the horizontal cascade structure of atmospheric fields

45
6/3/09 1 The Horizontal cascade structure of atmospheric 1 fields determined from aircraft data 2 S. Lovejoy 1 , A. F. Tuck 2 , D. Schertzer 3,4 3 1 Physics, McGill University, 3600 University st., Montreal, Que. H3A 2T8, 4 [email protected] 5 2 NOAA Earth System Research Laboratory, Chemical Sciences Division, 325 6 Broadway, Boulder CO 80305-3337, USA, 7 3 CEREVE, Université Paris Est, France, 8 4 Météo France, 1 Quai Branly, Paris 75005, France 9 10

Upload: others

Post on 19-Apr-2022

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: The Horizontal cascade structure of atmospheric fields

6/3/09 1

The Horizontal cascade structure of atmospheric 1

fields determined from aircraft data 2

S. Lovejoy1, A. F. Tuck2, D. Schertzer3,4 3

1 Physics, McGill University, 3600 University st., Montreal, Que. H3A 2T8, 4

[email protected] 5

2 NOAA Earth System Research Laboratory, Chemical Sciences Division, 325 6

Broadway, Boulder CO 80305-3337, USA, 7

3 CEREVE, Université Paris Est, France, 8

4Météo France, 1 Quai Branly, Paris 75005, France 9

10

Page 2: The Horizontal cascade structure of atmospheric fields

6/3/09 2

11

Abstract: 12

Aircraft measurements of the power spectra of the horizontal wind field typically 13

find a transition from ≈ k-5/3 to ≈ k-2.4 at scales somewhere around 40 km (k is a 14

wavenumber). In the usual interpretation this represents a transition between an isotropic 15

3-D and an isotropic 2-D turbulence; we have recently argued that the turbulence is so 16

highly anisotropic that it has different exponents in the horizontal and vertical. When, 17

coupled with non-flat aircraft trajectories this can explain the break as a transition from a 18

roughly horizontal spectrum at small scales to the spurious appearance of the vertical 19

spectrum at large scales. If the atmosphere indeed has wide range horizontal scaling, then 20

it is important to test out the multiplicative cascade models that predict its statistical 21

behaviour. In this paper, we do this by analyzing wind, temperature, pressure and 22

humidity data from the Winter Storm 2004 experiment using 24 aircraft legs each 1120 23

km long and at 280 m resolution. We analyse both the turbulent fluxes and the 24

fluctuations showing that in spite of the non-flat trajectories that there is good evidence of 25

roughly planetary scale multiplicative cascades. By carefully determining the scale-by-26

scale effects of intermittency on the aircraft altitude and measurements we estimate the 27

corresponding scaling exponents. We argue that our results should finally permit the 28

emergence of a long needed consensus about the basic scale-by-scale statistical properties 29

of the atmosphere. They also point to the urgent need to development anisotropic scaling 30

models of turbulence. 31

Page 3: The Horizontal cascade structure of atmospheric fields

6/3/09 3

32

1. Introduction: 33

1.1 The biases in the wind statistics: 34

Data analysis requires both a theoretical framework in which the physical 35

quantities can be defined and understood as well as a more technical sampling/measuring 36

framework which assesses their degree of correspondence with the data. In the case of 37

aircraft measurements of atmospheric turbulence, it is usual to interpret the measurements 38

in an isotropic or quasi isotropic framework in which basic exponents are the same in the 39

horizontal and vertical directions. However, starting in the 1980’s studies of the vertical 40

structure of the atmosphere have almost invariably concluded that the horizontal wind is 41

scaling in the vertical, but with a different exponent than in the horizontal. This 42

empirically observed scaling stratification has thus been interpreted either with quasi-43

linear gravity wave models (Gardner, 1994), (Gardner et al., 1993), (Dewan, 1997), 44

(Koch et al., 2005) or the 23/9D strongly turbulent model, (Schertzer and Lovejoy, 45

1985), (Lovejoy et al., 2008b). 46

In the last few years, these results have prompted the examination of the 47

consequences of such anisotropic but scaling turbulence for the interpretation of aircraft 48

data; both for stratospheric aircraft following isomachs (Lovejoy et al., 2004) and for 49

tropospheric aircraft following isobars (Lovejoy et al., 2009c). The perhaps surprising 50

conclusions were that they are best explained by single wide range anisotropic scaling 51

regime. A corollary is that the usual interpretation - that the atmosphere has two (or 52

more) isotropic scaling regimes (e.g. small scale, 3D isotropic and large scale, 2-D 53

Page 4: The Horizontal cascade structure of atmospheric fields

6/3/09 4

isotropic) is untenable. The key finding was that when measuring the wind field, rather 54

than detecting successively two different isotropic turbulence exponents, the aircraft first 55

detects at small scales the correct horizontal exponent, and then - at larger scales where 56

the departure of the aircraft from a flat trajectory becomes important - it spuriously 57

measures the different vertical exponent. In terms of traditional spectral exponents β, the 58

horizontal has β ≈ 5/3 (near the Kolmogorov value) while the vertical has β ≈ 2.4 (as 59

accurately estimated by drop sondes (Lovejoy et al., 2007)). As pointed out in (Lovejoy 60

et al., 2009c), the latter is almost exactly the same as value of the large scale exponents 61

found in all the major aircraft campaigns including GASP (Nastrom and Gage, 1985), 62

MOZAIC (Cho and Lindborg, 2001) and others (Gao and Meriwether, 1998), 63

(Bacmeister et al., 1996). 64

If the atmosphere is anisotropic and scaling over wide ranges, then it is 65

fundamentally important to characterize the anisotropic scaling regimes for as many of 66

the fields as possible, i.e. to determine the type of scaling (including the determination of 67

the basic turbulent parameters), as well as the limits to the scaling – the inner and outer 68

scales. The goal of this paper is therefore to go a step beyond the reinterpretation to see 69

how the data can be quantitatively used to characterize atmospheric turbulence while 70

attempting to avoid spurious aircraft induced effects. 71

The basic model accounting for the wide range anisotropic scaling of the fields is 72

anisotropic multiplicative cascades i.e. anisotropic extensions of the explicit 73

phenomenological cascade models that were developed through the 1960’s, 70’s and 80’s 74

((Novikov and Stewart, 1964), (Yaglom, 1966), (Mandelbrot, 1974), (Schertzer and 75

Lovejoy, 1987)). In recent papers, we have investigated the vertical cascade structure 76

Page 5: The Horizontal cascade structure of atmospheric fields

6/3/09 5

using drop sondes (Lovejoy et al., 2009b), the horizontal cascade structure using satellite 77

data (Lovejoy et al., 2009a) and vertical cross-sections using lidar backscatter data 78

((Lilley et al., 2004), (Lilley et al., 2008)). In addition, meteorological reanalyses as well 79

as numerical meteorological models (Stolle et al., 2009) have recently been shown with 80

high accuracy to have multiplicative cascade structures over almost their entire spatial 81

ranges and up to about 10 days in the time domain. What is missing from a fairly 82

complete spatial characterization of atmospheric cascades is the direct verification of the 83

multiplicative cascade structure in the horizontal on the standard meteorological fields 84

including the horizontal wind, the temperature, humidity, potential temperature; this is 85

our task here. 86

87

2. The data, its Intermittency: 88

The data analysed were from the Winter Storm 2004 experiment and involved 10 89

aircraft flights over a roughly a 2 week period over the northern Pacific. Each flight 90

dropped 20 - 30 drop sondes which were analyzed to determine the vertical cascade 91

structure (Lovejoy et al., 2009b). For air traffic reasons, the Gulfstream 4 plane flew 92

along either the 162, 178, or 196 mb isobars, (to ±0.11mb i.e. the pressure level was ≈ 93

constant to within ±0.068%), it has a radar altimeter which is reliable over the ocean, but 94

which is only used to anchor the GPS estimates, these geometric altitudes and were used 95

here. Each had one or more roughly constant straight and constant altitude legs more 96

than four hundred kilometers long between 11.9, 13.7 km altitude. The data were 97

sampled every 1 s and the mean horizontal aircraft speed with respect to the ground was 98

