57
1 Modulation of Mean Wind and Turbulence in the 1 Atmospheric Boundary Layer by Baroclinicity 2 3 Mostafa Momen 1,2* , Elie Bou-Zeid 1 , Marc B. Parlange 3 , and Marco Giometto 2 4 1 Department of Civil and Environmental Engineering, Princeton University, Princeton, NJ, USA 5 2 Department of Civil Engineering and Engineering Mechanics, Columbia University, New York, 6 NY, USA 7 3 Department of Civil Engineering, Monash University, Melbourne, Australia 8 9 * Corresponding author address: Mostafa Momen, 618A SW Mudd, Department of Civil 10 Engineering and Engineering Mechanics, Columbia University, New York, NY 10027. Email: 11 [email protected] 12

Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

1

Modulation of Mean Wind and Turbulence in the 1

Atmospheric Boundary Layer by Baroclinicity 2

3

Mostafa Momen1,2*, Elie Bou-Zeid1, Marc B. Parlange3, and Marco Giometto2 4

1Department of Civil and Environmental Engineering, Princeton University, Princeton, NJ, USA 5

2Department of Civil Engineering and Engineering Mechanics, Columbia University, New York, 6

NY, USA 7

3Department of Civil Engineering, Monash University, Melbourne, Australia 8

9

* Corresponding author address: Mostafa Momen, 618A SW Mudd, Department of Civil 10

Engineering and Engineering Mechanics, Columbia University, New York, NY 10027. Email: 11

[email protected] 12

Page 2: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

2

Abstract 13

This paper investigates the effects of baroclinic pressure gradients on mean flow and turbulence 14

in the diabatic atmospheric boundary layer (ABL). Large-eddy simulations are conducted where 15

the direction of the baroclinicity, its strength, as well as the surface buoyancy flux, are 16

systematically varied to examine their interacting effects. Results indicate that the thermal wind 17

vector, which represents the vertical change in the geostrophic wind vector resulting from 18

horizontal temperature gradients, significantly influences the velocity profiles, the Ekman turning, 19

and the strength and location of the low-level jet (LLJ). For instance, cold advection and positive 20

(negative) geostrophic shear increased (decreased) friction velocity and the LLJ elevation. Given 21

the baroclinicity strength and direction under neutral conditions, a simple analytical model is 22

proposed and validated here to predict the general trends of baroclinic mean winds. The baroclinic 23

effects on turbulence intensity and structure are even more intricate, with turbulent kinetic energy 24

(TKE) profiles displaying an increase of TKE magnitude with height for some cases. When a fixed 25

destabilizing surface heat flux is added, a positive geostrophic shear favors stream-wise aligned 26

roll-type structures, which are typical of neutral ABLs. Conversely, a negative geostrophic shear 27

promotes cell-type structures, which are typical of strongly unstable ABLs. Furthermore, 28

baroclinicity increases shear in the outer ABL and tends to make the outer flow more neutral by 29

decreasing the Richardson flux number. These findings are consequential for meteorological 30

measurements and the wind-energy industry, among others: baroclinicity alters the mean wind 31

profiles, the TKE, coherent structures, and the stability of the ABL and its effects need to be 32

considered. 33

34

Page 3: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

3

1 Introduction 35

Modeling geophysical flows to accurately predict the weather and climate requires a 36

comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 37

imposes the bottom boundary conditions on larger scale atmospheric flows; it is the main source 38

of thermal energy as well as trace gases, and the main sink of momentum and kinetic energy for 39

the whole atmosphere. ABL flows are mainly driven by horizontal pressure gradients and 40

buoyancy forces, with two additional important components: the friction and Coriolis forces. In 41

real-world ABLs, the pressure-gradient force can vary in time (unsteadiness) and height 42

(baroclinicity). However, the large majority of previous studies of ABL flows have focused on 43

barotropic conditions, with a time-constant horizontal pressure gradient (see Blackadar 1957, 44

Deardorff 1972, Ekman 1905, Monin & Obukhov 1954). This idealization has also been the focus 45

of most books on the subject, since it facilitated the development of a theoretical understanding of 46

ABL behavior and turbulence closure models. Nevertheless, with the increasing need for high 47

accuracy models for the thermo-fluid dynamics of the ABL in a very wide range of applications – 48

including wind energy, urban meteorology, and pollutant dispersion – the need to incorporate the 49

additional layers of complexity (that characterizes real-world conditions) becomes pressing. Some 50

of these complexities – including heterogeneity in the surface properties (e.g. complex terrains 51

Chow and Street 2009), clouds and moisture flux in the ABL (Holtslag et al. 2013), capping 52

inversion layer (Pedersen et al. 2014), and unsteadiness (see e.g. Momen and Bou-Zeid 2015, 53

2016b, 2017a; Van de Wiel et al. 2010) – are being increasingly examined in the literature. In 54

comparison, the departure from barotropic conditions has received less attention. 55

Page 4: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

4

A barotropic environment arises when horizontal density variations are only induced by 56

horizontal pressure variations. On the other hand, in a baroclinic environment, density varies due 57

to both horizontal pressure and temperature changes. As a consequence, the gradients of pressure 58

and density are not necessarily aligned under baroclinic conditions. The effects of this 59

misalignment manifest themselves in the momentum conservation equation as a modification in 60

the pressure gradient force term – ρ−1 ∇p. This can alter the mean acceleration and velocity in 61

various regions of the flow, thus affecting the turbulence structure and statistics. There are other 62

impacts of baroclinicity on vorticity generation (Kundu & Cohen 2008), on large-scale horizontal 63

instability in geophysical cold filaments (Gula et al. 2014), and other large-scale atmospheric 64

instabilities (Vallis 2006) that we do not delve into in this study. For instance, baroclinicity 65

influences the vorticity generation of flows (Kundu & Cohen 2008) that could be important for 66

studying the dynamics of gravity currents (Cantero et al. 2008; Momen et al. 2017b). Since air 67

density inherently depends on temperature, baroclinicity will emerge whenever temperatures in 68

the atmosphere vary on horizontal scales commensurate with those of density and pressure. Such 69

conditions occur quite often (Wyngaard 2010) especially at mid-latitudes, for example, from 70

strong horizontal gradients in temperature that form across fronts (Rizza et al. 2013) and across 71

land-water or ice-water boundaries (Floors et al. 2015). However, despite observational evidence 72

and theoretical support that point to the prevalence of baroclinic conditions during such critical 73

meteorological events (Sheppard et al. 1952; Venkatesh and Csanady 1974; Hoxit 1974; Arya and 74

Wyngaard 1975), there has been a remarkably limited number of studies examining the impact of 75

such conditions on the ABL. Bridging this knowledge gap is particularly critical since baroclinic 76

conditions might influence important processes and modulate the development of these events, 77

and since baroclinicity is often ignored when observational data are being analyzed. 78

Page 5: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

5

A few numerical studies have examined some aspects of the baroclinic ABLs. In particular, 79

large-eddy simulations (LES) were employed to investigate the mean and turbulence profiles in 80

the ABL and determine the unique characteristics of baroclinic flows. Brown (1996) performed 81

baroclinic LES to evaluate the performance of two simple closures and showed that the presence 82

of shear in the geostrophic winds does not significantly impact this performance. However, he 83

demonstrated that with greater shear in the geostrophic winds of a hybrid baroclinic case (negative 84

shear + cold advection, these will be further explained later) the shear production of turbulent 85

kinetic energy (TKE) becomes negligible at around 0.25zi (zi is the height of the base of the 86

inversion layer) and the transport terms are no longer negligible in the TKE budget above this 87

elevation. Strong baroclinicity might therefore challenge the validity of diagnostic stress-TKE 88

relationships, or 1.5-order closure schemes, which are based on the assumption that production 89

and dissipation are the leading order terms in the TKE budget equation. Furthermore, Brown 90

(1996) showed that an increase of geostrophic shear leads to significant changes in the stress and 91

velocity profiles. Baroclinicity, in conjunction with stability and earth’s rotation, were shown to 92

be the main governing parameters for determining the equilibrium depths of neutral and stable 93

ABLs; Zilitinkevich and Esau (2003) presented diagnostic equations to capture these equilibrium 94

heights and validated their results against LES for neutral and stable ABLs. 95

Furthermore, baroclinicity is known to modify not only wind and temperature profiles, but also 96

the second-order turbulence statistics in the ABL (Sorbjan 2004). For instance, Sorbjan (2004) 97

used LES to improve the convective scaling introduced by Deardorff (1970) in unstable ABLs by 98

recognizing the presence of two different regimes in the mixed layer. The first regime was 99

identified in the core part of the mixed layer, which is controlled by surface heating and can be 100

described by the convective scaling of Deardorff (1970), while the second regime characterizes 101

Page 6: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

6

the top of the mixed layer and is controlled by thermal stratification and shear. The latter does not 102

obey Deardorff’s scaling. Pedersen et al. (2013) also investigated the effect of unsteady and 103

baroclinic forcing on the wind profiles with the help of LES and compared their results with two 104

case studies of the daytime ABL. In both cases, they found that including height- and time- 105

variations in the applied pressure gradient forcing improves agreements with measured wind 106

speeds. Previous studies also indicate that baroclinic conditions will modify entrainment at the top 107

of the ABL (Brown 1996; Hess 2004) and the cross-isobaric and cross-isothermal wind angles, 108

thus affecting Ekman pumping and frontogenesis (Sheppard et al. 1952; Venkatesh and Csanady 109

1974; Arya and Wyngaard 1975), as well as the strength of low level jets (Floors et al. 2015). 110

These results have highlighted the stark differences between barotropic and baroclinic ABLs, and 111

motivate the need for additional studies to better understand these differences and their 112

implications for accurate modeling of real-world ABLs. 113

A horizontally-homogeneous pressure gradient can be represented as an equivalent horizontal 114

geostrophic wind, with components Ug and Vg. However, while in a barotropic ABL these 115

components are independent of height z, in a baroclinic ABL they can vary independently with z 116

( ) ( )( ) ( ) ( )

1 1, , , , , ,g g

p pU z t V z t z t z t

f y f x

, (1) 117

where f is the Coriolis parameter (here taken as 1.394×10−4 s−1 at 73oN; its absolute value is not 118

particularly relevant since we non-dimensionalize all of our results); ρ is the fluid density; p is the 119

pressure; x, y, and z are positions along the horizontal and vertical directions; and the overbar 120

denotes Reynolds averaging (here surrogated for by space and/or time averaging as appropriate). 121

Differentiating Eq. (1) with respect to z and using the hydrostatic approximation and the ideal gas 122

law (Brown 1996), the equations of the thermal wind, defined as the vertical gradient of the 123

Page 7: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

7

geostrophic wind, can be written in terms of the time-average (equivalent to a Reynolds average 124

here and denoted by an overbar) of the potential temperature θ: 125

,g g

r r

U Ug

z f y z

= − +

(2.a) 126

,g g

r r

V Vg

z f x z

= + +

(2.b) 127

where θr (K) is the Boussinesq reference potential temperature, corresponding to the reference 128

density ρr (kg/m3). The last terms on the right-hand side of the above equations are negligible if 129

the vertical temperature gradient is close to the adiabatic lapse rate (Arya & Wyngaard 1975). 130

Therefore, the dominant terms in the geostrophic shear are typically the horizontal temperature 131

gradients.

