momen&bou-zeid(2017) qjrms submitted - princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw...

45
The interacting effects of unsteady pressure gradients and static 1 stability in the atmospheric boundary layer 2 3 Running Head: The effects of unsteady pressure gradients and stability in the ABL 4 5 Mostafa Momen a,b and Elie Bou-Zeid b* 6 a Department of Earth System Science, Stanford University, Stanford, CA, US 7 b Department of Civil and Environmental Engineering, Princeton University, Princeton, NJ, US 8 * Correspondence to: E. Bou-Zeid, E414-CEE-EQuad, Princeton University, Princeton, NJ 9 08544, USA. Email: [email protected] 10 11 Keywords: Atmospheric Boundary Layer, Turbulence, Unsteady pressure gradient, Stability, 12 Reduced model, Large-eddy simulation 13

Upload: others

Post on 08-Mar-2021

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

The interacting effects of unsteady pressure gradients and static 1

stability in the atmospheric boundary layer 2

3

Running Head: The effects of unsteady pressure gradients and stability in the ABL 4

5

Mostafa Momena,b and Elie Bou-Zeidb* 6

aDepartment of Earth System Science, Stanford University, Stanford, CA, US 7

bDepartment of Civil and Environmental Engineering, Princeton University, Princeton, NJ, US 8

*Correspondence to: E. Bou-Zeid, E414-CEE-EQuad, Princeton University, Princeton, NJ 9

08544, USA. Email: [email protected] 10

11

Keywords: Atmospheric Boundary Layer, Turbulence, Unsteady pressure gradient, Stability, 12

Reduced model, Large-eddy simulation 13

Page 2: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

2

Abstract 14

This study investigates the response of the Atmospheric Boundary Layer (ABL) under steady 15

surface heating to changes in the horizontal pressure gradients. Both stabilizing and destabilizing 16

surface heat fluxes are considered. Using large-eddy simulations (LES), the maxima of the wind 17

speeds in the ABL (e.g. low-level jets in stable conditions) after increasing the pressure gradient 18

are shown to be related to the strength of the pre-existing turbulent vertical coupling in the system, 19

with the velocity overshoots decreasing when surface heat flux increases from the most stable to 20

the most unstable cases. The profiles of temperature, heat flux and turbulent kinetic energy are 21

found to be much more sensitive to unsteady pressure forcing in stable ABLs, compared to their 22

unstable counterparts where the turbulence is dominant and filters the pressure gradient variations. 23

To better understand the fundamental physics that control these responses, a reduced 24

analytical model for mean wind is further developed in this paper to be applicable under all 25

stabilities. The model is directly derived from the Reynolds-averaged Navier-Stokes equations and 26

validated against LES. The dynamics of the system according to this model are analogous to a 27

damped oscillator where the inertial, Coriolis, and friction forces correspond to the mass, spring 28

and damper, respectively. The generalization of the model to the non-neutral ABL requires 29

considering the effects of buoyancy on the model, and we show that this primarily impacts the 30

damping term corresponding to friction forces and vertical coupling. 31

32

Page 3: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

3

1. Introduction 33

Atmospheric and oceanic flows continuously vary with time. The winds in the atmosphere 34

fluctuate over a spectrum ranging from the synoptic (~ days), to the meso (~ hours), to the micro 35

(~ seconds) time scales. Many interacting and complex physical mechanisms underlie this 36

variability, which makes its observational study quite challenging. Models on the other hand can 37

start with one simple feature, and then sequentially add interacting and complicating features 38

(potentially by aggregating many individual studies) to the problem to better understand the 39

interacting dynamics of multiple processes. In the context of the velocity fields in the atmospheric 40

boundary layer (ABL), the timeline of this gradual build-up of complexity is illustrated in Table 41

1. This table lists some previous efforts to model the ABL over flat homogenous terrains under 42

steady-state conditions, as well as under variable static stability and variable driving mean 43

horizontal pressure gradient (topography and heterogeneity add further layers of complexity that 44

are out of the scope of this paper). Here, we focus on studies that aimed to develop simplified 45

models of the dynamics. For example, multiple studies investigated a steady ABL that is governed 46

by the equilibrium of the Coriolis, a constant friction, and the pressure gradient forces without 47

accounting for stability effects (i.e. the neural Ekman boundary layer, see rows 1, 3, 10, and 11 of 48

Table 1), although some also considered the baroclinic change of the pressure gradient with height. 49

However, the real-world winds vary in time due to the changes in the static stability of the 50

ABL and/or due to a variable pressure gradient. The diurnal variability of the surface buoyancy 51

flux can lead to the formation of the nocturnal low-level jets (see e.g. Banta et al. 2006; Bain et al. 52

2010; Mahrt 2013) that was first modelled by Blackadar (1957) using the inertial oscillation 53

concept. Some extensions to this mechanism were later suggested by many other researchers (rows 54

6, 13, 14, 15, and 16 of Table 1). The fluctuations in the geostrophic forcing (which is simply 55

Page 4: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

4

representing the large-scale horizontal pressure gradients driving the flow) cause variations in the 56

flow velocity as well (see large-eddy simulation studies of atmospheric and oceanic cases by Hsu 57

et al., 2000; Lohmann et al., 2006; Radhakrishnan and Piomelli, 2008; Gayen et al., 2010), and 58

various reduced models were proposed to explain and predict such variations (see rows 4, 12, and 59

16 of Table 1). 60

Nevertheless, as Table 1 illustrates, the previous analytical solutions in the unsteady 61

framework have some limitations including i) constraining assumptions about the structure of the 62

eddy-viscosity (thus limiting their applicability range), ii) intricate solutions of the governing 63

PDEs that are not always straightforward to acquire or interpret physically, or iii) lack of validation 64

against more accurate benchmarks such as large-eddy simulations (LES). Furthermore, no 65

previous work analyzes the unsteady pressure forcing cases under the full range of static stabilities 66

observed in the ABL. This paper aims to bridge this gap and propose a reduced model that is 67

applicable in a very wide array of conditions, including broad ranges of ABL stabilities and eddy 68

viscosity with time and height variabilities, by providing thorough LES validations. 69

In Momen and Bou-Zeid (2016), we developed a model for the unsteady pressure gradient 70

effects on neutral ABL mean wind velocities using an analogy to a damped-oscillator where the 71

inertial, Coriolis, and friction forces reflect the mass, spring and damper respectively. Next, in 72

Momen and Bou-zeid (2017a), we generalized the model by introducing a time-dependent 73

damping coefficient that was shown to be necessary to capture the gradual decrease in the surface 74

heat–flux in neutral to stable ABL transition (under steady pressure forcing). In both papers, we 75

validated the solutions with accurate LES results. Nonetheless, while we previously investigated 76

variations in the surface buoyancy flux and the pressure gradients, the two were considered 77

separately. Therefore, the influence of static stability, even when steady, on the response of the 78

Page 5: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

5

ABL to an unsteady pressure gradient was not considered despite being the analogue of many real-79

world conditions. 80

In this paper, we add this layer of complexity, brought about by a steady surface buoyancy, 81

to the investigation of ABLs under unsteady pressure gradients. We also extend the damped 82

oscillator modelling framework to parametrize the stability effects on turbulent friction in the 83

system, thus allowing the model to be applied without requiring LES. We also examine the 84

turbulent kinetic energy (TKE) budget dynamics in the diabatic unsteady ABLs, which have been 85

rarely studied before. 86

In particular, our research questions that address the physics and the mathematical modelling of 87

the diabatic ABLs (stable and unstable) are respectively as follows: 88

1) How does the interaction of buoyancy and variable pressure forcing influence mean and 89

turbulence in the ABL? 90

2) What modifications are required to the reduced damped-oscillator model of unsteady ABLs 91

to account for the effect of static stability? 92

We briefly describe the details of our LES in the next section. In Section 3, we investigate the 93

mean and turbulence dynamics of the unsteady diabatic ABLs to answer question 1. In Section 4, 94

we model wind variations using the damped-oscillator concept to answer question 2. It will be 95

shown that a time-invariant damping coefficient (though a function of height) can capture the LES 96

results well when the surface heat fluxes are steady. We also suggest a parametrization for the 97

damping term that is a function of the stability in the ABL. Finally, a summary of our findings is 98

given in Section 5. 99

Page 6: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

6

Table 1. A summary of previous studies of atmospheric boundary layer winds over homogenous terrain 100 that proposed reduced analytical models of the dynamics. 101

102

103

Pressure gradient

Buoyancy flux Validation Comments Reference

1 Steady ________ None The first model with height- and

time-constant viscosity Ekman (1905)

2 Steady Step change,

Stable None

Inertial oscillations of low-level jets with no frictional effects

Blackadar (1957)

3 Steady ________ None Ekman solutions for exponential

thermal winds

Mahrt and Schwerdtfeger

(1970)

4 Unsteady ________ None Analytical solutions for diurnally

periodic winds Bonner and Paegle

(1970)

5 Steady ________ None Suddenly applied uniform shear

stress Madsen (1977)

6 Steady Step changes Wangara

observations A two-layer slab model based on the

idea of Blackadar (1957) Thorpe and

Guymer (1977)

7 Steady Steady Observation

(Leipzig profile) Incorporated a surface layer at the

bottom of the Ekman layer Yordanov et al.

