline graphic. Analysis limited to mean fluctuation energy 300<E<700 km2s−2 and mean normalized cross helicity |σC|≤0.75.

At the heart of this analysis is the question, ‘What is meant by linear third-moment expressions when applied to a correlation length of data?’ The equations describing the cascade are derived for an ensemble average of many such realizations. It has been suggested that the scaling properties of incremental measurements for the HD cascade show better convergence for smaller data sets, and there is the suggestion that these expressions are applicable to much smaller samples of data than the lengthy ensembles used here [51,52]. A possible explanation for this is that averaging of small-scale fluctuations within a given realization (a sample that we take to have 1 h duration) possesses itself a degree of averaging that accomplishes some of the theoretical goals of the derivation. The implication is that when these third-moment expressions are applied to small samples the result is an estimate for the local cascade rather than the global ensemble average cascade.

In the remainder of this paper, we consider only those 1 h samples that pass the R2≥0.7 test of Inline graphic. Figure 2 shows the distributions of ϵ as computed from Inline graphic, Inline graphic and Inline graphic at a lag of five samples (=320 s). We discuss below why we select this lag value for our statistics. Vertical solid red lines represent the mean of each distribution. Each mean is positive representing an average cascade from large to small scales. Vertical dashed red lines give the computed standard deviation. While the distribution functions are nearly Gaussian in form, they have ‘hot tails’ where the distributions are consistently overpopulated relative to the Gaussian fits for values in excess of the standard deviation.

Figure 2.

Figure 2.

Distribution of ϵin as defined in equation (4.1) and computed from Inline graphic for lag 5 (320 s) using 1 h samples. Same for ϵout, ϵT and ϵC. Blue curve is least-squares best fit of Gaussian to the distribution functions. Vertical solid (dashed) lines are the means (standard deviations) of the data. While the best-fit Gaussian does reproduce the centre of the distribution functions, there is a strong abundance of points on the tails of the distributions making the standard deviation of the distributions significantly larger than the standard deviation of the best-fit Gaussians. (Online version in colour.)

While it is difficult to convey from figure 2, the mean of each distribution is positive as is generally expected for the average energy cascade rate. The standard deviations are large when compared with the averages. Both means and standard deviations for each of the four panels are listed in table 1.

Table 1.

Means and standard deviations for figure 2.

mean s.d.
PDF(ϵin) 4.08×103 1.29×105
PDF(ϵout) 1.56×104 1.85×105
PDF(ϵT) 9.83×103 6.31×104
PDF(ϵC) 5.75×103 1.47×105

A fundamental property of the third-moment expressions for ϵ (equations (3.3) and (4.1)–(4.3)) is that the functions D3 are linear in lag. We contend that linearity of the observed hourly estimates is not a coincidence, but results from well-determined functions and a valid assessment of ϵ which should be independent of scale in the inertial range. Two properties of the resulting estimates for ϵin, ϵout and ϵT are (i) that the standard deviation of each distribution function is significantly larger than the mean and (ii) negative values of ϵ are readily obtained. This points to two important aspects of turbulence in general and solar wind turbulence in particular. First, the predictions of K41a, IK and others address only the mean energy cascade rate for the ensemble. Individual realizations as represented by the statistically independent hourly samples we use here may bear little or no resemblance to the mean value of the distribution. Second, negative cascade rates are possible for individual hourly samples even if it seems impossible that they could persist indefinitely.

If we average the individual estimates for D3 derived from the 4787 1 h samples with R2≥0.7, we can obtain the underlying forms of these functions relating to the means of the above distributions. Figure 3 shows the resulting average third-moment functions Inline graphic, Inline graphic and Inline graphic. The greater slope of Inline graphic is consistent with the larger average value of ϵout when compared with Inline graphic and ϵin. Linearity is far from perfect but is within the computed uncertainty of the estimates shown in the figure.

Figure 3.

Figure 3.

Computed average functions Inline graphic, Inline graphic and Inline graphic for the 4787 1 h samples used in this study with R2≥0.7. Third moments are plotted as functions of increments in spatial lag L which is equivalent to a temporal lag τ if the wind speed is 400 km s−1. This represents the fixed grid upon which we interpolate each result to allow for varying wind speed. Example uncertainties as represented by the error-of-the-mean are shown on the figure. (Online version in colour.)

We contend that the large standard deviation relative to the mean of each distribution function for ϵ is an indication of intermittency. Intermittency has come to mean many things and is most often analysed as a ratio of higher order structure functions representing the filling factor, or natural uniformity, of the turbulence. A complimentary and perhaps more fundamental definition of intermittency is the statistical variation of the turbulent cascade relative to predictions for the mean. In this way, third moments provide a direct measure of the variability of the cascade that we can reasonably presume to vary from one sample to another so long as those samples are separated by a correlation length. It is a measure of the variability of the local dynamics of the turbulence. Here, we say nothing of correlation times. Using the Taylor frozen-in-flow assumption and taking all samples from 1 AU, we have no measure of the duration over which a sample shows the computed cascade dynamics. We assume that back transfer of energy from small to large scales cannot persist unless energy flows from one component to another (exchange of pseudo-energy within the cascade) and we have no reliable measure of this possible dynamic at this time. We can only claim that at a time and a point in space the energy cascade of one of the pseudo-energies may be negative in some instances.

(b). Pseudo-energy scalings

The cross helicity can be rewritten as

(b). 4.4

but solar wind studies are better organized using Zin,out. Therefore, we redefine a modified form of cross helicity according to

(b). 4.5

which is equivalent to σCr=−(BR/|BR|)σC so that σCr>0 represents a dominance of outward-propagating fluctuations. We hereafter drop the superscript ‘r’ and adopt equation (4.5) as our definition of the normalized cross helicity.

Comparison of the distribution functions in figure 2 reveals that the standard deviation for ϵT is smaller than for ϵin,out. This seems unlikely if the cascades of the two pseudo-energies are independent and uncorrelated. Figure 4 shows scatter plots using the same energy range as in figure 2 for selected ranges of σC. Three different ranges of |σC| are shown. Top to bottom they are 0≤|σC|<0.25, 0.25≤|σC|<0.5 and 0.5≤|σC|<0.75.

Figure 4.

Figure 4.

Scatter plots for ϵin,out for three ranges of |σC| given above the panels. In each plot, 1 h samples are used to produce estimates of ϵ at lag 5 (320 s). σC>0 are represented by black diamonds and samples with σC<0 are represented by blue triangles. As expected, there are fewer samples with σC<0 than with σC>0. Dashed red lines represent ϵin=ϵout and ϵin=−ϵout. Black and blue lines are derived from equation (4.8) using the average value of σC for each population with the additional factor −1 to account for the observed anti-coincidence. We have seen similar results when using 3 h and 12 h samples. (Online version in colour.)

Two clear conclusions can be drawn from figure 4. First, each of the six populations shown in figure 4 appears to scale linearly in ϵin versus ϵout. When |σC| is small the linear trend of the two populations appears to nearly overlay one another. As |σC| increases the linear trend of the two populations appears to separate, or fan, so that the σC>0 samples (intervals dominated by outward propagation) show a more shallow slope than the σC<0 samples (intervals dominated by inward propagation).