132

This phenomenon has led many researchers, e.g. Hess (1973) and Arya & Wyngaard (1975), to 133

develop theoretical frameworks and numerical models using baroclinicity parameters based on the 134

horizontal temperature gradients in Eqs. (2) such as: 135

(3) 136

where u* is the friction velocity at the surface, zi the ABL depth, and subscript 0 refers to quantities 137

evaluated at z = z0 (the aerodynamic roughness length). Equivalently, these parameters could be 138

presented as the magnitude of the vector and its orientation angle 139

. Similarity or approximate analytical models (based on the solution of the 140

Ekman equations with height-varying geostrophic wind) that account for these two parameters 141

have been tested and have shown improved skill in matching observational case studies (Venkatesh 142

and Csanady 1974; Arya and Wyngaard 1975; Berger and Grisogono 1998; Tan 2001; Hess 2004; 143

Page 8: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

8

Djolov et al. 2004; Nieuwstadt 1983) or have even explained observations that simply cannot exist 144

in a barotropic ABL such as wind blowing towards higher pressure (Sun et al. 2013). 145

Nevertheless, the emergence of two new parameters is in fact a major barrier to the 146

comprehensive understanding of baroclinic ABLs. Because of constraints in real-world 147

observations, field experimental studies have only investigated a limited range of baroclinic 148

regimes, without offering a full picture that can lead to a comprehensive understanding. Simulation 149

efforts, particularly those focusing on idealized LES that systematically vary Mx,0 and My,0, would 150

be better suited to provide a comprehensive rigorous grasp on the subject. However, there have 151

only been few such studies (e.g. Brown 1996; Sorbjan 2004; Brown et al. 2006) that systematically 152

probed the parameter space of this problem and the effects on mean and turbulence behavior, 153

compared to tens of studies on barotropic ABLs conducted each year. Other efforts have applied 154

LES to specific barotropic/baroclinic observational periods (e.g. LES of diurnal cycle with varying 155

geostrophic forcing Kumar et al. 2010; Basu et al. 2008; Pedersen et al. 2013). These studies could 156

use the directly measured pressure gradient field or large-scale horizontal temperature gradient to 157

obtain the real-world geostrophic wind profiles, mainly because of limitations in measurement 158

data, which resulted in some uncertainty. They had to resort to mesoscale simulations (or strong 159

assumptions about the baroclinicity in the measured data) to deduce these parameters. More 160

importantly, they did not aim to develop comprehensive theories for the baroclinic ABL by 161

investigating the full parameter space. Furthermore, it is important to note that these efforts, despite 162

their pioneering significance, were not able to fully probe the interaction of baroclinicity and 163

various stability conditions (for example these studies did not examine the stable baroclinic ABL). 164

This could be arguably because baroclinicity adds two new dimensionless parameters (Eq. (3)) to 165

Page 9: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

9

the problem compared to barotropic conditions. Another reason for overlooking stable 166

stratifications in earlier studies might be related to the higher resolution required in such 167

simulations. Therefore, important aspects of the problem, such as the influence of baroclinicity on 168

the Ekman turning, the low-level jet, TKE budget profiles, and large-scale turbulence structures, 169

are yet to be probed. 170

This gap in knowledge motivates the present study where we investigate the baroclinic ABL by 171

preforming a suite of LES including stability effects. In particular, we intend to address the 172

following questions: 173

1) What are the most significant changes in the ABL mean wind and turbulence characteristics 174

resulting from baroclinicity? 175

2) How do M0 and β jointly regulate the effects of the baroclinicity, and what are the 176

distinctions between weakly and strongly baroclinic ABLs? 177

3) How does static stability modulate baroclinic effects on the ABL? 178

The LES technique and numerical details of the code are described in the next section. In 179

Section 3, we examine the impacts of baroclinicity on the mean and turbulence characteristics of 180

neutral ABLs. Then, we develop an analytical model for neutral baroclinic ABLs that predicts the 181

velocity profiles, given baroclinicity strength and direction, and validate its solutions against LES 182

results for five different baroclinic scenarios and a barotropic case. In Section 4, we continue our 183

analysis by examining the interacting effects of baroclinicity and stability on the mean and 184

turbulence behavior. A summary and concluding remarks follow in Section 5. 185

Page 10: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

10

2 Details of the Numerical Simulations 186

Governing equations 187

We use the LES technique, which is well established by now, and thus we will only detail the 188

aspects of the numerical simulations that are unique to the baroclinic ABL. Since our LES code 189

(detailed in Bou-Zeid et al. 2005) uses pseudo-spectral numerical schemes in the horizontal 190

directions, with embedded periodic boundary conditions, some modifications to the equations are 191

required to avoid solving for the resolved potential temperature θ, the mean of which will be 192

spatially variable in the horizontal directions and hence non-periodic under baroclinic conditions. 193

In line with Sorbjan (2004), we first define the baroclinic horizontally-varying, time-averaged 194

mean field Θ(x, y, z) that can be assumed to have constant horizontal gradients ∂Θ/∂x and ∂Θ/∂y 195

(horizontal temperature variability scale ≫ turbulent eddy scale). Then we define its horizontal 196

mean ⟨Θ⟩p (x-y planar averaging is denoted by angle brackets with subscript p) that varies in z, and 197

the corresponding spatial perturbation Θ′′ = Θ − ⟨Θ⟩p; note that ∂Θ/∂x = ∂Θ′′/∂x and 198

∂Θ/∂y = ∂Θ′′/∂y. We can now define the modified turbulent instantaneous temperature 199

θ* = θ − Θ′′ = θ + ⟨Θ⟩p − Θ, which will be horizontally periodic. By time averaging the previous 200

relation we obtain ⟨θ*⟩t =⟨Θ⟩p (time averaging is denoted by angle brackets with subscript t). The 201

budget equation of θ* can then be easily derived from that of θ (see Eq. 7 below) after noting that 202

the advective terms are related by 203

, (4) 204

where xi is the position vector and is the velocity vector. Here, we have 205

assumed that ∂Θ ′′/∂ z = 0, i.e. the spatial perturbations of the time-averaged baroclinic 206

Page 11: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

11

temperatures are constant with height, which is equivalent to assuming that the horizontal 207

temperature gradient is height invariant. We note again that θ* is periodic and can thus be solved 208

for numerically using our pseudo-spectral code. 209

The buoyancy force appearing in the momentum equation will now be proportional to the 210

potential temperature perturbation from the mean of the isothermal (of constant potential 211

temperature) surface since this is the surface of neutral buoyancy for the parcel. In barotropic 212

environments, this is taken as the mean over a horizontal plane; however, that cannot be applied 213

to a baroclinic environment. The perturbation from the mean over the isothermal surface on which 214

the parcel lies, i.e. from Θ, is needed. The buoyancy perturbation temperature is therefore 215

θ′ = θ – Θ = θ* – ⟨Θ⟩p, but since ⟨θ*⟩t =⟨Θ⟩p, then θ′ = θ* – ⟨θ*⟩t ≡ θ*′. 216

At this stage, the modified governing equations of the LES can be written for the filtered 217

(denoted by a tilde) quantities as the continuity equation 218

, (5) 219

the momentum-conservation equation 220

, (6) 221

and the thermal energy conservation for potential temperature perturbation (θ̃*)

222

, (7) 223

where is the deviatoric part of the sub-grid scale (SGS) stress tensor 224

; p* is a modified pressure defined as ; θr is the reference 225

Page 12: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

12

temperature (the same for θ* and θ); g is the gravitational constant; and is the 226

SGS heat-flux vector. Both the SGS deviatoric stress and the SGS heat flux are modeled using an 227

eddy-diffusivity parametrization; specifically, a dynamic Lagrangian scale-dependent 228

Smagorinsky model is used to determine the SGS eddy-viscosity νT (Bou-Zeid et al. 2005) and 229

model τij. Then a constant SGS Prandtl number, Prsgs = 0.4, is invoked to model 230

(see Kumar et al. 2010). The LES equations are solved in non-231

dimensional form, using a-priori known parameters such as zi, and G0 ≡ (Ug,02 + Vg,0

2)1/2, the 232

magnitude of the geostrophic wind at the surface (a reference value of 8 m/s is taken for G0 here). 233

For consistency, we use a fixed value of zi = 800 m in stable cases and zi = 1000 m in other 234

simulations (representing the base of the capping inversion layer that we impose) as a fixed length 235

scale for non-dimensionalizing the heights in all plots to facilitate the comparison of cases in a 236

dimensional framework. However, the effective ABL depth might be lower than zi, particularly 237

under stable conditions and when there is a negative geostrophic shear in neutral cases; we will 238

use this effective height for some cases in the reduced-order model and evaluate its performance. 239

particularly under stable conditions. Hence, the results reported in the paper are all normalized 240

with these fixed scales. Further details on the LES model, the SGS model, and the diabatic 241

simulation setup can be found in Albertson and Parlange (1999), Bou-Zeid et al. (2005), and 242

Momen and Bou-Zeid (2017a,b). The approach used for initializing the code, the domain size, and 243

the boundary conditions are described in Appendix A. We note that the results of the code have 244

previously been validated for many applications, e.g. for barotropic neutral, unstable and stable 245

ABLs (Bou-Zeid et al. 2005; Kumar et al. 2010; Huang and Bou-Zeid 2013), and the new minor 246

changes in the forcing and the temperature equation should not impact the skill of the code in 247

Page 13: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

13

numerically integrating the equations. Validating the code against observational data in baroclinic 248

conditions is thus not deemed necessary, and could in fact be misleading. To have a consistent 249

comparison with observations, inputs such as the surface heat flux over ~ 100 km2, the mesoscale 250

horizontal temperature gradients (~ 100 km), and geostrophic wind or horizontal pressure gradient 251

data are needed, but these are not typically provided all together. 252

Baroclinicity parameters 253

Following the approach of Arya & Wyngaard (1975), we define two baroclinicity parameters 254

based on the magnitude of the two normalized components of the thermal wind vector at the 255

surface (strength parameter) and its orientation angle. However, we rescale the velocities by G0 256

instead of u*, since this is consistent with the normalization of our LES code equations and since 257

we are more interested in outer layer profiles (see Section 3.a), yielding 258

2 2

00 00

gi gzS

U V

zG z

+

, (8.a) 259

0

1

0

tang gV U

z z −

. (8.b) 260

Note that S0 and are here expressed in terms of surface parameters. In line with previous studies 261