(1983)

8 Steady Unsteady, Unstable

None

Oscillatory character of mixed layers with a zeroth–order model; only numerical simulations were

validated with Cabauw observations

Byun and Arya (1986)

9 Steady ________ Observation

(Leipzig profile) A gradually varying eddy-viscosity

in height Miles (1994)

10 Steady ________ None A gradually varying eddy-viscosity in

height Grisogono (1995)

11 Steady ________ None A height-varying eddy-viscosity and

a baroclinic ABL Tan (2001)

12 Unsteady ________ None Provides exact solutions subject to a prescribed eddy-viscosity profile

Lewis and Belcher (2004)

13 Steady Step change,

Stable Cabauw

observations Included the frictional effects in the

Blackadar mechanism Van de Wiel et al.

(2010)

14 Steady Step change,

Stable None Impulsively varying eddy-viscosity

Shapiro and Fedorovich (2010)

15 Steady Neutral and

convective limit None

Used the concept of inertial oscillations

Schröter et al. (2013)

16 Unsteady Step change,

Stable None

Both Holton and Blackadar mechanisms with 2 eddy-viscosities

Du and Rotunno (2014)

17 Unsteady ________ 3D LES Steady but height variable frictional

effects, proposed neutral parametrization

Momen and Bou-Zeid (2016)

18 Unsteady Arbitrary

variability, Stable

3D LES Height and time-variable frictional effects parametrization

Momen and Bou-zeid (2017a)

19 Unsteady Steady, Stable,

neutral and unstable

3D LES Height and time-variable eddy-viscosity, proposing neutral and

non-neutral parametrization

Current work

Page 7: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

7

2. Details of the large-eddy simulations 104

2.1. Governing equations 105

The basic premise in LES is that the largest eddies contain most of the energy and are responsible for 106

the majority of the transport of momentum and scalars and consequently they are the most 107

dynamically important scales. Hence, the LES technique consists of directly resolving only the large 108

high-energy scales, while parametrizing eddies smaller than a cut-off (filter or grid) scale, to reduce 109

the computational cost of high Reynolds number (Re) simulations significantly. The mean pressure 110

gradient in our simulations is represented as a horizontal geostrophic wind with components Ug 111

and Vg as: 112

U

gz,t( ),Vg

z,t( )( ) = 1

G-

1

f r¶p

¶yz,t( ), 1

f r¶p

¶xz,t( )æ

èçöø÷

, (1) 113

where f is the Coriolis parameter (1.394×10–4 Hz here), and G is the magnitude of the geostrophic 114

wind vector (here the numerical value of G is taken as 8 m s–1) which, along with the ABL depth 115

zi, is used to normalize the governing equations and all the results of the paper. 116

The governing equations of the code are: the resolved mass conservation equation (an 117

equivalent Poisson equation for pressure is actually solved), the resolved momentum equation 118

(neglecting the molecular terms due to a very high Re regime) written in rotational form to ensure 119

conservation of mass and kinetic energy (Orszag and Pao, 1975), and the resolved thermal energy 120

conservation equation written in terms of the potential temperature, which are respectively given 121

as 122

, (2) 123

Page 8: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

8

,

(3) 124

,

(4) 125

where xi is the position vector; is the resolved velocity vector; 126

is the deviatoric part of the sub-grid scale (SGS) stress tensor; p* is a modified 127

pressure; θ is potential temperature with its reference value θr; is the deviation of the local θ 128

from its horizontal average; g represents the gravitational acceleration; and is the 129

SGS heat flux vector. Since the focus here is not on the LES, the details of the code such as the 130

imposed boundary and initial conditions and the grid configuration are discussed in Appendix A. 131

132

2.2. Suite of large-eddy simulations 133

We impose seven surface heat fluxes and change the pressure gradient for each simulation in two 134

ways: one by sinusoidally varying it and the other by instantly doubling it (step change). Table 2 135

shows all the performed large-eddy simulations. To study the ABL behaviour under the sinusoidal 136

pressure-gradients, we impose a frequency for the forcing variability, τf, that is equal to the natural 137

frequency of the inertial oscillations in the ABL, which is the Coriolis parameter characterizing 138

the response time of the mean flow as τABL = 2π/f ≈ 12.5 hr (Tennekes and Lumley, 1972). The 139

pressure forcing of these resonant frequency cases, expressed as (Ug, Vg), is thus given as a 140

sinusoidally-varying function of time 141

( ) ( ) ( ) ( )cos 2 / , sin 2 /g f g fU t t V t t = = , (5) 142

where τf is the forcing time scale. The pressure forcing of the step changes is given as 143

Page 9: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

9

. (6) 144

Recall that these are the values normalized by the dimensional geostrophic wind velocity 145

G = 8 m s–1. We also note that these very long times of constant stability may not occur regularly 146

in the real ABLs due to the diurnal cycle (perhaps they happen only in the poles or in some 147

engineering applications). In addition, such a long-lived consistent variability in the pressure 148

forcing is unlikely. However, the long simulations are aimed at ensuring convergence of the 149

computed statistics, which will be applicable to shorter occurrences of these conditions when they 150

actually arise in nature (e.g. a step change in the pressure gradient, which represents a front, could 151

happen under a stable or an unstable ABL for several hours). 152

153 Table 2. Suite of the large-eddy simulations. L is the Obukhov length and zi is the initial (steady-state) 154

inversion height that is listed in Table 3. 155

Surface

heat flux

# Acronym Run

Time

dt (s) Nx,y,z zi/L Lz Imposed

Inversion base

Sponge

Layer

–40 w/m2

1 Step–40 40 hr 0.2 1603 7.2 1600m 1200m at 1200m

2 Sin–40 25 hr 0.08 1603 7.2 1600m 1200m at 1200m

–20 w/m2

3 Step–20 40 hr 0.2 1603 2.8 1600m 1200m at 1200m

4 Sin–20 25 hr 0.08 1603 2.8 1600m 1200m at 1200m

–10 w/m2

5 Step–10 40 hr 0.2 1603 1.5 1600m 1200m at 1200m

6 Sin–10 25hr 0.08 1603 1.5 1600m 1200m at 1200m

0 w/m2

7 Step 0 40 hr 0.2 1283 0 1000m NO NO

8 Sin 0 40 hr 0.1 1283 0 1000m NO NO

+25 w/m2

9 Step+25 40 hr 0.2 1283 –7.0 1700m 1250m NO

10 Sin+25 40 hr 0.1 1283 –7.0 1700m 1250m NO

+100 w/m2

11 Step+100 40 hr 0.2 1283 –20.4 1700m 1250m NO

12 Sin+100 40 hr 0.1 1283 –20.4 1700m 1250m NO

+400 w/m2

13 Step+400 40 hr 0.2 1283 –61.2 1700m 1250m NO

14 Sin+400 40 hr 0.1 1283 –61.2 1700m 1250m NO

156

Ug z,t( ) = 1 t < 20 hr2 t ³ 20 hr

ìíî

, Vg z,t( ) = 0

Page 10: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

10

3. Mean and turbulence evolution 157

In this section, we aim to answer our first question: how does the diabatic ABL respond to an 158

unsteady pressure forcing? To that end, we first compare the mean wind and TKE dynamics under 159

various stability conditions. Then, we investigate the implications on the temperature and heat flux 160

profiles. Finally, we examine the behaviour of the different terms of the TKE budget in such flows. 161

3.1. Wind velocity and TKE profiles 162

The evolution of the mean velocity field in the ABL for three step-change cases, the most stable, 163

the neutral, and the most unstable cases, are exhibited as they illustrate the general trends. Figure 164

1(a) shows the u-velocity time series at z/zi = 0.65. It is clear that the stable case overshoots more 165

strongly and peaks earlier, while the velocity in the unstable case fluctuates less and peaks later 166

than the other cases as the forcing varies. This response is a direct result of the fact that before the 167

jump (at t < 20 hr) the stabilizing heat fluxes, in the stable case, damp the turbulence and decrease 168

the friction and vertical coupling in the system. Hence, the velocity increases more since the 169

turbulent stresses that would hold it back are weaker and take a longer time to increase, after the 170

shear increases, to balance the newly imposed pressure gradient. On the other hand, in the unstable 171

case, the destabilizing heat fluxes produce higher turbulence and friction and hence lower vertical 172

gradients are required to balance the pressure gradient. Moreover, the stresses develop faster due 173

to the shorter eddy turn over times, which leads to a slightly delayed peak. Consequently, the 174

overshoot in the velocity is lower than other cases. The neutral case stands in between these two 175

limits. 176

We investigate the space-time responses of the ABLs under the step-change pressure 177

gradient to fully unravel their dynamics. Figure 1(b), (c), and (d) display these graphs respectively 178

in the most stable (see also Video S1), neutral, and most unstable (see also Video S2) conditions. 179