280 m/s. In addition, we checked that the distance covered on the ground between 99

Page 6: The Horizontal cascade structure of atmospheric fields

6/3/09 6

measurements was constant to within ±2% so that the horizontal velocity was nearly 100

constant (in addition, using interpolation, we repeated the key analyses using the actual 101

ground distance rather than the elapsed time and found only very small differences). This 102

is the same data set analyzed in (Lovejoy et al., 2009c) where it is described in more 103

detail. 104

The basic scale by scale relations between the trajectory and the fields was 105

investigated using both spectral and cross-spectral analysis in (Lovejoy et al., 2009c). In 106

particular, it was shown that while there were strong phase relations and strong 107

coherencies between the altitude and the wind and between the pressure and the wind - 108

indicating a strong influence of the trajectory on the wind statistics at large scales (and 109

visa versa at small scales) - on the contrary, the corresponding phases and coherencies of 110

the altitude and pressure with the temperature and with the humidity were very low. 111

This is consistent with the systematic finding that the latter show no scale breaks, that 112

they have wide range scaling e.g. (Hovde et al., 2009; Tuck et al., 2004); see also (Lilley 113

et al., 2004) for aerosol lidar backscatter and Tuck et al., 2004 for ozone and nitrous 114

oxide concentrations. 115

The coherency and phase analysis suggests that the main effect of the trajectory 116

fluctuations is on the wind field and that its influence will be smaller for the other 117

atmospheric fields. Going beyond phases and coherencies, let’s consider the detailed 118

statistics of the trajectories. For scales less than about 4 km, the aircraft inertia smoothes 119

them out; the main variations are smooth and are associated with various roll modes and 120

the pilot/autopilot response. However, at scales from about 4 to 100 km (and from ≈3 to 121

about 300 km for the stratospheric ER-2 trajectories along isomachs), the trajectory is 122

Page 7: The Horizontal cascade structure of atmospheric fields

6/3/09 7

fractal with the altitude lagging behind the wind fluctuations. Finally at scales >≈ 40 km 123

the aircraft follows the isobars quite closely so that the wind lags behind the pressure. 124

These conclusions were reached by considering the average for 24 aircraft legs flying 125

between roughly 11 and 13 km, each 1120 km long as well as through a leg by leg 126

analysis which showed considerable variability in the scale at which these transition 127

occurred. According to this analysis we may anticipate that there will be a strong effect of 128

the variability (indeed, intermittency) of the aircraft altitude on the measurements – at 129

least for scales less than about 40 km where the wind leads the altitude, imposing its 130

strong intermittency on the aircraft. 131

In order to demonstrate this and to quantify the intermittency of the trajectory we 132

performed a standard multifractal analysis. Since we will perform similar analyses on the 133

meteorological fields we give a general explanation of the method, using the example of 134

the familiar Kolmogorov law of turbulence. 135

Consider wind fluctuations Δv over distances Δx with underlying turbulent energy 136

flux φ; we therefore have: 137

Δv = φΔxH (1) 138

The form 1, is a generic relation between a turbulent fluctuation Δv and turbulent 139

flux φ, the Kolmogorov law is the special case where φ = ε1/3 and H = 1/3. If over a 140

range of scales, φ is the result of a multiplicative cascade process, then it obeys the 141

following statistics: 142

M q =λ

λeff

⎝⎜

⎠⎟

K q( )

; λ =Lref

L; λeff =

Lref

Leff

(2) 143

Page 8: The Horizontal cascade structure of atmospheric fields

6/3/09 8

Where M q = φλq / φλ

q is the normalized qth moment at resolution L (scale ratio λ), Lref 144

is a convenient reference scale (taken below as the largest great circle distance on the 145

earth, Lref = 20,000 km), Leff is the “effective” scale at which the cascade begins, and K(q) 146

is the scaling exponent characterizing the intermittency. In order to test eq. 2 on the data, 147

it is sufficient to use the (absolute) fluctuations at the smallest available scales l; 148

(corresponding to the large scale ratio Λ); and then to estimate the normalized flux as: 149

φΛ

φΛ

=Δv l( )Δv l( ) ; Λ =

Ll

(3) 150

i.e. it is not necessary to know H or even the physical nature of the flux φ so that our 151

results are independent of any specific theory of turbulence. The fluctuations can be 152

estimated either by differences (when 0< H <1) or (equivalently) by wavelets. The 153

normalized flux at lower resolution λ < Λ is then obtained by straightforward spatial 154

averaging of the fine scale (Λ) resolution normalized fluxes. Note that we do not 155

determine the direction (i.e. from large to small or small to large) of the cascade. 156

When this method is applied to the aircraft altitude, we obtain figure 1a which 157

shows that eq. 2 is well verified with lines converging to log10λeff ≈ 2.9 corresponding to 158

Leff ≈ 30 km. This shows that for scales <≈ 30 km, the altitude has very strong 159

intermittency which is of the type theoretically predicted by turbulent cascades. This 160

result is compatible with the phase and coherence analysis summarized above to the 161

effect that for scales < ≈ 40 km that the phase of the wind leads the phase of the altitude. 162

The simplest explanation is that the altitude is intermittent because it varies in response to 163

the strong wind intermittency. At the larger scales the phase changes sign and the 164

pressure leads the wind so that (due to the autopilot) the aircraft starts to closely follow 165

Page 9: The Horizontal cascade structure of atmospheric fields

6/3/09 9

the isobars which are much smoother (the response time of the autopilot is much shorter 166

than the 2- 3 minutes needed to cover ≈ 40 km but apparently, the turbulence is too 167

strong at smaller scales). 168

We now seek to quantify the accuracy with which the altitude intermittency 169

follows the cascade form. First, we may use the logarithmic slopes of Mq to estimate 170

K(q), and then use a parametric form of K(q) in order to reduce it to a manageable (finite) 171

number of parameters. This is most conveniently done by exploiting the existence of 172

stable, attractive cascade processes - the result of a kind of multiplicative central limit 173

theorem -which gives rise to “universal multifractals” (Schertzer and Lovejoy, 1987), 174

(Schertzer and Lovejoy, 1997). For universal multifractals K(q) has of the form: 175

K q( )=C1

α − 1qα − q( ); q ≥ 0; 0 ≤ α ≤ 2 (4) 176

The parameter C1 (0< C1 < D) characterizes the intermittency of the mean: C1 = K’(1), it 177

thus characterizes the intermittency of the mean field (D is the dimension of the 178

observing space, here D = 1). The second parameter is the multifractal index α which 179

characterises the degree of multifractality, it quantifies how rapidly (with intensity, with 180

q), the statistics deviate from the monofractal case α = 0). In table 1 we give the 181

estimates of the parameters for the altitude and various other fields. 182

To quantify the accuracy, we may characterize the deviations by the mean 183

absolute residuals for the statistical moments Mq of order q from 0 to 2 for all points 184

between the scale of the grid and the scales in the figure (about 100 times larger): 185

( ) ( )10 10log ( ) log /q effM K q λ λΔ = − (5) 186

Page 10: The Horizontal cascade structure of atmospheric fields

6/3/09 10

To convert Δ to a percent deviation, use δ = 100(10Δ-1), which we find for the altitude 187

(fitting between 1 km and 20 km) is ≈ ±2.6% (see table 1 for the mean δ averaged over 188

the q ≤ 2). 189

190

191

192

193

Fig. 1 a: The normalized moments q = 0.2, 0.4, … 3, for aircraft altitude, z; log10λ 194

corresponds to 20,000 km (= Lref). 195

196

Page 11: The Horizontal cascade structure of atmospheric fields

6/3/09 11

T Logθ h vlong vtrans p z

C1 0.052 0.052 0.040 0.033 0.046 0.031 0.23

α 2.15 2.20 2.10 2.10 2.10 2.2 2.15

Leff(km) 5000. 10000. 10000. 105 25000. 1600. 30.