(Brown 1996; Sorbjan 2004), we assume the second vertical derivative of the geostrophic wind to 262

be zero. This is again equivalent to assuming that the horizontal temperature gradient is height 263

invariant (see Eq. 2) and results in S (z) = S0 for any z. The subscripts denoting surface values of 264

the gradients are nevertheless included here for clarity. 265

Page 14: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

14

Suite of large-eddy simulations 266

Table 1 lists the simulations conducted in the paper for various values of S0, β, and surface 267

buoyancy flux. To understand the cases setup, we note that based on the definition of geostrophic 268

velocity, Eq. (1), it is always perpendicular to the horizontal pressure gradient and under barotropic 269

conditions the horizontal pressure and density gradients are parallel to each other . 270

If the horizontal temperature gradient happens to be aligned with the surface pressure gradient 271

(sometimes referred to as equivalent barotropic), the effect of is simply to 272

increase (positive shear) or decrease (negative shear) with height, but does not cause directional 273

shear. In such cases, is parallel to the isotherms and causes no heat advection (there could still 274

be heat advection by other processes such as ageostrophic winds). On the other hand, if 275

, then (note that , i.e. geostrophic wind is parallel to the isobars, is true at all 276

heights) and is thus perfectly normal to the isotherms, resulting in a highly effective scalar 277

advection by the surface geostrophic wind (cold advection if and point in the same 278

direction or warm advection if they point in opposite directions). In the cold and warm advection 279

cases, Vg is increased and decreased, respectively, while Ug remains constant. However, since there 280

are no physical constraints on the relative orientation between and (e.g. see FIG. 1c), 281

most real cases are expected to be hybrids of these 4 idealized scenarios (positive and negative 282

geostrophic shear and cold and warm advection). A schematic illustration of the explanation above 283

is featured in FIG. 1. 284

Page 15: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

15

Such baroclinic cases can often occur in nature. For instance, FIG. 1(c) depicts a schematic of 285

cold air and warm air advection systems in the US that would indicate baroclinicity. The wind 286

flows parallel to isobars and leads to advection of cold air toward the Great Plains and Gulf coast 287

(cold advection). Warm air is also pushed toward the northeastern US. The thermal boundary 288

between the cold and warm air (blue line) forms a cold front (boundary moving towards the warm 289

side) and similarly a warm front (boundary moving towards cold side) is found in New England. 290

Then, a low-pressure (storm) center is formed where warm and cold fronts meet. Baroclinicity is 291

also present in coastal regions. 292

Averaging in time and over x-y planes (the latter operation should only be applied to the 293

temperature perturbation θ* and not to the original temperature θ) is used as a surrogate for 294

Reynolds averaging; temporal averaging is done over the last τABL to get the converged statistics, 295

corresponding to an inertial period of the flow. The inertial time scale of ABLs is often 296

characterized by the Coriolis frequency (Tennekes and Lumley 1972) as τABL ≡ 2π/f 297

(≈ 12.5 hours here). 298

Furthermore, to ensure that our baroclinic numerical setup is implemented correctly, we 299

compared our simulations with other available numerical results as a check on the consistency of 300

the simulations and to ascertain that our findings can be compared to those other studies. We note 301

that these numerical results are not necessarily correct (refer to discussion on validation above), 302

but we do these comparisons for consistency checks and the good match does support the validity 303

of the simulation accuracy. In Appendix B, we show that our wind-velocity profiles compare 304

favorably with the positive-shear and hybrid cases of Brown (1996), where a similar SGS model, 305

forcing, and domain size were used. 306

Page 16: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

16

Table 1 about here 307

Figure 1 about here 308

Although we incorporate the horizontal temperature advection terms in the thermal energy Eq. 309

(7), we observed that their effects are small in our domain (see supplemental material). As Brown 310

(1996) also notes, the advective heat flux associated with horizontal temperature gradients is 311

negligible compared to the total heat flux (advective + surface), and under diabatic conditions does 312

not have a significant effect on the flow. As shown in table 1, the horizontal temperature gradient 313

is on the order of 10–5 K/m. 314

3 The effects of baroclinicity on the neutral ABL 315

In this section, we examine the influence of baroclinicity strength (S0) and direction (β) on the 316

mean and turbulence characteristics of the ABL under neutral conditions. 317

Implications for the mean velocity 318

The ABL adapts itself to the changes in the imposed pressure gradients with height. FIG. 2 319

shows these effects for all the neutral baroclinic cases versus a corresponding barotropic 320

simulation. As this figure indicates, not only the strength of the baroclinicity, but also its direction 321

can significantly change the ABL mean profiles compared to their reference barotropic 322

counterparts (FIG. 2(f)). Some of the resulting wind profiles have trends that were rarely examined. 323

For example, the presence of a negative shear in the geostrophic wind (β=135o and 180o, in FIG. 324

2(c) and FIG. 2(d)) results in jet-like velocity profiles that can decrease with height (to be discussed 325

further in Section 4.a). In general, the velocity vectors in the ABL tend to the magnitude and 326

Page 17: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

17

direction of the geostrophic wind vectors, in particular away from the surface where the turbulent 327

friction forces induced by the surface are reduced. 328

Figure 2 about here 329

Baroclinicity also strongly influences the Ekman turning. The hodographs of the mean 330

velocities of all neutral cases are plotted in FIG. 3. FIG. 3(f) exhibits the much-studied barotropic 331

ABL (the classic Ekman spiral). Contrasting the barotropic hodograph (added as a gray line in 332

subplots) with the baroclinic ones highlights how profoundly baroclinicity impacts the mean 333

velocity field. The hodographs are closer to the barotropic case near the surface, as could have 334

been anticipated given that the dominant term in the near-surface region is the turbulent stress 335

divergence and pressure effects play a lesser role. In addition, the imposed pressure gradients are 336

closer to the barotropic value near the wall, whereas they depart more significantly from their 337

barotropic reference away from the surface. More importantly perhaps, baroclinicity can result in 338

a significant increase in the v component (up to four time larger than the barotropic reference), 339

which at the surface crosses the isobars in our considered cases. This implies a similar impact on 340

Ekman pumping and significant implications at larger – meso to synoptic – scales (see Section 4.a 341

for more details). Even if the velocities are normalized by the geostrophic wind magnitude at 342

higher levels, G(z), the differences between the barotropic and baroclinic cases remain substantial. 343

This is illustrated in a graph in the supplemental material similar to FIG. 3, but where we normalize 344

wind velocities with their corresponding geostrophic wind at each height, G(z), instead of G0. 345

Figure 3 about here 346

Page 18: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

18

We note that an alternative velocity scale that could be used for normalization is the friction 347

velocity, u*. This would result in a better collapse of profiles in the surface layer i.e. z/zi ⪅ 0.2 (an 348

example is shown in the supplemental material). However, it will lead to larger discrepancies 349

between cases in the regions away from the wall, which are the main focus of this paper. On the 350

other hand, measuring the real-world G(z) is typically more difficult than G0, and normalizing by 351

G(z) impacts the profile shapes, thus making profiles more difficult to compare with other papers 352

in the literature. Hence, we primarily use G0 to normalize our results in the paper. 353

Assessment of a simple model for baroclinic mean winds 354

The development and validation of simple theoretical models describing the vertical structure of 355

the mean wind in barotropic ABLs have been a main focus of ABL research. Most such models 356

(Mahrt and Schwerdtfeger 1970; Grisogono 1995) are based on a priory prescribed, spatially-357

variable, eddy-viscosity profiles, which allow for an analytical treatment. The predictive capability 358

of many of these models has not been evaluated for the full parameter space of baroclinic ABLs. 359

In this section, the skills of a simple one-dimensional eddy-viscosity model are quantified via 360

comparison with LES results for the neutrally-stratified baroclinic ABL. Applying the time and 361

horizontal spatial averaging operation to Eq. (6), and modeling turbulent fluxes via K-theory (Stull 362

1988), we obtain the following equations for the steady-state, neutrally-stratified, baroclinic ABL 363

d

dzn

T

tot du

dz

æ

èç

ö

ø÷ - f V

g ,0+

¶Vg(z)

¶zz0

z

ò dz - væ

èç

ö

ø÷ = 0, (9.a) 364

d

dzn

T

tot dv

dz

æ

èç

ö

ø÷ + f U

g ,0+

¶Ug(z)

¶zz0

z

ò dz - uæ

èç

ö

ø÷ = 0. (9.b) 365

Page 19: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

19

where the vertical gradients of geostrophic winds in our considered cases can be represented in 366

terms of S0, Ug,0, Vg,0, and β as and 367

. Ug,0, Vg,0, S0, and β are prescribed according to 368

Table 1, to allow for a proper comparison with the corresponding LES cases. Equations (9) are a 369

system of second-order ordinary differential equations (ODEs) for the u and v velocities, functions 370

of z in the interval [z0, Lz], where Lz is the height of the domain (see Appendix A). (m2/s) is 371

the total kinematic eddy viscosity, a height-dependent, dimensional model parameter that needs to 372

be prescribed to account for the vertical momentum diffusion due to turbulent motions. The first 373

term in the equation represents the turbulent stress, the second term is the surface level pressure 374

gradient forcing and the third is the change in pressure gradient forcing with height, while the last 375

term represents the Coriolis force. A similar eddy viscosity to O’Brien’s profile (O’Brien 1970), 376

which was proposed for use in models of planetary boundary layers in both barotropic and 377

baroclinic situations (see e.g. Nieuwstadt 1983), is adopted here 378

(10.a) 379

, (10.b) 380

where κ is the von Kármán constant (here 0.4) and is a small constant diffusivity that 381

does not influence the solution in the ABL significantly. Here, we do not impose a specific value 382

for u*; it is found in an iterative approach by applying the log-law at the first two nodes (u* = κz 383

Page 20: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

20

du/dz) where baroclinic departure from that law is insignificant (see profiles normalized by u* in 384

supplementary material). The chosen functional form of complies with the LES setup, where 385

a strong inversion layer is applied at z > zi, resulting in negligible magnitudes at such heights. 386

Then, we substitute the proposed eddy-viscosity formula of Eq. (10) into Eqs. (9) and solve the 387

system of ODEs iteratively to obtain the velocity profiles. Note that in this reduced model, we do 388

not tune any parameter when comparing with LES, and the system of equations is closed. 389

Within the constant-flux layer, i.e. z ⪅ 0.1 zi, baroclinicity effects are weak, nT

tot ≈ u* κ z + O(z2) 390

is a reasonable assumption (not shown), and model results show a good agreement with the 391

corresponding LES profiles for both velocity components (u, v). This can be seen from FIG. 4, 392

which features a comparison between predictions from the proposed analytical model and LES 393

profiles for all six considered cases (see Table 1). 394

Figure 4 about here 395

Further aloft (i.e., z ⪆ 0.1zi), the dominant balance in Eqs. (9) is between the Coriolis forcing 396

and the geostrophic wind vector (the second and third term on the LHS of Eqs. 9) for the positive 397

geostrophic shear, cold and warm advection, and barotropic cases. As a result, at such heights, the 398

momentum-flux divergence term (first term on the LHS of Eqs. 9) is relegated to a minor 399

momentum-modulation role, and the governing equations are closed. It hence comes as no surprise 400

that the simple reduced model is in a very good agreement with LES velocity profiles ( FIG. 401