Page 11: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

11

In addition to the observation that the velocity is usually higher in the stable case and lower in the 180

unstable one as illustrated in Figure 1(a), one can also note here that in the unstable case (Figure 181

1(d)) the profiles are more uniform with height due to strong vertical mixing. In the stable ABLs, 182

the atmospheric strata are more decoupled resulting in stronger vertical variability as expected. As 183

a result, there exists a layer near the surface where the time variation of the velocity is stronger in 184

the unstable case than stable one due to strong vertical coupling causing a high downward flux of 185

momentum (compare 1(b) and 1(d) close to the surface); this layer was found to extend up to the 186

height of the turbulent stable boundary before the jump (e.g. ≈ 160 m in Step–40 case). 187

Table 3 lists the friction velocity u* and the height of the ABL zi before and after the 188

pressure-gradient jump in the step-change simulations. For the stable cases, we calculated zi by 189

locating the height where the TKE decreases to 5% of its magnitude near the surface. In the 190

unstable ABL, the imposed inversion fixes zi, which remains constant. One can see that the u* after 191

the jump increases in all the cases, which is expected since the pressure gradient, velocity, and 192

hence the shear rise after the jump. Furthermore, the friction velocities increase monotonically 193

with increasing surface heat flux. The height of the ABL, zi, also grows after the jump in the stable 194

conditions, and as expected it decreases as the stabilizing surface heat flux strengthens. 195

196

Page 12: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

12

197

Figure 1. Evolution of u under various stabilities: (a) times series in the middle of the ABL; and space-198

time graphs for cases (b) Step–40, (c) Step 0, and (d) Step+400. 199

Page 13: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

13

Table 3. Friction velocity and ABL heights under two pressure-gradient values and different stability 200

conditions. 201

Surface heat flux

#

Name

G = 8 m/s G = 16 m/s

u* (m/s) zi (m) u* (m/s) zi (m)

–40 w/m2 1 Step–40 0.21 160 0.49 320

–20 w/m2 3 Step–20 0.25 210 0.52 420

–10 w/m2 5 Step–10 0.27 280 0.54 560

0 w/m2 7 Step 0 0.31 1000 0.60 1000

+25 w/m2 9 Step+25 0.36 1250 0.67 1250

+100 w/m2 11 Step+100 0.4 1250 0.7 1250

+400 w/m2 13 Step+400 0.43 1250 0.76 1250

202

Inertial oscillations in the stable ABL occur near zi, where the TKE is low. The horizontal 203

winds oscillate with the Coriolis frequency and form the much-studied low-level jets (LLJ, see 204

Blackadar, 1957; Shapiro and Fedorovich, 2010; Van de Wiel et al., 2010). The mean and TKE 205

time series and profiles for the stable ABL simulations are shown in Figure 2. The left panels of 206

Figure 2 exhibit the dimensional LES data, while the right panels show the same data normalized 207

with zi for elevation and with the geostrophic velocity scale (G = 8 m/s) for velocities, as shown 208

in Table 2. The formation of a supergeostrophic velocity can be observed in the stable cases (Figure 209

2(e) and 2(f)). As we increase the stabilizing heat flux, the height of the ABL and of the LLJ 210

decrease, as illustrated in Figure 2(e). Furthermore, the friction in the system decreases with 211

increasing stability; consequently, the increase in the velocities, i.e. the magnitude of the LLJ, is 212

higher in the strongly stable cases. The level of the LLJ almost doubles after the jump. The 213

resonance phenomenon (due to imposing a forcing with a time scale close to the inertial time scale) 214

Page 14: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

14

is clear from Figure 2(g) and 2(h) showing super-geostrophic velocities. The overshoots in the 215

Step–40 case are larger than the Step–20 and Step–10 since the turbulent friction and vertical 216

coupling are weaker in this case. 217

The right panels of this Figure show that after normalizing the profiles and time series, they 218

collapse well (Figure 2(f) and 2(h)) except perhaps velocities of some cases near the wall. Since 219

G is the same for all cases, this collapse is mainly due to the rescaling of the elevation with zi. The 220

few deviations could be due to the fact that for these cases, immediately after the jump, turbulence 221

is unsteady while we are normalizing with the steady-state variables and several inertial 222

oscillations are required for the flow to reach a full (quasi) steady-state condition. We can account 223

for unsteadiness by rescaling based on a time-dependent variable. For example, a proper 224

characteristic velocity near the wall is friction velocity rather than the geostrophic velocity. In 225

supplementary Appendix S1, we normalize the velocity profiles in the stable step-change 226

simulations, and the profiles collapses improve near the wall (especially for z/zi < 0.15). 227

The TKE in the stable step-change simulations, depicted in Figure 2(c) and 2(d), changes 228

along with the mean variations shown in Figure 2(a) and 2(b). Another interesting feature of the 229

TKE time series shown in Figure 2(c) and 2(d) is the lag between the two heights (compare solid 230

and dashed lines at t ≈ 20 hr). As this Figure indicates, there is ~ 1 hr lag between the lower and 231

higher elevations at t ≈ 20 hr. This lag is due to the fact that shear production first generates TKE 232

near the surface, and then that TKE is transported upward. The lag is more pronounced in the 233

stable ABLs than the unstable ones, indicating that the layers are more decoupled, with slower 234

TKE transport, under stabilizing heat fluxes. 235

236

Page 15: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

15

237 Figure 2. M and TKE time-series, and M profiles for 6 stable cases under step (a, b, c, d, e) and sinusoidal 238

change (g, h) for dimensional data (left panel) and normalized data with G and zi (right panel). 239

Page 16: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

16

In the unstable ABL simulations, the velocities are, in general, more uniform due to higher 240

mixing (Figure 3(a), (b), (e), and (f))). The TKE increases with increasing destabilizing heat flux, 241

in these cases as shown in Figure 3(c) and 3(d). As a result, the friction in the system is higher in 242

the stronger unstable cases and thus after the jump the overshoot is slightly lower for the 243

simulations with stronger heat fluxes (Figure 3(a)). This damping is more obvious in Figure 3(b) 244

for the resonant-frequency cases, where the stronger unstable cases show a weaker increase in the 245

velocity magnitude as time evolves. 246

In addition, since the velocity profiles are more uniform in the unstable step-change cases, 247

we generally obtain better collapses of the velocity time series and profiles as Figure 3(a) and 3(e) 248

indicate. Nevertheless, for the resonant frequency cases, which are always unsteady, the velocities 249

in the three cases increasingly diverge for longer times (Figure 3(f)) if we use the constant 250

characteristic velocity scale G for non-dimensionalizing velocities. Although we do not investigate 251

the formulation of improved characteristic velocities for this case, we suggest that a combination 252

of u* and w* might yield better results. The weaker unstable case, which has lower turbulent 253

friction, exhibits higher peaks than the other cases as expected (Figure 3(f)). 254

255

Page 17: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

17

256

Figure 3. LES results of the step-change (left panels) and sinusoidal change (right panels) in the pressure 257

gradient under three unstable cases for: (a,b) the magnitude of the velocity time series, (c,d) the TKE time 258

series, and (e,f) the time evolution of the velocity profiles. 259

Page 18: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

18

3.2. Temperature and heat-flux profiles 260

Since static stability was shown to influence the response of the ABL to a change in the pressure 261

gradient, if that response results in changes in temperature and heat-flux profiles, feedbacks will 262

set in and lead to an increased dynamical complexity. Figure 4(a) indicates that in the steady part 263

of the Step–40 simulation, before the jump in the pressure gradient (i.e. at t ≤ 20 hr), the 264

temperature decreases due to the negative surface heat flux but the gradients that control the static 265

stability do not change markedly. However, in the transition period after the jump, the ABL warms 266

up because of the increase in the shear and higher mixing that brings hotter air from aloft closer to 267

the surface. After the ABL reaches the new equilibrium (e.g. ~35 hr as shown in Figure 4(a) and 268

4(b)), it starts cooling again and the vertical mixing decreases. The rate of the pre-jump cooling 269

and post-jump heating of the ABL at each height could be inferred from the profiles of the heat 270

flux depicted in Figure 4(b), where a positive vertical gradient of the flux indicates cooling at that 271

height and a negative gradient heating. In the stably stratified ABL, the heat-flux profiles are often 272

represented according to the similarity relationship:

¢w ¢q z( )¢w ¢q surface

= 1-z

h

æèç

öø÷

g

(Stull 1988). We obtain a 273

γ ≈ 1.3 for Step–40 case before the jump, which is consistent with the range indicated in Stull 274