Second, there is an anti-correlation between ϵin and ϵout: when one is positive the other is negative. Nothing from the cascade laws we have discussed so far explains this.

Let us first address the observation that the slopes of ϵin versus ϵout depend on σC. If we adopt equations (3.4) and (3.5) as expressions for the cascade of the pseudo-energies Ein,out, we can rewrite them as

(b). 4.6

and

(b). 4.7

and from this we can write

(b). 4.8

As an example, if σC=0 so that Eout=Ein, equation (4.8) predicts that ϵout/ϵin=1. When σC=0.5 (−0.5) equation (4.8) predicts that ϵout/ϵin=31/2 (3−1/2). Using the average value of σC for each of the six subsets of σC shown in figure 4, we can apply equation (4.8). In the process, we adopt a factor of −1 in equation (4.8) for reasons not yet established. The resulting lines are reproduced on the figure.

The existence of negative values of ϵ in figure 4 would seem to preclude the systematic applicability of equations (4.6)–(4.8). However, this does not mean these expressions fail to describe ϵ when averaged over the entire ensemble. For reasons not yet completely determined, equations (4.6)–(4.7) do reliably describe the ratio ϵout/ϵin for each hourly interval if we adopt an additional factor of −1.

(c). Back transfer

We now move to examine the second conclusion from figure 4: the observation that ϵin and ϵout are anti-correlated. This will account for the factor −1 above.

Consider the third-moment expressions equations (4.1) and (4.2). Both contain terms like |Zin,out|2 that vary in magnitude, but not sign. It is the terms Zin,outR that vary in sign and potentially facilitate the back transfer of energy. Comparison of ϵin/ϵout leads us to consider Inline graphic or its inverse, depending on the direction of the mean magnetic field. If |ΔvR(L)|>|ΔbR(L)| then the ratio ϵin/ϵout>0. However, if |ΔvR(L)|<|ΔbR(L)| then the ratio ϵin/ϵout<0 and this provides some possible insight into the back transfer of one of the pseudo-energies. This suggests that the back transfer of pseudo-energy may be related to the Alfvén ratio RA=Ev/Eb<1 and this is something we can test.

We compare the presence of a back transfer for one or another pseudo-energies with RA computed from the fluctuations on the largest scales in the hourly samples (a bulk r.m.s. calculation). We justify using RA computed from the entirety of a 1 h sample rather than at smaller scales because (i) we are not certain that the critical dynamic resides at the smaller scales and (ii) RA tends to be approximately preserved across the inertial range scales until the very smallest scales [127]. To determine the presence of back transfer, we compute the quantity Rϵ≡|(ϵoutϵin)/(ϵout+ϵin)|. Rϵ<1 requires that both ϵin and ϵout possess the same sign while Rϵ>1 requires that ϵin and ϵout possess different signs (the anti-correlation we report above).

Figure 5a plots RA versus Rϵ for all of the 1 h samples used in this study that had R2>0.7. There is a strong tendency for the data to follow the two axes such that Rϵ<2 (Rϵ>2) when RA>1 (RA<1). On average, the solar wind at 1 AU possesses Inline graphic [128130], so anti-correlated pseudo-energy cascades are a common feature in the solar wind.

Figure 5.

Figure 5.

(a) Scatter plot for ratio of kinetic to magnetic energy RAEv/Eb computed from large-scale bulk parameters versus parameter Rϵ quantifying the anti-correlated cascade of pseudo-energy. (b) RA computed from the second-order structure functions at same lag as in figures 2 and 4. In both panels, Rϵ is also computed using same lag as in figures 2 and 4. The vertical dashed lines mark Rϵ=1. There is a clear tendency for Rϵ>1 when Inline graphic in both panels. (c) Scatter plot of RA computed from bulk properties as used in panel (a) and [Δv(τ)]2/[Δb(τ)]2 computed for lag=5. (Online version in colour.)

As we stated above, we are not sure at what scale the critical dynamic resides. We can compute the ratio of magnetic to kinetic energy at the same scale where the back transfer is measured (lag=5) by computing the ratio of the second-order structure functions [Δv(τ)]2/[Δb(τ)]2. Figure 5b shows the result of this analysis compared with Rϵ for the same samples used to produce the panel on the left. The two plots are virtually identical.

Figure 5c shows a scatter plot of the two definitions for RA employed in panels (a,b) of the same figure. There is a strong correlation until RA>0.7 (approximately) when the two definitions diverge, although not systematically. Regrettably, this obscures any potential answer to the question of which spatial scale possesses the dynamic that results in the anti-correlation of the pseudo-energy cascades.

(d). Consistency across the inertial range

So far we have examined only the lag=5 results. We chose this value because there are approximately 12 such lags within each 1 h sample, and the resulting averaging that occurs within the 1 h sample provides a degree of confidence in the result. However, we have argued that the observed back transfer of energy is the result of having low RA values (magnetically dominated Z±). This argument alone does not explain which pseudo-energy back transfers.

There is the question of whether what we have shown here is a consistent dynamic across the inertial range scales or merely a local dynamic that is part of the general ebb and flow of energy that forms the cascade. Figure 6 compares the computed cascade rates at lags of 5 and 15. While agreement is not perfect, due at least in part because the lag=15 values are the result of less averaging than the lag=5 values, the linear trend is obvious.

Figure 6.

Figure 6.

We compare the computed cascade rates for lags 5 and 15 for each of the 1 h samples that passed the R2≥0.7 test. Comparison for ϵin, ϵout, ϵT and ϵC. General agreement between the computed ϵ at lags 5 and 15 suggest that the conclusions of the above analysis hold throughout the inertial range, or at least over a wide range of lags within the inertial range. (Online version in colour.)

The conclusion we draw from this is that the anti-correlation in cascade rates ϵin,out reported here represent behaviour that spans the inertial range scales. There may be unresolved variations and systematics. We cannot say there is not. What we can assert is that the conclusions drawn here appear to represent the entirety, or significant fraction, of the inertial range scales.

Our previous paper [126] showed that sequential values of ϵ differed by values comparable to the mean. This means that while a given 1 h sample may show back transfer of ϵin it is just as likely that the next 1 h sample will show back transfer in ϵout. While this does not require that the cascade change its character on timescales comparable to the correlation time (in the plasma frame), it is suggestive of it. Observations reported here are probably just snapshots in time that change with the evolving dynamics.

(e). Variation of cascade on the correlation scale

We have shown that the apparent rate of energy cascade within 1 h samples of solar wind turbulence at 1 AU exhibit large variability (standard deviation) relative to the mean ensemble values. We have further shown that the cascade associated with pseudo-energies Z±, or Zout and Zin, are anti-correlated and exhibit negative cascade rates implying transport of energy from small to large spatial scales. We would like to measure the timescale over which this behaviour changes because a persistent transport of energy from small to large spatial scales would seem impossible to maintain within the normal view of homogeneous turbulence unless there is an unresolved small-scale source driving the cascade. This is not possible using single-spacecraft techniques. However, we can perform an analysis of the variability of the cascade from one 1 h sample to the next and apply the normal turbulence assumption that small-scale structures last for shorter periods of time than large-scale structures.