4(a,b,e,f)). 402

Page 21: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

21

In the considered cases with substantial negative geostrophic shear, the reduction with height 403

of the geostrophic wind vector modifies the dominant balance above the constant-flux layer, in 404

favor of the stress-divergence term. In such conditions, a more accurate parameterization of 405

turbulent fluxes is hence required along with a more refined definition of the effective top of the 406

turbulent layer, which influences the viscosity profiles. Rather than using the height of the base of 407

the capping inversion layer, zi, we use an effective ABL depth zi,eff = 0.6 zi in Eqs. (10) when 408

applied to NS135 and NS180 cases (consistent with where TKE is significantly reduced, which 409

will be discussed in the next section). The agreement of the reduced model (dotted lines) with the 410

LES results is improved through the use of this effective depth. Nonetheless, the peak values of 411

the jets were still not completely captured in the reduced model, which could be due to the chosen 412

parametrization for turbulent fluxes. As described in the following section, the imposed negative 413

geostrophic shear strongly impacts the momentum transfer coefficient above the constant-flux 414

layer, reducing its magnitude via a reduction of the turbulence shear production. The O’Brien’s 415

eddy-viscosity model has a prescribed vertical structure and, not surprisingly, predicts a larger 416

eddy viscosity at such heights, resulting in more diffused velocity profiles when compared to the 417

reference LES ones. 418

Nevertheless, the important features of the baroclinic ABL are well captured across the 419

considered cases (even without an accurate determination of the effective ABL depth), suggesting 420

that this simple predictive model might be sufficient in many applications such as for estimating 421

the baroclinicity effects on meteorological measurements and wind power applications. While we 422

do not extend the model here to stable and unstable cases, this could be achieved via an alternative 423

parameterization for νT that accounts for stability (and potentially a non-local flux term under 424

Page 22: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

22

unstable conditions). Further work will be required to account for the interacting effects of 425

baroclinicity and stability on momentum transfer coefficients characterizing the ABL; these new 426

eddy-viscosity parameterizations (e.g. considering a local eddy-viscosity formulation) will 427

improve the predictive capabilities of such reduced-order models when stress divergence terms are 428

relevant in the momentum-balance equations (e.g. in the NS180 case). To develop better 429

turbulence closure models for baroclinic ABLs, one needs to understand how baroclinicity 430

modulates the turbulence. Therefore, in the following section, we make a first step toward such 431

future parameterizations, by quantifying the impact of baroclinicity and stability on turbulence 432

statistics. 433

Implications for turbulence 434

Baroclinicity also strongly influences the turbulence characteristics of atmospheric flows, notably 435

the TKE profiles. We report the total TKE (resolved + SGS) profiles: 436

, (11) 437

where is the resolved and the SGS component of TKE (primes 438

represent the turbulent fluctuations from the Reynolds average denoted by the overbar), the SGS 439

contribution is estimated as 440

, (12) 441

where is the trace of the Leonard stress tensor, and and are the widths of a 442

test and a grid filter, respectively (see Knaepen et al. 2002; Salesky & Chamecki 2012); we use 443

= 2 . 444

Page 23: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

23

In a steady barotropic neutral environment, the TKE decreases with height since its shear 445

production also decreases (unsteady barotropic cases could be more complicated see e.g. Momen 446

and Bou-Zeid 2016b). Our simulations illustrate that this is not necessarily the case in steady 447

baroclinic ABLs. FIG. 5 displays the TKE budget and TKE profiles in the neutral baroclinic cases. 448

While the TKE decreases with height near the surface in all cases, it has a non-monotonic trend 449

further aloft in some runs. For instance, it features an increase above the surface layer in the NS270 450

case (FIG. 5(e)). 451

The origin of the trends in the TKE profiles can be elucidated by examining the various terms 452

in its budget equation. By assuming no subsidence, 0w = , and periodic statistical homogeneity in 453

both horizontal directions, the conventional form of the resolved TKE budget can be obtained as 454

(see Momen & Bou-Zeid 2017a): 455

, (13) 456

where is the strain-rate tensor. Note that all ui, p and Sij in Eq. (13) and 457

following equations are the resolved components but we henceforth omit the tildes for notational 458

simplicity. We compute the TKE budget by defining the following expressions for the terms in 459

Eq. (13) as: 460

(14.a) 461

(14.b) 462

(14.c) 463

Page 24: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

24

where SP, RT, and εSGS are respectively the shear production, resolved turbulent transport, and 464

SGS dissipation (calculated, for better accuracy, using the last expressions on the left-hand side of 465

Eqs. (14) to avoid having to compute a perturbation during the LES run). The other terms were 466

found to be minor and are neglected here (the buoyancy term is not significant in the neutral ABLs 467

away from top inversion layer, but will be included in the later analysis of the diabatic ABLs). 468

Positive geostrophic shear (FIG. 5(a)) significantly increases the magnitude of the budget terms. 469

In the positive-shear case, SP >> RT inside the ABL (0 < z/zi <1) since the imposed pressure 470

gradient produces a constant shear in height. However, in the barotropic case (FIG. 5(f)), 1 ⪅ 471

RT/SP ⪅ 1.5 for 0.7 < z/zi < 1. On the other hand, in the cases with a negative geostrophic shear 472

(FIG. 5(c,d)) the magnitude of the shear production decreases relative to the barotropic reference. 473

In summary, the magnitude of the TKE budget terms in all the neutral cases is altered by the 474

baroclinicity direction and strength. Their shapes are equally influenced. For instance, we see 475

increasing values of TKE profiles with height at some levels in NS270 (FIG. 5(e)). We also observe 476

an inflection point in the SP and εSGS profiles of NS270, which is not seen in the barotropic case. 477

Figure 5 about here 478

4 The interacting effects of baroclinicity and stability 479

In this section, we describe how baroclinicity impacts the ABL mean and turbulence behavior 480

in the presence of buoyancy forces. To that end, we only show the strong baroclinic ABLs (S0 = 1) 481

under the imposed surface heat fluxes (+200 w/m2 and −20 w/m2) to distinguish their effects more 482

clearly from the barotropic cases. 483

Page 25: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

25

Implications for the mean velocity 484

Baroclinicity strongly affects the shape of the wind speed profiles in the diabatic ABLs, as 485

illustrated in FIG. 6. Unconventional velocity profiles are clearly observed in this figure. For 486

example, negative geostrophic shear can lead to velocity profiles that decrease with height (FIG. 487

6(c) and 6(d)) e.g. over more than half of the ABL depth in neutral cases. The positive-shear case 488

also highly increases the magnitude of the wind speeds (more than 50% compared to the barotropic 489

environment with the same ground-level geostrophic wind). 490

Away from the surface, z/zi > 0.3, the turbulent friction force is decreased. Thus, in a neutral 491

ABL, the dominant balance would be determined by Coriolis and pressure gradient forces, 492

resulting in geostrophic winds in that region. Hence, it is expected that the wind velocities will 493

align with the geostrophic wind vector in the outer layer (see FIG. 6). Buoyancy forces alter this 494

balance. As FIG. 6 shows, stable cases have the closest matches between the geostrophic and actual 495

wind speed profiles, while winds in the unstable cases result in strong departures from their 496

geostrophic counterparts. This is attributed to the negative stabilizing heat fluxes in the stable cases 497

that damp the turbulence in the ABL and reduce friction. Hence, the main balance between Coriolis 498

force and geostrophic wind in the outer layer is more dominant, with minimal frictional effects. 499

Nevertheless, in the unstable cases, friction is increased, and the vertical mixing is enhanced. 500

Therefore, the profiles are more uniform, and they depart from the geostrophic winds more than 501

other cases. 502

Figure 6 about here 503

Baroclinicity also impacts the much-studied low-level jet (LLJ) characteristics in the stable 504

ABL. The LLJ is defined as a region inside ABL where the wind speed is significantly faster than 505

Page 26: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

26

wind speeds above it (Stull 1988). In the stable barotropic ABL, a weak LLJ is formed at the height 506

where the TKE is significantly weakened (FIG. 6(f)). In order to assess the LLJ behavior, we 507

define the LLJ location, ZLLJ, and its strength, SLLJ, as: 508

. (15) 509

The strength of LLJ for a barotropic case is SLLJ,SB ≈ 1.1. This strength is decreased when there is 510

a positive shear (SLLJ,SS0 = 1.05 FIG. 6(a)) and intensified when there is a negative shear 511

(SLLJ,SS135 = 1.11 FIG. 6(c), and SLLJ,SS180 = 1.14 FIG. 6(d). We also observe jet-like profiles in 512

neutral and unstable conditions (albeit weaker than under stable conditions) when there is a 513

negative shear (FIG. 6(c) and 6(d)). The LLJ location, ZLLJ, is also influenced by the baroclinicity 514

direction (β). A positive geostrophic shear and a warm advection tend to elevate it while a negative 515

shear lowers its height. The LLJ location and strength are very important for the wind-energy 516

industry. LLJs boost the wind power production in wind farms and lead to higher economic 517

benefits for wind-farm owners (see e.g. Nunalee and Basu 2013 for LLJ implications in wind-518

energy industry). Hence, our results could be useful for wind-energy application (such as site 519

selection) to better understand how the LLJ location and strength are modulated by baroclinicity 520

in the atmosphere. Note that these results are not only valid for the wind speed profiles non-521

dimensionalized by the geostrophic wind vector at the surface, G0, but also, they show similar 522

characteristics if normalized by the height-varying geostrophic wind vector, G(z), as shown in the 523

supplemental material (although the LLJ is not typically defined in this way in the literature). 524

The Ekman spirals are similarly altered significantly by the presence of the baroclinicity. FIG. 525

7 displays the hodographs for baroclinic and barotropic ABLs under stable and unstable 526

conditions. The wind profiles in unstable cases do not converge to the geostrophic values within 527

Page 27: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

27

the ABL, which lie on the dotted line, due to the significant friction (turbulent stresses) at all 528

heights. They converge to geostrophic values after the capping inversion layer starts and damps 529

the shear stress. As expected, the hodographs in unstable conditions are contracted since the 530

mixing and vertical coupling are high, leading to more uniform velocity profiles. On the other 531

hand, in stable conditions eccentric hodographs are observed due to weaker mixing and vertical 532

coupling. The velocity vectors change more significantly in baroclinic cases with height compared 533

to their barotropic counterpart. As noted in the introduction, these modifications will alter the 534

Ekman pumping process and influence the genesis of storms and fronts. Understanding the genesis 535

of storms and their associated precipitation by baroclinicity are crucial for models that aim to 536

parameterize hydrometeorological processes and land-atmosphere interactions (see e.g. Ramos da 537