(1988) as γ = 1-3. 275

A similar trend for the strongly stable ABL under the sinusoidal change in the pressure 276

gradient is found as Figure 4(c) and 4(d) exhibit. However, because the increase in the shear and 277

consequently the mixing is continuously rising in this case, due to forcing at the resonant 278

frequency, the ABL keeps heating until it reaches the point where the maximum magnitude of the 279

velocity and the rate of mixing become constant. 280

Figure 4(e) and 4(f) show the temperature and heat-flux profiles for the strongly unstable 281

case under step-change and sinusoidal forcing. It is clear that doubling the pressure gradient in this 282

Page 19: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

19

case does not have a significant impact on the heating of the ABL or on the shape of the 283

temperature profiles because the mixing is already very high in the unstable ABL. Furthermore, 284

the heat-flux profiles and temperature gradients are similar under the step-change and sinusoidal-285

change forcings. The temperatures are almost the same at the beginning of Step+400 and Sin+400 286

(Figure 4(e)), and after about two inertial oscillations (~25 hr) the sinusoidal case becomes only 287

slightly warmer, perhaps due to slightly higher shear turbulent kinetic energy production. In the 288

convective ABL, before and after the jump, the heat fluxes are changing almost linearly in height, 289

consistent with the similarity relationship:

¢w ¢q z( )¢w ¢q surface

= 1-lz

zi

(Stull, 1988) where a value of λ ≈ 1.2-290

1.5 is reported. In our LES, we obtain a value of 1.4 for Step+400 based on Figure 4(f). 291

In summary, the static stability inferred from the temperature gradients or heat-flux profiles 292

is strongly altered by the changes in the pressure gradients under stable conditions, which then 293

feeds back to the response of the velocity and stress to these changes. On the other hand, the static 294

stability profiles in the unstable ABL are more resilient and remain unchanged when the pressure 295

gradients vary (Figure 4(e) and 4(f)). 296

Page 20: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

20

297 Figure 4. Potential temperature and heat-flux profile at different times for 3 cases: (a,b) Step–40, (c,d) Sin–298

40, (e,f) Step+400 and Sin+400. In the stable sinusoidal forcing (c,d), the times 4 hr and 9 hr correspond to 299

the local minima in the wind-speed oscillations, while times 14 and 19 hr are near local maxima in the time 300

series. The times in the step change simulations are before and after the jump in the forcing that occurs at t 301

= 20 hr. We also selected the same times in the unstable sinusoidal simulation as those for the step-change 302

to show the effect of the type of variability of the pressure-gradient in the dynamics. The differences 303

between the two variabilities are minor indicating that in unstable cases the scalar quantities are not 304

influenced significantly by the change in the pressure gradient. 305

Page 21: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

21

3.3. TKE budget 306

We now investigate the response of the different terms of the TKE budget to variable pressure 307

gradients. We neglect the SGS kinetic energy since it only makes a small contribution relative to 308

its resolved counterpart over most of the domain (see Momen and Bou-zeid 2017b). Moreover, we 309

assume no subsidence w=0 and statistical homogeneity in both horizontal directions due to domain 310

periodicity. This allows us to simplify the budget equations to the following conventional form for 311

the resolved TKE ¢ ¢= 0.5 i ie u u : 312

q q

¢¢ ¢ ¢ ¢ ¶¢¶ ¶ ¶ ¶ ¶¢ ¢ ¢ ¢ ¢= - - - - - ¶

¢¶ ¶

¢

¶ ¶ ¶3i i

i ijjue u v w w p w eu w v w g S

t z z z z w, (7) 313

where the overbar denotes the ensemble average, and the prime denotes the turbulent fluctuations 314

from that average. We compute the TKE budget in terms of the resolved variables and define the 315

following expressions for the terms in Eq. (7) as 316

(8) 317

where P, BP, RT, and ε are respectively the shear production, buoyant production/destruction, 318

resolved turbulent transport, and SGS dissipation (which is a TKE cascade to the SGS scales). 319

Here, we only calculate and show the resolved turbulent transport term and neglect other transport 320

terms (SGS and pressure transport) since they make smaller contributions than the resolved 321

turbulent transport term. Furthermore, we focus primarily on the production and dissipation terms, 322

Page 22: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

22

both from shear and buoyancy. Please see supplementary Appendix S2, or Momen and Bou-Zeid 323

(2017b) for further details about the calculation of these terms and the TKE budget closure in the 324

LES code. 325

After the jump in the pressure forcing, the shear production increases in Step–40 case as 326

Figure 5(a) shows. This rise in the shear happens in other cases (Figure 5(b)) as well when the 327

magnitude of the velocity increases. The SGS dissipation increases accordingly to adapt to the 328

higher produced TKE in the flow. 329

Figures 5(c) and 5(d) display the TKE budget terms in the unstable conditions. Lenschow 330

et al. (1980) measured these terms in the convective ABL. Our results before the jump in Figure 331

5(c) are in a good qualitative agreement with their TKE budget measurements in the unstable ABL. 332

While again the changes in the stable case are more pronounced than in the unstable case, the 333

results in the unstable case do not collapse well at various times particularly near the wall. 334

Although one might expect them to collapse if normalization with u* is carried out. 335

The buoyancy term destroys the turbulence in the stable ABL and produces TKE in the 336

unstable ABL as Figure 5 confirms. The magnitude of this term is prescribed at the surface through 337

the boundary conditions and thus the surface value is not affected by the change in the pressure 338

gradient. This is the physically correct boundary condition to set since in both stable and unstable 339

boundary layer the surface energy budget is controlled by the radiative balance of the surface. This 340

net radiative balance Rn, along with the small contribution of the ground heat flux G, set the 341

available energy Rn–G that is to be distributed between the sensible (H) and the latent (LE) heat 342

fluxes. Given a value of Rn–G, which to first order will remain constant (despite small changes in 343

outgoing longwave radiation due to changes in the surface temperature if the wind speed 344

increases), the turbulent transfer efficiencies for both sensible and latent heat will increase about 345

Page 23: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

23

equally and thus the distribution of the available energy between H and LE will also, to first order, 346

remain unchanged. Therefore, H would remain approximately constant at the surface in the real 347

world ABL. Nevertheless, at higher elevations in the stable ABL (Figure 5(a) and 5(b)), the heat 348

flux responds to the variations of the pressure gradient and increases to destroy the additional 349

produced turbulence. However, in the unstable ABL (Figure 5(c) and 5(d)), buoyancy production 350

is not remarkably influenced by the increase of the shear since the buoyant production under strong 351

unstable conditions is always dominant over the shear production except very close to the wall. 352

This is in agreement with the observation in the previous subsection that the heat flux is not 353

sensitive to the pressure gradient fluctuations. 354

As Figure 5 shows, the resolved transport term does not play a significant role in the stable 355

ABL since the buoyant destruction reduces the turbulence at each level, damps the vertical 356

velocity, and does not allow strong vertical transport of the TKE produced near the surface to 357

higher elevations, which is another indication of vertical decoupling in such flows. However, it 358

becomes quite important in the convective ABLs where it adjusts itself to the increase in the shear 359

production while the buoyant production remains almost constant. 360

Furthermore, based on Figure 5(a), we can see certain heights in the stable boundary layer 361

where the production term becomes negative. This negative shear production is called global 362

backscatter, denoting energy transfer from the turbulence to the mean flow and is also observed 363

when the flow undergoes abrupt transition in space or time (Bou-Zeid et al. 2009). The height of 364

this negative production increases after the jump. If we compare the location of these negative 365

production layers before and after the step change in the pressure gradient, we observe that it 366

occurs close to the LLJ height. 367

368

Page 24: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

24

369

Figure 5. TKE budget profile for 4 cases. In step-change simulations, t1 is when the flow is steady right 370

before the jump, and t2 represents a time after the jump when the mean flow is unsteady and has not reached 371

the full equilibrium. In the sinusoidal changes, both times indicate unsteady conditions in which t1 is close 372

to a trough and t2 is close to a peak in the velocity time series. The first node is not shown here because it 373

is highly influenced by the wall model. Note that all the profiles are time-averaged over τABL/4 around the 374

times shown in the Figure. 375

376

The time series analyses of the TKE budgets also show that the variability in the pressure 377

gradients strongly influences the TKE budget terms in the stable ABLs, while its effects are minor 378

in unstable conditions, particularly at higher elevations (z/zi > 0.3). Figure 6 displays the TKE 379

budget time series for all stable (top panel) and unstable (bottom panel) simulations. It is clear 380

Page 25: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

25

again that the resolved transport terms are small in stable cases (since the layers are decoupled), 381

but significant in unstable conditions as expected. 382

Furthermore, the buoyant production variability is not influenced by doubling the pressure 383

gradient in all unstable ABLs regardless of the amount of the imposed heat flux. However, in stable 384

cases buoyant destruction increases up to 5-12 times at their peak after the step-change in the 385

pressure gradient, compared to their steady-state value before the change. The variability of 386

buoyant production in the most stable case is more than the weakest stable simulation. 387