We compute the difference between cascade rates from one 1 h sample to the next. Where data are missing or samples failed to pass our selection criteria, we omit pairings and use only available, sequential 1 h samples. We define Δ1 hϵ(t)≡ϵ(t+1 h)−ϵ(t) and compute the distribution functions for ϵin, ϵout and ϵT. These are shown in figure 7. The width of the distributions are comparable to those of the ϵ values, themselves, suggesting that there is little correlation between the ϵ values computed for successive 1 h samples. While this does not provide a timescale over which the cascade rate changes for a given sample of plasma, it does suggest that timescale is shorter than it would be if sequential 1 h samples were strongly correlated.

Figure 7.

Figure 7.

(ac) Distribution functions for the difference between sequential 1 h sample values of ϵin, ϵout and ϵT. Note the width of the distributions are comparable to the width of the distributions for the ϵ values, themselves, suggesting that there is little correlation between sequential hourly values. (d) Scatter plot of Δ1 hϵin and Δ1 hϵout showing the persistent anti-correlation of these values. (Online version in colour.)

Figure 7 also shows a scatter plot of Δ1 hϵin and Δ1 hϵout. It comes as no surprise that these 1 h differences are anti-correlated when the values of ϵ are themselves anti-correlated. To maintain that anti-correlation, they must change sign in unison.

5. Summary and conclusion

Third-moment expressions have now been used by numerous authors in simulation and data analysis [18,20,23,29,30,103,117119,123126,131138]. The technique is not without controversy. The role of expansion is not well quantified [117,139,140] and the techniques for addressing large-scale shear are difficult to apply [117,119]. However, the technique appears to possess considerable potential for unlocking the dynamics of interplanetary turbulence and enabling the solar wind to fullfil its potential as a very large wind tunnel for the study of MHD turbulence.

In this analysis, we have excluded transients, such as shocks and their drivers, and regions of shear. We have not taken into account expansion. Simply put, the published expressions for expansion effects do not agree and our efforts to derive expansion terms have not produced results that we are prepared to support. We do not believe that expansion effects can account for the broad distribution and negative values of ϵ in any general sense or the associated back transfer of energy shown here.

We have shown that 1 h samples of the solar wind taken at 1 AU can be used to produce third-moment estimates for the inertial range cascade that possess linear scaling with lag and appear to offer reasonable estimates for the cascade of the pseudo-energies. As before [20,23,30,126], the average over many samples produces a net forward transfer of both pseudo-energies consistent with the observed solar wind heating. The estimates derived from individual 1 h samples show evidence of intermittent dynamics with a high degree of variability relative to the mean cascade rates. We do not believe the reported variability is a measurement error. We do believe it to be a fundamental measurement of the underlying intermittency of the turbulence.

To that end, we have argued that the transfer of pseudo-energies associated with the inward- and outward-propagating components are often, and perhaps generally, anti-correlated at 1 AU. Where one is seen to forward transfer to small scales the other is seen to back transfer energy to larger scales. Which component back transfers energy is not determined by this analysis and it can be either. We believe the best explanation lies with the dominance of the dynamics by the magnetic term as measured by the Alfvén ratio RA<1. This provides possible insights into the third-moment expressions that facilitates anti-correlation between the two cascades.

Whenever one discusses back transfer or the preferential transfer of one pseudo-energy over another, several terms are often raised. The first term is ‘inverse cascade’ [141,142]. Inverse cascade is generally taken to mean the back transfer of one quantity while another quantity forward-transfers to smaller scales. It is often described according to absolute equilibrium ensemble theory [35], and it represents a systematic movement of two quantities to separate spatial scales. The second term is ‘dynamic alignment’ [143]. Dynamic alignment has been proposed as a mechanism for producing strong correlations between the magnetic field and velocity fluctuations in the solar wind [113,144]. The third term is ‘selective decay’ [145,146]. All three processes represent a systematic transport, accumulation and dissipation of one statistical quantity over another. The anti-correlated spectral transfer we show here shows no such systematic transport. It is a random exchange between the forward- and back transfer of two pseudo-energies without any apparent systematic determination evident.

The correlation length for spacecraft data at 1 AU in the solar wind is approximately 1 h. This is roughly the scale that undergoes one turbulent lifetime in the time it takes for the wind to reach 1 AU from its solar source. The energy contained within the inertial range is roughly equal to the accumulation of energy at the average cascade rate over the time it takes to reach 1 AU. So, the inertial range has been fully depleted approximately one time by the time it reaches 1 AU. The scales used here, 320 and 960 s, represent the first decade within the measured inertial range. The dissipation scale, where the spectrum breaks and steepens and dissipation sets in, is seen at 2–5 s. This means there are two more decades of inertial range that remain unstudied by this analysis. It also means that if the back transfer reported here varies between pseudo-energies on the plasma-frame timescale of the turbulent fluctuations at 320 and 960 s, then it is possible for the back transfer to draw energy from the inertial range without depleting it. We do not propose that the back transfer reported here persists over long periods of time. We have already shown that individual cascade rates change substantially from 1 h of data to the next [126]. Back transfers of energy should be possible so long as they do not persist for long periods of time.

Acknowledgements

The authors thank the ACE/SWEPAM team for providing the thermal proton data used in this study. J.T.C. has recently obtained his B.Sc. degree in physics at the University of New Hampshire. J.E.S. is a graduate student at the University of Colorado at Boulder.

Funding statement

J.T.C. and C.W.S. are supported by Caltech subcontract 44A-1062037 to the University of New Hampshire in support of the ACE/MAG instrument. B.J.V. is supported by NSF/SHINE grant no. AGS1357893.

