Вы находитесь на странице: 1из 199

Large-eddy Simulation Study of Turbulent Flows over

Idealized Urban Roughness within a Convective Boundary


Layer

By
Chan Ming Chung
BEng. (HKU)

A thesis submitted in partial fulfillment of the requirement for


the degree of Doctor of Philosophy at
The University of Hong Kong
May 2016
Abstract of thesis entitled

Large-eddy Simulation Study of Turbulent Flows over Idealized


Urban Roughness within a Convective Boundary Layer
Submitted by
Chan Ming Chung

for the degree of Doctor of Philosophy


at The University of Hong Kong
in May 2016

Convective boundary layer over urban area is rarely studied in the literature due to the technical
difficulty of field observation and laboratory experiments. Whereas, the importance of its
understanding to many applications is undeniable. Therefore, large-eddy simulation (LES) is
employed to study the turbulent flow over idealized urban roughness under a range of
convective conditions. Unlike many studies of buoyant instability with urban structures, this
LES study simulates explicitly both the motions at roughness and boundary-layer scales,
facilitating the analysis on the multi-scale processes that are important under convective
condition. The near-wall region is divided into inertial sublayer (ISL) and roughness sublayer
(RSL). This study focuses on the Monin-Obukhov similarity theory (MOST), which states that
the flow in ISL is scaled by local height rather than boundary layer depth, and the flow
properties in RSL under convective condition. It is found that MOST describes the LES results
approximately well, which means the applicability of MOST over idealized urban roughness is
basically supported by this study. However, unsatisfactory results at slightly unstable state and
near ISL upper limit, and the sensitivity test of the MOST statistics on the domain horizontal
dimensions, suggests that boundary layer eddies can modulate the ISL flow statistics. Therefore,
MOST is not strictly valid and this explains why most MOST data of field studies are usually
unsatisfactory. LES result shows that the flow statistics in RSL deviates from those predicted
by MOST, but they are found to be different from those of vegetative roughness. It is found
that mesh effect is important in the vicinity of roughness, but it does not influence the flow in
ISL and above since the turbulent processes are ‘local’ under neutral condition and ‘top-down’

1
under convective condition. The horizontal inhomogeneity of flow statistics in RSL are
enhanced by buoyant instability, especially the flux statistics. Functional forms of mean
velocity and temperature profiles are derived analytically from the proposed MOST functions
and compared well with LES results. The calculated aerodynamic and temperature roughness
lengths show an increasing trend with instability at strong convective condition, which is found
to be due to the enhanced turbulence at roughness scale. Therefore, the convectional view that
roughness lengths are only functions of roughness morphology is not valid. The turbulence
kinetic energy (TKE) budget analysis shows that the local buoyant production is much weaker
than the shear production and the transport terms are enhanced in convective condition. The
wake production term explicitly calculated to be negative by LES, rather than positive
documented in some field studies. The convective coherent structures are found to be multi-
scaled and formed collectively near surface and the two-point correlation of vertical velocity
fluctuation is non-local. Similar to vegetative roughness, the sweep-ejection pair eddy is also
found above urban roughness. The results imply the enhanced turbulence near wall in
convective condition is governed by a ‘top-down’ mechanism rather than a ‘local’ one, which
invalidates the Townsend’s hypothesis, while momentum and heat are actually transported by
different organized motions.

Word count: 494

2
Declaration

I declare that the thesis entitled “Large-eddy Simulation Study of Turbulent Flows over
Idealized Urban Roughness within a Convective Boundary Layer” represents my own work,
except where due acknowledgement is made, and that it has not been previously included in a
thesis, dissertation or report submitted to this University or to any other institution for a degree,
diploma or other qualifications.

Signature:
Chan Ming Chung

i
Acknowledgements

I would like to express my deepest gratitude to my PhD supervisor, Dr. Chun-Ho Liu of the
Department of Mechanical Engineering, The University of Hong Kong, for his kind support,
advice and encouragement throughout my study. I would like to express my sincere gratitude
to my colleagues Colman Wong, Wai-chi Cheng, Pieta Leung, Yat-Kiu Ho, Chi-To Ng and
Pluto Chui as well as my dearest friends Yik-Sau Tang and Anthony Yuen. Their suggestions
and feedbacks are valuable to my study. Also thanks are expressed to W.K. Kwan, Lilian Y.L.
Chan, Vincent K.W. Lo and W.K. Yeung for their technical supports. I also wish to express my
special gratitude to Prof. Ricchard K.K. Yuen for his support and advice during my last year of
PhD study. Last but not least, I would like to thank my mother, my father and my sisters for all
their love and care during this journey.

Chan Ming Chung


May 2016

ii
Contents

Declaration i

Acknowledgements ii

Contents iii

List of Figures vii

List of Tables x

Nomenclature xi

1 Introduction 1
1.1 Overview 1
1.1.1. Background of scientific field 1
1.1.2. Some important theories, concepts and potential problems 3
1.1.3. Research trend towards understanding flow mechanisms 5
1.1.4. Research methods 6
1.2 Research Objectives and Thesis Outline 9

2 Literature Review 11
2.1 Monin-Obukhov Similarity Theory (MOST) 11
2.1.1. MOST functions of mean gradients (𝜙𝑚 and 𝜙ℎ ) 15
2.1.2. MOST functions of second moments of fluctuation (𝜙𝑤 and 𝜙𝜃 ) 17
2.2 Layer Decomposition of Convective Urban Boundary Layer (CUBL) 18
2.2.1. Properties of inertial sublayer (ISL) 19
2.2.2. Properties of roughness sublayer (RSL) 21

iii
2.2.3. Properties of mixed layer (ML) 23
2.3 Mean Profiles in ISL and Roughness Parameterization 25
2.3.1. Derivation of mean profiles 26
2.3.2. Parameterization of roughness properties on mean wind 27
2.3.3. Parameterization of roughness properties on mean temperature 31
2.4 Interaction of Near Wall Region and Outer Layer 32
2.4.1. Townsend’s Hypothesis 32
2.4.2. Evidences on violation of MOST and Townsend’s hypothesis 34
2.5 Mechanism of Turbulent Processes 35
2.5.1. Turbulence kinetic energy (TKE) budget 36
2.5.2. Coherent structures 37

3 Numerical Model Configuration 41


3.1 Computational Domain 41
3.2 Fundamental Equations and Boundary Conditions 43
3.2.1. Basic transport equations 44
3.2.2. Boussinesq approximation 45
3.2.3. Boundary and initial conditions 47
3.2.4. Quasi-steady state and wall units 48
3.2.5. Non-dimensionalization 49
3.3 Large-eddy Simulation (LES) Modeling 50
3.4 Discretization and Numerical Methods 52
3.4.1. Domain meshing 53
3.4.2. Numerical methods 55
3.5 Processing and Post-processing of Numerical Models 56
3.5.1. Processing of numerical models 56
3.5.2. Post-processing of field data 57
3.6 Summary of numerical models 58
3.7 Model Validation by Wind Tunnel Experiment 59

4 Examination of Monin-Obukhov Similarity Theory (MOST) by Large-eddy 64


Simulation (LES)
4.1 MOST Statistics in ISL 64

iv
4.1.1. Mean streamwise velocity gradient 𝜙𝑚 66
4.1.2. Mean temperature gradient 𝜙ℎ 68
4.1.3. Second moment of vertical velocity fluctuation 𝜙𝑤 70
4.1.4. Second moment of temperature fluctuation 𝜙𝜃 72
4.2 Mixing Length in ISL Derived from MOST 74
4.3 Sensitivity Study on the Effect of Horizontal Domain Dimensions on MOST 76
Quantities at Local Free Convection State
4.4 Concluding Remarks 81

5 Properties of Roughness Sublayer (RSL) at Different Instability Conditions 83


and Roughness Configurations
5.1 Deviation from MOST Statistics in RSL and the Effect of Mesh Resolution at 83
Roughness Aspect Ratio 1
5.1.1. Neutral stratification 85
5.1.2. Unstable stratification 88
5.2 Deviation from MOST Functions in RSL at Different Instability Levels and 90
Roughness Aspect Ratios
5.2.1. Mean velocity gradient 𝜙𝑚 91
5.2.2. Second moment of vertical velocity fluctuation 𝜙𝑤 92
5.3 Horizontal Inhomogeneity of Flow Statistics in Near Wall Region at Different 94
Instability Levels and Roughness Aspect Ratios
5.3.1. Streamwise variation of vertical profiles of mean streamwise velocity 95
𝑢̅ and Reynolds stress ̅̅̅̅̅̅
𝑢′𝑤′
5.3.2. Horizontal variability of single-point ensemble statistics and the 101
corresponding RSL height 𝑧∗
5.4 Concluding Remarks 103

6 Parameterization of Surface Roughness on Inertial Sublayer (ISL) Flow 105


under Unstable Condition
6.1 Analytical Derivation of Mean Profiles Based on the Proposed MOST 105
Functions
6.2 Mean Streamwise Velocity and Temperature Profiles 107
6.3 Aerodynamic and Temperature Roughness Lengths 110

v
6.4 Relationships of Roughness Lengths with Buoyancy Induced Turbulence at 104
Roughness Scale
6.5 Concluding Remarks 118

7 Mechanism of Turbulent Transports at Near-Wall Region 120


7.1 Turbulent Kinetic Energy (TKE) Budget at Roughness Scale 120
7.2 Horizontally Averaged Turbulent Kinetic Energy (TKE) Budget above 127
Building Roof-level
7.3 Visualization of Instantaneous Turbulent Fields 131
7.4 Comparison of Momentum and Heat Transports 145
7.5 Two-point Correlation and Integral Length Scale 149
7.6 Conditional Average 154
7.7 Concluding Remarks 160

8 Conclusion 162

A Appendix 165
A.1 Steps of Mathematical Integration of the Proposed MOST functions 165
A.2 Derivation of Turbulent Kinetic (TKE) Budget Equation 166
A.3 Derivation of Horizontally Averaged Turbulent Kinetic (TKE) Budget 167
Equation

Reference 169

vi
List of Figures

1.1 Flow chart of this thesis with objectives 10


3.1 Computational domain of the standard case at roughness aspect ratio 1 42
3.2 A periodic roughness unit at roughness aspect ratio 1 43
3.3 Boundary conditions for velocity and temperature 48
3.4 Mesh inside canyon cavity of model A1 53
3.5 Wind tunnel setting by Uehara et al. 60
3.6 Normalized vertical profiles (comparison between LES and wind tunnel 61
results)
4.1 Normalized mean streamwise velocity gradient against local stability level 67
4.2 Normalized mean temperature gradient against local stability level 69
4.3 Normalized second moment of vertical velocity fluctuation against local 71
stability level
4.4 Normalized second moment of temperature fluctuation 𝜙𝜃 against local 73
stability level
4.5 Mixing length against the effective height 75
4.6 Deviation of mixing lengths between LES results and MOST functions (x- 76
axis) against the effective height
4.7 Second moment versus the dimensionless height (variable instability levels) 77
4.8 Second moment versus the dimensionless height (variable domain extents) 79
4.9 MOST statistics (variable domain extents) 80
5.1 Vertical profiles or MOST statistics at near-wall region 84
5.2 Ratio of LES calculated dimensionless mean velocity gradient and MOST 91
function at various instability levels
5.3 Ratio of LES calculated dimensionless second moment of vertical velocity 93
fluctuation and MOST function at various instability levels
5.4 Profiles of ensemble mean streamwise velocity at different streamwise (x-) 96
positions

vii
5.5 Profiles of ensemble averaged Reynolds stress at different streamwise (x-) 97
positions
5.6 2D contours of Reynolds stress and momentum flux by mean advection 98
5.7 Vertical profiles of horizontal variability 99
5.8 Ratio of LES calculated dimensionless second moment of vertical velocity 102
fluctuation and MOST function at various instability levels
6.1 Profiles of normalized mean streamwise velocity and mean temperature 108
6.2 Profiles of normalized mean streamwise velocity corrected by the stability 109
correction function
6.3 Aerodynamic roughness length and temperature roughness length against 111
instability
6.4 Comparison between roughness lengths and streamwise profiles of standard 115
deviation of vertical velocity
6.5 Contours of instantaneous vertical velocity fluctuation on x-z plane 117
7.1 2D contours of TKE budget terms of model A1 at roughness scale 122
7.2 2D contours of TKE budget terms of model A0.5 at roughness scale 125
7.3 Vertical profiles of horizontally averaged TKE budget terms for model A1 128
above roughness roof-level
7.4 Vertical profiles of horizontally averaged TKE budget terms for model A0.5 129
above roughness roof-level
7.5 Instantaneous fluctuation flow fields in x-z plane at neutral condition for 132
model A1
7.6 Instantaneous fluctuation flow fields in x-z plane at unstable condition for 133
model A1
7.7 Instantaneous vertical velocity fluctuation fields in y-z plane at neutral 134
condition for model A1 at different heights
7.8 Instantaneous vertical velocity fluctuation fields in y-z plane at unstable 135
condition for model A1 at different heights
7.9 Instantaneous fields of fluxes in x-z plane at neutral condition for model A1 138
7.10 Instantaneous fields of fluxes in x-z plane at unstable condition for model 139
A1
7.11 Streamwise filtered instantaneous fields in y-z plane at neutral condition for 141
model A1 at roof-level