388

389 Figure 6. TKE time series for stable conditions (top panel) and unstable conditions (bottom panel). The 390 stable cases are more sensitive to changing the pressure gradients than the unstable simulations. In unstable 391 conditions, the budget unsteadiness decreased at higher elevations and the level of the variability before 392 and after the change in the pressure gradient was almost the same at z/zi > 0.3. 393

Page 26: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

26

4. Reduced model for the emerging dynamics 394

In this section, we investigate how we can model the mean velocity under variable pressure 395

gradients under different stabilities. To that end, we will extend a simplified model, which was 396

previously developed by Momen and Bou-Zeid (2016) for neutral ABLs, by parametrizing the 397

effect of constant stabilizing/destabilizing heat fluxes on the frictional effects that modulate the 398

mean velocity variations. 399

400

4.1. The damped-oscillator model 401

We start with the Reynolds-averaged Navier-Stokes (RANS) equations: 402

, (9) 403

where u and v are the mean horizontal velocity components at height z and time t. The non-404

linear advective terms are negligible due to the horizontal homogeneity and small subsidence w. 405

Note that in this section all the velocity variables (except where a perturbation is explicitly denoted 406

by a prime) are the averaged velocities and overbars will be henceforth omitted for notational 407

simplicity. 408

The turbulent stress gradients can be modelled using the mean velocities as 409

( )

( )

( , ) ( , ),

( , ) ( , ),

u w

v

z t u z

w

tz

z t v z tz

¶¶¶ ¢ ¢ =¶

¢ ¢ = (10) 410

where 1/α is a time scale in our problem. The exact expression of α depends on the best similarity 411

theory that describes the flow dynamics (e.g. log-law, defect-law, or Rossby-number similarity), 412

and it should be stressed that the model in Eq. (10) is only accurate if the turbulence is in (or at 413

Page 27: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

27

least close to) quasi-equilibrium with the mean flow (i.e. there exists a similarity theory that relates 414

the stresses to the mean velocities). This parametrization of the stress terms is a very classic model 415

of their divergence as being proportional to the velocity at the same height (e.g. Haurwitz 1947; 416

Schmidt 1947; McNider 1982; Baker et al. 1987; Polavarapu 1995; Pu and Dickinson 2014; Du 417

and Rotunno 2014; Momen and Bou-zeid 2017a). Note that α(z,t) in its general form is both a 418

function of height and time. But here we will postulate that its time-averaged value is 419

sufficient for the current model for cases where the buoyancy forcing is steady in time (an 420

assumption that will be validated later) and we will hereafter use that averaged value (omitting the 421

overbar and writing it as α(z)). 422

Let us define the complex variable A u + i v; substituting Eq. (10) into Eq. (9) and 423

differentiating once with respect to time we get: 424

. (11) 425

The dynamics of the system are therefore analogous to a damped oscillator (such as a mass-spring-426

damper or MSD system) where the inertial, Coriolis, and friction forces are equivalent to the mass, 427

spring and damper, respectively. 428

When a steady buoyancy is superposed on the unsteady pressure gradient, the same model 429

structure can be maintained, but the damping term in the model, corresponding to friction forces, 430

now needs to account for stability since the change in the stability modifies the similarity theory 431

(used in Eq. 10) that best describes the equilibrium profiles and needs to be accounted for (we will 432

validate this assumption in Section 4.3). Therefore, adding a constant stability can only change the 433

damper or α of our equation of motion (11). However, for the reverse case with constant pressure 434

gradient under variable surface heat flux and stability, the analytical model needs to be extended 435

Page 28: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

28

to allow a time-variable damper coefficient, which was done in Momen and Bou-zeid (2017a) who 436

solved this problem for time-varying coefficients α. A linear decrease in time of the heat flux with 437

a constant pressure gradient was examined in that work, and time-variable α solutions were 438

validated against the large-eddy simulations and showed better agreement with LES, for the LLJ 439

attributes, compared to previous models of Blackadar (1957) and Van de Wiel et al. (2010). In this 440

paper, a time-constant α is sufficient, and this leads to the following homogeneous solution of the 441

ODE 442

, (12) 443

where C0 is determined from initial conditions. The full general solution of (11) also requires 444

particular solutions associated with the forcing term (see Momen and Bou-Zeid, 2016). 445

446

4.2. The time scales and vertical coupling in the diabatic ABL 447

The derived equations show that the two important time scales in our problem are: 1) the natural 448

inertial time scale of the ABL τABL = 2π/fc , and 2) the turbulence time scale (e.g. the time scale of 449

the largest turbulent eddies in the ABL which is τt = zi/u*). If the turbulence intensity (taken as its 450

exact definition of TKE normalized by the mean flow velocity) is high in the flow, the 451

characteristic velocity of the eddies will increase and their turn-over time will decrease; therefore, 452

the friction and the damper coefficient α will increase. Note that α in our model, based on Eq. (10), 453

indicates how strong the vertical coupling in the system is. Turbulence intensity and vertical 454

coupling are related: under unstable conditions the upward transport of TKE from the layers near 455

the surface reduces the turbulence intensity in these layers and simultaneously increases it further 456

aloft, in the layers that are a net recipient of TKE transport; on the other hand, the reverse applies 457

in the stable ABL. This leads to αunstable> α neutral > αstable in the outer layer and 458

Page 29: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

29

αunstable< αneutral < αstable in the surface layer. But since the turbulence intensity is generally higher 459

under unstable conditions, the overall solutions are damped more in an unstable regime than a 460

stable regime. Moreover, this damping in the unstable regimes increases as the heat flux increases, 461

and in the stable regimes weakens as the negative heat fluxes decrease (increases in magnitude). 462

Therefore, as illustrated in Figures 1, 2(a), and 3(a), after the jump occurs in the geostrophic 463

forcing, the most stable cases show higher overshoot because of the weaker damping. 464

This argument becomes clearer when α is directly calculated from LES for all different 465

stability scenarios. To do so, we define the total Reynolds stresses as: 466

R13 - uw-u w( ) - xz

R23 - vw- v w( ) - yz

, (13) 467

where the first term on the right-hand side of Eq. (13) is the resolved component and the second 468

term denotes the SGS part of the total corresponding Reynolds stress. Figure 7 shows the calculated 469

dimensional α for different stabilities, after the jump in pressure gradient, based on Eqs (10) and 470

(13). As this figure indicates, the directly computed α increases monotonically away from the 471

surface when the surface heat flux rises from the most negative to the most positive, while near 472

the wall the opposite trend is observed. The opposing trends at various heights in the variability of 473

α in the surface and outer layers are directly related to the trends of stress gradient normalized by 474

the mean wind speed (as the direct expression of α in Eq. 14 will later show). Furthermore, this 475

figure shows that α becomes less variable in height as the heat flux increases, also implying higher 476

mixing and hence uniformity of profiles under unstable conditions. Note that we will non-477

dimensionalize α in the next subsection based on G and the physical zi, but here we show the 478

dimensional values for better comparison of all the cases with different zi. 479

480

Page 30: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

30

481 Figure 7. Dimensional α directly calculated by LES for all step-change simulations. 482

483

4.3. A similarity model for the turbulent friction 484

In this subsection, we parametrize the α coefficient of our damped-oscillator model based on 485

characteristic length and velocity scales that vary with the stability of the ABL. This will allow us 486

to apply our model in any setup without requiring LES. 487

For the neutral ABL, we previously proposed a formulation for the eddy-viscosity and 488

showed its relation with α in Momen and Bou-Zeid (2016). This analytical expression for eddy-489

viscosity consisted of a characteristic length scale, and a characteristic velocity scale (the 490

geostrophic velocity). The model was in a good agreement with LES, with some deviation near 491

the wall since the assumed Boussinesq analogy (implying αx = αy) is not an accurate approximation 492

in that region (it is simple to relax this assumption; but for the purpose of illustrating the model 493

skill it is not critical and thus we will invoke this analogy in the present paper as well); in addition, 494

the geostrophic velocity was used in that paper, which is not an appropriate velocity scale near the 495

wall. 496

Page 31: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

31

Now, we intend to develop a model for α that accounts for the stability effects. We model 497

this coefficient in two regions using different scales: first in the surface layer (z/zi < 0.25) where 498

the effect of the wall and friction velocity is more important and second in the outer layer (z/zi > 499

0.25) where the free stream conditions are more significant. Considering the definition of α, we 500

can approximate it as: 501

. (14) 502

where lm denotes a characteristic mixing length. 503

For the surface layer, lm (similar to Delage, 1974; Lacser and Arya, 1986; Estournel and 504