References

  • 1.Coleman PJ., Jr 1966. Hydromagnetic waves in the interplanetary plasma. Phys. Rev. Lett. 17, 207–211. ( 10.1103/PhysRevLett.17.207) [DOI] [Google Scholar]
  • 2.Coleman PJ., Jr 1968. Turbulence, viscosity, and dissipation in the solar wind plasma. Astrophys. J. 153, 371–388. ( 10.1086/149674) [DOI] [Google Scholar]
  • 3.Belcher JW, Davis L., Jr 1971. Large-amplitude Alfvén waves in the interplanetary medium, 2. J. Geophys. Res. 76, 3534–3563. ( 10.1029/JA076i016p03534) [DOI] [Google Scholar]
  • 4.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. ( 10.1007/s11214-011-9793-9) [DOI] [Google Scholar]
  • 5.Matthaeus WH, Lamkin SL. 1986. Turbulent magnetic reconnection. Phys. Fluids 29, 2513–2534. ( 10.1063/1.866004) [DOI] [Google Scholar]
  • 6.Lazarian A, Eyink G, Vishniac E, Kowal G. 2015. Turbulent reconnection and its implications. Phil. Trans. R. Soc. A 373, 20140144 ( 10.1098/rsta.2014.0144) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Schwartz SJ, Feldman WC, Gary SP. 1981. The source of proton anisotropy in the high-speed solar wind. J. Geophys. Res. 86, 541–546. ( 10.1029/JA086iA02p00541) [DOI] [Google Scholar]
  • 8.Zhou Y, Matthaeus WH. 1990. Models of inertial range spectra of interplanetary magnetohydrodynamic turbulence. J. Geophys. Res. 95, 14 881–14 892. ( 10.1029/JA095iA09p14881) [DOI] [Google Scholar]
  • 9.Matthaeus WH, Oughton S, Pontius DH, Jr, Zhou Y. 1994. Evolution of energy containing turbulent eddies in the solar wind. J. Geophys. Res. 99, 19 267–19 287. ( 10.1029/94JA01233) [DOI] [Google Scholar]
  • 10.Verma MK, Roberts DA, Goldstein ML. 1995. Turbulent heating and temperature evolution in the solar wind. J. Geophys. Res. 100, 19 839–19 850. ( 10.1029/95JA01216) [DOI] [Google Scholar]
  • 11.Tu CY, Marsch E. 1995. MHD structures, waves and turbulence in the solar wind. Norwell, MA: Kluwer Academic Publishers; (Reprinted from Space Sci. Rev., 73(1–2), 1995.) [Google Scholar]
  • 12.Zank GP, Matthaeus WH, Smith CW. 1996. Evolution of turbulent magnetic fluctuation power with heliospheric distance. J. Geophys. Res. 101, 17 093–17 107. ( 10.1029/96JA01275) [DOI] [Google Scholar]
  • 13.Matthaeus WH, Zank GP, Smith CW, Oughton S. 1999. Turbulence, spatial transport, and heating of the solar wind. Phys. Rev. Lett. 82, 3444–3447. ( 10.1103/PhysRevLett.82.3444) [DOI] [Google Scholar]
  • 14.Smith CW, Matthaeus WH, Zank GP, Ness NF, Oughton S, Richardson JD. 2001. Heating of the low-latitude solar wind by dissipation of turbulent magnetic fluctuations. J. Geophys. Res. 106, 8253–8272. ( 10.1029/2000JA000366) [DOI] [Google Scholar]
  • 15.Isenberg PA, Smith CW, Matthaeus WH. 2003. Turbulent heating of the distant solar wind by interstellar pickup protons. Astrophys. J. 592, 564–573. ( 10.1086/375584) [DOI] [Google Scholar]
  • 16.Isenberg PA. 2005. Turbulence-driven solar wind heating and energization of pickup protons in the outer heliosphere. Astrophys. J. 623, 502–510. ( 10.1086/428609) [DOI] [Google Scholar]
  • 17.Breech B, Matthaeus WH, Minnie J, Oughton S, Parhi S, Bieber JW, Bavassano B. 2005. Radial evolution of cross helicity in high latitude solar wind. Geophys. Res. Lett. 32, L06103 ( 10.1029/2004GL022321) [DOI] [Google Scholar]
  • 18.MacBride BT, Forman MA, Smith CW. 2005. Turbulence and the third moment of fluctuations: Kolmogorov's 4/5 law and its MHD analogues in the solar wind. In Solar Wind 11 (eds Fleck B, Zurbuchen TH.), ESA SP-592, pp. 613–616. The Netherlands: European Space Agency. [Google Scholar]
  • 19.Smith CW, Isenberg PA, Matthaeus WH, Richardson JD. 2006. Turbulent heating of the solar wind by newborn interstellar pickup protons. Astrophys. J. 638, 508–517. ( 10.1086/498671) [DOI] [Google Scholar]
  • 20.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, 1644–1660. ( 10.1086/529575) [DOI] [Google Scholar]
  • 21.Breech B, Matthaeus WH, Minnie J, Bieber JW, Oughton S, Smith CW, Isenberg PA. 2008. Turbulence transport throughout the heliosphere. J. Geophys. Res. 113, A08105 ( 10.1029/2007JA012711) [DOI] [Google Scholar]
  • 22.Breech B, Matthaeus WH, Cranmer SR, Kasper JC, Oughton S. 2009. Electron and proton heating by solar wind turbulence. J. Geophys. Res. 114, A09103 ( 10.1029/2009JA014354) [DOI] [Google Scholar]
  • 23.Stawarz JE, Smith CW, Vasquez BJ, Forman MA, MacBride BT. 2009. The turbulent cascade and proton heating in the solar wind at 1 AU. Astrophys. J. 697, 1119–1127. ( 10.1088/0004-637X/697/2/1119) [DOI] [Google Scholar]
  • 24.Isenberg PA, Smith CW, Matthaeus WH, Richardson JD. 2010. Turbulent heating of the distant solar wind by interstellar pickup protons in a decelerating flow. Astrophys. J. 719, 716–721. ( 10.1088/0004-637X/719/1/716) [DOI] [Google Scholar]
  • 25.Ng CS, Bhattacharjee A, Munsi D, Isenberg PA, Smith CW. 2010. Kolmogorov versus Iroshnikov–Kraichnan spectra: consequences for ion heating in the solar wind. J. Geophys. Res. 115, A02101 ( 10.1029/2009JA014377) [DOI] [Google Scholar]
  • 26.Breech B, Cranmer SR, Matthaeus WH, Kasper JC, Oughton S. 2010. Heating of the solar wind with electron and proton effects. Solar Wind 12, 214–217. [Google Scholar]
  • 27.Oughton S, Matthaeus WH, Smith CW, Breech B, Isenberg PA. 2011. Transport of solar wind fluctuations: a two-component model. J. Geophys. Res. 116, A08105 ( 10.1029/2010JA016365) [DOI] [Google Scholar]
  • 28.Zank GP, Dosch A, Hunana P, Florinski V, Matthaeus WH, Webb GM. 2012. The transport of low-frequency turbulence in astrophysical flows. I. Governing equations. Astrophys. J. 745, 35 ( 10.1088/0004-637X/745/1/35) [DOI] [Google Scholar]
  • 29.Coburn JT, Smith CW, Vasquez BJ, Stawarz JE, Forman MA. 2012. The turbulent cascade and proton heating in the solar wind during solar minimum. Astrophys. J. 754, 93 ( 10.1088/0004-637X/754/2/93) [DOI] [Google Scholar]
  • 30.Coburn JT, Smith CW, Vasquez BJ, Stawarz JE, Forman MA. 2013. The turbulent cascade and proton heating in the solar wind during solar minimum. In Solar wind 13 (eds Zank GP, et al.). AIP Conference Proceedings, vol. 1539, pp. 147–150. Melville, NY: AIP. [Google Scholar]