viii
7.12 Streamwise filtered instantaneous fields in y-z plane at neutral condition for 142
model A1 at mid-ISL
7.13 Streamwise filtered instantaneous fields in y-z plane at unstable condition 143
for model A1 at roof-level
7.14 Streamwise filtered instantaneous fields in y-z plane at unstable condition 144
for model A1 at mid-ISL
7.15 Vertical profiles of flux quadrant ratios above roughness roof-level of model 147
A1 at different instability levels
7.16 Vertical profiles of horizontally averaged correlation coefficient between 148
momentum flux and heat flux above roughness roof-level for model A1 at
different instability levels
7.17 Two-point correlations in x-z plane at neutral condition of reference position 150
at mid-ISL
7.18 Two-point correlations in x-z plane at unstable condition of reference 151
position at mid-ISL
7.19 Two-point correlations of vertical velocity fluctuation in x-z plane at 152
different reference positions
7.20 Vertical profiles of integral length scales of vertical velocity fluctuation 153
7.21 Conditional averages of flow velocity vector and temperature on x-z plane 155
given local pressure maxima
7.22 Conditional averages of flow velocity vector and temperature on y-z plane 157
under neutral condition given local pressure maxima at roof-level
7.23 Conditional averages of flow velocity vector and temperature on y-z plane 158
under unstable condition given local pressure maxima at roof-level
7.24 Conditional averages of flow velocity vector and temperature on x-z plane 159
under unstable condition given local maximum and local minimum of
vertical velocity at roof-level

ix
List of Tables

2.1 Commonly used functional forms of dimensionless mean gradients 16


2.2 Decomposition of urban boundary layer and characteristics of each region 19
3.1 Mesh configurations for single periodic roughness units in Coarse (C) mesh 54
3.2 Mesh element number for the standard cases where 𝐿𝑥 = 24ℎ in Coarse (C) 54
mesh
3.3 Mesh element sizes in Coarse (C) mesh 55
3.4 Mesh element sizes in wall unit in Coarse (C) mesh 55
3.5 Model configuration list 59
6.1 Values of numerical constants in Equation (6.1) 100

x
Nomenclature

Abbreviations

A Aspect ratio of LES model


ABL atmospheric boundary layer
BD Businger-Dyer
C Coarse mesh (LES)
CBL Convective boundary layer
CFD computational fluid dynamics
CUBL Convective urban boundary layer
DNS direct numerical simulation
F Fine mesh (LES)
FC Local free-convection
ISL inertial sublayer
LES Large-eddy simulation
MOST Monin-Obukhov similarity theory
ML Mixed-layer
RANS Reynolds-averaged-Navier-Stoke
RSL roughness sublayer
SGS subgrid scale
SL Surface layer
TKE turbulence kinetic energy
UBL urban boundary layer
VC Very Coarse mesh (LES)
X Domain streamwise length of LES model
Y Domain spanwise length of LES model

xi
Roman Symbols

𝑎 Empirical constant of Brutsaert relation


𝑎𝑖 Empirical constant in linaer MOST functions
̅
𝜕𝑘
𝐴 Mean advection term in TKE budget (−𝑢̅𝑗 𝜕𝑥 )
𝑗

𝐴𝑓 Total building frontal area


𝐴𝑝 Total building plan area
𝐴𝑇 Total lot area
𝑏 Roughness cavity width
𝑏𝑖 Empirical constant of Businger-Dyer relations
𝐵 ̅̅̅̅̅̅)
Buoyant production term in TKE budget (𝛼𝑔𝑤′𝜃′
𝑐𝑝 Heat capacity of air
𝐶𝑖 Empirical constant of local free-convective power funtions
𝐶𝑘 SGS model constant (0.07)
𝐶𝑚 Drag coefficient
𝐶𝑇 Temperature transfer coefficient
𝐶𝜀 SGS model constant (1.05)
𝑑 Roughness displacement height
𝐷𝑇 Thermal diffusivity
𝑓𝑖 A single-point statistic
𝑔 Gravitational acceleration
𝐺 Filter function in LES
ℎ Reference length unit, height of building for model A1
ℎ𝑟 Building height
ℎ𝑢𝜏
ℎ+ Roughness height in wall unit ( 𝜈
)

ℎ̂ Specific enthalpy
𝐻 Height of the free-stream region (7ℎ)
𝑘 Turbulent kinetic energy
𝑘𝑠𝑔𝑠 SGS turbulent kinetic energy
𝑙𝑚 Prandtl’s mixing length
𝑙𝜈 Viscous length (𝜈/𝑢𝜏 )

xii
𝐿 Obukhov length
𝐿𝑥 Domain streamwise length
𝐿𝑦 Domain spanwise lengths
𝐿𝑧 Overall domain height (8ℎ)
𝐿𝑤𝑤 Integral length scale of vertical velocity fluctuation
𝑛𝑖 Power exponents of local free-convective power funtions
𝑝 Kinematic pressure
𝑃 Thermodynamic pressure
′ ′ 𝜕𝑢̅̅̅𝑖
𝑃𝑠 ̅̅̅̅̅̅
Shear production term in TKE budget (−𝑢 𝑖 𝑢𝑗 𝜕𝑥 )
𝑗

̅̅̅"
𝜕𝑢
𝑃𝑤 ̅̅̅̅̅̅
Wake production in horizontal averaged TKE budget (− 〈𝑢 ′ ′ 𝑖
𝑖 𝑢𝑗 " 𝜕𝑥 〉)
𝑗

𝑃0 Reference pressure in Boussinesq approximation


𝑞𝑗 SGS heat flux ((𝜃𝑢𝑗 )𝑟 − 𝜃 𝑟 𝑢𝑖𝑟 )
𝑞0 Surface heat flux
𝑟𝜙𝜙 Two-point correlation of field variable 𝜙 in x-z plane
𝑅 Residual term in TKE budget
𝑅𝑢′ 𝑤′ ,𝜃′ 𝑤′ Correlation coefficient between momentum and heat fluxes
𝑅0 Specific gas constant
𝑅′ Residual term in horizontal averaged TKE budget
𝑡 Time
𝑇 Thermodynamic temperature
𝜕
𝑇𝑑 Dispersive transport in horizontal averaged TKE budget (− 𝜕𝑥 〈𝑘̅"𝑢̅𝑗 "〉)
𝑗

𝜕 ̅̅̅̅̅𝑖′ ))
𝑇𝑝 Viscous diffusion term in TKE budget (− (𝑝′𝑢
𝜕𝑥𝑖
𝜕
𝑇𝑡 ̅̅̅̅̅𝑗′ ))
Turbulent transport term in TKE budget (− 𝜕𝑥 (𝑘𝑢
𝑗

𝜕2 𝑘
̅
𝑇𝑣 Viscous diffusion term in TKE budget (𝜈 𝜕𝑥 )
𝑗 𝜕𝑥𝑗

𝑇0 Ambient temperature in absolute scale


𝑢 Streamwise velocity (𝑢1 )
𝑢𝑖 Velocity component
𝑢𝑖 + Velcoity wall unit (𝑢𝑖 /𝑢𝜏 )
𝑢𝑓 Local free-convective velocity

xiii
𝑢𝜏 Friction velocity
𝑈 Mean streamwise velocity
𝑈0 , Initial streamwise velocity in LES
𝑣 Spanwise velocity (𝑢2 )
𝑉 Whole domain volume
𝑤 Vertical velocity (𝑢3 )
𝑤∗ Deardorff’s convective velocity
𝑥 Streamwise coordinate
𝑥𝑖 Spatial coordinate
𝑥𝑟 Reference x-coordinate in two-point correlation
𝑥𝑖 + Spatial wall unit (𝑥𝑖 /𝑙𝜈 )
𝑦 Spanwise coordinate
𝑧 Local height, measured from ground level
𝑧𝑒𝑓𝑓 Effective height (𝑧 − 𝑑)
𝑧𝑖 Boundary layer depth (𝐿𝑧 − 𝑑 in LES)
𝑧ℎ Temperature roughness length
𝑧𝑟 Reference z-coordinate in two-point correlation
𝑧𝑟𝑒𝑓 Reference height in Boussinesq approximation
𝑧0 Aerodynamic roughness length
𝑧∗ Roughness sublayer (RSL) height
𝑧𝑢𝜏
𝑧+ Local height in wall unit ( 𝜈
)

Greek Symbols

𝛼 Thermal expansion coefficient


𝛼𝑃 Isothermal compressibility
𝛽 Buoyancy parameter
𝛾𝑖 Empirical constant of MOST functions by Wilson
𝛿𝑖𝑗 Kronecker delta
Δ Geomtric mean filter and mesh width in LES ((Δ1 Δ2 Δ3 )1/3 )
Δ𝑖 Filter width in LES

xiv
Δ𝑡 Incremental time step in LES calculation
∆𝑃 Overall pressure variation throughout domain
∆𝑇 Overall temperature variation throughout domain
Δ𝜃 Potential difference between domain top and wall in LES
̅̅̅̅̅̅̅̅
𝜕𝑢 ′ 𝜕𝑢′
𝜀 Viscous dissipation term in TKE budget (𝜈 𝜕𝑥 𝑖 𝜕𝑥 𝑖 )
𝑗 𝑗

𝜀1 Dimensionless number for Boussinesq approximation (𝛼∆𝑇)


𝜀2 Dimensionless number for Boussinesq approximation (𝛼𝑃 ∆𝑃)
𝜁 Local stability parameter (𝑧/𝐿 or 𝑧𝑒𝑓𝑓 /𝐿)
𝜁0 Stability parameter based on roughness length (𝑧0 /𝐿)
𝜁ℎ Stability parameter based on roughness length (𝑧ℎ /𝐿)
𝜃 Potential temperature
𝜃𝑓 local free-convective temperature
𝜃𝑤 Wall surface temperature
𝜃𝜏 Friction temperature
𝜃0 Reference temperature
𝜃∗ Deardorff’s convective temperature
𝜃+ Temperature wall unit ((𝜃 − 𝜃𝑊 )/𝜃𝜏 )
Θ Mean temperature
𝜅 von Karman constant
𝜆 Binary constant (0 or 1) for Π term in momentum equation
𝜆𝑓 Dimensionless frontal area (𝐴𝑓 /𝐴𝑇 )
𝜆𝑝 Dimensionless plan area (𝐴𝑝 /𝐴𝑇 )
Π Background pressure gradient
𝜌 Air density
𝜌0 Reference air density in Boussinesq approximation
𝜎𝑢 Standard deviation of streamwise velocity fluctuation (√̅̅̅̅̅
𝑢′𝑢′)
𝜎𝑤 ̅̅̅̅̅̅)
Standard deviation of vertical velocity fluctuation (√𝑤′𝑤′
𝜎𝜃 Standard deviation of temperature fluctuation (√̅̅̅̅̅
𝜃′𝜃′)
𝜏𝑖𝑗 SGS stress ((𝑢𝑖 𝑢𝑗 )𝑟 − 𝑢𝑖𝑟 𝑢𝑖𝑟 )
𝑑
𝜏𝑖𝑗 Deviatoric stress
𝜏0 Surface shear stress

xv
𝜈 Kinematic viscosity
𝜈𝑠𝑔𝑠 SGS turbulent viscosity
𝜙ℎ MOST normalized mean temperature gradient
𝜙𝑚 MOST normalized mean streamwise velocity gradient
𝜙𝑤 MOST normalized vertical velocity variance
𝜙𝜃 MOST normalized temperature variance
Ψ𝑖 Stability correction function in mean profile

Subscripts

ℎ Mean temperature gradient (MOST)


𝑚 Mean velocity gradient (MOST)
𝑞1 1st quadrant (outward interaction for 𝑢′𝑤′ and hot updraft for 𝜃′𝑤′)
𝑞2 2nd quadrant (ejection for 𝑢′𝑤′ and outward interaction for 𝜃′𝑤′)
𝑞3 3rd quadrant (inward interaction for 𝑢′𝑤′ and cold downdraft for 𝜃′𝑤′)
𝑞4 4th quadrant (sweep for 𝑢′𝑤′ and inward interaction for 𝜃′𝑤′)
𝑠 hydrostatic state
𝑡𝑟 Transition local stability
𝑤 Second moment of vertical velocity fluctuation (MOST)
𝜃 Second moment of temperature fluctuation (MOST)