Silva and Avissar 2006; Momen et al. 2017a). This occurs due to variations in the cross-isobaric 538

mass transport and mass flux out of the boundary layer in low and high-pressure systems. For 539

instance, the cross-isobaric mass transport, , in our barotropic cases 540

(CMTSB = 0.07, CMTUB = 0.19, see FIG. 7f), is significantly increased in the positive geostrophic 541

shear (CMTSS0 = 0.61, CMTUS 0= 0.77, see FIG. 7a) and slightly decreased in the negative 542

geostrophic shear (CMTSS180 = 0.05, CMTUS180 = 0.15, see FIG. 7d). 543

Figure 7 about here 544

Implications for turbulence 545

The friction velocity is highly modulated by the baroclinicity strength and through the 546

interaction of baroclinicity and stability. Table 2 indicates that increasing surface heat flux from 547

stable to neutral and to unstable increases the friction velocity as expected since the turbulence 548

Page 28: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

28

intensity rises. Furthermore, as the strength of the baroclinicity in neutral cases increases the 549

changes in the friction velocity compared to the barotropic case become higher. In general, 550

irrespective of the static stability in the diabatic cases and the baroclinicity strength in neutral 551

simulations, the friction velocity in the positive geostrophic shear and cold-advection cases 552

increases compared to the barotropic environment, while it decreases in the cases with negative 553

geostrophic shear. These variations in the friction velocity are more pronounced in the unstable 554

cases than in the neutral and stable cases since under unstable conditions mixing and vertical 555

coupling are high and the surface is thus more influenced by the imposed geostrophic wind shear 556

at higher elevations. On the other hand, in stable cases the surface is more decoupled from the 557

middle and top of the ABL and hence it is not significantly affected by the changes in the 558

geostrophic wind vector away from the surface. 559

Table 2 about here 560

The ratio of the shear production over the buoyant production or destruction, which is the 561

negative inverse of the flux Richardson number (–Rif –1), is calculated to infer how it is modulated 562

by baroclinicity. Rif is a dimensionless number that indicates the dominance of buoyant versus 563

shear production of turbulence in the ABL and Rif ≡ − BP /SP, where 564

is the total buoyant production/loss. Here, to get 565

the BP, we add the SGS contribution to the resolved component. Additionally, this number can 566

characterize the local stability of the flow. In the unstable barotropic simulation (FIG. 8(f)), SP is 567

almost two times BP very close to the surface, while above z/zi > 0.1, BP is dominant. In the stable 568

barotropic case (FIG. 8(f)), SP reaches a minimum in proximity of the LLJ where it becomes 569

Page 29: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

29

slightly negative – often corresponding to a global energy backscatter region – that indicates 570

energy transfer from small to large scales (conversion of TKE to mean kinetic energy, see e.g. 571

Bou-Zeid et al. 2009; Gayen et al. 2010). This negative shear production occurs right above the 572

LLJ, where dU/dz < 0 and (see e.g. the TKE budget of stable cases in the supplemental 573

material). 574

Near the surface and for a constant surface buoyancy flux, baroclinicity alters the stability of 575

the flow. Positive geostrophic shear and cold advection tend to increase the Rif –1 magnitude, which 576

implies that the ABL is closer to neutral conditions. Conversely, a negative geostrophic shear tends 577

to decrease the Rif –1 magnitude, which implies that the ABL becomes more unstable. This fact is 578

also clear from the zi /L values shown in Table 2 (compare these cases with their barotropic 579

counterpart). However, away from the surface, 0.3 < z/zi <1, baroclinicity always increases shear 580

production in the ABL and tends to make the outer ABL flow more neutral throughout the 581

considered cases by increasing |Rif –1|. This can be related to the geostrophic shear that results in 582

wind shear and turbulence production even far from the surface in all baroclinic cases, but not in 583

the barotropic one. Note that this outer region in the stable ABL corresponds to the flow above the 584

LLJ, which is completely decoupled from the surface and often not considered part of the stable 585

ABL. 586

In the stable baroclinic cases, the departures of the Rif –1 from the barotropic reference very 587

close to the surface are weaker than at higher elevations. That is because in stable environments 588

the vertical profiles are more decoupled from the wall. Therefore, the surface value is mainly 589

controlled by the imposed pressure gradient at/near the wall, which is equal to the barotropic 590

geostrophic value, rather than the pressure gradient aloft. However, at higher heights, 0.3 < z/zi <1, 591

Rif –1 changes substantially compared to the barotropic case. In the stable barotropic ABL, the ratio 592

Page 30: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

30

of the shear over buoyancy is near zero at higher elevations, 0.3 < z/zi <1. However, in the 593

baroclinic environments, this ratio becomes < −1 and the ABL is hence shear-dominated above 594

the LLJ. 595

Figure 8 about here 596

The baroclinic unstable environments show significant Rif –1 departures from their barotropic 597

counterpart near the surface since mixing is high and thus the surface is more influenced by 598

changes at higher elevations. In the unstable positive-shear case (FIG. 8(a)), the shear over buoyant 599

production ratio Rif –1 becomes about three times larger than the barotropic case (FIG. 8(f)) near 600

the surface. At higher elevations, 0.1 < z/zi < 0.8, unlike the barotropic unstable case and all stable 601

cases, shear increases but the buoyancy term remains dominant. 602

The TKE profiles and budgets, illustrated in FIG. 9, show that in the positive-shear case US0, 603

the TKE becomes relatively constant in the middle of the domain due to the persistent imposed 604

geostrophic shear. All the TKE budget terms are magnified in this case (FIG. 9(a)), compared to 605

the barotropic case (FIG. 9(f)). Moreover, RT / εSGS becomes smaller than under the barotropic 606

condition indicating that the relative importance of turbulent transport over dissipation decreases 607

and production-dissipation balance is more applicable (so the produced turbulence at the surface 608

in case US0 is transported upwards less than in case UB). As in the neutral cases, the TKE displays 609

some unusual vertical trends in the baroclinic diabatic ABLs such as increasing magnitude with 610

height, FIG. 9(d) for 0.15 < z/zi < 0.35, under unstable condition, as well as in stable cases SS180 611

and SS270 shown in the supplemental material. 612

In summary, baroclinicity significantly impacts the magnitude of the TKE and TKE budget 613

terms, changes their relative ratios in the ABL, and alters their profiles. These results indicate that 614

Page 31: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

31

it is imperative to consider baroclinicity effects in future ABL models and turbulence closures in 615

order obtain a more accurate representation of real-world ABLs. This can be done by employing 616

S0 and β parameters (Eq. 8) in such models, in addition to conventional parameters such as friction 617

velocity, aerodynamic roughness length, surface heat flux, and ABL height. Local eddy-viscosity 618

models might be able to account indirectly for the effect of S0 and β if they can dynamically capture 619

the resulting changes in mean wind magnitude and direction related to baroclinicity (i.e. as in the 620

analytical model we presented for the neutral cases), but for non-local models this might be more 621

challenging. 622

Figure 9 about here 623

Implications on the large turbulence structures 624

The coherent turbulence structures are significantly influenced by the baroclinicity since it 625

increases or reduces shear production in the ABL. In particular, here we present the effects on 626

convective ABLs. In such flows, cellular- or roll- type convections in the velocity fields are 627

formed. Certain environments can produce wide or narrow mixed-layer rolls and they modulate 628

large-scale motions and thermals, potentially in a two-way coupling (Young et al. 2002). These 629

thermal convection structures, under different stabilities, can lead to instability in the Ekman 630

boundary layer (Lilly 1966). Most of the previous studies focused on the role of the stability 631

parameters like zi/L (Deardorff 1972; Salesky et al. 2016) in favoring rolls or cells regimes. We 632

show here how these coherent ABL structures are also modulated by baroclinicity, at a constant 633

surface heat flux. 634

FIG. 10 displays the normalized instantaneous vertical velocity field in three convective cases 635

at two different heights. FIG. 10(a,b) shows the barotropic simulations whereas FIG. 10(c,d) is the 636

Page 32: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

32

warm-advection case and FIG. 10(e,f) depicts the negative geostrophic shear case. FIG. 10(a,b,c,d) 637

exhibits the signature of the well-documented streamwise-elongated, high and low momentum 638

bulges (red strips of high velocity, often referred to as the large-scale motions) whereas FIG. 639

10(e,f) indicates convective cells in the negative geostrophic shear case (see also the supplemental 640

material). Salesky et al. (2016) found that the transition from rolls to cells occurs in the regions 641

where zi/L ≈ –15 to –20. From Table 2, one can note that the UB and US270 cases depicted in FIG. 642

10 have very similar values of zi/L ≈ –14. This explains the similarity of their roll structures. Case 643

US180 on the other hand has zi/L = –70.2, placing it in the cell region according to Salesky et al. 644

(2016); this is indeed confirmed in FIG. 10. Hence, baroclinicity can lead to the formation of 645

convective cells even when the heat flux and geostrophic wind values at the surface suggest the 646

presence of rolls. This modulation appears to be consistent with the modification of zi/L and the 647

transition from rolls to cells seems to occur at zi/L values comparable to the barotropic ABL. 648

Nevertheless, even at the same zi/L of a barotropic counterpart, baroclinicity will change the 649

direction of the mean flow with height, with apparent consequences on the large-scale motions. 650

For instance, FIG. 10(c,d) indicates that the rolls are inclined laterally (compare it with the 651

barotropic case shown in FIG. 10(a,b)) due to the imposed geostrophic wind vector in the negative 652

y-direction. 653

Figure 10 about here 654

5 Conclusions 655

We investigated the baroclinic environment, which emerges when the isobars intersect the 656

isotherms and isopycnals, and which occurs very frequently in the real-world atmospheric 657

boundary layer due to the prevalent strong horizontal temperature gradients. Unlike the barotropic 658

Page 33: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

33

case, both the magnitude and direction of the pressure gradient vector can change in height in a 659

baroclinic environment. We introduced two parameters to quantify the baroclinicity strength (S0) 660

and direction (β), similar to Arya & Wyngaard (1975), based on the thermal wind vector 661

components. Each of them can have various distinctive impacts on the mean and turbulence 662

characteristics. Twenty-three large-eddy simulations were conducted to study these effects. The 663

main findings of this paper include: 664

• Baroclinicity strongly influences the Ekman turning and mean hodographs under all 665

stabilities, even when normalized by the geostrophic wind magnitude at each elevation 666

G(z). Both weak (S0 = 0.25) and strong (S0 = 1.0) baroclinicity result in significant 667

departures from the barotropic case and these deviations grow with increasing S0. 668

• We proposed a reduced-order model to capture the effects of the baroclinicity strength and 669

direction on the mean velocity profiles of neutral cases using a prescribed cubic eddy-670

viscosity profile. Despite the simplicity of the model, it captures the main LES trends 671

qualitatively well. Thus, it could be used in many applications, for example, to assess 672

baroclinic effects on meteorological measurements or to estimate wind power changes due 673

to baroclinicity. 674

• The velocity vectors in the ABL follow the trends of the magnitude (S0) and direction (β) 675

of the geostrophic wind vectors, particularly away from the surface layer (as turbulent 676

stresses become less dominant in such regions). The matches between the wind velocity 677

vectors and geostrophic wind values also depend on the stability of the ABL. They match 678

closely in stable ABLs, where friction forces are damped due to stabilizing heat fluxes; on 679

the contrary, larger departures are in unstable conditions, where friction is dominant and 680

vertical mixing leads to more uniform wind profiles. The resulting velocity profiles are 681