  • 31.Bruno R, Carbone V. 2013. The solar wind as a turbulence laboratory. Living Rev. Solar Phys. 10, 2 ( 10.12942/lrsp-2013-2) [DOI] [Google Scholar]
  • 32.Benzi R, Frisch U. 2010. Turbulence. Scholarpedia 5, 3439 ( 10.4249/scholarpedia.3439) [DOI] [Google Scholar]
  • 33.von Karman T, Howarth J. 1938. On the statistical theory of isotropic turbulence. Proc. R. Soc. Lond. A 164, 192–215. ( 10.1098/rspa.1938.0013) [DOI] [Google Scholar]
  • 34.Kolmogorov AN. 1941. Dissipation of energy in locally isotropic turbulence. Dokl. Akad. Nauk SSSR 32, 16–18. (Reprinted in Proc. R. Soc. Lond. A434, 15–17, 1991.) [Google Scholar]
  • 35.Kraichnan RH, Montgomery DC. 1980. Two dimensional turbulence. Rep. Prog. Phys. 43, 547–619. ( 10.1088/0034-4885/43/5/001) [DOI] [Google Scholar]
  • 36.Behannon KW. 1975. Observations of the interplanetary magnetic field between 0.46 and 1 AU by the Mariner 10 spacecraft. PhD thesis, Catholic University of America, Washington, DC, USA. [Google Scholar]
  • 37.Denskat KU, Beinroth HJ, Neubauer FM. 1983. Interplanetary magnetic field power spectra with frequencies from 2.4×10−5 Hz to 470 Hz from Helios-observations during solar minimum conditions. J. Geophys. 54, 60–67. [Google Scholar]
  • 38.Goldstein ML, Roberts DA, Fitch CA. 1994. Properties of the fluctuating magnetic helicity in the inertial and dissipation ranges of solar wind turbulence. J. Geophys. Res. 99, 11 519–11 538. ( 10.1029/94JA00789) [DOI] [Google Scholar]
  • 39.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, 4775–4787. ( 10.1029/97JA03394) [DOI] [Google Scholar]
  • 40.Leamon RJ, Smith CW, Ness NF, Wong HK. 1999. Dissipation range dynamics Kinetic: Alfvén waves and the importance of βe. J. Geophys. Res. 104, 22 331–22 344. ( 10.1029/1999JA900158) [DOI] [Google Scholar]
  • 41.Smith CW, Hamilton K, Vasquez BJ, Leamon RJ. 2006. Dependence of the dissipation range spectrum of interplanetary magnetic fluctuations upon the rate of energy cascade. Astrophys. J. Lett. 645, L85–L88. ( 10.1086/506151) [DOI] [Google Scholar]
  • 42.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 ( 10.1029/2007JA012559) [DOI] [Google Scholar]
  • 43.Riazantseva MO, Budaev VP, Zelenyi LM, Zastenker GN, Pavlos GP, Safrankova J, Nemecek Z, Prech L, Nemec F. 2015. Dynamic properties of small-scale solar wind plasma fluctuations. Phil. Trans. R. Soc. A 373, 20140146 ( 10.1098/rsta.2014.0146) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.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, 165003 ( 10.1103/PhysRevLett.103.165003) [DOI] [PubMed] [Google Scholar]
  • 45.Monin AS. 1959. Theory of locally isotropic turbulence. Dokl. Akad. Nauk. SSR 125, 515–518. [Google Scholar]
  • 46.Monin AS, Yaglom AM. 1975. Statistical fluid mechanics: mechanics of turbulence, vol. 2 Cambridge, MA: MIT Press. [Google Scholar]
  • 47.Frisch U. 1995. Turbulence. New York, NY: Cambridge University Press. [Google Scholar]
  • 48.Kolmogorov AN. 1941. The local structure of turbulence in incompressible viscous fluid for very large Reynolds numbers. Dokl. Akad. Nauk SSSR 30, 301–305. (Reprinted in Proc. R. Soc. London A434, 9–13, 1991.) [Google Scholar]
  • 49.Politano H, Pouquet A. 1998. von Kármán–Howarth equation for magnetohydrodynamics and its consequences on third-order longitudinal structure and correlation functions. Phys. Rev. E 57, R21–R24. ( 10.1103/PhysRevE.57.R21) [DOI] [Google Scholar]
  • 50.Politano H, Pouquet A. 1998. Dynamical length scales for turbulent magnetized flow. Geophys. Res. Lett. 25, 273–276. ( 10.1029/97GL03642) [DOI] [Google Scholar]
  • 51.Mineveau C, Sreenivasan KR. 1991. The multifractal nature of turbulent energy dissipation. J. Fluid Mech. 224, 429–484. ( 10.1017/S0022112091001830) [DOI] [Google Scholar]
  • 52.Stolovitzky G, Kailasnath P, Sreenivasan KR. 1992. Kolmogorov's refined similarity hypotheses. Phys. Rev. Lett. 69, 1178–1181. ( 10.1103/PhysRevLett.69.1178) [DOI] [PubMed] [Google Scholar]
  • 53.Cranmer SR, Asgari-Targhi M, Miralles MP, Raymond JC, Strachan L, Tian H, Woolsey LN. 2015. The role of turbulence in coronal heating and solar wind expansion. Phil. Trans. R. Soc. A 373, 20140148 ( 10.1098/rsta.2014.0148) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Goldstein ML, Wicks RT, Perri S, Sahraoui F. 2015. Kinetic scale turbulence and dissipation in the solar wind: key observational results and future outlook. Phil. Trans. R. Soc. A 373, 20140147 ( 10.1098/rsta.2014.0147) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Howes GG. 2015. A dynamical model of plasma turbulence in the solar wind. Phil. Trans. R. Soc. A 373, 20140145 ( 10.1098/rsta.2014.0145) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Matthaeus WH, Wan M, Servidio S, Greco A, Osman KT, Oughton S, Dmitruk P. 2015. Intermittency, nonlinear dynamics and dissipation in the solar wind and astrophysical plasmas. Phil. Trans. R. Soc. A 373, 20140154 ( 10.1098/rsta.2014.0154) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Oughton S, Matthaeus WH, Wan M, Osman KT. 2015. Anisotropy in solar wind plasma turbulence. Phil. Trans. R. Soc. A 373, 20140152 ( 10.1098/rsta.2014.0152) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Totten TL, Freeman JW, Arya S. 1995. An empirical determination of the polytropic index for the free-streaming solar wind using Helios 1 data. J. Geophys. Res. 100, 13–17. ( 10.1029/94JA02420) [DOI] [Google Scholar]
  • 59.Richardson JD, Paularena KI, Lazarus AJ, Belcher JW. 1995. Radial evolution of the solar wind from IMP 8 to Voyager 2. Geophys. Res. Lett. 22, 325–328. ( 10.1029/94GL03273) [DOI] [Google Scholar]
  • 60.Richardson JD, Smith CW. 2003. The radial temperature profile of the solar wind. Geophys. Res. Lett. 30, 1206 ( 10.1029/2002GL016551) [DOI] [Google Scholar]