Dimensionless Groups

𝐴𝑅 Roughness aspect ratio (ℎ/𝑏)


𝑃𝑟 Molecular Prandtl number (𝜈/𝐷𝑇 )
𝑃𝑟𝑡 turbulent Prandtl number
𝑃𝑟𝑠𝑔𝑠 SGS turbulent Prandtl number
𝑈0 ℎ
𝑅𝑒0 Reynolds number based on ℎ and 𝑈0 ( )
𝜈
𝑢𝜏 𝐿 𝑧
𝑅𝑒𝜏 Friction Reynolds number ( )
𝜈
𝑢𝜏 𝑧0
𝑅𝑒 ∗ Roughness Reynolds number ( )
𝜈

xvi
𝛼𝑔ℎΔ𝜃
𝑅𝑖0 Richardson number based on ℎ, 𝑈0 and Δ𝜃 (− )
𝑈0 2
𝛼𝑔𝐿𝑧 Δ𝜃
𝑅𝑖τ Friction Richardson number (− )
𝑢𝜏 2
̅ 2ℎ −𝜃𝑤 )
𝛼𝑔(2ℎ)(𝜃
𝑅𝑖2ℎ Richardson number at 2ℎ ( ̅2ℎ 2
)
𝑢

Other Symbols and Operators

•̂ Normalized of field variable by 𝑈0 , ℎ and ∆𝜃


•𝑟 Resolved-scale field variable
•𝑠 SGS field variable
•̅ Ensemble mean (averaged in t- and y-domains in LES)
•′ Deviation from ensemble mean
〈•̅〉 Horizontal mean (averaged in x-domain in LES)
•′′ Deviation from horizontal mean
•+ Wall unit, normalized by 𝑢𝜏 , 𝜃𝜏 and 𝑙𝜐
𝐷(•) Dispersion of • from the horizontal mean

xvii
Chapter 1

Introduction

In recent decades, the total amount of urban area (e.g. cities and towns) and the density
of human beings living in urban areas are increasing rapidly. On the other hand, the demands
for better living environment and more accurate whether prediction are also increasing. As a
result, the field of urban boundary layer (UBL), which studies how the flow within an
atmospheric boundary layer (ABL) is influenced by an underlying urban area, has drawn a
widespread attention from various groups of people, including researchers, meteorologists,
urban planners, the government, environmentalists and some dwellers. This study is aimed to
understand how the flow at the lowest portion of the UBL, the surface layer (SL), which is
mostly concerned due to direct interaction with human activities, is influenced by the urban
surface heating due to solar radiation and anthropogenic heating by utilizing a state-of-the-art
research method – Large-eddy simulation (LES). An overview of the present study, the research
objectives, the significance of research outputs and an outline of thesis will be given in this
chapter.

1.1 Overview

1.1.1 Background of scientific field

This study belongs to the field of UBL, which is a special case of ABL and is
encompassed in a broader field – micrometeorology – that studies the lowest 100 to 1000m of
the earth atmosphere. With the addition of urban roughness structures, the flow behavior at the
surface layer (SL), defined as the lowest 10% of the UBL or ABL, may be very different from
the rural one with a fairly smooth surface. A fundamental objective of UBL studies is to develop
1
tractable models that describe the flow and transport processes, which are the fundamental
component of many applications, e.g. numerical weather prediction (NWP), urban energy
balance analysis, concentration and flux estimation of pollutants or other constituents, urban
planning, traffic control, etc. However, accurate while simple models require insight of the
physics governing the processes that the models represent. This motivates many observational
and experimental studies, as well as this study, in order to provide a deep understanding of the
physical processes of turbulent transports.
Unstable (or convective) flow is less understood compared to neutrally stratified flow
(neutral flow in short) due to the additional complexities introduced by the heating and
buoyancy force field. The importance of stability effect should not be underestimated. Heating
and buoyancy are ubiquitous wherever on the earth as solar radiation is the most fundamental
driving force of the atmospheric dynamics and it covers a significant proportion of time in a
day. Convective boundary layer (CBL), which studies the unstable buoyancy effect on the ABL,
is a quite convectional field and has a long history of progress. However, the roughness effect,
or more specifically urban roughness, is comparatively less understood, but significant progress
has been achieved. The combination of thermal instability and urban roughness lead to the
convective urban boundary layer (CUBL), which is far less understood despite its importance.
Thus, it is the aim of this study to fill this knowledge gap.
Flow over roughness is a broad topic. The flow over urban roughness bears a strong
similarity from the engineering flows (channel flow, pipe flow and boundary layer flow) over
artificial roughness and, therefore, the physical insights from the engineering flow can
complement the research of UBL. However, it should be careful when a direct comparison of
the two flows is performed since there are some important distinctions between them, namely,
the Reynolds number effect, the presence of buoyancy stability in UBL and the effect of
planetary rotation. In addition to UBL and engineering flows, it also includes flows over natural
obstructions, such as vegetation and gravel bed. Conceptual advance of UBL is sometimes be
made from this broad knowledge base of flows over different types of roughness. On the other
hand, comparing flows over roughness to the flows over smooth surfaces, such as ABL and the
engineering canonical flows (channel flow, pipe flow and boundary layer flow), which have
already been undergone a long history of research, also provides many meaningful physical
insights on the flow processes related to roughness. If CUBL are being considered, i.e. both
unstable buoyancy and roughness are important factors, it may also be meaningful to compare
the flow to the convectional CBL. Some of these meaningful researches and useful concepts
from these will be briefly described in the Chapter 2.
2
1.1.2 Some important theories, concepts and potential problems

One of the most important achievements of this scientific field is the Monin-Obukhov
similarity theory (MOST) (Monin and Obukhov [1], Obukhov [2]), which forms the
fundamental component of NWP and will be reviewed in the Chapter 2. It organizes the
statistical relationship among different local flow variables in the SL of ABL, including the
surface fluxes of momentum, heat and scalars, stability level and local height. The functional
dependence can be theoretically derived using dimensional analysis, while the exact functional
form of the expression for a specific flow statistic, named MOST function, has to be found out
by experimental or observational data. Based on MOST, the MOST functions should be
“universal”. An essential assumption of MOST is that the governing turbulent process is “local”,
which means that the local flow properties are only affected by the local flow conditions. This
implies that both the larger and the smaller scale motions are much weaker than the motions at
the scale of local height. Thus, the length scales other than the local height is excluded. In terms
of turbulence kinetic energy (TKE) budget, the “local” statement aforementioned can be
interpreted as a turbulent process in which the TKE production is much stronger than the TKE
transports such that the TKE generated are dissipated locally. This is consistent to the
Townsend’s hypothesis that the turbulence in SL can be viewed as a local equilibrium process
(Townsend [3]). Townsend also hypothesized that the turbulence could be decomposed into an
active part that accounts for the flux producing motions and an inactive part that is larger in
scale and produces no vertical flux, while it is the active part but not the inactive part that
explains the MOST. Since the popular field observation study of unstably stratified SL, the
Kansas 1968 experiment (Izumi [4]), that supported MOST, MOST has long been a mainstream
opinion of the field of ABL (see the historical survey of MOST by Foken [5]).
On the other hand, when roughness scale is much smaller than the boundary layer scale
and assuming that MOST is valid, the MOST functions of the mean gradients (of velocity,
temperature or other scalars if their surface sources is sufficiently homogenous) can be
integrated mathematically to derive the mean profiles and each mean profile expression now
contains a constant - roughness length (of momentum, temperature or specific scalar transfers)
- that describes the dynamical property of the surface as a whole (see the introduction of Kanda
et al. [6] for physical characteristics of roughness lengths and the review of Kramm et al. [7]
for the integration of MOST functions). The well-known logarithmic law of the wall (log-law
in short) can be deduced from the integrated form of MOST as a specific case in neutral

3
condition. The aerodynamic roughness length (a.k.a. roughness length of momentum transfer)
has long been considered a function of surface geometry and dimensions only and is
independent of the buoyancy effect. However, some researchers questioned its validity and
provided evidences disproving it, e.g. Zilitinkevich et al. [8] showed by field measurement data
of vegetation canopy that the aerodynamic roughness length could be modified by buoyancy
effect. This issue on urban type roughness will also be addressed in this study through LES
data.
Recently, however, questions have been raised concerning the validity of the
assumption that the local turbulent processes dominate the flow statistics. An increasing
number of publications provide evidence that boundary layer scaled motions, which has long
been considered inactive in the sense of Townsend’s hypothesis, are able to influence the local
statistics in SL through certain modulating mechanisms and account for significant proportion
of vertical fluxes of momentum, heat and scalars (Smedman et al. [9], McNaughton et al. [10]
and Laubach and McNaughton [11]). Some researchers also suggest that this is the reason that
the universality of the MOST functions derived from field measurement data is usually far from
satisfactory as the data points are usually too scattered in many studies when plotted following
the style of MOST functions. More studies with higher quality of data have to be conducted in
order to clarity this issue and find the mechanisms of how the non-local processes, that make
MOST fail, occur.
On the other hand, MOST is only valid in the region sufficiently away from the
roughness but still within the SL so that the dynamical processes due to roughness is negligible
compared to the dynamical processes due to shear as in convectional SL. This leads to a
commonly accepted decomposition of the SL into two parts: (i) inertial sublayer (ISL) in which
the flow characteristics are similar to those in standard SL over smooth surface, and (ii)
roughness sublayer (RSL) in which the flow characteristics are influenced by the roughness
properties and thus deviate from those of the standard SL (Raupach et al. [12]). Two to five
times the mean roughness element height is usually quoted as the RSL height in the literature.
In contrast to the convectional SL such as in rural areas, the roughness elements of most urban
areas are non-negligible compare to the ISL height and thus there is a non-negligible RSL
height that should be taken into account explicitly in modelling of UBL. It is possible that the
urban roughness is so large that the RSL may completely replace the ISL so that the ISL may
not exist (Rotach [13]). Due to the direct influence of urban roughness, MOST is not valid in
RSL since MOST excludes any length scales of physical roughness. Since the RSL is important
in the UBL modelling, its flow characteristics and the corresponding mechanisms are worth to
4
be investigated in order for the models developed in the future to include the physical properties
of RSL. A prominent property of RSL is the horizontal inhomogeneity of different ensemble
statistics. As this inhomogeneity reflects the influence of roughness, some researchers, e.g.
Florens et al. [14], suggested that the RSL height could be calculated as the height at which
certain ensemble statistics become horizontal homogenous, though the degree of homogeneity
and the choice of ensemble statistics are taken arbitrarily. Next, it will be asked how the
buoyancy effect modifies the RSL height as well as the flow properties in RSL. Unfortunately,
the studies concerning this question are rare. It has to be studied in order to properly account
for both the effects of roughness and buoyancy in UBL modelling.

1.1.3 Research trend towards understanding flow mechanisms

Historically, the researchers of ABL, including UBL and CBL as well, were interested
in the single-point statistics, including time series, spatial profiles, probability densities and,
the more detailed one, spectra, due to the limit in measurement technology. This limited the
investigation of the phenomenon to statistical analysis, in which turbulent fluctuations are
inevitably treated as random processes. Therefore, it was very difficult to understand the
mechanisms driving the phenomenon and the models developed were mostly empirical in
nature and thus their validity are limited to a narrow range of condition.
With the advance of measurement and computational simulation technologies,
collection of high resolution data of complete spatial and temporal flow field were becoming
feasible. From the past several decades, the researchers shifted their interest to the mechanisms
behind those statistical data. They discovered that large proportion of turbulent transports and
the motions that contained most energy were from the coherent motions that were of large scale
and occurs repeatedly. The scientific field then began to treat the turbulence not as a completely
random process, but of certain organizing characteristics that are possibly predictable. The
studies of flow mechanisms started from the kinematics, which studies the spatial structures of
these coherent motions (coherent structures in short), e.g. see the review of Robinson [15] for
some basic coherent structures in wall boundary layer flow and the LES study of Khanna and
Brasseur [16] in unstable ABL, and those present in CUBL will be briefly reviewed in Chapter
2. The knowledge of kinematics are nowadays relatively well established. The next step is to
probe the dynamics of these coherent motions, i.e. how the coherent structures are generated,
evolve and interact with other classes of coherent structures. It is believed that these knowledge

5
is essential in understanding why some of the existing theories, e.g. MOST, fail and in
developing physically based models of atmospheric flows, including CUBL, for more accurate
and efficient prediction.