Page 34: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

34

rarely observed in barotropic environments. For example, a negative geostrophic shear (β 682

= 1800) yields a mean velocity profile that decreases with height over more than half of the 683

ABL depth in neutral cases, while a positive geostrophic shear (β = 00) strongly increases 684

the wind speed magnitude throughout the ABL. 685

• The friction velocity values, u*, are strongly impacted by the baroclinicity as well. 686

Irrespective of the static stability in diabatic simulations and baroclinicity strength in 687

neutral cases, positive geostrophic shear and cold advection increased the friction velocity 688

while negative geostrophic shear decreased u*, compared to their barotropic counterparts. 689

• The low-level jets are highly modulated by baroclinicity. Their strength is decreased under 690

positive geostrophic shear (β = 00) and intensified under negative geostrophic shear (β = 691

1350 and β = 1800). Furthermore, positive geostrophic shear and warm advection tend to 692

increase their height, whereas negative geostrophic shear lowers it. 693

• Baroclinicity generally increases shear in the outer ABL (since there always exists a 694

vertical gradient in geostrophic wind under baroclinic conditions) and tends to make the 695

core of the ABL more neutral by decreasing |Rif| away from the surface. Near the surface, 696

these trends are reversed when there is a negative geostrophic shear. 697

• In some baroclinic cases, increasing TKE with height is observed, which is not seen in a 698

comparable barotropic environment. Baroclinicity impacts the turbulence by producing 699

shear at higher elevations. Therefore, it alters the TKE budget balance. The positive 700

geostrophic shear increases the magnitude of the TKE budget terms and negative 701

geostrophic shear decrease them. Baroclinicity also modifies the relative ratio of the TKE 702

budget terms in the ABL and influences the shape of the profiles. 703

Page 35: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

35

• The large turbulence structures are altered by baroclinicity. For instance, the positive-shear 704

case led to a roll-type convection in the vertical velocity field while the negative-shear case 705

led to a cellular-type convection, although the imposed heat flux and geostrophic wind 706

values at the surface were the same in both cases. This modification is consistent with the 707

previously-reported dependence of the structural properties of barotropic ABL turbulence 708

on zi/L. In addition, the direction of these large structures tends to become aligned with the 709

imposed shear in the ABL. 710

Our results demonstrate that baroclinicity can substantially change the characteristics of mean 711

flow and turbulence in the ABL when compared to the classic barotropic environment, and 712

therefore it must be accounted for in numerical/theoretical models (perhaps by using appropriate 713

S0 and β). This will result in better agreements with real-world measurements. It should also be 714

considered when experimental observations are analyzed. Further studies are required to better 715

understand how baroclinicity modulates other aspects such as turbulent stresses and anisotropy, 716

turbulent transport efficiency, and conventional similarity theories in the ABL, and to more 717

thoroughly quantify our findings using field measurements. 718

719

Acknowledgments. The authors acknowledge support from the Princeton Environmental Institute and 720

the National Oceanographic and Atmospheric Administration's (NOAA), through the Cooperative Institute 721

for Climate Science (CICS) of Princeton University, under grant number 344-6127. Furthermore, MM and 722

MGG acknowledge the Department of Civil Engineering and Engineering Mechanics at Columbia 723

University for support via startup-funds. The simulations were performed on the computing clusters of the 724

National Centre for Atmospheric Research under project numbers UPRI0007 and UPRI0016. 725

726

APPENDIX A 727

Initialization and boundary conditions of the simulations 728

Page 36: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

36

The simulations are first integrated to reach a steady-state condition on a 64×64×64 numerical 729

grid for eight inertial time scales (τABL). Next, all the cases are interpolated into numerical grids 730

with 192×192×192 grid nodes. The warm-up procedure is then continued for 2τABL in the high-731

resolution mode, to ensure development of the smallest resolved scales, equilibrium of the mean 732

forces, and convergence of the flow statistics. Finally, statistics are computed over the last τABL. 733

Averaging over τABL (about 12.5 hours here) is enough to fully converge statistics (Monin and 734

Yaglom 1975), especially considering that we also average in the horizontal directions over the 735

whole domain. The initial values of velocities in our code are set as the solutions to the steady 736

neutral barotropic Ekman layer suggested by Businger (1982): 737

, (A1) 738

where, , and νm = 5 m2 s−1 is a constant eddy viscosity. Thus, upon running the code 739

for various baroclinic cases, we found that after ≈ 8 inertial cycles (τABL) the solutions of baroclinic 740

cases reached a steady state for this specific initialization (equation A1). We also checked the 741

momentum balance of several high-resolution cases and verified this steady-state condition. We 742

note that this long initialization is mainly done so that the computed statistics do not depend on 743

initial conditions. This steady-state condition might not be prevalent in real-world ABLs because 744

of many complicating factors that lead to the unsteadiness of the airflow. Nevertheless, 745

understanding the steady-state baroclinicity effects on the spatial structures of the mean and 746

turbulence is a necessary first step before proceeding to understanding all the complicating features 747

of unsteady baroclinic ABLs. 748

Page 37: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

37

In unstable and neutral simulations, the height of the domain is Lz = 1500 m with a horizontal 749

size of Lx = Ly = 2πLz. In stable simulations, we set Lz = 1100 m to obtain a higher resolution. 750

Boundary conditions are periodic in the horizontal directions. For the bottom boundary, a local 751

equilibrium logarithmic-wall model based on spatially—filtered velocities is used, to evaluate 752

surface stresses (see Bou-Zeid et al. 2005), and a constant surface heat flux is imposed for the 753

scalar equation. The aerodynamic roughness length z0 is set to 0.1m for all the considered cases. 754

The surface value of the geostrophic wind at the wall, G0, is always equal to the barotropic 755

geostrophic value (8 m/s here). Under neutral conditions, normalization by G0 makes its value 756

irrelevant; however, this dimensional value is important in diabatic simulations (relative to the 757

surface heat flux) and affects stability parameters like Obukhov length and Richardson number. 758

For the top boundary, we use a stress-free and impermeable (w = 0) boundary condition for the 759

velocity field and an inversion layer with 0.03 K/m slope is imposed above zi = 800 m in the stable 760

cases and 1000 m in the neutral and unstable runs. The inversion layer imposes a limit on the 761

height and growth of the ABL. We note that in this study we did not aim to investigate the 762

dynamical evolution of the ABL height, since we focus on the steady-state baroclinic ABLs under 763

various imposed surface heat flux conditions. In stable cases, we also impose a sponge layer with 764

a Rayleigh damping method (Israeli and Orszag 1981) at z = 825 m to damp all perturbations thus 765

preventing the reflection of internal waves. 766

767

APPENDIX B 768

Comparison of the results of our baroclinic LES setup with other codes 769

Page 38: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

38

We compared our LES results with two neutral cases of Brown (1996). Brown (1996) used a 770

non-uniform vertical grid with resolution 40×40×32 while we use uniform grids with resolution of 771

128×128×128. His boundary conditions were periodic in the horizontal directions, with a rough 772

wall-model at the bottom, and a rigid lid with zero stress and heat flux at the top. Brown (1996) 773

used the static Smagorinsky model (imposed non-dynamic coefficient) for parameterizing the SGS 774

components and hence we also use the static Smagorisnky model in these two simulations to allow 775

direct comparison. The results shown in the supplemental material indicate a good agreement 776

between our LES and his numerical results. We note that in other simulations in the paper, we use 777

a scale-dependent Lagrangian dynamic Smagorinsky SGS model (Bou-Zeid et al. 2005). This 778

model reproduces the logarithmic region in high-Reynolds number simulations, which do not 779

resolve the viscous sublayer, better than other SGS models like the static of scale-invariant 780

dynamic Smagorinsky models. Furthermore, the velocity profiles of this model match better with 781

an experimental data compared to other SGS models (see Bou-Zeid et al. 2005 for SGS models 782

comparison). 783

Note that Sorbjan (2004) only shows his results at t/τABL ≈ 0.25; However, we believe that in 784

order to reach equilibrium and overcome the initial-condition effects, the simulations have to be 785

integrated at least about two times the inertial time scale (2τABL). We concur with Sorbjan (2004) 786

results for earlier times in our simulations, but we do not show them here since they have not 787

reached the full equilibrium with the imposed geostrophic forcing. 788

Brown (1996) indicated that the time series of u* were close to achieving equilibrium after t/τABL 789

≈ 1.6 and so he ran the simulations up to this time and we compare our results with his cases here. 790

These comparisons demonstrate that our baroclinic setup matches the results of other available 791

codes well, which serves as a check on the simulation setup and validity. 792

Page 39: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

39

793

Page 40: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

40

REFERENCES: 794

Albertson, J. D., and M. B. Parlange, 1999: Natural integration of scalar fluxes from complex terrain. Adv. Water 795

Resour., 23, 239–252, doi:10.1016/S0309-1708(99)00011-1. 796

Arya, S. P. S., and J. C. Wyngaard, 1975: Effect of baroclinicity on wind profiles and the geostrophic drag law for the 797

the convective planetary boundary layer. J. Atmos. Sci., 32, 767–778, doi:10.1175/1520-798

0469(1975)032<0767:EOBOWP>2.0.CO;2. 799

Basu, S., J. Vinuesa, and A. Swift, 2008: Dynamic LES Modeling of a Diurnal Cycle. J. Appl. Meteorol. Climatol., 800

47, 1156–1174, doi:10.1175/2007JAMC1677.1. 801

Berger, B. W., and B. Grisogono, 1998: The baroclinic, variable eddy viscosity Ekman layer: an approximate 802

analytical solution. Boundary-Layer Meteorol., 87, 363–380, doi:10.1023/A:1001076030166. 803

Blackadar, A., 1957: Boundary layer wind maxima and their significance for the growth of nocturnal inversions. Bull. 804

Am. Meteorol. Soc., 38, 283–290. 805

Bou-Zeid, E., C. Meneveau, and M. Parlange, 2005: A scale-dependent Lagrangian dynamic model for large eddy 806

simulation of complex turbulent flows. Phys. Fluids, 17, 025105, doi:10.1063/1.1839152. 807

http://link.aip.org/link/PHFLE6/v17/i2/p025105/s1&Agg=doi (Accessed February 6, 2014). 808

Bou-Zeid, E., J. Overney, B. D. Rogers, and M. B. Parlange, 2009: The effects of building representation and 809

clustering in large-eddy simulations of flows in urban canopies. Boundary-Layer Meteorol., 132, 415–436, 810

doi:10.1007/s10546-009-9410-6. 811

Brown, A. R., 1996: Large-eddy simulation and parametrization of the baroclinic boundary-layer. Q. J. R. Meteorol. 812