δ(%) 0.5 2.0 0.5 0.4 0.8 0.5 2.6

Table 1: Horizontal cascade parameter estimates. These are estimated over the 197

range 100 km down to 2 km except for z which is over the range 20 km to 0.5 km. The 198

regressions force the lines to pass through a common external scale. We have not given 199

error estimates on the parameters due to the significant systematic (trajectory) effects 200

discussed below. Note that the aircraft α estimates are a bit too big since the theoretical 201

maximum is α = 2. They were estimated with the double trace moment technique which 202

depends largely on the statistics of the weaker events and these could be affected by 203

aircraft turbulence. In our experience, values of α > 2 are always accompanied by error 204

bars that encompass values < 2. 205

206

207

Page 12: The Horizontal cascade structure of atmospheric fields

6/3/09 12

3. Scaling analyses: 208

3.1 Trace moments and the cascade structure of the fields: 209

Based on the analysis of the aircraft altitude and cross-spectra, we have argued 210

that the aircraft trajectory fluctuations and their coupling with the fields must be taken 211

into account for quantitative analysis of the statistics. However, before we attempt a 212

more refined analysis, we present the basic cascade (flux) and then fluctuation analyses. 213

As indicated in section 2 (c.f. eq. 3), the dimensionless, normalized flux is estimated 214

using absolute fluctuations at the smallest scales. In the case of the horizontal wind in the 215

≈k-5/3 regime, it is presumably the energy flux to the one third power, but knowledge of 216

the exact physical nature of the flux is not necessary to test the prediction of cascade 217

processes, eq. 2. First we decompose the wind into longitudinal and transverse 218

components anticipating that the altitude fluctuations will affect each somewhat 219

differently. Fig. 1b shows the results along with those of the pressure. Although the 220

basic structure is reasonably close to (near) planetary scale cascades, we can see evidence 221

of the three regimes discussed above; at scales smaller than about 4 km, the variability is 222

a little too large, logMq is quite linear from 4 – 40 km, and then flattens out a bit at scales 223

> 40 km. Qualitatively this can be understood by the action of inertially smoothed 224

trajectories at small but nonzero slopes (scales < ≈ 4 km), by turbulent wind induced 225

altitude fluctuations for scales between 4 and 40 km and then at larger scales an excess of 226

variability since the aircraft starts to follow the isobaric slopes and responds to the 227

(larger) vertical intermittency. For the pressure, on the contrary the variability at the 228

largest scales is low since the aircraft doesn’t stray far from the isobars. We may also 229

note that for the wind, the outer scales are somewhat larger than the planetary scales 230

Page 13: The Horizontal cascade structure of atmospheric fields

6/3/09 13

(table 1), especially for the longitudinal component. In (Lovejoy et al., 2009c), we have 231

already argued on the basis of cross-spectral analysis that the longitudinal component was 232

the most affected by the trajectory; it is presumably partially responsible for the large 233

value of Leff. In comparison, the value of Leff for the transverse component – although 234

perhaps still too large - is about the same as that reported for satellite radar reflectivities 235

from precipitation (Lovejoy et al., 2008a); the fact that it is larger than the largest great 236

circle distance (20,000 km) simply means that even at planetary scales there is residual 237

variability due to the interactions of the wind with other atmospheric fields. In 238

comparison the outer scale of the pressure is somewhat smaller than planetary scales; this 239

is not surprising since the aircraft was attempting to fly along isobars. 240

241

Page 14: The Horizontal cascade structure of atmospheric fields

6/3/09 14

Fig. 1b: Same as fig. 1a but for the fields strongly affected by the trajectories: the 242

longitudinal wind (top left a), the transverse wind (top right, b), pressure (c, lower left 243

right). 244

245

Turning our attention to the analyses of the relatively unbiased temperature, 246

humidity and log potential temperature fields (fig. 1c), we again see evidence for the 247

three regimes and we note that the cascade structure is somewhat more closely followed 248

with external scales somewhat smaller than 20,000 km. Overall, we conclude that the 249

basic predictions of the cascade theories are well respected; in particular there is no 250

evidence for a break anywhere near the meso-scale (≈ 10 km). The outer scales of these 251

fields are in fact very close to those of visible, infra red and passive microwave radiances 252

as determined by satellite (Lovejoy et al., 2009a). Note that the log equivalent potential 253

temperature is not shown here or below because the humidity is sufficiently low at the 254

aircraft altitude that the graph is nearly indistinguishable from that of the log potential 255

temperature. 256

257

258

259

Page 15: The Horizontal cascade structure of atmospheric fields

6/3/09 15

260

Fig. 1c: Same as 1b but for the fields that are relatively unaffected by the trajectory: 261

temperature (top left), relative humidity (top right), log potential temperature (lower left 262

right). 263

3.2 Fluctuation analysis using structure functions: 264

In section 3.1 we discussed the statistics of the turbulent flux, we now turn our 265

attention to the fluctuations, here estimated using differences: e.g. 266

v Δx( )= v x + Δx( )− v x( ) . Taking the ensemble average of eq. 1 and defining the qth 267

order structure function (Sq): 268

Sq Δx( )= Δv Δx( )q ∝ Δxξ q( ); ξ q( )= qH − K q( ) (6) 269

Page 16: The Horizontal cascade structure of atmospheric fields

6/3/09 16

The results for altitude are given in fig. 2a. We again see evidence for three 270

regimes; the large scale regime very nearly corresponds to linear variations (H = 1), i.e. to 271

a constant slope with the aircraft nearly exactly following the isobars. As we shall see in 272

more detail in the next section - where we estimate H as a function of scale from the 273

logarithmic slopes of fig. 2a - H decreases at the smallest scale from a value ≈1 274

corresponding to inertially smoothed trajectories, to a minimum value ≈ 0.5 at around 4 275

km, and then increases to the maximum value 1 for scales > 40 km where the trajectories 276

are close to isobars and are again, relatively smooth. Note that in fig. 2a, b, c we plot the 277

nondimensional structure functions obtained by dividing the fluctuations by the mean at 278

the smallest scale. 279

280

Page 17: The Horizontal cascade structure of atmospheric fields

6/3/09 17

Fig. 2 a: Structure functions of order q = 0.2, 0.4, ..2.0 for the aircraft altitude. 281