1.1.4 Research methods

The research of ABL, including UBL and CBL, is conventionally conducted through
observations in real condition. It is the main source of data to understand the phenomena and
develop empirical model and it is still a mainstream research tool nowadays. However, field
observation studies suffer from a number of problems. Firstly, as discussed by Wyngaard [17]
and Wyngaard [18], the fluid motions in real ABL contain a very extensive range of scales that
lead to high stochastic variability. On the other hand, strict horizontal homogeneity and
steadiness, as required in ensemble averaging and some ABL theories, are almost impossible
to achieve due to the surface topology, synoptic perturbation of the atmosphere and the diurnal
variation. These factors altogether make the convergence of ensemble statistics of flow
measurement difficult. It is believed that this is the reason of the unsatisfactory collapse of data
when plotted in MOST style. Secondly, the meteorological condition, e.g. direction of wind,
solar radiation, cloud cover and boundary layer height, cannot be controlled. As a result,
together with the large time scale of the change of condition in ABL, field measurement of
certain desired conditions usually takes a very long time, from several months to several years,
to complete in order to collect sufficient data for ensemble averaging. Thirdly, due to the
extensive spatial scale, it is difficult to measure the whole boundary layer, though remote
sensing technique able to measure the whole boundary layer starts to be popular recently, and
it is also difficult to take simultaneous measurement at multiple points. As a result, it is difficult
to study the flow patterns, like the coherent structures, in detail.
To solve the aforementioned problems, reduced scale experiments such as wind tunnel
and water channel can be used instead to simulate the ABL under a carefully controlled
condition. Compared to field measurements, high degrees of homogeneity and steadiness can
be attained and the time needed to collect sufficient data is much shorter. The data resolution
is also much higher in reduced scale experiment. The quality of data is ensured by all of these
advantages. Hot wire anemometry is a traditional technique to measure flow velocity at single
point. With the advance of new technology like particle image velocimetry (PIV), multiple-
point measurement can now be done more easily so that the flow patterns can be investigated

6
in a deeper way. Nonetheless, the reduced scale experiments are limited due to their high cost
and several technical problems. To study a flow over roughness in an equilibrium condition, in
which the variation of flow properties due to the internal boundary layer generated by the first
row roughness has been died out and the flow properties are homogenous in the streamwise
direction, a very long roughness fetch before test section is required (Wieringa [19]). To achieve
this condition, a long wind tunnel or water channel is needed. An important limitation of
reduced scale experiment is that it is hard to simulate both the windy and unstable condition.
Because of it small length scale, the temperature has to be very high in order to satisfy the
physical similarity (i.e. the Richardson number) equivalent to that of the real CBL or CUBL,
which makes reduced scale experiments infeasible to study buoyancy effects, as explained by
Kanda [20].
With the advance of computational power in the past several decades, the field of
computational fluid dynamics (CFD) grew rapidly and the cost of simulations of atmospheric
flows is nowadays much lower. The resolution of simulated flow field is normally much higher
than those of the reduced scale experiments and field observational studies. Both the spatial
and temporal fields in numerical simulation are complete in the sense that all the flow
properties are known at all locations and all times, which certainly cannot be achieved in the
two aforementioned types of study. Numerical simulation allows the selection of fundamental
physics to be simulated through assignment of the fundamental equations (e.g. Navier-Stoke
equation) and the boundary conditions of the numerical model. This allows the study to focus
on physics due to a small number of governing factors, such as compressibility, buoyancy,
planetary rotation, without assuming other unimportant factors are negligible as in physical
experiment and field measurement. In addition, to use an idealized numerical domain
representative of the target atmospheric one, the numerical study is thus an idealized study,
which avoid the problems of data being contaminated by unwanted factors. Using periodic
boundary condition, a hypothetical domain of infinite horizontal extent and the quasi-steady
state of flow can be simulated without the need of a very large domain or a very long fetch for
the flow development. More importantly, for the studies of CBL or CUBL with mean wind,
numerical simulation avoids the problem of too high temperature in reduced scale experiment.
The level of instability is free to prescribe in numerical simulation, given that the fundamental
equations are still justified within the range of instability level. For example, Boussinesq
approximation is commonly employed to simplify the original set of compressible equations
to an incompressible ones to reduce the difficulty of numerical solution. Boussinesq
approximation is used in this study and its validity when applied in atmospheric flows will be
7
briefly described in Chapter 3.
Basically, there are three categories of numerical simulation: (i) Reynolds-averaged-
Navier-Stoke (RANS) simulation, (ii) direct numerical simulation (DNS) and (iii) large-eddy
simulation (LES). In DNS, all scales of turbulent motions are solved explicitly from the
primitive governing equations without any simplifying models embedded to the solution
processing. Therefore, the solution process of DNS is tedious and its volume of calculated
solution data is huge when the Reynolds number is high since the flow contains an extensive
range of length scales of turbulent motions (e.g. for atmospheric flows Reynolds number is at
the order ~108 and length scale ranges from 1km to 1mm). This limits DNS to the studies at
low to moderate Reynolds number as in engineering flows (e.g. channel flow, pipe flow and
boundary layer) and not suitable to atmospheric flows, although understanding of some
important flow physics in atmospheric flows can be gained from these low to moderate
Reynolds number flow studies.
To overcome the problems of handling large amount of motion scales, RANS method
could be used, in which the primitive equations are ensemble averaged such that only the mean
properties are solved while the turbulence are models. The solution processing of RANS are
much faster. However, the accuracy of RANS is widely known to be poor due to the non-
universality of the turbulence model, which is normally limited to certain types of turbulent
flow and calibrated within narrow range of condition. Since the turbulence is modeled rather
than being simulated, RANS is normally used to predict the mean flow but not well suitable to
be used in study of turbulence properties, such as turbulence statistics and coherent structures.
In order to simulate explicitly the turbulent motions while avoiding the restrictive
computational requirements as in DNS, LES could be employed. In LES, the primitive
equations are filtered such that only the large-scale (or resolved scale) motions, represented by
the filtered flow variables, are solved explicitly while the unresolved small-scale part, a.k.a.
subgrid scale (SGS) component, are represented by SGS model. The accuracy of mean flow
properties calculated by LES (requiring ensemble averaging) is much higher than that of RANS.
It is because the flow properties, both the mean fields and the turbulence fields, are critically
determined by the large-scale turbulence, which dominates the fluxes due to its anisotropic
nature, rather than the SGS motions, which are much more random and isotropic. It shows that
the accuracy of simulation is critically linked to the correctness of the physical natures being
simulated. Numerical studies showed that LES performed well over RANS especially in critical
phenomena, such as flow separation and reattachment, recirculation and shear layers (see
Chapter 13 of Pope [21]). Since the flow of CUBL contains different types of turbulence, it is
8
difficult to model all these types of turbulence using a single turbulence model in RANS. LES
is therefore chosen in this study, together with the aim to study the turbulence nature which
certainly cannot be done using RANS.

1.2 Research Objectives and Thesis Outline

In this study, LES is employed to simulate the flow over idealized urban type roughness
under convective condition due to urban surface heating of varying levels, with the neutral case
as the reference case. Both the turbulence generated due to the interaction of the bulk flow with
the roughness elements and the buoyancy induced eddies spanning the whole boundary layer
are essential elements of CUBL flows and therefore are explicitly simulated by LES in this
study. The main focuses of this study include MOST in CUBL, the interaction of the large
convective eddies with the near surface flow and the properties of RSL in convective condition.
The objectives of this study are:
(i) To study the validity and applicability of MOST for different single-point statistics
under the idealized conditions of LES;
(ii) To study how the flow statistics deviate from those predicted by MOST and how the
properties of horizontal inhomogeneity change in RSL under a range of convective conditions;
(iii) To study how the aerodynamic and temperature roughness lengths change with
buoyant instability and the reason behind;
(iv) To investigate mechanisms of momentum and heat transports by considering TKE
budgets and coherent structures from instantaneous flow visualization, two-point correlation
and conditional averaging; and
(v) To study to what extent the Townsend’s hypothesis, which states that the large
convective eddies and the active turbulence in near surface (ISL and RSL) do not interact, is
valid.
The thesis structure is outlined as follows. The literature review of existing theories and
concepts concerning CBL and flow over roughness will be given in Chapter 2. The
methodology to construct the LES models, including domain setting, the selection of primitive
equations and corresponding input parameters, basics of LES and SGS modeling and numerical
aspects, post-processing methods as well as comparison between LES and wind tunnel results
will be discussed in Chapter 3. The single-point statistics in the forms of MOST functions,
Prandtl’s mixing length and sensitivity of MOST statistics to LES domain dimensions will be

9
Figure 1.1: Flow chart of this thesis with objectives.

studied in Chapter 4. The influence of roughness on RSL flow properties and RSL height will
be studied through the deviation of flow statistics from the MOST predicted ones and the
horizontal inhomogeneity of flow statistics in Chapter 5. The variations of ISL mean velocity
and temperature profiles as well as aerodynamic and temperature roughness lengths with
buoyant instability are studied in Chapter 6. Chapter 7 will study the mechanism of turbulent
transports at near-wall region through TKE budget analysis, instantaneous flow visualization,
quadrant analysis, correlation between momentum and heat fluxes, two-point correlation,
integral length scale and conditional averaging. Conclusion of this study will be given in
Chapter 8. The flow chart of this thesis to address the objectives is shown in Figure 1.1.

10
Chapter 2

Literature Review

Literature reviews of the existing theories, concepts as well as their recent advancement
that are related to this study are given in this chapter. This chapter reveals the uncertainties,
unclearness and inadequacy of these theories and concepts that await clarification and
refinement. These constitute the motivation and set the stage of the present study. This study
will center on MOST and its related concepts such as Townsend’s hypothesis. Therefore,
MOST, including its derivation and the functional forms of MOST functions, will be reviewed
first. The layer deposition of CUBL commonly accepted by the literature and fundament to this
study will then presented next. The parameterization of urban surface by simple parameters
such as roughness lengths and those concepts concerning RSL will be reviewed then. In order
to understand the flow mechanisms that complement the basic statistical study, the budgets of
TKE and fluxes and knowledge of coherent structures are reviewed as well.

2.1 Monin-Obukhov Similarity Theory (MOST)

Before reviewing flows over roughness, the concepts of conventional surface layer (SL)
that MOST develops on without explicitly taking roughness effect in account will be reviewed
first. SL is the lowest layer of a conventional atmospheric boundary layer (ABL) that connects
the land surface and the atmosphere aloft. It is the layer that is essential to the exchanges of
momentum (surface drag), temperature and other scalar constituents with the atmosphere,
which strongly affects the local meteorology and impacts to human beings, and thus its
modelling is an essential component to numerous applications, e.g. numerical weather
prediction (NWP), surface drag and fluxes estimations, determination of surface aerodynamic
properties and air pollution estimation. Depending on the meteorological condition, SL usually

11
accounts for only about 10% (or more in neutral condition) of the ABL in which the momentum
and heat fluxes are relatively vertically uniform due to its thinness. An important attributes of
SL is that its flow statistics strongly depend on the local height 𝑧 rather that the larger boundary
layer height 𝑧𝑖 . This offers an advantage to derive a local relation of the statistical flow
quantities in SL, which gives rise to the Monin-Obukhov similarity theory (MOST). We will
first consider MOST in SL without direct influence of any roughness effects.
MOST, originally proposed and derived by Monin and Obukhov [1] (see also Obukhov
[2]), is based on dimensional analysis. It assumes that a single-point statistics 𝑓𝑖 under the
quasi-steady and horizontal homogenous conditions in SL only depends on a small number of
variables (first proposed by Obukhov [22]), including the surface shear stress 𝜏0 , air density 𝜌,
the local height 𝑧, the surface heat flux 𝑞0 (positive when heat transfers upwards), the specific
1
heat capacity of air 𝑐𝑝 , the gravity 𝑔 and the thermal expansion coefficient 𝛼 (≈ 𝑇 , where 𝑇0
0

is the ambient temperature in absolute scale):


𝑓𝑖 = 𝑓𝑖,1 (𝜏0 , 𝜌, 𝑧, 𝑞0 , 𝑐𝑝 , 𝑔, 𝛼) (2.1)
It is also assumed that other factors such as boundary layer height 𝑧𝑖 , Coriolis force due to
planetary rotation, roughness effect and humidity effect (or other active scalars) are negligible
here, though the humidity effect can be included in MOST by using virtual temperature in place
of the potential temperature (see chapter 10 of Wyngaard [17] for reference and the humidity
effect will not be considered in this study). To further simplify the derivation, a number of
variables are defined. Firstly, friction velocity 𝑢𝜏 is defined to represent the surface stress and
can be approximated by the Reynolds stress averaged in both temporal and horizontal
directions within SL:

𝜏0 ̅̅̅̅̅̅〉𝑆𝐿
𝑢𝜏 = √ ≈ 〈𝑢′𝑤′ (2.2)
𝜌

Overbar means ensemble average and angle bracket means horizontal average (defined in
Chapter 3). The buoyancy parameter is defined to replace the gravity and thermal expansion
coefficient:
𝛽 = 𝛼𝑔 (2.3)
The friction temperature is defined to replace the heat flux:
𝑞0
𝜃𝜏 = − (2.4)
𝜌𝑐𝑝 𝑢𝜏
while the vertical temperature flux is used instead of the heat flux and can also be approximated
in SL as:

12
𝑞0 ̅̅̅̅̅̅〉𝑆𝐿
≈ 〈𝜃′𝑤′ (2.5)
𝜌𝑐𝑝
The potential temperature 𝜃 is used instead of the usual temperature 𝑇 , which is explained
along with Boussinesq approximation in Chapter 3. Furthermore, the Obukhov length 𝐿
(proposed by Obukhov [22]) is defined to represent the strength of buoyancy effect relative to
shear:
𝜌𝑐𝑝 𝑢𝜏 3
𝐿=− (2.6)
𝜅𝛽𝑞0
where 𝜅 is the von Karman constant that conventionally defined based on the log-law and it is
empirically found to be ~0.4 (its nature was reviewed by Hogstrom [23]). Finally, the MOST
relation is reduced to a much simpler form:
𝑓𝑖 = 𝑓𝑖,2 (𝑢𝜏 , 𝑧, 𝐿, 𝜃𝜏 ) (2.7)
By dimensional analysis, the functional relation can be written in dimensionless form:
𝑓𝑖 𝑧
= 𝜙𝑖 ( ) = 𝜙𝑖 (𝜁) (2.8)
𝑓𝑖,𝑀𝑂𝑆𝑇 𝐿
𝑧
𝜁 = 𝐿 is the local stability parameter specifying the relative strength of buoyancy to mechanical

shear and is negative in unstable (convective) condition. 𝑓𝑖,𝑀𝑂𝑆𝑇 is a reference scale, formed
from the combination of (𝑢𝜏 , 𝑧, 𝜃𝜏 ), to normalize 𝑓𝑖 . 𝜙𝑖 (𝜁) is the MOST function (or named
stability correction function in some studies), which is universal if MOST is valid. Therefore,
𝜙𝑖 is a constant at neutral condition (𝜁 = 0).
Since MOST is a “local” argument, the nature of the flow statistics that MOST describes
should also be local. Within a considerable range of instability, the most important flow
𝑑𝑈
statistics that are known to satisfy MOST well include mean streamwise velocity gradient ,
𝑑𝑧
𝑑Θ ̅̅̅̅̅̅〉 and
mean temperature gradient , second moment of vertical velocity fluctuation 〈𝑤′𝑤′
𝑑𝑧
̅̅̅̅̅〉 ( 𝑈 = 〈𝑢̅〉 and Θ = 〈𝜃̅〉 for conciseness).
second moment of temperature fluctuation 〈𝜃′𝜃′
Their dimensionless MOST functions are defined as:
𝜅𝑧 𝑑𝑈 (2.9)
𝜙𝑚 (𝜁) =
𝑢𝜏 𝑑𝑧
𝜅𝑧 𝑑Θ (2.10)
𝜙ℎ (𝜁) =
𝑃𝑟𝑡 𝜃𝜏 𝑑𝑧
̅̅̅̅̅̅〉
〈𝑤′𝑤′
𝜙𝑤 (𝜁) = (2.11)
𝑢𝜏 2
̅̅̅̅̅〉
〈𝜃′𝜃′
𝜙𝜃 (𝜁) = (2.12)
𝜃𝜏 2
𝜙𝑚 (0) = 1 at neutral state so as to be consistent with the log-law. 𝑃𝑟𝑡 is the turbulent Prandtl

13
number, which is defined such that 𝜙ℎ (0) = 1 at neutral state. For standard SL, 𝑃𝑟𝑡 −1 is found
to range from 1 to 1.4 (reviewed by Foken [5]). The subscripts m and h stand for momentum
and heat.
Nonetheless, some statistics such as second moments of streamwise and spanwise
̅̅̅̅̅〉 and 〈𝑣′𝑣′
velocity fluctuations, 〈𝑢′𝑢′ ̅̅̅̅̅〉, are widely known not to satisfy MOST. This problem

will be explained later when the Townsend’s hypothesis and inactive turbulence are reviewed.
Dimensional analysis only stipulates the dependence of the variables, but it does not provide
the functional form of 𝜙𝑖 (𝜁) , which can only be found empirically, through observations,
experiments or numerical solution of the fundamental transport equations, unless further
assumption(s) is made.
The MOST functions can be derived theoretically if we assume that the shear effect is
negligible such that 𝑢𝜏 is irrelevant and is taken out from the list of dimensional variables.
When this condition is valid, usually at strong unstable level such that the local stability
parameter 𝜁 is high enough, the local free-convection state is said to be attained. Since the
definitions of 𝐿 and 𝜃𝜏 also contain 𝑢𝜏 , an alternative set of variables have to be sought. Here,
the local free-convective velocity and temperature scales are defined, respectively, as:
𝑞0 1 𝑞0
𝑢𝑓 = (𝜅𝛽 𝑧)3 , 𝜃𝑓 = − (2.13)
𝜌𝑐𝑝 𝜌𝑐𝑝 𝑢𝑓
Subsequently, the MOST relations can now be written as:
𝑓𝑖 = 𝑓𝑖,3 (𝑧, 𝑢𝑓 , 𝜃𝑓 ) (2.14)
By dimensional analysis, the dimensionless form can be written as:
𝑓𝑖
= 𝐶𝑖 (2.15)
𝑓𝑖,𝐹𝐶
𝑓𝑖,𝐹𝐶 is a reference scale from the list (𝑧, 𝑢𝑓 , 𝜃𝑓 ) to normalize 𝑓𝑖 and the subscript FC stands for
the local free-convective condition. 𝐶𝑖 is an empirical constant. This relation is much simpler
since only the empirical constant 𝐶𝑖 rather than the function 𝜙𝑖 (𝜁) needs to be found from
empirical data. To reconcile the MOST relations, the form of (2.15) can be re-written into the
commonly used form of (2.8) through transformation of variables:
𝑧 1 𝑧 1
𝑢𝑓 = 𝑢𝜏 (− )3 , 𝜃𝑓 = 𝜃𝑓 (− )−3 (2.16)
𝐿 𝐿
As a result, each MOST function 𝜙𝑖 (𝜁) under the local free-convective state now has a specific
functional form, which is a power function with a definite index:
𝜙𝑚 (𝜁) = 𝐶𝑚 (−𝜁)−1/3 (2.17)

14
𝜙ℎ (𝜁) = 𝐶ℎ (−𝜁)−1/3 (2.18)
𝜙𝑤 (𝜁) = 𝐶𝑤 (−𝜁)1/3 (2.19)
𝜙𝜃 (𝜁) = 𝐶𝜃 (−𝜁)−1/3 (2.20)
where 𝐶𝑚 , 𝐶ℎ , 𝐶𝑤 and 𝐶𝜃 are empirical constants.
Although the form of 𝜙𝑖 (𝜁) are much simpler in local free-convective condition, the
range of 𝜁 within which the local free-convective condition is valid is still vague. Kader and
Yaglom [24] suggested that local free-convective state occurs when 𝜁 < −0.12 . However,
some observational studies showed that the lower bound of 𝜁 at which the MOST function
starts to satisfy the local free-convective form is different for different statistics. For example,
by reviewing past observational data, McNaughton et al. [10] suggested that the local free-
convection scaling of scalar statistics (e.g. temperature gradient, temperature variance and
temperature dissipation rate) can be applied down to a much lower 𝜁 (down to |𝜁|~0.04) than
purely dynamical quantities (e.g. vertical velocity variance and TKE dissipation rate).
Another theory to derive the MOST functions were proposed by Kader and Yaglom
[24]. They applied the “directional dimensional analysis”, which treated the vertical and
horizontal length as separate dimensions such that there was an additional length scale
compared to the original MOST derivation. Based on this argument, Kader and Yaglom
proposed three sublayers corresponding to certain ranges of local stability 𝜁 : (i) dynamic
sublayer ( 0 < −𝜁 < 0.04 ) at which TKE is generated mainly by shear, (ii) dynamic–
convective sublayer ( 0.12 < −𝜁 < 1.2 ) at which the shear produces TKE in streamwise
direction and buoyancy produces TKE in vertical direction while the energy exchange between
them through pressure redistribution is insignificant, and (iii) free-convection sublayer (−𝜁 >
2) at which TKE is generated mainly by buoyancy. This three-layer model provides not just the
functional forms, but also the physical interpretation based on turbulence energy consideration.
The derived functions of Kader and Yaglom and its problems asserted by other researchers will
discussed in following sub-sections.

2.1.1 MOST functions of mean gradients (𝝓𝒎 and 𝝓𝒉 )

Since the emergence of MOST, many researchers have proposed different functional
forms approximating the “universal” MOST functions. Throughout the history, nearly all
MOST functions proposed were from regression of field measurement data. Table 2.1 shows
some of the popular functional forms of 𝜙𝑚 and 𝜙ℎ .

15
𝜙𝑚 𝜙ℎ
Businger et al. (1 − 15𝜁)−1/4 (1 − 9𝜁) −1/2
, 𝑃𝑟𝑡 = 0.74
[25]
Dyer and (1 − 16𝜁)−1/4 (1 − 16𝜁)−1/2 , 𝑃𝑟𝑡 = 1
Hicks [26]
Kader and 1 , −0.04 < 𝜁 < 0 1 , −0.04 < 𝜁 < 0
Yaglom [24] {0.5|𝜁|−1/3 , −1.2 < 𝜁 < −0.12 −
1
{0.34|𝜁| 3 , −1.2 < 𝜁 < −0.12 ,
0.38|𝜁|1/3 , 𝜁 < −2 1
0.28|𝜁|−3 , 𝜁 < −2
, 𝑃𝑟𝑡 = 0.85
Hogstrom [23] (1 − 19𝜁)−1/4 (1 − 11.6𝜁)−1/2 , 𝑃𝑟𝑡 = 0.95
Wilson [27] (1 + 3.6|𝜁|2/3 )−1/2 (1 + 7.9|𝜁|2/3 )−1/2 , 𝑃𝑟𝑡 = 0.95
Webb [28] 1 + 4.5𝜁 , −0.03 < 𝜁 < 0 -

Table 2.1: Commonly used functional forms of dimensionless mean gradients.