Soc., 122, 1779–1798, doi:10.1002/qj.49712253603. http://doi.wiley.com/10.1002/qj.49712253603. 813

Brown, A. R., A. C. M. Beljaars, and H. Hersbach, 2006: Errors in parametrizations of convective boundary-layer 814

turbulent momentum mixing. Q. J. R. Meteorol. Soc., 1859–1876, doi:10.1256/qj.05.182. 815

Businger, J. A., 1982: (“Equations and Concepts”, Atmospheric Turbulence and Air Pollution Modelling, F.T.M. 816

Nieuwstadt and H. van Dop (eds.). D. Reidel Publishing Company, Dordrecht, Holland,. 817

Cantero, M. I., S. Balachandar, M. H. García, and D. Bock, 2008: Turbulent structures in planar gravity currents and 818

Page 41: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

41

their influence on the flow dynamics. J. Geophys. Res. Ocean., 113, 1–22, doi:10.1029/2007JC004645. 819

Chow, F. K., and R. L. Street, 2009: Evaluation of turbulence closure models for Large-Eddy simulation over complex 820

terrain: Flow over Askervein Hill. J. Appl. Meteorol. Climatol., 48, 1050–1065, doi:10.1175/2008JAMC1862.1. 821

Deardorff, J. W., 1970: Convective Velocity and Temperature Scales for the Unstable Planetary Boundary Layer and 822

for Rayleigh Convection. J. Atmos. Sci., 27, 1211–1213, doi:10.1175/1520-823

0469(1970)027<1211:CVATSF>2.0.CO;2. 824

Deardorff, J. W., 1972: Numerical Investigation of Neutral and Unstable Planetary Boundary Layers. J. Atmos. Sci., 825

29, 91–115, doi:10.1175/1520-0469(1972)029<0091:NIONAU>2.0.CO;2. 826

Djolov, G. D., D. L. Yordanov, and D. E. Syrakov, 2004: Baroclinic planetary boundary-layer model for neutral and 827

stable stratification conditions. Boundary-Layer Meteorol., 111, 467–490, 828

doi:10.1023/B:BOUN.0000016507.94623.e5. 829

Ekman, V. W., 1905: On the Influence of the Earth’s Rotation on Ocean-Currents. Ark. Mat. Astron. Fys., 2, 1–52. 830

Floors, R., A. Pena, and S.-E. Gryning, 2015: The effect of baroclinicity on the wind in the planetary boundary layer. 831

Q. J. R. Meteorol. Soc., 141, 619–630, doi:10.1002/qj.2386. 832

Gayen, B., S. Sarkar, and J. R. Taylor, 2010: Large eddy simulation of a stratified boundary layer under an oscillatory 833

current. J. Fluid Mech., 643, 233, doi:10.1017/S002211200999200X. 834

http://www.journals.cambridge.org/abstract_S002211200999200X (Accessed March 16, 2014). 835

Grisogono, B., 1995: A generalized Ekman layer profile with gradually varying eddy diffusivities. Q. J. R. Meteorol. 836

Soc., 121, 445–453, doi:551.5 11.61:551.554. 837

Gula, J., Molemaker, M. J., McWilliams, and J. C., 2014: Submesoscale Cold Filaments in the Gulf Stream. J. Phys. 838

Oceanogr., 44, 2617–2643, doi:10.1175/JPO-D-14-0029.1. 839

Hess, G. D., 1973: On Rossby-Number Similarity Theory for a Baroclinic Planetary Boundary Layer. J. Atmos. Sci., 840

30, 1722–1723, doi:10.1175/1520-0469(1973)030<1722:ORNSTF>2.0.CO;2. 841

http://journals.ametsoc.org/doi/abs/10.1175/1520-0469(1973)030%3C1722:ORNSTF%3E2.0.CO;2. 842

Hess, G. D., 2004: The planetary boundary layer, capped by an inversion: extension to baroclinic and stable conditions. 843

Page 42: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

42

Aust. Meteorol. Mag., 53, 53(2), 95-101, June 2004. http://www.scopus.com/inward/record.url?eid=2-s2.0-844

4344655822&partnerID=tZOtx3y1. 845

Holtslag, A. A. M., and Coauthors, 2013: Stable atmospheric boundary layers and diurnal cycles: Challenges for 846

weather and climate models. Bull. Am. Meteorol. Soc., 94, 1691–1706, doi:10.1175/BAMS-D-11-00187.1. 847

Hoxit, L., 1974: Planetary Boundary Layer Winds in Baroclinic Conditions. J. Atmos. Sci., 31, 1003–1020. 848

Huang, J., and E. Bou-Zeid, 2013: Turbulence and vertical fluxes in the stable atmospheric boundary layer. Part I: A 849

large-eddy simulation study. J. Atmos. Sci., 70, 1513–1527, doi:10.1175/JAS-D-12-0167.1. 850

Israeli, M., and S. A. Orszag, 1981: Approximation of radiation boundary conditions. J. Comput. Phys., 41, 115–135, 851

doi:10.1016/0021-9991(81)90082-6. 852

Knaepen, B., O. Debliquy, and D. Carati, 2002: Subgrid-scale energy and pseudo pressure in large-eddy simulation. 853

Phys. Fluids, 14, 4235–4241, doi:10.1063/1.1514643. 854

Kumar, V., G. Svensson, a. a. M. Holtslag, C. Meneveau, and M. B. Parlange, 2010: Impact of Surface Flux 855

Formulations and Geostrophic Forcing on Large-Eddy Simulations of Diurnal Atmospheric Boundary Layer 856

Flow. J. Appl. Meteorol. Climatol., 49, 1496–1516, doi:10.1175/2010JAMC2145.1. 857

http://journals.ametsoc.org/doi/abs/10.1175/2010JAMC2145.1 (Accessed March 5, 2014). 858

Kundu, P. K., and I. M. Cohen, 2008: Fluid Mechanics. Academic Press, Boston, 872 pp. 859

Lilly, D. K., 1966: On the Instability of Ekman Boundary Layer. J. Atmos. Sci., 23, 481–494. 860

Mahrt, L. J., and W. Schwerdtfeger, 1970: Ekman spirals for exponential thermal wind. Boundary-Layer Meteorol., 861

1, 137–145, doi:10.1007/BF00185735. 862

Momen, M., and E. Bou-Zeid, 2015: Analytical damped-oscillator models for unsteady atmospheric boundary layers. 863

Bulletin of the American Physical Society, Division of Fluid Dynamics, American Physical Society (APS), Vol. 864

60 of, American Physical Society http://meetings.aps.org/Meeting/DFD15/Event/254408. 865

Momen, M., and E. Bou-Zeid, 2016a: Large-eddy simulations and damped-oscillator models of the unsteady Ekman 866

boundary layer. J. Atmos. Sci., 73, doi:10.1175/JAS-D-15-0038.1. 867

Momen, M., and E. Bou-Zeid, 2016b: Turbulence dynamics in unsteady atmospheric flow. Bulletin of the 69th 868

Page 43: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

43

American Physical Society, Division of Fluid Dynamics, American Physical Society (APS), Vol. 61 of 869

http://absimage.aps.org/image/DFD16/MWS_DFD16-2016-001285.pdf. 870

Momen, M., and E. Bou-Zeid, 2017a: Mean and turbulence dynamics in unsteady Ekman boundary layers. J. Fluid 871

Mech., 816, 209–242, doi:https://doi.org/10.1017/jfm.2017.76. 872

Momen, M., and E. Bou-Zeid, 2017b: Analytical reduced models for non-stationary diabatic atmospheric boundary 873

layers. Bound. Layer Meteorol., 164, 383–399, doi:https://doi.org/10.1007/s10546-017-0247-0. 874

Momen, M., J. D. Wood, K. A. Novick, R. Pangle, W. T. Pockman, N. G. McDowell, and A. G. Konings, 2017a: 875

Interacting Effects of Leaf Water Potential and Biomass on Vegetation Optical Depth. J. Geophys. Res. 876

Biogeosciences, 122, doi:10.1002/2017JG004145. 877

Momen, M., Z. Zheng, E. Bou-Zeid, and H. A. Stone, 2017b: Inertial gravity currents produced by fluid drainage from 878

an edge. J. Fluid Mech., 827, doi:10.1017/jfm.2017.480. 879

Monin, A. S., and A. M. Obukhov, 1954: Basic laws of turbulent mixing in the surface layer of the atmosphere. 880

Contrib. Geophys. Inst. Acad. Sci. USSR, 24, 163–187. 881

Monin, A. S., and A. M. Yaglom, 1975: Statistical Fluid Mechanics. MIT Press, Cambridge, MA, 769 pp. 882

Nieuwstadt, F. T. M., 1983: On the solution of the stationary, baroclinic Ekman-layer equations with a finite boundary-883

layer height. Boundary-Layer Meteorol., 26, 377–390, doi:10.1007/BF00119534. 884

Nunalee, C. G., and S. Basu, 2013: Mesoscale modeling of coastal low-level jets: impliactions for offshore wind 885

resource estimation. Wind Energy, 17, 657–669, doi:10.1002/we. 886

http://onlinelibrary.wiley.com/doi/10.1002/we.1608/full. 887

O’brien, J. J., 1970: A note on the vertical structure of the eddy exchange coefficient in the planetary boundary layer. 888

J. Atmos. Sci., 27, 1213–1215, doi:doi: 10.1175/1520-0469(1970)027. 889

Pedersen, J. G., M. Kelly, S. E. Gryning, and B. Brümmer, 2013: The effect of unsteady and baroclinic forcing on 890

predicted wind profiles in large eddy simulations: Two case studies of the daytime atmospheric boundary layer. 891

Meteorol. Zeitschrift, 22, 661–674, doi:10.1127/0941-2948/2013/0477. 892

Pedersen, J. G., S. E. Gryning, and M. Kelly, 2014: On the Structure and Adjustment of Inversion-Capped Neutral 893

Page 44: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

44

Atmospheric Boundary-Layer Flows: Large-Eddy Simulation Study. Boundary-Layer Meteorol., 153, 43–62, 894

doi:10.1007/s10546-014-9937-z. 895

Ramos da Silva, R., and R. Avissar, 2006: The Hydrometeorology of a Deforested Region of the Amazon Basin. J. 896

Hydrometeorol., 7, 1028–1042, doi:10.1175/JHM537.1. 897

Rizza, U., M. M. Miglietta, O. C. Acevedo, V. Anabor, G. A. Degrazia, A. G. Goulart, and H. R. Zimmerman, 2013: 898

Large-eddy simulation of the planetary boundary layer under baroclinic conditions during daytime and sunset 899

turbulence. Meteorol. Appl., 20, 56–71, doi:10.1002/met.1284. 900

Salesky, S. T., and M. Chamecki, 2012: A Similarity Model of Subfilter-Scale Energy for Large-Eddy Simulations of 901

the Atmospheric Boundary Layer. Boundary-Layer Meteorol., 145, 69–91, doi:10.1007/s10546-011-9618-0. 902