Guedalia, 1987; Stull, 1988) can be written as: 505

,

1 1

m sl z L

= , (15) 506

where lm,s represents the mixing length in the surface layer, κ is the Von-Karman constant (here 507

0.4), β is an empirical constant, and L is the Obukhov length: 508

L =

-u*3q

r

g ¢w ¢q. (16) 509

Note that one can also add a term to Eq. (15) to account for the free stream conditions like 510

geostrophic velocity and Coriolis effect as in other works; however, we do not add that term here 511

since this part of the model is strictly for the surface layer where the effect of the free stream 512

conditions and Coriolis forcing are weak. In order to apply the model of Eq. (14), we also need to 513

use an analytical expression for the profile of M. We use the Monin-Obukhov similarity theory 514

(Monin and Obukhov 1954), which is valid in the surface layer, to find the necessary relationship 515

Page 32: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

32

for the wind profile with stability effects. Substituting the Monin-Obukhov profile and the mixing 516

length expression into Eq. (14) yields: 517

=u

*

log z z0( ) y z L( )

1

zL

æèç

öø÷

, (17) 518

where ψ(z/L) is the conventional stability correction function based on the Obukhov length. Here 519

we use the Businger-Dyer relationship for this function as: 520

2

1

4.7 01 12 ln ln 2 tan 2 02 2

z L z Lz

x x x z LL

, (18) 521

where x=(1–15z/L)1/4. The ψ(z/L) expression for the unstable case in Eq. (18) is found by Paulson 522

(1970). 523

In the outer layer, our LES results indicate that the turbulent stresses change almost linearly 524

(supplementary Appendix S3). Therefore, the local u*2 decreases linearly with height. In addition, 525

the variation of M is usually less than the turbulent stresses in the outer layer; hence, a first-order 526

approximation that assumes M is a constant would be sufficient; it yields a linearly decreasing α 527

with z. To ensure that the surface and outer layer parametrizations match, we compute α(z/zi=0.25) 528

from the surface layer and then decrease it linearly in height as: 529

( ) ( ) ( )

,

0.250.25 1 i

im o

z zz z

l

-=

, (19) 530

where now the mixing length in the outer layer (lm,o) depends on the stability and an outer 531

turbulence mixing length l∞ that applies far from the surface following 532

,

1 1

m ol l L

= . (20) 533

Page 33: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

33

We use l∞ =7m based on the results of Huang et al. (2013). β is an empirical constant that can be 534

found from experiments or numerical simulations. Our LES results indicate that the values of 535

β = 47 in the inner layer, and β = 24 for the outer layer yield good results for all the simulations 536

conducted in the paper (the validations and model comparisons with LES will be shown in Section 537

4.4). 538

4.4. Validating the turbulent friction parametrization and the damped-oscillator model 539

We compare the parametrization of α in Eqs. (17) and (19) directly to the one computed from the 540

LES via Eq. (14) to test the model a priori. In addition, we validate the predictions of our proposed 541

analytical model in Eq. (12) a posteriori in two modes: 1) by optimizing α based on the LES results 542

(in a least square sense) which gives us the best match between the velocity of our analytical model 543

and the LES, and 2) by using the analytical parametrization proposed in Section 4.3, which would 544

be useable without any LES results. We call the first solutions the Optimized MSD system 545

(OMSD), and the second approach the Parametrized MSD (PMSD). Note that the OMSD is only 546

used to check whether the current structure of our MSD model is able to capture the LES results 547

or not, without embedding the influence of the possible errors due to the parametrization of α. On 548

the other hand, in the second approach, we use our analytical model independent of LES data by 549

using the introduced parametrization for α. These PMSD solutions include the error of our MSD 550

model as well as the errors associated with the modelling of αPMSD, hence indicating the benefit of 551

doing a direct validation of the α expression against the one obtained directly from LES αLES. We 552

also compare these values to the ones that result from the optimization, αOMSD. 553

We validated our α parametrization as well as our reduced model for six cases: Step–40, 554

Sin–20, Step–10, Step+100, Step+400, Sin+400, but since the skill of the model was similar in all 555

cases here we only show Step-40 and Step+100 cases. Figure 8(a) displays αOMSD, αPMSD and a 556

Page 34: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

34

range of αLES for the whole 20 hours steady-state simulation and Figure 8(b) exhibits a similar plot 557

after the jump in pressure gradient for Step–40 simulation. The range of the LES is found by 558

obtaining the maximum and minimum values of α at each height throughout the indicated period. 559

Figure 8(d) and 8(e) show these calculated α for the Step+100 case. The validations for the Step–560

10 and Step+25simulations are depicted in the supplementary Appendix S4. The obtained αPMSD 561

and αOMSD are mostly in the LES range, meaning that our developed parametrization αPMSD and the 562

optimized αPMSD, which we determine by matching the velocity variations at each height from our 563

reduced model against LES, agree well with the direct LES calculations from the stress terms. The 564

αOMSD in some cases (e.g. in Figure 8(d)) falls slightly outside of the range near the wall, probably 565

due to the fact the optimization is correcting the structural model errors related to the Boussinesq 566

analogy in that region. This figure also indicates a good agreement between our developed αOMSD 567

and αPMSD, implying that our analytical parametrization matches the best α of our reduced model, 568

which yields lowest prediction errors. 569

We now use these αOMSD and αPMSD to predict the wind variations using our damped-570

oscillator model. We add the particular solution of the step-change forcing (see Momen and Bou-571

Zeid, 2016) to the homogenous solution of Eq. (12) to generate our reduced model predictions as 572

shown in Figure 8(c) and 8(f). The results exhibit three heights, both in the surface and the outer 573

layers. Figure 8(c) and 8(f) display the validation of the damped-oscillator model against the LES 574

in both the optimized and parametrized modes for stable and unstable conditions, respectively. 575

Both α models agree well with LES, though αPMSD is slightly less accurate than the αPMSD in these 576

a-posteriori tests. 577

Page 35: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

35

578

Figure 8. α from LES, PMSD (a-priori analysis), OMSD (a-posteriori analysis) for 2 cases in the stable 579

(a,b) and unstable ABL (d,e). The validation of the reduced model for the stable (c) and unstable (f) ABL. 580

Page 36: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

36

581

Figure 9. M profiles and error analysis for OMSD and PMSD models in which (a) and (c) show the 582

predictions of our reduced model in two modes against LES results after the jump for the stable and unstable 583

ABL respectively, and (b) and (d) exhibit the relative root mean square error of our model predicting the 584

velocity variations at t = 20-30hours for stable and unstable conditions respectively. 585

We now examine the performance of the model in predicting the velocity profiles after the 586

sudden jump in the pressure gradient in one stable and one unstable case. Figure 9(a) and 9(c) 587

show that our models capture the LES velocity profiles quite well despite the fact that the 588

performance of the parametrized model is slightly inferior to the OMSD. This is more obvious in 589

the error analysis that is done in the right panel of Figure 10. The relative errors of our reduced 590

Page 37: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

37

models for the period of t = 20-30 hr, starting right after the jump, are less than 15% for all heights, 591

which confirms the very satisfactory skill of our damped-oscillator models in predicting the mean 592

ABL response to changes in the pressure forcing. The performance of the PMSD model is less or 593

equal to the performance of the OMSD at different heights and it could be improved with better 594

parametrizations of α, which is related to the turbulence closure problem and is a very wide 595

research topic we cannot delve deeply into in this paper. We also varied αOMSD by ±25 % to 596

investigate the sensitivity of our solutions to this coefficient; the results are shown in 597

supplementary Appendix S5. Generally, we found that underestimating α causes higher errors than 598

overestimation because of higher overshoots of velocity that occur with a lower α. 599

To sum up, our reduced damped-oscillator model can also be used in diabatic ABLs to 600

predict the wind-velocity variations. The stability only influences the damper of our model and 601

hence the suggested analytical parametrization agrees well with the LES results. 602

5. Conclusions 603

We studied the effects of an unsteady pressure gradient on diabatic ABLs using a suite of large-604

eddy simulations, and developed an analytical reduced model for the resulting wind-velocity 605

variations. Fourteen simulations were conducted to study the response of unsteady ABLs under 606

various stability conditions. We considered two scenarios: a step change and a sinusoidal variation 607

in the pressure gradient. Then, for each scenario, seven cases were run with different constant 608

surface heat fluxes, ranging from strongly unstable, to neutral, to strongly stable conditions. Our 609

major findings include: 610

The maximum wind speeds in the ABL after increasing the pressure gradient are related to 611

the rate of pre-existing vertical coupling and TKE in the system, and thus to stability. The 612

maximum overshoots decrease when the surface heat flux increases, i.e. under increasingly 613