Distances Δx are in km. Separate sets of regression lines are shown for both small and 282

large scales in order to emphasize the change with scale. 283

284

In fig. 2b we show the corresponding results for the longitudinal and transverse 285

wind and pressure. While the behaviour for the wind has essentially two scaling regimes 286

with a 4 - 40 km transition regime – unsurprisingly - the pressure has poor scaling; we 287

reserve a detailed analysis for the next section. In fig. 2c we show the corresponding plot 288

for those fields less affected by the trajectory fluctuations, the temperature, humidity and 289

log potential temperature; we see that the scaling is indeed very good. 290

291

292

293

294

Page 18: The Horizontal cascade structure of atmospheric fields

6/3/09 18

295

Fig. 2 b: Same as fig. 2a but for the fields strongly affected by the aircraft trajectory: the 296

longitudinal wind (upper left), the transverse wind (upper right) and the pressure (lower 297

left). 298

299

300

301

302

Page 19: The Horizontal cascade structure of atmospheric fields

6/3/09 19

303

304

Fig. 2 c: Same as for 2a except for those fields relatively unaffected by the aircraft 305

trajectories: the temperature (upper left), the humidity (upper right) and the log potential 306

temperature (lower left). 307

308

4. Refined analysis: 309

4.1 horizontal exponent estimates 310

In section 3 we considered the raw flux statistics; no attempt was made to 311

quantitatively correct or take into account the effects of the trajectory, we turn to this task 312

now. From the “raw” flux and fluctuation analyses, we saw that the aircraft trajectories 313

Page 20: The Horizontal cascade structure of atmospheric fields

6/3/09 20

are highly intermittent for scales < 40 km (fig. 1a) after which they begin to accurately 314

follow isobars and have nearly constant slopes (fig. 2a). We also saw that for the fields 315

not strongly affected by the trajectory (temperature, humidity, log potential temperature), 316

that the overall scaling was very good. However, if we want to quantitatively 317

characterize the corresponding horizontal scaling and estimate K(q) and hence the 318

parameters H, C1, α, it is not clear what is the optimum range of scales that should be 319

used. In order to obtain a more exact picture of the scale by scale variations in the 320

statistics, we first consider scale by scale estimates of the basic flux and intermittency 321

parameters using the logarithmic derivatives of the second flux exponent (K(2)) and first 322

order (q = 1) structure function exponent H. K(2) was chosen to characterize the 323

amplitude of the intermittency rather than the more fundamental C1 because it is directly 324

related to the spectral exponent β = 1+2H - K(2) and so can be directly used to determine 325

the contribution of intermittency corrections to the spectrum (note that since α ≈ 2 we 326

have C1 ≈ K(2)/2, eq. 4). 327

Page 21: The Horizontal cascade structure of atmospheric fields

6/3/09 21

328

Fig. 3a: Second order trace moments (right column) and their logarithmic derivatives 329

(estimates of K(2), an intermittency index), left column. Top row: wind, longitudinal 330

(blue) transverse (red). The region within the dashed lines has particularly low K(2) (≈ 331

0.05). Middle row: pressure (blue), altitude (red). Bottom, temperature (blue), humidity 332

(red), log potential temperature (gold), reference lines K(2) = 0.12. Log10λ = 0 333

corresponds to 20,000 km (= Lref). 334

335

Fig. 3a shows the results for the second moment of the flux M2; the right column 336

for the moments themselves, the left for the corresponding logarithmic derivatives 337

estimated from the corresponding right hand graphs by performing regressions over an 338

Page 22: The Horizontal cascade structure of atmospheric fields

6/3/09 22

octave of scales centred on the scale indicated. Focusing on the top row for the wind, we 339

see that the logarithmic derivatives bring out the three regions quite clearly. At small 340

scales < ≈ 4 km (large λ), the intermittency is large corresponding to the strong 341

intermittency of the trajectory while at the largest scales >≈ 40 km (small λ) it is again 342

large due to the fact that the near constant slope means that the aircraft is picking up the 343

(large) vertical intermittency indeed, if we use the vertical value K(2) ≈ 0.17 estimated 344

from the drop sondes we find that it is not far from the value of the upper left (velocity) 345

graph in fig. 3a. Note that the estimate of the logarithmic slope is itself highly fluctuating 346

at low λ’s since the statistics get progressively worse at large scales due to the smaller 347

and smaller number of large scale structures. Detailed consideration of the scale by scale 348

estimates of K(2) for the altitude show that it is actually not so constant in spite of the 349

relatively straight appearance of the lines in fig. 1a. For the velocity, the region 4 - 40 km 350

is the least intermittent (in the sense of the lowest exponent K(2)); this is presumably the 351

optimum region for estimating the scaling exponents: at smaller scales, the aircraft 352

trajectory is too intermittent whereas at the larger scales, the intermittency increases 353

again since – at least for the wind - the aircraft is moving significantly in the vertical 354

along a near constant isobaric slope (in any case at the large scales, we obtain the vertical 355

rather than the horizontal exponents and the vertical intermittency is larger, see below). 356

We therefore used this regime to estimate the exponents (table 2 estimated by taking 357

regression over the range 4 – 40 km). We see that in conformity with the cross-spectral 358

analyses and our previous discussion, that the scale by scale intermittency (i.e. K(2) = 359

dlogMq/dlogλ) for T, logθ and h is nearly constant for scales >≈ 4 km confirming that the 360

main deviation from scaling for these fields is only at the smallest scales. We also 361

Page 23: The Horizontal cascade structure of atmospheric fields

6/3/09 23

observe that the exponents are nearly the same for all three fields. In table 2 we also 362

give the corresponding C1 estimates obtained by numerically estimating K’(1) using 363

regressions over the range 4 – 40 km. 364

365

366

Table 2: Parameter estimates over the “optimum” range 4 to 40 km. β is the spectral 367

exponent, Hst is from the first order structure function, Hβ = (β + K(2)-1)/2, and Hmean is 368

the average of the two. C1 = K’(1) from the trace moments and C1,st = ξ(1) - ξ’(1), C1mean 369

is the average. 370

371

Turning our attention to the fluctuation exponent H, (fig. 3b) we see (top row, 372

right, see the reference lines) that they reasonably follow successively the values H = 1/3, 373

0.75 (i.e. the standard Kolmogorov value and the observed vertical value respectively). 374

T logθ h

α 1.78 1.82 1.81

β 1.89 1.91 1.99

K(2) 0.12 0.12 0.11

C1 0.064 0.063 0.051

C1,st 0.040 0.042 0.028

C1mean 0.052±0.012 0.052±0.010 0.040±0.012

Hst 0.49 0.50 0.52

Hβ 0.51 0.52 0.50

Hmean 0.50±0.01 0.51±0.01 0.51±0.01

Page 24: The Horizontal cascade structure of atmospheric fields

6/3/09 24

However the scale by scale estimates (top, left) show that this mean behaviour hides an 375

increase in H from ≈ 1/3 at the smallest scales to H ≈ 0.75 at scales 40 -100 km. 376

According to the leg by leg analyses in (Lovejoy et al., 2009c), this fairly systematic 377

increase of ensemble mean H hides highly variable transition scales in individual legs; 378

the observed fairly continuous change in H is in fact the result of the transition scale 379

varying greatly from leg to leg. On the structure function (upper right) we can also see 380

that fitting a single line through the entire range is not so bad, and yields roughly the 381

mean of 1/3 and 0.75; this may explain the wind value H ≈ 0.56±0.02 from (Hovde et al., 382

2009) (using nearly the same data set as here). This can be compared to the stratospheric 383

estimates over (different) fractal isomach trajectories of 0.50±0.02, 0.52±0.03 384

(transverse, longitudinal wind respectively, (Lovejoy et al., 2004)) and 0.53±0.01 (Tuck 385

et al., 2004). Note that the latter finds significant differences in H for the wind for 386

stratospheric trajectories across or along the polar jet (0.45±0.14 and 0.55±0.12 387

respectively) but this could be a reflection of a systematic shift in the transition scale 388

from an H ≈ 1/3 to an H ≈ 0.75 regime (see however the interpretation in Tuck et al., 389