The simplest one is the linear form proposed by Webb [28], but it was empirically found
to be only valid for slightly unstable condition. Until nowadays, the most widely used
functional forms of 𝜙𝑚 and 𝜙ℎ are the Businger-Dyer (BD) relations:
𝜙𝑚 (𝜁) = (1 − 𝑏𝑚 𝜁)−1/4 (2.21)
𝜙ℎ (𝜁) = (1 − 𝑏ℎ 𝜁)−1/2 (2.22)
proposed by Businger et al. [25], Dyer and Hicks [26] and Dyer [29], with the empirical
constants derived from the famous Kansas experiment (Izumi [4] and Kaimal and Wyngaard
[30]), and are applicable from neutral to moderately unstable state ( −1 < 𝜁 < 0 ). The
popularity of BD relations is due to its simple functional forms that permit analytical
mathematical integration, which allows the mean profiles to be derived analytically (its usage
and derivation will be discussed in later section). Hogstrom [23], using the same functional
forms of BD relations, refined the empirical constants and turbulent Prandtl number based on
a number of field measurement data, but at the same time suggested that the proposed forms of
𝜙𝑚 and 𝜙ℎ were not exact due to the statistical uncertainty of order of 10% to 20%. Hogstrom
suggested there were five possible causes that might lead to the uncertainty. The first four were
from the experimental errors of field measurement while the last one was the limitation in the
validity of MOST. However, Hogstrom finally denied the last reason and insisted MOST should
be universal. The views concerning this possibility by different researchers will be discussed
in later sections. The problem of BD relations is that the indexes of 𝜙𝑚 and 𝜙ℎ (-1/4 and -1/2)
are different from those of the local free-convection argument (both are -1/3). As a result, their
behaviors predicted at strong instability level are very different and the BD relations are usually
regarded as lacking physical basis. Carl et al. [31] suggested from his field measurement data
16
that the index -1/3 was superior -1/4 for 𝜙𝑚 . On the other hand, Wilson [27] proposed the form:
𝜙𝑖 (𝜁) = (1 + 𝛾𝑖 |𝜁|2/3 )−1/2 , 𝑖 = 𝑚 𝑜𝑟 ℎ (2.23)
which approaches the -1/3 power asymptotically. Wilson suggested that this functional form
describes the slightly unstable state well while it agrees with the local free-convection form at
strong unstable state (i.e. the index tends to -1/3).
The dissimilarity of indexes between 𝜙𝑚 and 𝜙ℎ in BD relations, in conflict to the local
free-convection argument, is also of interest since the turbulent transports of momentum and
heat in SL have long been assumed equivalent, the so-called Reynolds analogy. Actually, many
studies and their proposed MOST functions suggest that 𝜙ℎ < 𝜙𝑚 in unstable Very condition,
which means heat is transported more efficiently than momentum.
In the literature, there are very less theoretical arguments explaining the forms of 𝜙𝑚
and 𝜙ℎ . The “directional dimensional analysis” of Kader and Yaglom [24] is a theoretical one
that suggested the power index of 𝜙𝑚 at strong instability (i.e. instability free-convection
sublayer: −𝜁 > 2) is 1/3, but this is very different from common measurement data as well as
all other proposed forms of 𝜙𝑚 , of which the asymptotic power index is negative (e.g. BD
relation and local free-convection). In fact, the “directional dimensional analysis” has not yet
been validated, though the field measurement data provided by Kader and Yaglom supported
their arguments. Nevertheless, at strong instability, 𝜙ℎ derived by Kader and Yaglom is
consistent to the local free-convection argument. On the other hand, Katul et al. [32] and Li et
al. [33] proposed theoretical arguments based on turbulent spectrum to explain the forms of
𝜙𝑚 and 𝜙ℎ , with the assumption that the most efficient eddies in contributing the transports
was of size 𝑧 , the local distance from the wall. This theory recovers the -1/4 power-law
(suggested by BD relations) at weak instability and the -1/3 power-law (suggested by local free
convection hypothesis) at strong instability for 𝜙𝑚 . Li et al. then explained, also by turbulence
spectrum, the enhanced efficiency of heat transport (i.e. 𝜙ℎ < 𝜙𝑚 ) was due to a “scale-
resonance” between turnover eddies and excursions in the instantaneous temperature profile.

2.1.2 MOST functions of second moments of fluctuation (𝝓𝒘 and 𝝓𝜽 )

The studies of 𝜙𝑤 and 𝜙𝜃 are relatively less compared to those of 𝜙𝑚 and 𝜙ℎ , which
may be due to its less importance in applications and less controversy on their functional forms.
Most studies agree that both 𝜙𝑤 and 𝜙𝜃 at sufficiently strong unstable condition satisfy those
derived from the local free-convection argument, such that the power indexes of 𝜙𝑤 and 𝜙𝜃 at

17
asymptotic state are 1/3 and -1/3 respectively, as discussed by Johansson et al. [34]. However,
the problem of 𝜙𝑤 and 𝜙𝜃 is that their coefficients 𝐶𝑤 and 𝐶𝜃 can deviate very much across
different studies. For example, Wyngaard et al. [35] found from the Kansas experiment that
𝐶𝑤 = 3.61 and 𝐶𝜃 = 0.90 while Kadar and Yaglom [24] suggested 𝐶𝑤 = 5.0 and 𝐶𝜃 = 1.06
from their measurement data. On the other hand, 𝜙𝑚 and 𝜙ℎ tend to a constant at near-neutral
state. Most studies suggest the neutral value of 𝜙𝑚 ~1.6. However, the neutral value of 𝜙𝜃 can
deviate substantially from different studies. For example, Kadar and Yaglom suggested
𝜙𝜃 (0)~8.41 while Johansson et al. suggested that 𝜙𝜃 (0)~4.
Compared to 𝜙𝑚 and 𝜙ℎ , the transition of 𝜙𝑤 and 𝜙𝜃 from the neutral constant
function to the local free-convective power function is much more rapid (i.e. the transition
point is sharp). Different MOST functional forms to concatenate both the neutral and strongly
unstable states are proposed by different researchers. For example, Panofsky et al. [36]
proposed two different functional forms of 𝜙𝑤 :
𝜙𝑤 = 1.6 + 2.9𝜁 2/3 (2.24)
𝜙𝑤 = 1.69(1 − 3𝜁)2/3 (2.25)
They suggested the later one is preferable to the former one. For 𝜙𝜃 , De Druin et al. [37]
proposed:
𝜙𝜃 = 8.41(1 − 28.4𝜁)−2/3 (2.26)

2.2 Layer Decomposition of Convective Urban Boundary Layer


(CUBL)

If the surface roughness is considered dynamically important, then the conventional SL,
which does not explicitly account for the effect of roughness on the turbulence properties, is
replaced by the inertial sublayer (ISL), the upper portion of SL that shows no influence of
roughness, and the roughness sublayer (RSL), the lower portion of SL that is directly influenced
by roughness (see Raupach et al. [12] for a detailed introduction). This decomposition into RSL
and ISL is in general sense that includes all types of roughness, e.g. urban type, vegetation,
gravel bed and engineering roughness. The main difference between UBL and ABL is that UBL
contains a significant depth of RSL. A convective urban boundary layer (CUBL), or just urban
boundary layer (UBL) in some studies, is commonly divided into, from bottom to top, (i) RSL,
ISL (a.k.a. constant-flux layer), mixed layer (ML) and interfacial layer. This is similar to the

18
Region Characteristics
Urban boundary layer (UBL)
Ground level to the Portion of planetary boundary layer whose characteristics are affected by the
top of the boundary presence of the urban area. Local to mesoscale phenomenon. Includes all other
layer layers.
Urban canopy layer (UCL)
Ground level to about Produced by microscale effects of site characteristics. Dynamic and thermal
roof level processes are dominated by the immediate surroundings. Flow and scalar structure
are generally very complex. Most clearly pronounced in areas of high building
density, but may be discontinuous in less developed suburban area.
Roughness sub-layer (RSL)
Ground to 𝑧∗ Also called the transition layer, interfacial layer or wake layer; includes UCL.
Mechanically and thermally influenced by length scales associated with the
roughness. Because of wake diffusion and differential source/sinks of momentum
and scalars, momentum and heat transfers are dissimilar, and therefore Reynolds
analogy may not be valid. As a result of local-scale advection, the turbulence field is
often not horizontally uniform, even in a time average, and must be considered three-
dimensionally.
Constant-flux layer (CFL)
𝑧∗ to ~0.1𝑧𝑖 Also called the inertial sub-layer. Mean profiles obey semi-logarithmic laws or their
diabatic extensions and MOS applies. Very little is known about it in urban area, in
part due to the height restrictions of measurement towers which, due to the existence
of a RSL with dimensions of tens of metres, often do not extend into the CFL. The
vertical extent of the urban CFL is also determined by the development of an internal
boundary layer which reacts to mesoscale land-use changes in upwind surface
characteristics. It is possible that under unstable conditions the depth of the RSL
exceeds the potential depth of the CFL and no such layer exists.
Mixed layer (ML)
~0.1𝑧𝑖 to 𝑧𝑖 Little is known about urban MLs, but turbulence properties are probably independent
of the surface roughness. It is topped by an entrainment layer which may be
substantial because of enhanced coupling between the rough and warm urban surface
and the ML and the relatively larger availability of turbulent kinetic energy (TKE).
𝑧∗ is the RSL depth

Table 2.2: Decomposition of urban boundary layer and characteristics of each region, suggested by
Roth [38].

decomposition proposed by Roth [38], which is shown in Table 2.2 together with the
characteristics of each layer. In this study, the urban canopy layer introduced by Roth,
extending from the ground level to roof level is included as a part of RSL as adopted by many
studies.

2.2.1 Properties of inertial sublayer (ISL)

In ISL, the roughness effect are negligible to the local turbulence properties and the
turbulence behaves like the SL one so that MOST can still be applied here. The upper limit of
19
ISL is the same as that of SL, i.e. ~10% of boundary layer depth in unstable condition or higher
in neutral condition. The only effects of roughness on the ISL flow are (i) the vertical shift of
the surface datum, of which the shifted height is defined as the displacement height 𝑑, and (ii)
the bulk decrease in mean wind velocity profile 𝑈(𝑧) , parameterized by the aerodynamic
roughness length 𝑧0 (similarly for temperature and other scalars, if there are any). Due to the
vertical shift, the vertical coordinate 𝑧 in MOST, measured from the ground level, should be
replaced by the effective height 𝑧𝑒𝑓𝑓 = 𝑧 − 𝑑 . It is this corrected height that is physically
meaningful since this is the length scale of the most significant turbulent eddies for turbulent
transports. MOST only describes the local properties (e.g. mean velocity gradient) rather than
the integrated properties (e.g. mean velocity). Therefore, the value of 𝑧0 , determined by the
surface roughness morphology, does not alter the local flow properties governed by MOST and
𝑧0 is a surface (boundary) properties (𝑧0 arises from the mathematical integration of MOST
function 𝜙𝑚 (𝜁), reviewed in next subsection).
A very important characteristics of ISL is that the local flow properties are independent
of the surface roughness properties: the wall similarity hypothesis (see the reviews of Raupach
et al. [12] and Jimenez [39]). This implies that the local flow properties in ISL only depend on
the local height 𝑧𝑒𝑓𝑓 whatever the size, shape, orientation and distribution of roughness
elements. MOST, which describes the local flow properties should be universal, does not take
any length scales other than the local height (𝑧 or 𝑧𝑒𝑓𝑓 ) into account and hence the wall
similarity hypothesis is implicit in MOST. Despite there are some challenges on this hypothesis
(e.g. Krogstad et al. [40] and Krogstad and Antonia [41]) and some discoveries from
measurements that the flow away from RSL shows some roughness effects (e.g. Leonardi et al.
[42] and Bhaganagar et al. [43]), this classical view is widely accepted by the literature for
different types of roughness. For urban type roughness, Roth [38] reviewed a large number of
field measurement studies and the data of vertical velocity and temperature fluctuations in both
neutral and unstable states generally satisfied MOST when sufficiently away from the
roughness surface. Moriwaki and Kanda [44] evaluates the applicability of MOST relations for
mean gradients 𝜙𝑚 and 𝜙ℎ over a low-storied residential area and showed that the measured
values of 𝜙𝑚 and 𝜙ℎ over RSL agreed to the convectional semi-empirical functions, the BD-
relations. Similarly, Wood et al. [45] and Al-Jiboori et al. [46] also supported the validity of
MOST in ISL under a range of unstable stratification level.