Page 38: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

38

unstable conditions. The height of the low-level jets in the wind profiles is found to be 614

constant when non-dimensionalized by the height of the ABL, zi (defined as the height at 615

which the TKE reaches ~5% of its surface value in stable ABLs). 616

The temperature profiles in the stable ABL vary significantly when the pressure gradient 617

changes. In particular, right after the sudden jump in the pressure gradient, the surface 618

warms because mixing in the stable ABL increases and hotter air parcels are transported 619

from higher levels to lower elevations. After about one inertial time scale, the stable ABL 620

starts cooling again and the vertical mixing reduces. On the other hand, in the unstable 621

ABL the temperature keeps increasing with almost the same rate after doubling the pressure 622

gradient since the mixing was already high in the convective ABLs. There are little 623

differences in the unstable ABL under a step change versus a sinusoidal variation in the 624

pressure gradients, underlining the dominant role of turbulence in modulating the 625

convective boundary layer. 626

In the stable ABL, the buoyancy destruction responds to the variations of the pressure 627

gradient and increases to destroy the additional produced turbulence. However, in the 628

unstable ABL, the buoyancy term is not significantly influenced by the increase of the 629

shear since the buoyant production is higher than the shear production at all elevations, 630

except perhaps near the wall, and it dominates the dynamics at all times under unsteady 631

conditions. 632

Using LES, we showed that, even in unsteady pressure-gradient ABLs, the resolved 633

transport term does not play a vital role in the stably stratified flows since the buoyant 634

destruction reduces the turbulence at each level, damps vertical velocity, and does not allow 635

for strong transport of the excess TKE produced near the surface to higher elevations, 636

Page 39: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

39

which is another facet of vertical decoupling in such flows. However, the transport 637

becomes more important in convective ABLs where it adapts to the variability of the shear 638

production (due to changes in the pressure gradient) while the buoyant production is almost 639

constant. 640

The production term becomes close to zero or negative in some regions of the stratified 641

ABLs, near the low-level jet location, indicating the existence of the uncommonly observed 642

global backscatter. 643

The findings confirm that turbulence is a dominant mechanism in the unstable ABL that 644

smoothes out other disturbances, including in this case with the effects of pressure-gradient 645

variations. On the other hand, in the stable ABL, the turbulence is weaker and other 646

mechanisms become more consequential for the dynamics (as e.g. suggested by Bou-Zeid 647

et al., 2010). 648

The extension of the damped-oscillator model introduced in Momen and Bou-Zeid (2016) 649

for neutral conditions to diabatic ABLs successfully captures the present LES results. In 650

that model, inertial, Coriolis, and friction forces represented the mass, spring and damper 651

effects respectively. The extension to diabatic conditions only alters the damper coefficient, 652

which characterizes how strong the vertical coupling is in the system. 653

When the friction force (i.e. the gradient of the momentum flux) increases, the damper 654

coefficient α will increase. Hence, it is expected that the damper coefficient results in 655

αunstable > αneutral > αstable away from the wall where the momentum flux divergence is higher 656

in the unstable ABL, while in the surface layer, the reverse trend is observed. This is 657

supported by the direct calculations from the LES. 658

Page 40: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

40

The frictional effects were parametrized using the MOST theory in the surface and a 659

linearly decreasing function via a mixing-length model in the outer layer. The damped-660

oscillator model was run using this parametrization independent of LES, and good 661

agreements with less than 15% relative root mean square error were found compared to the 662

LES results. 663

664

Acknowledgments. The authors acknowledge support from the National Oceanographic and Atmospheric 665

Administration's (NOAA) via the Cooperative Institute for Climate Science (CICS) of Princeton University 666

under grant number 344-6127, from the Grand Challenges Program of the Princeton Environmental 667

Institute, and from the Physical and Dynamic Meteorology Program of the National Science Foundation 668

under AGS-1026636. The simulations were performed on the computing clusters of the National Centre for 669

Atmospheric Research under project number P36861020. 670

671

Appendix A: Boundary and initial conditions and numerical grid configuration of the LES 672

code 673

The boundary conditions for solving Eq. (4) are an imposed kinematic surface buoyancy flux 674

¢w ¢q at the surface and a no flux boundary (dθ/dz=0) at the top of the domain. Periodic lateral 675

boundaries in the horizontal directions are used for all variables. For the momentum equation, a 676

wall shear stress boundary condition is imposed at the surface using a local log-law formulation 677

(Bou-Zeid et al., 2005), while the top boundary condition for momentum is zero stress. Both top 678

and bottom boundaries are impermeable and thus the vertical velocity, w, is set to zero there. 679

For the stable simulations, a sponge layer with a Rayleigh damping method (Israeli and 680

Orszag 1981) is used as the upper boundary condition to damp all the vertical perturbations and 681

Page 41: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

41

avoid reflections of waves from the top of the domain. In these cases, the domain height is 682

Lz = 1600 m with an imposed inversion base height of zi = 1200 m, which is also the height where 683

the sponge layer effect starts. This will fix the maximum attainable height of the stable ABL. While 684

this might appear too conservative for stable conditions, the need for such a high domain is 685

illustrated later in Section 3 (due to gradually varying and/or growing boundary layers in some 686

cases). In the neutral conditions, since we do not have a real inversion we set the inversion height 687

as zi = Lz = 1000 m in the code. Furthermore, the code does not solve the scalar Eq. (5) for neutral 688

simulations and neglects the buoyancy terms in the momentum equation. For the unstable 689

simulations, we set Lz = 1700 m and the inversion height to zi = 1250 m. The differences in zi will 690

not be important since all the results and equations are normalized with that height. 691

The simulations are integrated for 40 hours of physical time, which are equivalent to about 692

30 large eddy turnover times (ABL height / friction velocity), or more. For these long simulations, 693

if we do not control the inversion layer under unstable conditions it will be warmed up by the 694

surface heat flux and will be dissipated after some time. To avoid continuous growth of the ABL 695

height in the convective ABL and maintain a steady zi, we keep the θ profile in the capping 696

inversion layer steady by compensation for any heating and cooling by entrainment. Although this 697

setup might not reflect the real top boundary conditions in the ABL, similar approaches have been 698

previously adopted to ensure the statistical convergence of the result (Sescu and Meneveau 2014; 699

Shah and Bou-Zeid 2014) and key profiles have been shown to remain very similar to the real-700

world ABL. In this study, we conducted further tests that confirm that the top boundary conditions 701

have a minor influence on the flow dynamics inside the ABL with unsteady pressure gradient and 702

steady surface heat flux. 703

Page 42: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

42

In all the simulation we set Lx = Ly = 2πLz, the number of the numerical grid points in the 704

stable ABL is Nx = Ny = Nz = 160 (which is sufficient even for strong stabilities as illustrated in 705

Huang and Bou-Zeid 2013), and the number of the points in the unstable and neutral ABLs are 706

Nx = Ny = Nz = 128. 707

For each surface heat flux, a 50-hour, 64×64×64 nodes, steady simulation is performed to 708

start warm-up. Then, the outputs are linearly interpolated into a higher resolution domain (1283 or 709

1603) and used as initial conditions for the corresponding cases. More details on the LES model, 710

the used SGS model, and the unsteady simulation setup could be found respectively in Albertson 711

and Parlange (1999), Bou-Zeid et al. (2005), and Momen and Bou-zeid (2017b). 712

Supporting information 713

The manuscript includes two supporting videos and four supplementary appendices. 714

References: 715

Albertson JD, Parlange MB (1999) Natural integration of scalar fluxes from complex terrain. Adv Water Resour 716

23:239–252. doi: 10.1016/S0309-1708(99)00011-1 717

Bain CL, Parker DJ, Taylor CM, et al (2010) Observations of the Nocturnal Boundary Layer Associated with the West 718

African Monsoon. Mon Weather Rev 138:3142–3156. doi: 10.1175/2010MWR3287.1 719

Baker WE, Bloom SC, Woolen JS, et al (1987) Experiments with a Three-Dimensional Statistical Objective Anlysis 720

Scheme Using FGGE Data. Mon Weather Rev 115:272–296. 721

Banta RM, Pichugina YL, Brewer WA (2006) Turbulent Velocity-Variance Profiles in the Stable Boundary Layer 722

Generated by a Nocturnal Low-Level Jet. J Atmos Sci 63:2700–2719. 723

Blackadar A (1957) Boundary layer wind maxima and their significance for the growth of nocturnal inversions. Bull 724

Am Meteorol Soc 38:283–290. 725

Bonner W, Paegle J (1970) Diurnal Variations in Boundary Layer Winds Over the South-Central United States In 726

Summer. Mon Weather Rev 98:735–744. 727

Page 43: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

43

Bou-Zeid E, Higgins C, Huwald H, et al (2010) Field study of the dynamics and modelling of subgrid-scale turbulence 728

in a stable atmospheric surface layer over a glacier. 665:480–515. doi: 10.1017/S0022112010004015 729