2004 and Tuck 2008, which is consistent with the thermal wind equation). 390

Moving down to the second row in fig. 3b, we see that at small enough scales, the 391

altitude has Htr ≈ 1 presumably as a consequence of the inertial smoothing leading to near 392

linear behaviour (nearly constant slopes; the subscript “tr” is for “trajectory”). At larger 393

scales, the inertia effect is gradually overcome by turbulence so that at 4 km it reaches a 394

minimum near Htr ≈ 0.48 after which it systematically rises again to Htr ≈ 1 (for scales > 395

≈ 40 km) as the aircraft flies increasingly close to a constant isobaric slope. Up to this 396

scale, the altitude statistics are thus close to those of the ER-2 stratospheric trajectories 397

Page 25: The Horizontal cascade structure of atmospheric fields

6/3/09 25

except that the latter maintain a value Htr ≈ 0.58 along isomachs (over the range ≈ 3 km 398

until ≈ 300 km when the systematic vertical rise of ≈ 1 m/km (due to fuel consumption) 399

eventually implies Htr ≈ 1 (on average, for scales >≈ 300 km; (Lovejoy et al., 2004)). As 400

noted above, along this fairly wide fractal range, the ER-2 had an “anomalous” wind 401

exponent Hwind ≈ 0.5 which is nearly the mean of the horizontal and vertical values (1/3, 402

0.75) so that the effect of following a fractal trajectory with fairly well defined fractal 403

dimension (= 1+Htr) apparently leads the ER2 to sample both horizontal and vertical 404

exponents yielding overall an average of the two. 405

Consider now the pressure exponent. At the smallest scales, it starts off with a 406

value near 2/3 decreasing systematically to zero at scales >≈ 40 km indicating very low 407

fluctuations consistent with near isobaric trajectories (there are some large fluctuations in 408

the scale by scale estimates presumably due to poor statistics). The small scale value ≈ 409

2/3 is presumably a consequence of the dynamic pressure relation Δp ≈ ρΔv2/2 so that 410

Hpress = 2Hwind with Hwind ≈ 1/3. 411

Finally, on the bottom row, we show H estimates for T, logθ, h, again finding 412

relatively good scaling (constant exponents) for scales >≈ 4 km with H ≈ 0.50 for all three 413

(see table 2). For temperature, this is also close to the value 0.52±0.02 obtained from the 414

similar tropospheric data (Hovde et al., 2009) and also not far from stratospheric ER-2 415

analyses with Htemp ≈ 0.45±0.02, (Lovejoy et al., 2004), and Htemp ≈ 0.54±0.01 (Tuck et 416

al., 2004). This value is so close to a ratio of small integers (1/2) that one would expect 417

there to be a straightforward theoretical explanation (or at least dimensional analysis) 418

leading to this value but we are not aware of any adequate theory. The usual argument is 419

to consider temperature to be a passive scalar in which case Corrsin-Obukhov theory 420

Page 26: The Horizontal cascade structure of atmospheric fields

6/3/09 26

predicts H = 1/3 (see Tuck 2008 for arguments about the nature of atmospheric 421

temperature that imply it is not a passive scalar). Alternatively, consider the prediction of 422

Bolgiano-Obkuhov theory for isotropic buoyancy driven turbulence, which for the wind 423

predicts H = 3/5 which is near the correct vertical value. However the same theory 424

predicts Htemp = 1/5 which is even further from the empirical value (see (Monin and 425

Yaglom, 1975)). Further support for the non standard value H ≈ ½ for temperature 426

(corresponding to β ≈ 1.9) comes from aircraft temperature spectra in the lower 427

troposphere (Chigirinskaya et al., 1994) where β ≈ 1.9 fits very well over the range > 1 428

km in scale, and stratospheric potential temperature spectra where β ≈ 1.9 is also very 429

close to data for scales >≈3 km in scale (Bacmeister et al., 1996) (recall that K(2)≈ 0.1 430

and β =1+2H - K(2)). 431

432

433

434

Page 27: The Horizontal cascade structure of atmospheric fields

6/3/09 27

435

Fig. 3b: First order structure functions (right column) and their logarithmic derivatives 436

(estimates of H), left column. Top row: wind, longitudinal (blue) transverse (red), the 437

reference lines correspond to the exponents 1/3, 0.75. Middle row: pressure (blue), 438

altitude (red), the reference lines correspond to H = 1 i.e. constant mean slopes. Bottom, 439

temperature (blue), humidity (red), log potential temperature (gold), reference lines H = 440

0.5. 441

442

443

444

Page 28: The Horizontal cascade structure of atmospheric fields

6/3/09 28

Finally, we consider the spectra (fig. 3c). As indicated above, since the spectrum 445

is a second order statistic, we expect that the scale by scale analysis of the latter (again by 446

logarithmic derivatives) should at least roughly satisfy the equation β = 1+2H- K(2) 447

(actually, this equation is strictly only valid when the scaling is satisfied over wide ranges 448

so hence some deviations are to be expected). Starting with the wind (the top row), we 449

see the spectral version of the transition discussed earlier: at low wavenumbers (k <(40 450

km)-1), β ≈ 2.4 (= 2x0.75- 0.1) to β ≈ 1.6 (= 2x1/3- 0.05 for (40 km)-1 > k > (4 km)-1). In 451

the high wavenumber regime, we see a new feature which is a slight bump near k ≈ (1 452

km)-1, (magnified in the derivative on the left) which may be a residual signature of 453

various aircraft roll modes and autopilot feedbacks. Turning our attention to the middle 454

row (altitude, pressure), we see that the small scales are indeed particularly smooth with 455

the larger scales following near Kolmogorov (β ≈ 5/3) behaviours (probably the 456

maximum near (200 km)-1 is only a statistical fluctuation). Interestingly, the regime with 457

reasonable β ≈ 5/3 scaling seems to be a consequence of the partial cancellation of a 458

continuously varying H and K(2) (compare the corresponding plots of 3a, 3b). Finally, 459

the temperature, humidity and log potential temperature (bottom row) is seen to have 460

significant fluctuations (left), but the overall scaling (far right) is nevertheless fairly good. 461

We now consider the parameter estimates (table 2), taken from the “optimum” 462

regime 4 – 40 km. For both the C1’s and H’s there are two slightly different ways to 463

estimate them. H can be estimated from both structure functions: Hst = ξ(1) and also 464

from the spectral exponent β via: Hβ = (β-1+K(2))/2. The C1 can be estimated from the 465

flux moments: C1 = K’(1) and also from the structure function exponent: C1 = ξ(1) - ξ’(1). 466

In the table we have given all these estimates as well as the mean of the two as the best 467

Page 29: The Horizontal cascade structure of atmospheric fields

6/3/09 29

guess, and with the error as half the difference. From the table we see that the three 468

fields have nearly identical exponents, and that the important parameter H - which should 469

be determinable by dimensional analysis - is near the value 1/2. 470

In fig. 4 we show the moment scaling exponents K(q) estimated over the 4 - 40 471

km range. We see that T, logθ, h remain almost indistinguishable out to very large q 472

values (corresponding to rare extreme fluxes) whereas the longitudinal and transverse 473

components separate for q >≈3 (already a large value of q). The asymptotic linearity is 474

predicted as a consequence of a “multifractal phase transition”, (Szépfalusy et al., 1987), 475