  • 61.Vasquez BJ, Smith CW, Hamilton K, MacBride BT, Leamon RJ. 2007. Evaluation of the turbulent energy cascade rates from the upper inertial range in the solar wind at 1 AU. J. Geophys. Res. 112, A07101. [Google Scholar]
  • 62.Lamarche L. J., Vasquez BJ, Smith CW. 2014. Proton temperature change with heliocentric distance from 0.3 to 1 AU according to relative temperatures. J. Geophys. Res. 119, 3267–3280. ( 10.1002/2013JA019529) [DOI] [Google Scholar]
  • 63.Matthaeus WH, Zank GP, Oughton S, Mullan DJ, Dmitruk P. 1999. Coronal heating by magnetohydrodynamic turbulence driven by reflected low-frequency waves. Astrophys. J. 523, L93 ( 10.1086/312259) [DOI] [Google Scholar]
  • 64.Sreenivasan KR. 1995. On the universality of the Kolmogorov constant. Phys. Fluids 7, 2778–2784. ( 10.1063/1.868656) [DOI] [Google Scholar]
  • 65.Iroshnikov PS. 1964. Turbulence of a conducting fluid in a strong magnetic field. Sov. Astron. 7, 566. [Google Scholar]
  • 66.Kraichnan RH. 1965. Inertial range of hydromagnetic turbulence. Phys. Fluids 8, 1385–1387. ( 10.1063/1.1761412) [DOI] [Google Scholar]
  • 67.Matthaeus WH, Zhou Y. 1989. Extended inertial range phenomenology of magnetohydrodynamic turbulence. Phys. Fluids B 1, 1929–1931. ( 10.1063/1.859110) [DOI] [Google Scholar]
  • 68.Goldreich P, Sridhar S. 1995. Toward a theory of interstellar turbulence. II. Strong Alfvénic turbulence. Astrophys. J. 438, 763–775. ( 10.1086/175121) [DOI] [Google Scholar]
  • 69.Boldyrev S. 2005. On the spectrum of magnetohydrodynamic turbulence. Astrophys. J. 626, L37–L40. ( 10.1086/431649) [DOI] [Google Scholar]
  • 70.Boldyrev S. 2006. Spectrum of magnetohydrodynamic turbulence. Phys. Rev. Lett. 96, 115002 ( 10.1103/PhysRevLett.96.115002) [DOI] [PubMed] [Google Scholar]
  • 71.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, A101012. [Google Scholar]
  • 72.Saur J, Bieber JW. 1999. Geometry of low-frequency solar wind magnetic turbulence: evidence for radially aligned Alfvénic fluctuations. J. Geophys. Res. 104, 9975–9988. ( 10.1029/1998JA900077) [DOI] [Google Scholar]
  • 73.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, 20 673–20 683. ( 10.1029/JA095iA12p20673) [DOI] [Google Scholar]
  • 74.Bieber JW, Wanner W, Matthaeus WH. 1996. Dominant two-dimensional solar wind turbulence with implications for cosmic ray transport. J. Geophys. Res. 101, 2511–2522. ( 10.1029/95JA02588) [DOI] [Google Scholar]
  • 75.Smith CW. 2003. The geometry of turbulent magnetic fluctuations at high heliographic latitudes. In Proc. of the Tenth International Solar Wind Conference (eds Velli M, Bruno R, Malara R.), AIP Conference Proceedings, vol. 679, pp. 413–416. New York, NY: American Institute of Physics. [Google Scholar]
  • 76.Smith CW. 2003. Magnetic Helicity in the solar wind. In Proc. Thirty-Fourth COSPAR Scientific Assembly Advances in Space Research, vol. 32(10), pp. 1971–1980. UK: Elsevier. [Google Scholar]
  • 77.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. [Google Scholar]
  • 78.Shebalin JV, Matthaeus WH, Montgomery D. 1983. Anisotropy in MHD turbulence due to a mean magnetic field. J. Plasma Phys. 29, 525–547. ( 10.1017/S0022377800000933) [DOI] [Google Scholar]
  • 79.Matthaeus WH, Ghosh S, Oughton S, Roberts DA. 1996. Anisotropic three-dimensional MHD turbulence. J. Geophys. Res. 101, 7619–7629. ( 10.1029/95JA03830) [DOI] [Google Scholar]
  • 80.Ghosh S, Matthaeus WH, Roberts DA, Goldstein ML. 1998. The evolution of slab fluctuations in the presence of pressure-balanced magnetic structures and velocity shears. J. Geophys. Res. 103, 23 691–23 704. ( 10.1029/98JA02195) [DOI] [Google Scholar]
  • 81.Ghosh S, Matthaeus WH, Roberts DA, Goldstein ML. 1998. Waves, structures, and the appearance of two-component turbulence in the solar wind. J. Geophys. Res. 103, 23 705–23 715. ( 10.1029/98JA02194) [DOI] [Google Scholar]
  • 82.Higdon JC. 1984. Density fluctuations in the interstellar medium: Evidence for anisotropic magnetogasdynamic turbulence. I. Model and astrophysical sites. Astrophys. J. 285, 109–123. ( 10.1086/162481) [DOI] [Google Scholar]
  • 83.Howes GG, Cowley SC, Dorlund W, Hammett GW, Quataert E, Schekochihin AA. 2008. A model of turbulence in magnetized plasmas: implications for the dissipation range in the solar wind. J. Geophys. Res. 113, A05103. [Google Scholar]
  • 84.Howes GG, Dorland W, Cowley SC, Hammett GW, Quataert E, Schekochihin AA, Tatsuno T. 2008. Kinetic simulations of magnetized turbulence in astrophysical plasmas. Phys. Rev. Lett. 100, 065004 ( 10.1103/PhysRevLett.100.065004) [DOI] [PubMed] [Google Scholar]
  • 85.Howes GG, Dorland W, Cowley SC, Hammett GW, Quataert E, Schekochihin AA, Tatsuno T. 2008. Reply. Phys. Rev. Lett. 101, 149502 ( 10.1103/PhysRevLett.101.149502) [DOI] [PubMed] [Google Scholar]
  • 86.Sahraoui F, Goldstein ML, Robert P, Khotyaintsev YuV. 2009. Evidence of a cascade and dissipation of solar-wind turbulence at the electron gyroscale. Phys. Rev. Lett. 102, 231102 ( 10.1103/PhysRevLett.102.231102) [DOI] [PubMed] [Google Scholar]
  • 87.Sahraoui F, Goldstein ML, Belmont G, Canu P, Rezeau L. 2010. Three dimensional anisotropic k spectra of turbulence at subproton scales in the solar wind. Phys. Rev. Lett. 105, 131101 ( 10.1103/PhysRevLett.105.131101) [DOI] [PubMed] [Google Scholar]
  • 88.He J, Marsch E, Tu C, Yao S, Tian H. 2011. Possible evidence of Alfven-cyclotron waves in the angle distribution of magnetic helicity of solar wind turbulence. Astrophys. J. 731, 85 ( 10.1088/0004-637X/731/2/85) [DOI] [Google Scholar]
  • 89.Podesta JJ, Gary SP. 2011. 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 ( 10.1088/0004-637X/734/1/15) [DOI] [Google Scholar]
  • 90.Alexandrova O, Bale SD, Lacombe C. 2013. Comment on ‘Evidence of a cascade and dissipation of solar-wind turbulence at the electron gyroscale’. Phys. Rev. Lett. 111, 149001 ( 10.1103/PhysRevLett.111.149001) [DOI] [PubMed] [Google Scholar]
  • 91.Ghosh S, Siregar E, Roberts DA, Goldstein ML. 1996. Simulation of high-frequency solar wind power spectra using Hall magnetohydrodynamics. J. Geophys. Res. 101, 2493–2504. ( 10.1029/95JA03201) [DOI] [Google Scholar]