20
2.2.2 Properties of roughness sublayer (RSL)

In RSL, the local flow properties are directly influenced by the geometrical properties
of roughness (i.e. types, size, shape, orientation and distribution). Due to the turbulence
generated when the bulk flow interacts with the roughness elements, the TKE in RSL is stronger
and the vertical transports of momentum, heat, scalars and TKE by turbulent mixing are
therefore stronger than those predicted by direction extrapolation of MOST from ISL to RSL.
An important attribute that distinguishes RSL and ISL is the turbulent length scale,
which is defined through different manners, e.g. integral length scale by integrating a two-point
correlation coefficient in streamwise direction, spectral peak wavelength of turbulence
spectrum from and the length scale derived from the definition of turbulent diffusivity. Some
studies have shown that the length scale in RSL reflects certain properties of roughness (e.g.
the review of Finnigan [47] and Raupach et al. [12] for vegetation canopy, the wind tunnel
experiment of Reynolds and Castro [48] for urban type roughness and the review of spectral
measurements by Roth [38]), in contrary to the length scale in ISL that is proportional to 𝑧𝑒𝑓𝑓 .
However, these studies were conducted under neutral condition, but not unstable condition.
Apart from the inconsistency of the turbulence length scale to MOST, another attributes
of RSL for urban type roughness is the horizontal inhomogeneity of single-point statistics at
the vicinity of roughness due to the highly inhomogeneous nature of urban morphology.
(However, there may be no or negligible statistical inhomogeneity in some vegetative types of
roughness due to their homogenous nature and the penetrative property like porous materials.)
The experiment of open channel flow over cubes by Florens et al. [14] showed that the spatial
dispersions, which are the standard deviations of horizontal variation of ensemble mean from
the spatial mean, of different single-point statistics were highest near the roughness and
decayed in the upward direction. The spatial dispersions of ensemble statistics pose a problem
to the description of the flow since a one-dimensional (vertical) profile is the normal way of
description in most applications. A high spatial dispersion near roughness means that a high
number of sampling profiles is needed, but this is a difficult task in field measurements and
earlier experiments (e.g. 4-profile by Cheng and Castro [49]). It is still not known how the
unstable stratification would affect the degree of inhomogeneity. When describing the flow
using spatial averaged vertical (1D) profiles, most studies (e.g. Rotach [50] and Coceal et al.
[51]) find that the Reynolds stress in RSL is weaker than the extrapolation of stress profile from
the upper region, which is another important attribute of RSL. This can be explained by the

21
spatial averaged momentum equation, which shows that the decrease of Reynolds stress is due
to the dispersive stress appeared due to spatial inhomogeneity of ensemble mean velocity and
also the form (pressure) drag on the roughness elements if inside the roughness cavity region.
The next question is how to define the extent of RSL region, i.e. the RSL height 𝑧∗ .
There are different ways of definition based on the special properties of RSL as described above,
including the inconsistency of MOST profiles, spatial inhomogeneity of ensemble statistics and
the decrease in Reynolds stress. For example, based on horizontal inhomogeneity, Florens et
al. [14] defined 𝑧∗ as the height at which all the spatial dispersions of mean velocity
components, velocity fluctuation moments and Reynolds stress were less than a threshold value
of 5%. However, there are two shortcomings of this methods to define 𝑧∗ : (i) the threshold
value is arbitrary and (ii) the rate of convergence for each ensemble statistic is different (e.g.
the mean streamwise velocity converges with height much faster than the Reynolds stress). On
the other hand, Rotach [13] defined RSL as the region in which the spatial averaged mean flow
deviates from the smooth wall flow (i.e. the log law or MOST in neutral condition). However,
these two definitions seem not to be equivalent as many studies found that the measured values
of 𝑧∗ are different based on these two definitions. Raupach et al. [12] observed the log-law
layer was extend lower the layer of inhomogeneity. Also, Cheng and Castro [49] and Castro
[52] suggested the log-law profile in spatial average could be observed within the horizontally
inhomogeneous region. Nevertheless, 2-5 roughness height is usually quoted as the RSL height
in most publications (Raupach et al. [12], while Jimenez [39] suggested 2-3 roughness height)
and the all the definitions of 𝑧∗ seem to fall within this range. It seems that the definition of 𝑧∗
is a difficult problem, adding the numerous possibility of roughness types and configurations,
and will be an ongoing research problem in the future. The current state of knowledge about
RSL properties and 𝑧∗ is still mainly on neutral condition while those under unstable condition
are far less, although studies about buoyancy effects on urban environments inside roughness
cavity (or street cavity) are numerous (e.g. Uehara et al. [53], Li et al. [54], Cai [55] and Cheng
and Liu [56], who treated the flows as street cavity flows rather than flows over roughness).
Field observational studies on applicability of MOST within urban RSL include Yusup and Lim
[57], who suggested the concept of “local scaling of MOST” that MOST is still applicable in
RSL if the local Reynolds stress is used instead. Study of unstable RSL is important as this will
strongly influence the predictions of urban meteorology and pollution dispersion, especially
because the resolution of computation becomes fine enough to calculate the RSL flow in recent
years.

22
Another issue about RSL of urban area is that the height of urban roughness may be
very high such that the RSL completely occupies the SL and thus the ISL does not exist (Rotach
[13]). From the experiments on diamond-pattern wire mesh and sand-paper type grit, Amir and
Castro [58] suggested that when the ratio of roughness height to boundary layer depth exceeds
0.15, the effects of roughness extends into the whole boundary layer. In this case, the
assumption MOST and the concept of wall similarity are not applicable. Jimenez [39] proposed
that the boundary layer depth should be at least 80 times the roughness height in order for the
similarity laws (MOST and log-law) to be expected. Jimenez [39] also suggested that the flow
is better viewed as flow over obstacles, rather than rough wall flow, when the roughness height
is large such that the blockage fraction of boundary layer is large. Given that the mean building
heights of most metropolitan cities are not small, understanding the properties of RSL is
essential to the understanding the environment of those cities, especially in day-time on which
unstable condition is likely and human activities are more frequent.
There are already some modelling works of spatially averaged RSL profiles in the
literature, e.g. Belcher et al. [59] and Coceal and Belcher [60] on urban type roughness; Harman
and Finnigan [61], de Ridder [62] and Arnqvist and Bergstrom [63] on plant type roughness.
The modelling of RSL is currently still at the preliminary state and the results are usually
empirical in nature. Much more efforts are needed to add the physical nature of RSL to the
modelling, especially under unstable condition, and test these RSL models. Therefore,
understanding the RSL properties is an essential step to advance the modelling capability.

2.2.3 Properties of mixed layer (ML)

Above SL or ISL, the boundary layer flow properties depend on the boundary layer
depth 𝑧𝑖 . Even under slightly unstable condition, the whole boundary layer is dominated by the
large convective eddies generated by the buoyancy flux. In this case, the boundary layer is
categorized as CBL (or CUBL over urban areas), which is bounded atop by a very stable
interfacial layer (a.k.a. temperature inversion), and the large region between the interfacial
layer and SL is defined as the mixed layer (ML). The interfacial layer damps most of the
turbulent transporting motions from below and separates the CBL from the free atmosphere. In
normal day-time situation, due to the accumulation of thermal energy within the CBL, the CBL
will grow and the air from the free atmosphere will thus entrain down to the CBL. The CBL
growth is usually steady and, as a result, the heat flux profile is linear in vertical direction.

23
Seibert et al. [64] has reviewed the methods to determine 𝑧𝑖 in field measurement and its
parameterizations. Due to the strong large convective eddies in CBL, the flow in ML is well-
mixed and therefore the mean wind (if available) and temperature profiles are fairly uniform
across the ML. Wyngaard [17] (Chapter 11) has discussed the statistical aspects of CBL and
parameterization of their vertical profiles in ML.
The most important theory of ML statistics is the Deardorff scaling (Deardorff [65],
Deardorff [66], Deardorff [67]), which is derived from the dimensional analysis, as a
counterpart of MOST. In Deardorff scaling, the CBL depth 𝑧𝑖 is an important length scale of
the CBL turbulence since the strongest eddies in CBL are observed to be spanned the whole
CBL while the friction velocity 𝑢𝜏 is of no role since the turbulence generated by mechanical
shear is much weaker than the buoyancy in ML. To characterize the strength of turbulence due
to the buoyancy (heat) flux, Deardorff defined the convective velocity scale and the
corresponding temperature scale (not to confuse with the local-convective scales described in
the previous section) as:
𝑞0 1 𝑞0
𝑤∗ = (𝛽𝑧𝑖 )3 , 𝜃∗ = (2.27)
𝜌𝑐𝑝 𝜌𝑐𝑝 𝑤∗
Similar to MOST, the relation of a single-point statistic 𝑓𝑖 in ML can be written as:
𝑓𝑖 = 𝑓𝑖,4 (𝑧, 𝑧𝑖 , 𝑤∗ , 𝜃∗ ) (2.28)
The dimensionless form can be written as:
𝑓𝑖 𝑧
= 𝐹𝑖 ( ) (2.29)
𝑓𝑖,𝑀𝐿 𝑧𝑖
𝑧
𝑓𝑖,𝑀𝐿 is a reference scale, a combination of 𝑤∗ and 𝜃∗ , depending on the dimension of 𝑓𝑖 . is
𝑧𝑖
𝑧
a scaled height. 𝐹𝑖 (𝑧 ) is a function specifying the vertical profile of 𝑓𝑖 .
𝑖

Deardorff [67] showed by LES that the Deardorff scaling is applicable when the ratio
𝑧
− 𝐿𝑖 , characterizing the stability level of the whole boundary layer, is of order 4.5 or more and

the reduced scale experiment of Willis and Deardorff [68] using a convection chamber without
mean flow confirmed the validity of Deardorff scaling by 𝑧𝑖 , 𝑤∗ and 𝜃∗ . The field data obtained
in a CBL by Lenschow et al. [69] using an aircraft showed that the second moments of vertical
velocity fluctuation and the temperature fluctuation were fitted to the functions:
2
̅̅̅̅̅̅〉
〈𝑤′𝑤′ 𝑧 3 𝑧 2 (2.30)
= 1.8 ( ) (1 − 0.8 )
𝑤∗ 2 𝑧𝑖 𝑧𝑖
2
̅̅̅̅̅〉
〈𝜃′𝜃′ 𝑧 −3
= 1.8 ( ) (2.31)
𝜃∗ 2 𝑧𝑖

24
By using a saline convection tank to model the CBL at an improved accuracy compared to the
convectional convection tank, Hibberd and Sawford [70] showed that the streamwise and
vertical velocities and temperature fluctuation statistics satisfied the Deardorff scaling and the
fitted functions for 𝑤′ and 𝜃′ are:
2
̅̅̅̅̅̅〉
〈𝑤′𝑤′ 𝑧 3 𝑧 2 (2.32)
= 1.2 ( ) (1 − 0.9 )
𝑤∗ 2 𝑧𝑖 𝑧𝑖
2
̅̅̅̅̅〉
〈𝜃′𝜃′ 𝑧 −3 𝑧 2 𝑧
= 2.3 ( ) (1 − ) , 𝑤ℎ𝑒𝑛 < 0.4 (2.33)
2 𝑧 𝑧 𝑧
𝜃∗ 𝑖 𝑖 𝑖
However, the variation of CBL field data is very large, as compared to that of SL. There
are two basic reasons: (i) the time and length scales of CBL motions are much larger than the
SL one and comparable to the scales of synoptic and diurnal variations; and (ii) the ML flow
properties are more susceptible to a number of processes that do not account for by Deardorff
scaling, e.g. entrainment from free atmosphere, baroclinity and advection. Therefore, the
profile functions under Deardorff scaling are too simple and are far from universal, unless the
measurements are conducted under controlled experiments or the parameters of other
secondary processes are further classified. For CUBL, Zilitinkevich et al. [71] suggested that
the strength of large convective eddies also depends on the urban roughness, characterized by
the roughness length 𝑧0 , given the same 𝑤∗ and their suggested mechanism is through the
enhanced turbulent drag due to mechanical shear induced by roughness similar to the enhanced
drag on mean bulk flow by roughness under neutral condition. The dependence of the strength
of large convective eddies on the roughness is contradictory to the traditional view that the
outer layer flow is independent of roughness. More studies are needed to clarify this view.

2.3 Mean Profiles in ISL and Roughness Parameterization

The vertical profiles of mean velocity and temperature are perhaps the most important
achievement of UBL or ABL researches. In order to simulate the larger scale atmospheric flows,
the smaller scale flows in the near-surface region must be modelled. The conventional approach
is the logarithmic law of the wall in ISL or SL under neutral or near-neutral condition, together
with the parameters characterizing the aerodynamic properties of wall (or roughness) (Yaglom
[72]). This approach, supplemented by the buoyancy effects, is reviewed in this section.