(Schertzer et al., 1993) and depends on the finite size of the sample (the asymptotic slope 476

is simply the largest singularity present in the sample). 477

478

479

480

Page 30: The Horizontal cascade structure of atmospheric fields

6/3/09 30

481

Fig. 3c: Spectra (right column) and their logarithmic derivatives (estimates of -β), left 482

column. Top row: wind, longitudinal (blue) transverse (red). The thin reference lines 483

correspond to β = 5/3, 2.4. Middle row: pressure (blue), altitude (red) with reference 484

lines corresponding to β = 5/3. Bottom, temperature (blue), humidity (red), log potential 485

temperature (gold), reference lines β = 2. Vertical dashed lines indicate the region with 486

scales 4 – 40 km where the velocity intermittency is low. 487

488

Page 31: The Horizontal cascade structure of atmospheric fields

6/3/09 31

489

Fig. 4: The K(q) functions for altitude and pressure (upper, lower blue), temperature, 490

humidity, log potential temperature (lower, middle, upper green), longitudinal, transverse 491

wind (upper, lower, red). 492

493

4.2 Horizontal - Vertical comparison, vertical stratification, 494

estimating Hz, Del: 495

We have argued that atmospheric fields are compatible with wide range horizontal 496

scaling and have estimated the corresponding exponents. This work complements that of 497

(Lovejoy et al., 2009b) where high resolution drop sondes were used to estimate the 498

corresponding vertical cascades and exponents; the overall conclusion is thus that the 499

dynamics are scaling and turbulent but anisotropic. The simplest anisotropic turbulence 500

Page 32: The Horizontal cascade structure of atmospheric fields

6/3/09 32

model involves a unique scale function for all the fields. This would imply that the ratio 501

of horizontal and vertical components is Hhor/Hver = Hz = constant, so that for universal 502

multifractals αhor = αver and C1hor/C1ver= Hz. In (Lilley et al., 2008) there was an 503

extensive analysis of this for the lidar backscatter data (B) from passive pollutants. 504

Recall that the significance of Hz is that it controls the scaling of the aspect ratio of 505

structures in vertical sections. In particular, if we assume horizontal isotropy, then 506

volumes of structures vary as L2+Hz

where L is their horizontal extent, hence the “elliptical 507

dimension” Del controlling the rate of change of volumes of nonintermittent structures is 508

Del = 2+Hz. The 23/9 D model derives its name from Hz = 5/9, the result of Kolmogorov 509

scaling (H = 1/3) in the horizontal and Bolgiano-Obukhov scaling in the vertical (H = 510

3/5). In comparison, the popular quasi-linear gravity wave theories ((Gardner, 1994), 511

(Dewan, 1997)) have H = 1/3 in the horizontal and H = 1 in the vertical so that Hz = 1/3 512

and Del = 7/3. 513

Combining the results from the aircraft and the drop sondes and taking into account a 514

small apparent altitude dependence of the sonde intermittency exponents C1 (so as 515

estimate them at the 200 mb aircraft level), we obtain table 3. Note that the value for the 516

horizontal wind is given as the theoretical value 1/3 whereas the detailed leg by leg 517

analysis in (Lovejoy et al., 2009c) which fits two power laws (one for small, the other 518

for large scales) gives 0.26±0.07, 0.27±0.13 for the dominant small scale exponent 519

(transverse and longitudinal respectively). Given the strong coupling between the 520

aircraft trajectory and the wind, it seemed best to assume that these estimates support the 521

Kolmogorov value and it was used without an error bar. Similarly, the value of Hz for 522

Page 33: The Horizontal cascade structure of atmospheric fields

6/3/09 33

velocity is from the analysis of (Lovejoy et al., 2009c) which fits two power laws one 523

fixed at Hh = 1/3, the other (large scale one corresponding to Hv) varying. 524

It should be noted that although in table 3 we give the ratio of the C1 values, since 525

their values are small, their relative errors are large and consequently their ratios have 526

large uncertainties. Since the H’s are larger, the ratio Hz = Hh/Hv is more reliable than 527

C1h/C1v, indeed in the latter case the error is very hard to reliably estimate and is not 528

indicated except in the lidar case. The main conclusions are a) for T, log θ and v , the 529

exponent ratio Hz is close to 0.47, b) for T, log θ, B, Hz is within one standard deviation 530

of the 23/9 D model value Hz = (1/3)/(3/5) = 5/9, c) the value for the humidity is 531

somewhat larger. Note that finding virtually identical exponents for T and logθ is not 532

surprising since if the pressure is exactly constant, then the two have a one to one (albeit 533

nonlinear) relation. Finally, also using the Gulfstream 4 data, (Hovde et al., 2009) find H 534

= 0.45±0.03 for the humidity (see also Tuck 2008 for a possible explanation for the low 535

value). 536

If the ratios in table 3 are taken at face value then we are lead to the conclusion that 537

two or possibly three scale functions are required to specify the scale of atmospheric 538

structures. While this is certainly possible, let us for the moment underline the various 539

difficulties in obtaining the in situ estimates: the nontrivial vertical dropsonde outages, 540

the nontrivial aircraft trajectory fluctuations. In addition, detailed analysis of the altitude 541

dependence of the horizontal velocity exponent in (Lovejoy et al., 2007) indicates that 542

starting with the theoretical Boligano - Obukhov value 3/5 near the surface, that the 543

exponent increases somewhat with altitude to the value ≈ 0.75 at 10 - 12 km. Similarly, 544

Page 34: The Horizontal cascade structure of atmospheric fields

6/3/09 34

the humidity (and hence logθE) exponents may have both horizontal and vertical 545

variations which may account for their high Hz values (recall however that the values of 546

H are expected to be determined by dimensional analysis on fluxes; they are not expected 547

to have truly continuous variations). We should therefore regard these studies as only 548

early attempts to quantify the stratification. 549

550

551

T Logθ h v B

αh 1.78 1.82 1.81 1.94 1.83

αv 1.70 1.90 1.85 1.90 1.82

Hh 0.50±0.01 0.51±0.01 0.51±0.01 1/3 0.33±0.02

Hv 1.07±0.18 1.07±0.18 0.78±0.07 0.75±0.05 0.60±0.02

Hz=Hh/Hv 0.47±0.09 0.47±0.09 0.65±0.06 0.46±0.05 0.55±0.02

C1h 0.052±0.012 0.052±0.0100.040±0.012 0.04 0.076

C1v 0.072 0.071 0.091 0.088 0.11

Hz=C1h/C1v 0.72 0.71 0.44 0.45 0.69±0.2

Table 3: The above uses the estimate of the vertical Hv, C1v from sondes (table 2 in 552

(Lovejoy et al., 2009b)). The (horizontal) values for Hh for T, logθ, h are from table 2 553

(from 4 – 40 km, see above), for the lidar reflectivity B it is from (Lilley et al., 2008). 554

Finally, the C1 for v (velocity) is for the range 4- 40 km using trace moments. The 555

horizontal α values were for nonlinear fits of the K(q) for 0<q<3 (same range of scales). 556

557

Page 35: The Horizontal cascade structure of atmospheric fields

6/3/09 35

558

5. Conclusions: 559

Satellites and other remote methods make the determination of the scale by scale 560

statistical properties of atmospheric radiances relatively easy to establish. In comparison 561

our knowledge of the corresponding variability of the standard atmospheric fields (wind, 562

temperature humidity etc.) is quite meagre and is based almost entirely on problematic in 563

situ measurements. The fact that today’s in situ atmospheric data is often orders of 564

magnitude better than those available even twenty years ago - and is often easily 565

accessible - has unfortunately encouraged naïve applications. For example, in a series of 566

papers, we have argued that the extreme (and highly clustered) nature of data outages in 567

drop sonde profiles must be carefully taken into consideration when they are analysed 568