  • 92.Stawicki O, Gary SP, Li H. 2001. Solar wind magnetic fluctuation spectra: dispersion vs. damping. J. Geophys. Res. 106, 8273–8281. ( 10.1029/2000JA000446) [DOI] [Google Scholar]
  • 93.Krishan V, Mahajan SM. 2004. Magnetic fluctuations and Hall magnetohydrodynamic turbulence in the solar wind. J. Geophys. Res. 109, A11105 ( 10.1029/2004JA010496) [DOI] [Google Scholar]
  • 94.Galtier S. 2006. Wave turbulence in incompressible Hall magnetohydrodynamics. J. Plasma Phys. 72, 721–769. ( 10.1017/S0022377806004521) [DOI] [Google Scholar]
  • 95.Galtier S, Buchlin E. 2007. Multiscale hall-magnetohydrodynamic turbulence in the solar wind. Astrophys. J. 656, 560–566. ( 10.1086/510423) [DOI] [Google Scholar]
  • 96.Saito S, Gary SP, Li H, Narita Y. 2008. Whister turbulence: particle-in-cell simulations. Phys. Plasmas 15, 102305 ( 10.1063/1.2997339) [DOI] [Google Scholar]
  • 97.Gary SP, Smith CW. 2009. Short-wavelength turbulence in the solar wind: linear theory of whistler and kinetic Alfvén fluctuations. J. Geophys. Res. 114, A12105 ( 10.1029/2009JA014525) [DOI] [Google Scholar]
  • 98.Gary SP. 2015. Short-wavelength plasma turbulence and temperature anisotropy instabilities: recent computational progress. Phil. Trans. R. Soc. A 373, 20140149 ( 10.1098/rsta.2014.0149) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Bruno R, Trenchi L, Telloni D. 2014. Spectral slope variation at proton scales from fast to slow solar wind. Astrophys. J. Lett. 793, L15 ( 10.1088/2041-8205/793/1/L15) [DOI] [Google Scholar]
  • 100.Smith CW, Mullan DJ, Ness NF, Skoug RM, Steinberg J. 2001. Day the solar wind almost disappeared: magnetic field fluctuations, wave refraction and dissipation. J. Geophys. Res. 106, 18 625–18 634. ( 10.1029/2001JA000022) [DOI] [Google Scholar]
  • 101.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 ( 10.1088/2041-8205/787/2/L24) [DOI] [Google Scholar]
  • 102.Chen CHK, Leung L, Boldyrev S, Maruca BA, Dale SD. 2014. Ion-scale spectral break of solar wind turbulence at high and low beta. Geophys. Res. Lett. 41, 8081–8088. ( 10.1002/2014GL062009) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Marino R, Sorisso-Valvo L, Carbone V, Noullez A, Bruno R, Bavassano B. 2008. Heating the solar wind by a magnetohydrodynamic turbulent energy cascade. Astrophys. J. Lett. 677, L71–L74. ( 10.1086/587957) [DOI] [Google Scholar]
  • 104.Vasquez BJ, Smith CW, Stawarz JE. 2010. Proton heating rate in polar winds at solar minima. In SOHO 23: Understanding a Peculiar Solar Minimum (eds Cranmer SR, Hoeksema T, Kohl J.), vol. 428, pp. 269–273. ASP Conference Series San Fransisco, CA: National Society of the Pacific. [Google Scholar]
  • 105.Elsässer WM. 1950. The hydromagnetic equations. Phys. Rev. 79, 183 ( 10.1103/PhysRev.79.183) [DOI] [Google Scholar]
  • 106.Taylor GI. 1938. The spectrum of turbulence. Proc. R. Soc. Lond. A 164, 476–490. ( 10.1098/rspa.1938.0032) [DOI] [Google Scholar]
  • 107.Woltjer L. 1958. A theorem for force-free magnetic fields. Proc. Natl Acad. Sci. USA 44, 489–491. ( 10.1073/pnas.44.6.489) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Woltjer L. 1958. On hydromagnetic equilibrium. Proc. Natl Acad. Sci. USA 44, 833–841. ( 10.1073/pnas.44.9.833) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Moffatt HK. 1969. The degree of knottedness of tangled vortex line. J. Fluid Mech. 35, 117–129. ( 10.1017/S0022112069000991) [DOI] [Google Scholar]
  • 110.Frisch U, Pouquet A, Léorat J, Mazure A. 1975. Possibility of an inverse cascade of magnetic helicity in magnetohydrodynamic turbulence. J. Fluid Mech. 68, 769–778. ( 10.1017/S002211207500122X) [DOI] [Google Scholar]
  • 111.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. ( 10.1088/0004-637X/702/2/1604) [DOI] [Google Scholar]
  • 112.Hossain M., Gray PC, Pontius DH, Jr, Matthaeus WH. 1995. Phenomenology for the decay of energy-containing eddies in homogeneous, MHD turbulence. Phys. Fluids 7, 2886–2904. ( 10.1063/1.868665) [DOI] [Google Scholar]
  • 113.Dobrowolny M, Mangeney A, Veltri P. 1980. Fully developed anisotropic hydromagnetic turbulence in interplanetary space. Phys. Rev. Lett. 45, 144–147. ( 10.1103/PhysRevLett.45.144) [DOI] [Google Scholar]
  • 114.Stone EC, Frandsen AM, Mewaldt RA, Christian ER, Margolies D, Ormes JF, Snow F. 1998. The advanced composition explorer. Space Sci. Rev. 86, 1–22. ( 10.1023/A:1005082526237) [DOI] [Google Scholar]
  • 115.Smith CW, Acuña MH, Burlaga LF, L'Heureux J, Ness NF, Scheifele J. 1998. The ACE magnetic field experiment. Space Sci. Rev. 86, 613–632. ( 10.1023/A:1005092216668) [DOI] [Google Scholar]
  • 116.McComas DJ, Bame SJ, Barker P, Feldman WC, Phillips JL, Riley P, Griffee JW. 1998. Solar wind electron proton alpha monitor (SWEPAM) for the advanced composition explorer. Space Sci. Rev. 86, 563–612. ( 10.1023/A:1005040232597) [DOI] [Google Scholar]
  • 117.Wan M, Servidio S, Oughton S, Matthaeus WH. 2009. The third-order law for increments in magnetohydrodynamic turbulence with constant shear. Phys. Plasmas 16, 090703 ( 10.1063/1.3240333) [DOI] [Google Scholar]
  • 118.Wan M, Servidio S, Oughton S, Matthaeus WH. 2010. The third-order law for magnetohydrodynamic turbulence with shear: numerical investigation. Phys. Plasmas 17, 052307 ( 10.1063/1.3398481) [DOI] [Google Scholar]
  • 119.Stawarz JE, Vasquez BJ, Smith CW, Forman MA, Klewicki JC. 2011. Third moments and the role of anisotropy from velocity shear in the solar wind. Astrophys. J. 736, 44 ( 10.1088/0004-637X/736/1/44) [DOI] [Google Scholar]
  • 120.Borovsky JE, Denton MH. 2010. Solar wind turbulence and shear: a superposed-epoch analysis of corotating interaction regions at 1 AU. J. Geophys. Res. 115, A10101 ( 10.1029/2009JA014966) [DOI] [Google Scholar]
  • 121.Tessein JA, Smith CW, Vasquez BJ, Skoug RM. 2011. Turbulence associated with corotating interaction regions at 1 AU: inertial and dissipation range magnetic field spectra. Astrophys. J. 116, A10104 ( 10.1029/2011JA016647) [DOI] [Google Scholar]