Bou-Zeid E, Meneveau C, Parlange M (2005) A scale-dependent Lagrangian dynamic model for large eddy simulation 730

of complex turbulent flows. Phys Fluids 17:25105. doi: 10.1063/1.1839152 731

Bou-Zeid E, Overney J, Rogers BD, Parlange MB (2009) The effects of building representation and clustering in 732

large-eddy simulations of flows in urban canopies. Boundary-Layer Meteorol 132:415–436. doi: 733

10.1007/s10546-009-9410-6 734

Byun DW, Arya SPS (1986) A study of mixed-layer momentum evolution. Atmos Environ 20:715–728. doi: 735

10.1016/0004-6981(86)90186-1 736

Delage Y (1974) A numerical study of the nocturnal atmospheric boundary layer. Q J R Meteorol Soc 100:351–364. 737

doi: 10.1002/qj.49710042507 738

Du Y, Rotunno R (2014) A Simple Analytical Model of the Nocturnal Low-Level Jet over the Great Plains of the 739

United States. J Atmos Sci 71:3674–3683. doi: 10.1175/JAS-D-14-0060.1 740

Ekman VW (1905) On the Influence of the Earth’s Rotation on Ocean-Currents. Ark Mat Astron Fys 2:1–52. 741

Estournel C, Guedalia D (1987) A new parameterization of eddy diffusivities for nocturnal boundary-layer modeling. 742

Boundary-Layer Meteorol 39:191–203. doi: 10.1007/BF00121874 743

Gayen B, Sarkar S, Taylor JR (2010) Large eddy simulation of a stratified boundary layer under an oscillatory current. 744

J Fluid Mech 643:233. doi: 10.1017/S002211200999200X 745

Grisogono B (1995) A generalized Ekman layer profile with gradually varying eddy diffusivities. Q J R Meteorol Soc 746

121:445–453. doi: 551.5 11.61:551.554 747

Haurwitz B (1947) Comments on the sea-breeze circulation. J Meteorol. doi: 10.1175/1520-0469(1947)004,0001: 748

COTSBC.2.0.CO;2 749

Hsu BC, Lu X, Kwan M (2000) LES and RANS studies of oscillating flows over flat plate. J Eng Mech 126:186–193. 750

Huang J, Bou-Zeid E (2013) Turbulence and vertical fluxes in the stable atmospheric boundary layer. Part I: A large-751

eddy simulation study. J Atmos Sci 70:1513–1527. doi: 10.1175/JAS-D-12-0167.1 752

Huang J, Bou-Zeid E, Golaz J-C (2013) Turbulence and Vertical Fluxes in the Stable Atmospheric Boundary Layer. 753

Part I: A Large-Eddy Simulation Study. J Atmos Sci 70:1513–1527. doi: 10.1175/JAS-D-12-0167.1 754

Israeli M, Orszag SA (1981) Approximation of radiation boundary conditions. J Comput Phys 41:115–135. doi: 755

Page 44: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

44

10.1016/0021-9991(81)90082-6 756

Lacser A, Arya SPS (1986) A comparative assessment of mixing-length parameterizations in the stably stratified 757

nocturnal boundary layer (NBL). Boundary-Layer Meteorol 36:53–70. doi: 10.1007/BF00117458 758

Lenschow DH, Wyngaard JC, Pennell WT (1980) Mean-Field and Second-Moment Budgets in a Baroclinic, 759

Convective Boundary Layer. J. Atmos. Sci. 37:1313–1326. 760

Lewis DM, Belcher SE (2004) Time-dependent, coupled, Ekman boundary layer solutions incorporating Stokes drift. 761

Dyn Atmos Ocean 37:313–351. doi: 10.1016/j.dynatmoce.2003.11.001 762

Lohmann IP, Fredsøe J, Sumer BM, Christensen ED (2006) Large Eddy Simulation of the ventilated wave boundary 763

layer. J Geophys Res 111:C06036. doi: 10.1029/2005JC002946 764

Madsen OS (1977) A Realistic Model the Wind-Iduced Ekman Boundary Layer. J Phys Oceanogr 7:248–255. 765

Mahrt L (2013) Stably Stratified Atmospheric Boundary Layers. Annu Rev Fluid Mech 46:23–45. doi: 766

10.1146/annurev-fluid-010313-141354 767

Mahrt LJ, Schwerdtfeger W (1970) Ekman spirals for exponential thermal wind. Boundary-Layer Meteorol 1:137–768

145. doi: 10.1007/BF00185735 769

McNider RT (1982) A note on velocity fluctuations in drainage flows. J. Atmos. Sci. 39:1658–1660. 770

Miles J (1994) Analytical solutions for the Ekman layer. Boundary-Layer Meteorol 67:1–10. doi: 771

10.1007/BF00705505 772

Momen M, Bou-Zeid E (2017a) Analytical reduced models for non-stationary diabatic atmospheric boundary layers. 773

Bound Layer Meteorol In review. 774

Momen M, Bou-Zeid E (2017b) Mean and turbulence dynamics in unsteady Ekman boundary layers. J Fluid Mech 775

816:209–242. 776

Momen M, Bou-Zeid E (2016) Large Eddy Simulations and Damped-Oscillator Models of the Unsteady Ekman 777

Boundary Layer. J Atmos Sci 73:25–40. doi: 10.1175/JAS-D-15-0038.1 778

Monin AS, Obukhov AM (1954) Basic laws of turbulent mixing in the surface layer of the atmosphere. Contrib 779

Geophys Inst Acad Sci USSR 24:163–187. 780

Orszag SA, Pao YH (1974) Numerical computation of turbulent shear flows. Adv Geophys 18:225–236. 781

Paulson C a. (1970) The Mathematical Representation of Wind Speed and Temperature Profiles in the Unstable 782

Atmospheric Surface Layer. J. Appl. Meteorol. 9:857–861. 783

Page 45: Momen&Bou-Zeid(2017) QJRMS Submitted - Princeton...$%/ wr dq xqvwhdg\ suhvvxuh judglhqw zdv qrw frqvlghuhg ghvslwh ehlqj wkh dqdorjxh ri pdq\ uhdo zruog frqglwlrqv ,q wklv sdshu zh

45

Polavarapu SM (1995) Divergent wind analyses in the oceanic boundary layer. Tellus 47A:221–239. 784

Pu B, Dickinson RE (2014) Diurnal spatial variability of Great Plains summer precipitation related to the dynamics 785

of the low-level jet. J Atmos Sci 140113153922004. doi: 10.1175/JAS-D-13-0243.1 786

Radhakrishnan S, Piomelli U (2008) Large-eddy simulation of oscillating boundary layers: Model comparison and 787

validation. J Geophys Res Ocean 113:1–14. doi: 10.1029/2007JC004518 788

Schmidt F. (1947) An elementary theory of the land- and sea-breeze circulation. J Meteorol 4:9–20. doi: 10.1175/ 789

1520-0469(1947)004,0009:AETOTL.2.0.CO;2 790

Schröter JS, Moene AF, Holtslag AAM (2013) Convective boundary layer wind dynamics and inertial oscillations: 791

The influence of surface stress. Q J R Meteorol Soc 139:1694–1711. doi: 10.1002/qj.2069 792

Sescu A, Meneveau C (2014) A control algorithm for statistically stationary large-eddy simulations of thermally 793

stratified boundary layers. Q J R Meteorol Soc 140:2017–2022. doi: 10.1002/qj.2266 794

Shah S, Bou-Zeid E (2014) Very-Large-Scale Motions in the Atmospheric Boundary Layer Educed by Snapshot 795

Proper Orthogonal Decomposition. Boundary-Layer Meteorol 153:355–387. doi: 10.1007/s10546-014-9950-2 796

Shapiro A, Fedorovich E (2010) Analytical description of a nocturnal low-level jet. Q J R Meteorol Soc n/a-n/a. doi: 797

10.1002/qj.628 798

Stull R (1988) An Introduction to Boundary Layer Meteorology, 1st edn. Kluwer Academic Publishers, Dordrecht 799

Tan ZM (2001) An approximate analytical solution for the baroclinic and variable eddy diffusivity semi-geostrophic 800

Ekman boundary layer. Boundary-Layer Meteorol 98:361–385. doi: 10.1023/A:1018708726112 801

Tennekes H, Lumley J (1972) A First Course in Turbulence. MIT, Boston 802

Thorpe AJ, Guymer TH (1977) The nocturnal jet. Q J R Meteorol Soc 103:633–653. doi: 10.1002/qj.49710343809 803

Van de Wiel BJH, Moene a. F, Steeneveld GJ, et al (2010) A Conceptual View on Inertial Oscillations and Nocturnal 804

Low-Level Jets. J Atmos Sci 67:2679–2689. doi: 10.1175/2010JAS3289.1 805

Yordanov D, Syrakov D, Djolov G (1983) A barotropic planetary boundary layer. Boundary-Layer Meteorol 25:363–806

373. doi: 10.1007/BF02041155 807

808