(Lovejoy et al., 2009b). Recently in a companion paper, we have argued that due to 569

strong coupling over wide ranges of scale between the wind and the trajectories, that 570

aircraft measurements require a theory of anisotropic scaling turbulence for their 571

interpretation (Lovejoy et al., 2009c). Whereas the classical interpretation of aircraft 572

wind statistics assumes that the turbulence is isotropic and interprets the observed break 573

at around 40 km as a transition between two isotropic regimes, in the reinterpretation, it is 574

spurious and is simply a transition from horizontally to vertically dominated parts of the 575

trajectory. 576

In this paper, we have attempted to exploit this new understanding in order to 577

both demonstrate and to quantify the wide range horizontal scaling of the key 578

atmospheric fields: wind, temperature, humidity, potential temperature. The basic 579

theoretical framework for such wide range scaling is anisotropic cascades; we 580

Page 36: The Horizontal cascade structure of atmospheric fields

6/3/09 36

demonstrate the horizontal cascade structure by directly analyzing the raw turbulent 581

fluxes finding that they all follow a multiplicative cascade structure with external scales 582

of the order of 10,000 km. Although this basic structure was present in all the analyzed 583

fields, we noted the presence of the three different regimes predicted on the basis of their 584

cross-spectra with the aircraft altitude and with the pressure. The deviations were fairly 585

strong for the wind and pressure which were strongly coupled with the trajectory, but 586

relatively small for the temperature, humidity and potential temperature which were only 587

weakly coupled. A rough summary is that at scales <≈ 4 km the trajectories vary 588

smoothly but are affected by aircraft roll and pilot/autopilot controls, for scales ≈ 4 - 40 589

km the aircraft follow a rough, intermittent fractal trajectory; finally for scales ≥ 40 km 590

the aircraft very closely follow the isobars which have significant vertical slopes. These 591

regimes are for the ensemble statistics; the transition points vary considerably from 592

trajectory to trajectory. 593

Refined analyses of the statistics was performed by considering the scale by scale 594

intermittency (characterized by the second order moment of the flux) as well as the scale 595

by scale smoothness of the fluctuations (characterized by the first order structure 596

function). It was argued that the optimum scale range for parameter estimation was ≈ 4 -597

40 km where the intermittency in the altitude was not strong, the effect of the 598

pilot/autopilot, rolls was weak and the constraint of flying on an isobar was not yet so 599

strong as to impose the verticals statistics on the wind field. We therefore used this range 600

to give refined parameter estimates. Perhaps one of the most surprising results was the 601

near identity of the scaling parameters H, C1, α for temperature, log potential temperature 602

and humidity, and in particular the non standard result H = 0.51±0.01 for all three fields 603

Page 37: The Horizontal cascade structure of atmospheric fields

6/3/09 37

which is extremely close to the fraction ½ but which nevertheless (apparently) still belies 604

theoretical explanation (see however Tuck 2008). 605

Finally, we compared the estimates of the horizontal exponents with those from 606

drop sondes of the corresponding vertical exponents. Overall, the ratios Hz for v, T, logθ, 607

h were not far from the 23/9 D model prediction 5/9 (based on energy flux domination of 608

horizontal statistics and buoyancy variance flux domination of vertical statistics), yet 609

there were apparently systematic deviations: v, T, logθ had Hz ≈ 0.47 and for h, Hz ≈0.65. 610

For the moment there is no satisfactory theoretical explanation for these results, although 611

some clarification may come from a comparison with the structure of atmospheric 612

reanalyses and (hydrostatic) numerical models which not only have nearly perfect 613

cascade structures in the horizontal, but also at least some similar (and non-standard) 614

parameters. For example they notably show that along isobars, the wind spectral 615

exponent is almost exactly β = 2.4 i.e. the same as that observed here (Stolle et al., 616

2009). The fact that we find qualitatively similar wide range cascade structures in the 617

simulations and in the atmosphere opens up new avenues for statistically verifying the 618

models: their scale by scale statistics should be the quantitatively the same as those of the 619

measurements. 620

It is remarkable that in spite of the current golden age of meteorological 621

observations that there is still no scientific consensus about the atmosphere’s basic scale 622

by scale statistical properties, in particular those of the dynamical (wind) field. However, 623

if we reinterpret the (systematically observed) transition from ≈ 5/3 to ≈ 2.4 spectral 624

scaling in terms of a single anisotropic scaling turbulence then there are no longer 625

seriously obstacles to the emergence of a consensus. To complete this emerging “new 626

Page 38: The Horizontal cascade structure of atmospheric fields

6/3/09 38

synthesis” (Lovejoy and Schertzer, 2009) the analyses presented here must be extended 627

to the time domain, to different altitudes, should be rechecked with different sensors and 628

in particular to aircraft flying on isoheights. Perhaps most importantly, completion of the 629

synthesis requires that we abandon the dogma of isotropy (or quasi isotropy) and embark 630

upon the systematic development of anisotropic but scaling theories of turbulence. 631

632

6. Acknowledgements: 633

We thank the crew of the Gulfstream 4 who gathered the data. This work was 634

performed for scientific purposes only; it did not enjoy any specific funding. 635

636

7. References: 637

638

J.T. Bacmeister et al., Stratospheric horizontal wavenumber spectra of winds, potnetial 639

temperature, and atmospheric tracers observed by high-altitude aircraft, J. 640

Geophy. Res. 101(1996), pp. 9441-9470. 641

Y. Chigirinskaya, D. Schertzer, S. Lovejoy, A. Lazarev and A. Ordanovich, Unified 642

multifractal atmospheric dynamics tested in the tropics Part 1: horizontal scaling 643

and self organized criticality, Nonlinear Processes in Geophysics 1(1994), pp. 644

105-114. 645

J. Cho and E. Lindborg, Horizontal velocity structure functions in the upper troposphere 646

and lower stratosphere i: Observations, J. Geophys. Res. 106(2001), pp. 10223-647

10232. 648

Page 39: The Horizontal cascade structure of atmospheric fields

6/3/09 39

E. Dewan, Saturated-cascade similtude theory of gravity wave sepctra, J. Geophys. Res. 649

102(1997), pp. 29799-29817. 650

X. Gao and J.W. Meriwether, Mesoscale spectral analysis of in situ horizontal and 651

vertical wind measurements at 6 km, J. of Geophysical Res. 103(1998), pp. 6397-652

6404. 653

C. Gardner, Diffusive filtering theory of gravity wave spectra in the atmosphere, J. 654

Geophys. Res., 99(1994), p. 20601. 655

C.S. Gardner, C.A. Hostetler and S.J. Franke, Gravity Wave models for the horizontal 656

wave number spectra of atmospheric velocity and density flucutations, J. 657

Geophys. Res. 98(1993), pp. 1035-1049. 658

S.J. Hovde, A.F. Tuck, S. Lovejoy and D. Schertzer, Vertical Scaling of the Atmosphere: 659

Dropsondes from 13 km to the Surface, Quart. J. of the Royal Meteor. Soc.(2009). 660

S.E. Koch et al., Turbulence and gravity waves within an upper-level front J. Atmos. Sci. 661

62(2005), pp. 3885-3908 662

M. Lilley, S. Lovejoy, D. Schertzer, K.B. Strawbridge and A. Radkevitch, Scaling 663

turbulent atmospheric stratification, Part II: empirical study of the the 664

stratification of the intermittency, Quart. J. Roy. Meteor. Soc.(2008), p. DOI: 665

10.1002/qj.1202. 666

M. Lilley, S. Lovejoy, K. Strawbridge and D. Schertzer, 23/9 dimensional anisotropic 667

scaling of passive admixtures using lidar aerosol data, Phys. Rev. E 70(2004), pp. 668