Salesky, S. T., and E. Bou-Zeid, 2016: On the Nature of the Transition Between Roll and Cellular Organization in the 903

Convective Boundary Layer. Boundary-Layer Meteorol., 163, 1–28, doi:10.1007/s10546-016-0220-3. 904

Shah, S., and E. Bou-Zeid, 2014: Very-Large-Scale Motions in the Atmospheric Boundary Layer Educed by Snapshot 905

Proper Orthogonal Decomposition. Boundary-Layer Meteorol., 153, 355–387, doi:10.1007/s10546-014-9950-906

2. 907

Sheppard, P. A., H. Charnock, and J. R. D. Francis, 1952: Observations of the westerlies over the sea. 78, 563–582, 908

doi:10.1002/qj.49707833807. 909

Sorbjan, Z., 2004: Large-Eddy Simulations of the Baroclinic Mixed Layer. Boundary-Layer Meteorol., 57–80. 910

Stull, R., 1988: An Introduction to Boundary Layer Meteorology. 1st ed. Kluwer Academic Publishers, Dordrecht, 911

378 pp. 912

Sun, J., D. H. Lenschow, L. Mahrt, and C. Nappo, 2013: The Relationships among Wind, Horizontal Pressure 913

Gradient, and Turbulent Momentum Transport during CASES-99. J. Atmos. Sci., 70, 3397–3414, 914

doi:10.1175/JAS-D-12-0233.1. http://dx.doi.org/10.1175/JAS-D-12-0233.1. 915

Tan, Z. M., 2001: An approximate analytical solution for the baroclinic and variable eddy diffusivity semi-geostrophic 916

Ekman boundary layer. Boundary-Layer Meteorol., 98, 361–385, doi:10.1023/A:1018708726112. 917

Tennekes, H., and J. Lumley, 1972: A First Course in Turbulence. MIT, Boston, 43-44 pp. 918

Page 45: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

45

Vallis, G. K., 2006: Atmospheric and Oceanic Fluid Dynamics. Cambridge University Press, New York, 104-116, 919

261-263 pp. 920

Venkatesh, S., and G. T. Csanady, 1974: A baroclinic planetary boundary-layer model, and its application to the 921

Wangara data. Boundary-Layer Meteorol., 5, 459–473, doi:10.1007/BF00123492. 922

Van de Wiel, B. J. H., a. F. Moene, G. J. Steeneveld, P. Baas, F. C. Bosveld, and a. a. M. Holtslag, 2010: A Conceptual 923

View on Inertial Oscillations and Nocturnal Low-Level Jets. J. Atmos. Sci., 67, 2679–2689, 924

doi:10.1175/2010JAS3289.1. http://journals.ametsoc.org/doi/abs/10.1175/2010JAS3289.1 (Accessed May 27, 925

2014). 926

Wyngaard, J. C., 2010: Turbulence in the Atmosphere. Cambridge University Press, 393 pp. 927

Young, G., D. A. R. Kristovich, A. Sciences, M. Hjelmfelt, and R. Foster, 2002: Rolls, streets, waves, and more. Bull. 928

Am. Meteorol. Soc.,. 929

Zilitinkevich, S., and I. N. Esau, 2003: The effect of baroclinicity on the equilibrium depth of neutral and stable 930

planetary boundary layers. Q. J. R. Meteorol. Soc., 129, 3339–3356, doi:10.1256/qj.02.94. 931

932

Page 46: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

46

TABLE 1. Details of the large-eddy simulations: 23 cases with two baroclinicity strengths, five baroclinicity 933

directions, and under three stabilities. The case names are constructed as follows: for example, for case NS180, the 934 first letter denotes the surface heat flux (N for neutral, U for unstable, and S for stable, replaced with placeholder * in 935 the table), the second letter represents the strength of baroclinicity (W for weak and S for strong), and the last number 936 shows the baroclinicity direction. In the second column, the values in parenthesis are the equivalent horizontal 937 temperature gradient in terms of Kelvin (K) per 100 km. A non-dimensional stability criterion, zi /L, is reported for all 938 cases in Table 2. 939

Simulation name

S0

(|Δθ|/Δx ~

K/100 km)

β

Heat Flux (W m−2)

Stable

(S)

Neutral

(N)

Unstable

(U)

Barotropic (*B) 0 (0) – – 20 0 200

Weak baroclinicity + positive shear (*W0) 0.25 (0.85) 0 0

Weak baroclinicity + cold advection (*W90) 0.25 (0.85) 90 0

Weak baroclinicity + hybrid case (*W135) 0.25 (0.85) 135 0

Weak baroclinicity + negative shear (*W180) 0.25 (0.85) 180 0

Weak baroclinicity + warm advection (*W270) 0.25 (0.85) 270 0

Strong baroclinicity + positive shear (*S0) 1 (3.4) 0 – 20 0 200

Strong baroclinicity + cold advection (*S90) 1 (3.4) 90 – 20 0 200

Strong baroclinicity + hybrid case (*S135) 1 (3.4) 135 – 20 0 200

Strong baroclinicity + negative shear (*S180) 1 (3.4) 180 – 20 0 200

Strong baroclinicity + warm advection (*S270) 1 (3.4) 270 – 20 0 200

940

941

Page 47: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

47

TABLE 2. The impact of baroclinicity, static stability, and their interactions on the friction velocity defined 942

as: . is the Obukhov length based on 943

the values of the parameters at the surface, zi the imposed inversion height (1000 m for unstable cases and 800 m for 944 stable cases), and κ the von Kármán constant (here 0.4). 945

Simulation name

(baroclinicity strength and

direction, see Table 1)

u* (m/s)

zi /L

Stable (S) Neutral (N) Unstable (U) Stable (S) Neutral (N) Unstable (U)

*B 0.28 0.37 0.53 7.6 0.0 –14.1

*W0 0.41 0.0

*W90 0.40 0.0

*W135 0.37 0.0

*W180 0.35 0.0

*W270 0.35 0.0

*S0 0.34 0.51 0.73 4.3 0.0 –5.4

*S90 0.29 0.44 0.62 6.9 0.0 –8.8

*S135 0.26 0.36 0.47 9.5 0.0 –20.2

*S180 0.25 0.30 0.31 10.7 0.0 –70.2

*S270 0.27 0.32 0.52 8.5 0.0 –14.5

946

947

Page 48: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

48

948

FIG. 1. Schematic view of the two parameters of the ABL for the hybrid cases (NW135, SS135, NS135, and 949

US135), where β = 135o. G0 is the geostrophic wind vector at the surface, which is along the x-direction here. (a): x-y 950

plane view where the radius of the plotted circle shows the strength of the baroclinicity S0 =S(z), which controls the 951

magnitude of gradient of the geostrophic wind vector, while the angle β shows the relative orientation of the surface 952

geostrophic wind and its gradient (the thermal wind, always parallel to the isotherms). (b): x-y-z view depicting the 953

profiles of Ug(z) and Vg(z) shown by the blue arrows at several heights. (c): An example of cold/warm advection cases 954

in the US. Solid black lines show the isopleths of equal 850 mb height and the red dashed lines indicate isotherms 955

(courtesy of https://www.wunderground.com/). 956

¶Vg

¶z

æ

èçö

ø÷0

¶Ug

¶z

æ

èçö

ø÷0

β = 135o

R ∝S0 (=S(z))

θ1

θ2

θ3

θ4

θ4>θ3>θ2>θ1

Cold advection +

Negative shear,

β = 135o

Isotherm

x

y

G0 direction

x

y

z

G0

(a) (b)

(c)

Negative shear, S180

z= zi

Page 49: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

49

957

FIG. 2. Velocity profiles in strong (solid lines) and weak (dashed lines) neutral baroclinic ABLs. The dotted lines 958

represent the geostrophic-wind profiles for the strong baroclinic cases, the forcings for the weak simulations can be 959

inferred by dividing the slopes of the strong cases by 4. 960

Page 50: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

50

961

FIG. 3. The effect of baroclinicity on the velocity hodographs for all neutral cases. The velocity vectors are 962

normalized by the magnitude of the geostrophic wind at the surface, G0. The blue and red dotted lines respectively 963

depict the geostrophic-wind vectors for the strong and weak baroclinicity. The points with (u, v) = (0, 0) denote the 964

surface z = z0 for the velocity plots, while the green dots represent the surface value of the 965

geostrophic wind vector , which is plotted as dashed lines. 966

Page 51: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

51

967

FIG. 4. Comparison of the simple reduced-order model (dashed lines) u and v profiles with the corresponding LES 968

time-space averaged profiles (solid lines). Blue lines denote u and red lines denote v. The dotted lines in (c,d) represent 969

the reduced-order model by decreasing the effective ABL depth in Eqs. (10) for eddy-viscosity profiles by about 40%. 970

Page 52: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

52

971

FIG. 5. The effect of baroclinicity strength and direction on the TKE budget and total TKE profiles. All the TKE 972

budget terms are normalized with zi/G02 and the TKE profiles with 1/G0

2. Solid lines show strong baroclinic cases and 973

dashed lines indicate weak baroclinic cases. Budget terms are defined in Eqs. (13) and (14). 974

Page 53: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

53

975

FIG. 6. The interacting effect of stability and baroclinicity: dimensionless wind speed profiles of baroclinic and 976

barotropic ABLs under stable, neutral and unstable conditions. The dotted lines represent the corresponding 977

geostrophic wind magnitude in height. 978

Page 54: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

54

979

FIG. 7. Velocity hodographs for all stable and unstable cases. The unstable ABLs have more uniform profiles and 980

hence their hodographs are shorter, whereas the hodographs of the stable simulations are more spread due to stronger 981

vertical decoupling and lower mixing. The black dotted lines indicate the geostrophic-wind vectors (for strong 982

baroclinicity). The points with (u,v)=(0,0) denote the surface z=z0 for the velocity plots, while the green dots represent 983

the surface geostrophic velocity vector. 984

Page 55: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

55

985

FIG. 8. The ratio of the shear production over buoyant production in the presence of strong baroclinicity. The 986

magnitudes are largely influenced by the baroclinicity in both stable and unstable conditions and the profile shapes 987

are remarkably distinct from their barotropic counterparts (f). 988

Page 56: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

56

989

FIG. 9. The interacting effects of baroclinicity and stability on the TKE budget and TKE profiles. All the TKE 990

budget terms are non-dimensionalized by zi/G02 and the TKE profiles are rescaled with 1/G0

2. This figure shows all 991

the unstable cases; the profiles of the stables cases can be found in the supplemental material. 992

Page 57: Modulation of Mean Wind and Turbulence in the Atmospheric ... · 37 comprehensive understanding of the physics of the atmospheric boundary layer (ABL). The ABL 38 imposes the bottom

57

993

FIG. 10. The impact of the baroclinicity on the direction and formation of convective rolls and cells. The left 994

subplots depict the contours of the instantaneous vertical velocity fields normalized by the geostrophic wind 995

magnitude at the surface G0 at z/zi = 0.1, and the right subplots shows the results at z/zi = 0.3. 996