  • 122.Smith CW, Tessein JA, Vasquez BJ, Skoug RM. 2011. Turbulence associated with corotating interaction regions at 1 AU: inertial range cross-helicity spectra. Astrophys. J. 116, A10103 ( 10.1029/2011JA016645) [DOI] [Google Scholar]
  • 123.Stawarz JE, Smith CW, Vasquez BJ, Forman MA, MacBride BT. 2010. The turbulent cascade for high cross helicity states at 1 AU. Astrophys. J. 713, 920–934. ( 10.1088/0004-637X/713/2/920) [DOI] [Google Scholar]
  • 124.Smith CW, Stawarz JE, Vasquez BJ, Forman MA, MacBride BT. 2009. The turbulent cascade at 1 AU in high cross-helicity flows. Phys. Rev. Lett. 103, 201101 ( 10.1103/PhysRevLett.103.201101) [DOI] [PubMed] [Google Scholar]
  • 125.Forman MA, Smith CW, Vasquez BJ, MacBride BT, Stawarz JE, Podesta JJ, Elton DC, Malecot UY, Gagne Y. 2010. Using third-order moments of fluctuations in V and B to determine turbulent heating rates in the solar wind. In Solar Wind 12 (eds Maksimovic M, Issautier K, Meyer-Vernet N, Moncuquet M, Pantellini F.). AIP Conference Proceedings, vol. 1216, pp. 176–179. Melville, NY: AIP. [Google Scholar]
  • 126.Coburn JT, Smith CW, Vasquez BJ, Forman MA, Stawarz JE. 2014. Variable cascade dynamics and intermittency in the solar wind at 1 AU. Astrophys. J. 786, 52 ( 10.1088/0004-637X/786/1/52) [DOI] [Google Scholar]
  • 127.Podesta JJ, Roberts DA, Goldstein ML. 2007. Spectral exponents of kinetic and magnetic energy spectra in solar wind turbulence. Astrophys. J. 664, 543–548. ( 10.1086/519211) [DOI] [Google Scholar]
  • 128.Matthaeus WH, Goldstein ML. 1982. Measurement of the rugged invariants of magnetohydrodynamic turbulence in the solar wind. J. Geophys. Res. 87, 6011–6028. ( 10.1029/JA087iA08p06011) [DOI] [Google Scholar]
  • 129.Roberts DA, Klein LW, Goldstein ML, Matthaeus WH. 1987. The nature and evolution of magnetohydrodynamic fluctuations in the solar wind: voyager observations. J. Geophys. Res. 92, 11 021–11 040. ( 10.1029/JA092iA10p11021) [DOI] [Google Scholar]
  • 130.Bavassano B, Bruno R. 2000. Velocity and magnetic field fluctuations in Alfvenic regions of the inner solar wind: three-fluid observations. J. Geophys. Res. 105, 5113–5118. ( 10.1029/1999JA000336) [DOI] [Google Scholar]
  • 131.Sorisso-Valvo L, Marino R, Carbone V, Noullez A, Lepreti F, Veltri P, Bruno R, Bavassano B, Pietropaolo E. 2007. Observation of inertial range cascade in interplanetary space plasma. Phys. Rev. Lett. 99, 115001 ( 10.1103/PhysRevLett.99.115001) [DOI] [PubMed] [Google Scholar]
  • 132.Podesta JJ, Forman MA, Smith CW. 2007. Anisotropic forms of third-order moments and relationship to the cascade rate in axisymmetric magnetohydrodynamic turbulence. Phys. Plasmas 14, 092305 ( 10.1063/1.2783224) [DOI] [Google Scholar]
  • 133.Podesta JJ. 2008. Laws for third-order moments in homogeneous anisotropic incompressible magnetohydrodynamic turbulence. J. Fluid Mech. 609, 171–194. ( 10.1017/S0022112008002280) [DOI] [Google Scholar]
  • 134.Podesta JJ, Forman MA, Smith CW, Elton DC, Malécot Y. 2009. Accurate estimation of third-order moments from turbulence measurements. Nonlin. Proc. Geophys. 16, 99–110. ( 10.5194/npg-16-99-2009) [DOI] [Google Scholar]
  • 135.Carbone V, Marino R, Sorriso-Valvo L, Noullez A, Bruno R. 2009. Scaling laws of turbulence and heating of fast solar wind: the role of density fluctuations. Phys. Rev. Lett. 103, 061102 ( 10.1103/PhysRevLett.103.061102) [DOI] [PubMed] [Google Scholar]
  • 136.Podesta JJ. 2011. On the energy cascade rate of solar wind turbulence in high cross helicity flows. J. Geophys. Res. 116, A05101 ( 10.1029/2010JA016306) [DOI] [Google Scholar]
  • 137.Osman KT, Wan M, Matthaeus WH, Weygand JM, Dasso S. 2011. Anisotropic third-moment estimates of the energy cascade in wolar wind turbulence using multispacecraft data. Phys. Rev. Lett. 107, 165001 ( 10.1103/PhysRevLett.107.165001) [DOI] [PubMed] [Google Scholar]
  • 138.Marino R, Sorriso-Valvo L, D'Amicis R, Carbone V, Bruno R, Veltri P. 2012. On the occurrence of the third-order scaling in high latitude solar wind. Astrophys. J. 750, 41 ( 10.1088/0004-637X/750/1/41) [DOI] [Google Scholar]
  • 139.Gogoberidze G, Perri S, Carbone V. 2013. The Yaglom law in the expanding solar wind. Astrophys. J. 769, 111 ( 10.1088/0004-637X/769/2/111) [DOI] [Google Scholar]
  • 140.Hellinger P, S˘tverák S˘, Matteini L, Velli M. 2013. Proton thermal energetics in the solar wind: Helios reloaded. J. Geophys. Res. 118, 1351–1365. ( 10.1002/jgra.50107) [DOI] [Google Scholar]
  • 141.Kraichnan RH. 1967. Inertial ranges in two-dimensional turbulence. Phys. Fluids 10, 1417–1423. ( 10.1063/1.1762301) [DOI] [Google Scholar]
  • 142.Frisch U, Sulem PL. 1984. Numerical simulation of the inverse cascade in two-dimensional turbulence. Phys. Fluids 27, 1921–1923. ( 10.1063/1.864870) [DOI] [Google Scholar]
  • 143.Grappin R, Frisch U, Léorat J, Pouquet A. 1982. Alfvénic fluctuations as asymptotic states of MHD turbulence. Astron. Astrophys. 105, 6–14. [Google Scholar]
  • 144.Matthaeus WH, Goldstein ML, Montgomery DC. 1983. Turbulent generation of outward-traveling interplanetary Alfvénic fluctuations. Phys. Rev. Lett. 51, 1484–1487. ( 10.1103/PhysRevLett.51.1484) [DOI] [Google Scholar]
  • 145.Matthaeus WH, Montgomery DC. 1980. Selective decay hypothesis at high mechanical and magnetic Reynolds numbers. In Nonlinear dynamics (ed. Helleman RHG.). Annals of the New York Academy of Sciences, vol. 357, pp. 203–222. New York, NY: New York Academy of Sciences. [Google Scholar]
  • 146.Matthaeus WH, Stribling WT, Martinez D, Oughton S, Montgomery D. 1991. Selective decay and coherent vortices in two-dimensional incompressible turbulence. Phys. Rev. Lett. 66, 2731–2734. ( 10.1103/PhysRevLett.66.2731) [DOI] [PubMed] [Google Scholar]

Articles from Philosophical transactions. Series A, Mathematical, physical, and engineering sciences are provided here courtesy of The Royal Society

RESOURCES