036307-036301-036307. 669

S. Lovejoy and D. Schertzer, Towards a new synthesis for atmospheric dynamics: space-670

time cascades, Atmos. Res. submitted(2009). 671

Page 40: The Horizontal cascade structure of atmospheric fields

6/3/09 40

S. Lovejoy, D. Schertzer and V. Allaire, The remarkable wide range scaling of TRMM 672

precipitation, Atmos. Res. 10.1016/j.atmosres.2008.02.016(2008a). 673

S. Lovejoy et al., Atmospheric complexity or scale by scale simplicity? , Geophys. 674

Resear. Lett. 36(2009a), pp. L01801, doi:01810.01029/02008GL035863. 675

S. Lovejoy, D. Schertzer, M. Lilley, K.B. Strawbridge and A. Radkevitch, Scaling 676

turbulent atmospheric stratification, Part I: turbulence and waves, Quart. J. Roy. 677

Meteor. Soc.(2008b), p. DOI: 10.1002/qj.1201. 678

S. Lovejoy, D. Schertzer and A.F. Tuck, Fractal aircraft trajectories and nonclassical 679

turbulent exponents, Physical Review E 70(2004), pp. 036306-036301-036305. 680

S. Lovejoy, A.F. Tuck, S.J. Hovde and D. Schertzer, Is isotropic turbulence relevant in 681

the atmosphere?, Geophys. Res. Lett. L14802, 682

doi:10.1029/2007GL029359.(2007). 683

S. Lovejoy, A.F. Tuck, S.J. Hovde and D. Schertzer, The vertical cascade structure of the 684

atmosphere and multifractal drop sonde outages, J. Geophy. Res. 114(2009b), pp. 685

D07111, doi:07110.01029/02008JD010651. 686

S. Lovejoy, A.F. Tuck, D. Schertzer and S.J. Hovde, Reinterpreting aircraft 687

measurements in anisotropic scaling turbulence, Atmos. Chem. Phys. Discuss., 688

9(2009c), pp. 3871-3920. 689

B.B. Mandelbrot, Intermittent turbulence in self-similar cascades: divergence of high 690

moments and dimension of the carrier, Journal of Fluid Mechanics 62(1974), pp. 691

331-350. 692

A.S. Monin and A.M. Yaglom, Statistical Fluid Mechanics, MIT press, Boston Ma 693

(1975). 694

Page 41: The Horizontal cascade structure of atmospheric fields

6/3/09 41

G.D. Nastrom and K.S. Gage, A climatology of atmospheric wavenumber spectra of 695

wind and temperature by commercial aircraft, J. Atmos. Sci. 42(1985), pp. 950-696

960. 697

E.A. Novikov and R. Stewart, Intermittency of turbulence and spectrum of fluctuations in 698

energy-disspation, Izv. Akad. Nauk. SSSR. Ser. Geofiz. 3(1964), pp. 408-412. 699

D. Schertzer and S. Lovejoy, The dimension and intermittency of atmospheric dynamics. 700

In: B. Launder, Editor, Turbulent Shear Flow 4, Springer-Verlag (1985), pp. 7-33. 701

D. Schertzer and S. Lovejoy, Physical modeling and Analysis of Rain and Clouds by 702

Anisotropic Scaling of Multiplicative Processes, Journal of Geophysical Research 703

92(1987), pp. 9693-9714. 704

D. Schertzer and S. Lovejoy, Universal Multifractals do Exist!, J. Appl. Meteor. 705

36(1997), pp. 1296-1303. 706

D. Schertzer, S. Lovejoy and D. Lavallée, Generic Multifractal phase transitions and self-707

organized criticality. In: J.M. Perdang and A. Lejeune, Editors, Cellular 708

Automata: prospects in astronomy and astrophysics, World Scientific (1993), pp. 709

216-227. 710

J. Stolle, S. Lovejoy and D. Schertzer, The stochastic cascade structure of deterministic 711

numerical models of the atmosphere, Nonlin. Proc. in Geophys. 712

(submitted)(2009). 713

P. Szépfalusy, T. Tél, A. Csordas and Z. Kovas, Phase transitions associated with 714

dynamical properties of chaotic systems, Physical Review A 36(1987), p. 3525. 715

Page 42: The Horizontal cascade structure of atmospheric fields

6/3/09 42

A.F. Tuck, S.J. Hovde and T.P. Bui, Scale Invariance in jet streams: ER-2 data around 716

the lower-stratospheric polar night vortex, Q. J. R. Meteorol. Soc. 130(2004), pp. 717

2423-2444. 718

A.M. Yaglom, The influence on the fluctuation in energy dissipation on the shape of 719

turbulent characteristics in the inertial interval, Sov. Phys. Dokl. 2(1966), pp. 26-720

30. 721

722

723

Page 43: The Horizontal cascade structure of atmospheric fields

List of Figure Captions:

Fig. 1 a: The normalized moments q = 0.2, 0.4, … 3, for aircraft altitude, z; log10λ

corresponds to 20,000 km (= Lref).

Fig. 1b: Same as fig. 1a but for the fields strongly affected by the trajectories: the

longitudinal wind (top left a), the transverse wind (top right, b), pressure (c, lower left

right).

Fig. 1c: Same as 1b but for the fields that are relatively unaffected by the trajectory:

temperature (top left), relative humidity (top right), log potential temperature (lower left

right).

Fig. 2 a: Structure functions of order q = 0.2, 0.4, ..2.0 for the aircraft altitude. Distances

Δx are in km. Separate sets of regression lines are shown for both small and large scales

in order to emphasize the change with scale.

Fig. 2 b: Same as fig. 2a but for the fields strongly affected by the aircraft trajectory: the

longitudinal wind (upper left), the transverse wind (upper right) and the pressure (lower

left).

Page 44: The Horizontal cascade structure of atmospheric fields

Fig. 2 c: Same as for 2a except for those fields relatively unaffected by the aircraft

trajectories: the temperature (upper left), the humidity (upper right) and the log potential

temperature (lower left).

Fig. 3a: Second order trace moments (right column) and their logarithmic derivatives

(estimates of K(2), an intermittency index), left column. Top row: wind, longitudinal

(blue) transverse (red). The region within the dashed lines has particularly low K(2) (≈

0.05). Middle row: pressure (blue), altitude (red). Bottom, temperature (blue), humidity

(red), log potential temperature (gold), reference lines K(2) = 0.12. Log10λ = 0

corresponds to 20,000 km (= Lref).

Fig. 3b: First order structure functions (right column) and their logarithmic derivatives

(estimates of H), left column. Top row: wind, longitudinal (blue) transverse (red), the

reference lines correspond to the exponents 1/3, 0.75. Middle row: pressure (blue),

altitude (red), the reference lines correspond to H = 1 i.e. constant mean slopes. Bottom,

temperature (blue), humidity (red), log potential temperature (gold), reference lines H =

0.5.

Fig. 3c: Spectra (right column) and their logarithmic derivatives (estimates of -β), left

column. Top row: wind, longitudinal (blue) transverse (red). The thin reference lines

correspond to β = 5/3, 2.4. Middle row: pressure (blue), altitude (red) with reference

lines corresponding to β = 5/3. Bottom, temperature (blue), humidity (red), log potential

Page 45: The Horizontal cascade structure of atmospheric fields

temperature (gold), reference lines β = 2. Vertical dashed lines indicate the region with

scales 4 – 40 km where the velocity intermittency is low.

Fig. 4: The K(q) functions for altitude and pressure (upper, lower blue), temperature,

humidity, log potential temperature (lower, middle, upper green), longitudinal, transverse

wind (upper, lower, red).