25
2.3.1 Derivation of mean profiles

The profiles of mean velocity 𝑈(𝑧) and temperature 𝛩(𝑧) in ISL under neutral or
stratified conditions can be derived by mathematically integrating the mean gradient 𝜙𝑚 (𝜁)
and 𝜙ℎ (𝜁) once their functional forms are known. The procedures of analytical integration of
some popular forms of 𝜙𝑚 (𝜁) and 𝜙ℎ (𝜁) were demonstrated by Kramm et al. [7]. The key
steps of the derivation are briefly reviewed here. By integrating 𝜙𝑚 (𝜁) and 𝜙ℎ (𝜁) in the
vertical direction within ISL:
𝑈(𝑧) 1 𝑧 (2.34)
= [ln ( ) − Ψ𝑚 (𝜁) + Ψ𝑚 (𝜁0 )]
𝑢𝜏 𝜅 𝑧0
𝛩(𝑧) − 𝜃𝑤 𝑃𝑟𝑡 𝑧 (2.35)
= [ln ( ) − Ψℎ (𝜁) + Ψℎ (𝜁ℎ )]
𝜃𝜏 𝜅 𝑧ℎ
𝑧0 and 𝑧ℎ are the aerodynamic and temperature roughness lengths, respectively, and they
appears here as integration constants. 𝜃𝑤 is the wall surface temperature, but its definition is
not strict forward and requires further comments (together with the discussion of 𝑧ℎ in a later
subsection). Ψ𝑚 (𝜁) and Ψℎ (𝜁) are the stability correction functions for mean velocity and
temperature profiles, defined as:
𝜁
1 − 𝜙𝑖 (𝜁) (2.36)
Ψ𝑖 (𝜁) = ∫ 𝑑𝜁 , 𝑖 = 𝑚 𝑜𝑟 ℎ
0 𝜁
𝑧0 𝑧ℎ
And 𝜁0 = and 𝜁ℎ = are the stability parameters concerning the dynamical effect of
𝐿 𝐿

roughness relative to buoyancy effect. Within ISL, 𝑧 is usually much larger than 𝑧0 and 𝑧ℎ
(equivalently 𝜁 ≫ 𝜁0,ℎ ) and so Ψ𝑚 (𝜁0 ) and Ψℎ (𝜁ℎ ) are sometimes neglected in many literature.
The mean profiles are reduced to the log-law equations in neutral condition, in which
Ψ𝑖 (𝜁) = 0 since 𝜁 = 0 and 𝜙𝑖 (𝜁) = 0 . Under unstable condition, Ψ𝑖 (𝜁) can be derived
analytically if the functional form of 𝜙𝑖 (𝜁) contains elementary functions and is sufficiently
simple. For example, Paulson [73] demonstrated the integrated form of the BD-relations:
−2 −1 2
1 + 𝜙𝑚 1 + 𝜙𝑚 𝜋 (2.37)
Ψ𝑚 (𝜁) = 𝑙𝑛 [( )( ) ] − 2𝑡𝑎𝑛−1 (𝜙𝑚 −1 )
+
2 2 2
1 + 𝜙ℎ−1 (2.38)
Ψℎ (𝜁) = 2𝑙𝑛 ( )
2
where 𝜙𝑚 and 𝜙ℎ are given by equations (2.21) and (2.22). The integrated stability correction
function from the functional forms of 𝜙𝑚 and 𝜙ℎ proposed by Wilson [27], equation (2.23),
are:
1 + √1 + 𝛾𝑖 |𝜁|2/3 (2.39)
Ψ𝑖 (𝜁) = 3𝑙𝑛 ( ), 𝑖 = 𝑚 𝑜𝑟 ℎ
2

26
In situation where roughness size is not negligible compared to the local height 𝑧, the
effective height 𝑧𝑒𝑓𝑓 = 𝑧 − 𝑑 replaces 𝑧 in the above mean profile equations and the local
𝑧𝑒𝑓𝑓 𝑧−𝑑
stability is replaced by 𝜁 = = . These mean profile equations and the associated
𝐿 𝐿

parameters then provide a simple way to describe the mean flow properties near surface and
constitute an essential component of surface parameterization for the computation of larger
scale models.
For practical applications, the mean profiles Equations (2.34) and (2.35) are usually
reformulated into bulk transfer coefficients (e.g. demonstrated by Kanda and Moriizumi [74]):

𝜏0 = 𝜌𝑢𝜏2 = 𝜌𝐶𝑑 [𝑈(𝑧)]2 (2.40)

𝑞0 = −𝜌𝑐𝑝 𝑢𝜏 𝜃𝜏 = 𝜌𝑐𝑝 𝐶𝑇 [𝑈(𝑧)][𝜃𝑤 − 𝛩(𝑧)] (2.41)


and the transfer coefficients (𝐶𝑑 : drag coefficient and 𝐶𝑇 : temperature transfer coefficient) can
be expressed in terms the stability correction functions Ψ𝑖 and roughness lengths 𝑧0 and 𝑧ℎ :
𝑢𝜏2 2
𝑧 2
(2.42)
𝐶𝑑 = = 𝜅 ⁄[ln ( ) − Ψ𝑚 (𝜁) + Ψ 𝑚 (𝜁0 )]
[𝑈(𝑧)]2 𝑧0
𝑢𝜏 𝜃𝜏
𝐶𝑇 =
[𝑈(𝑧)][𝛩(𝑧) − 𝜃𝑤 ] (2.43)
𝑧 𝑧 2
= 𝜅 2 ⁄{𝑃𝑟𝑡 [ln ( ) − Ψ𝑚 (𝜁) + Ψ𝑚 (𝜁0 )] [ln ( ) − Ψℎ (𝜁) + Ψℎ (𝜁ℎ )]}
𝑧0 𝑧ℎ
However, the bulk coefficients are difficult for detailed analysis of momentum and
heat transfers. It is more revealing to analyze the results in the forms of mean profiles Equations
(2.34) and (2.35) so that the effects of the stability correction functions Ψ𝑖 and the roughness
lengths 𝑧0 and 𝑧ℎ can be decoupled and they are investigated separately.

2.3.2 Parameterization of roughness properties on mean wind

Aerodynamic roughness length 𝑧0 and displacement height 𝑑 (or zero-plane


displacement) are traditional parameters to describe the nature of surface roughness. For
general roughness types. In addition to the roughness configurations, 𝑧0 and 𝑑 may depend on
the flow Reynolds number. Three flow types are traditionally classified (Nikuradse [75]): (i)
dynamically smooth (ℎ+ < 5) in which viscous effect is most important, (ii) transitional (5 <
ℎ+ < 70) in which both viscous and roughness effects are important and (iii) fully rough (ℎ+ >
70) in which the viscous effect is negligible to the above bulk flow (i.e. Reynolds number
ℎ𝑢𝜏
independent), while ℎ+ = is the mean roughness height in wall units (introduced in Chapter
𝜈

3). 𝑧0 of different roughness types has long been intensively studied by field measurements and

27
experiments. Raupach et al. [12] have reviewed and discussed the parameterizations of 𝑧0 for
laboratory and vegetation roughness in these three flow regimes. However, the displacement
height 𝑑 is usually ignored because the position taken by most field measurements and
applications are far away from the roughness while the size of roughness is usually small as in
the case of vegetative and engineering roughness, such that 𝑧 ≫ ℎ . In real urban area, the
roughness elements, i.e. buildings, are usually non-negligible in size and thus the inclusion of
𝑑 is important for accurate flow description, especially at the lower portion of ISL. In addition,
the Reynolds number is practically very high and the fully rough flow regime is normally
satisfied. However, when studying rough wall flows using reduced scale experiment or wall
resolved numerical simulation, the Reynolds number effect has to be checked with care. Uehara
et al. [76] used wind-tunnel to study how the viscosity affects the flow in the vicinity of urban
𝑧𝑢𝜏
roughness and suggested both the dimensionless height 𝑧 + = and roughness Reynolds
𝜈
𝑧𝑜 𝑢𝜏
number 𝑅𝑒 ∗ = are important to characterize the viscous effect. Snyder and Castro [77]
𝜈

suggested that 𝑅𝑒 ∗ > 1 in order for 𝑧𝑜 of shape-edged obstacles with significant separation
between obstacles to be Reynolds number independent, less stringent than the conventional
requirement that 𝑅𝑒 ∗ > 2 ~ 5. Studies normally suggest that form drag on roughness elements
dominates when Reynolds number is high enough while the viscous stress is negligible.
Conventionally, for urban type roughness, 𝑧0 and 𝑑 are considered as properties of
physical geometrical configurations, such as the size, shape, distribution and orientation of
roughness elements. The approaches to parameterize 𝑧0 and 𝑑 include the direct measurement,
land-use methods and morphometric methods (Britter and Hanna [78]). Land-use methods are
rough estimations based on the descriptive land use type. This method is highly subjective and
not quite accurate.
In direct measurement method, both from fields or experiments, the mean velocity
profile is measured at different heights within ISL. The parameters 𝑧0 , 𝑑 and friction velocity
𝑢𝜏 , and sometimes the von Karman constant 𝜅, are then calculated by data regression based on
an assumed mean profile equation, usually the log-law under neutral condition. A problem of
this approach is that there will be uncertainty on the parameters originated from the data scatter
or deviation from the assumed profile (Bauer et al. [79]). The errors of the parameters to be
calculated from the measured data are interrelated, which means the uncertainty on one
parameter will pass to the others, known as artificial correlation, as demonstrated by Andreas
et al. [80]. The more the parameters are regressed from the measured data, the larger the error
will be. Therefore, it advantageous to reduce the number of parameters to be calculated by

28
explicitly finding some of the parameters through other different ways. For example, the
friction velocity 𝑢𝜏 can be explicitly measured from Reynolds stress measurement (e.g. the
eddy covariance method) or surface drag measurement (e.g. drag plate). Also, Jackson [81]
proposed by considering vegetation type roughness that the displacement height 𝑑 can be
regarded as the level at which the mean drag on the surface appears to act. However, Cheng
and Castro [49] concluded from his wind tunnel studies that it is likely that Jackson’s method
does not work for the urban type roughness. Other studies including Coceal et al. [82], Leonardi
and Castro [83] also could not support the validity of Jackson’s method. Nevertheless, as there
are still no other method to estimate the displacement height 𝑑, the objective estimation from
Jackson is still widely used. In addition, the value of von Karman constant 𝜅 is usually assumed
rather than calculated from data. As discussed by Hogstrom [23], the value of 𝜅 in ABL is
independent of the effects of roughness and the planetary rotation and close to the value of 0.40.
In contrast, Nagib and Chauhan [84] suggests that 𝜅 is different for different types of flow (e.g.
𝜅 = 0.37 for channel flow; 𝜅 = 0.41 for pipe flow; 𝜅 = 0.384 for zero-pressure boundary
layer). Nevertheless, the uncertainly of 𝜅 normally leads to less impact on the mean velocity
profile estimation than from the other parameters.
Using data to calculate 𝑧0 and 𝑑 may be unwieldy in many sites. The morphometric
method, which relates 𝑧0 and 𝑑 to the geometrical configurations of roughness, may be used
instead. The primary factor on 𝑧0 and 𝑑 is the building height to street width ratio (aspect ratio),
ℎ/𝑏, ℎ is the building height and 𝑏 is the street width. For example, the DNS study of Leonardi
et al. [42] finds that the maximum form drag (i.e. maximum 𝑧0 ) occurs when the aspect ratio
is 1/7. The explanation to this is given by Oke [85], who classified the flow around roughness
elements into three regimes: (i) isolated roughness flow (ℎ/𝑏 < 0.3); (ii) wake interference
flow (0.3 < ℎ/𝑏 < 0.7) ; and (iii) skimming flow (ℎ/𝑏 > 0.7) . In isolated roughness flow
regime, the roughness elements are far apart so that the flow over them reattaches on the ground
surface before separating from the ground again. In skimming flow regime, the flow skims atop
the roughness while one or multiple flow recirculation(s) is formed inside the cavity. The
behavior of wake interference flow regime is amid the other two, where smaller recirculation(s)
appears at the lower cavity. It is believed that the strongest drag in isolated roughness flow
regime is due to the direct flow impingement on the roughness frontal surface that causes the
flow diverges around the roughness. For engineering roughness, a similar classification is made
by Perry et al. [86]: (i) k-type roughness corresponding to isolated roughness flow and (ii) d-
type roughness corresponding to skimming flow. However, a single geometric ratio like aspect

29
ratio is usually insufficient to parameterize 𝑧0 and 𝑑. Some propose that a pair of geometric
ratios is better to describe the surface, namely, 𝜆𝑝 = 𝐴𝑝 /𝐴𝑇 and 𝜆𝑓 = 𝐴𝑓 /𝐴𝑇 , where 𝐴𝑝 is the
total building plan area, 𝐴𝑓 is the total building frontal area and 𝐴𝑇 is the total lot area.
Grimmond and Oke [87] reviewed the formulas proposed by different researchers and
conducted a sensitivity analysis on these formulas. They found that those morphometric
estimates did not agree satisfactorily to the even highest-quality measurements. They
concluded that none of the descriptors used by these formulas were able to capture fully the
geometry appropriate for aerodynamic parameterizations. Nonetheless, they suggested that the
methods proposed by Bottema [88] and Raupach [89] were probably the best among those
methods they reviewed. Britter and Hanna [78] suggested that 𝜆𝑓 is more important to drag as
it represents the surface directly facing the wind. As such, Hanna and Britter [90] have proposed
the following simple formulas for simple estimation using only 𝜆𝑓 as parameter:
𝑧0 𝜆𝑓 𝑖𝑓 𝜆𝑓 < 0.15 (2.44)
={
ℎ 0.15 𝑖𝑓 𝜆𝑓 > 0.15
3𝜆𝑓 𝑖𝑓 𝜆𝑓 < 0.05
𝑑 (2.45)
= {0.15 + 5.5(𝜆𝑓 − 0.05) 𝑖𝑓 0.05 < 𝜆𝑓 < 0.15

0.7 + 0.35(𝜆𝑓 − 0.15) 𝑖𝑓 0.15 < 𝜆𝑓 < 1.0
The above morphometric method actually only considers the distribution. Other roughness
factors like building shape, building height variability and orientation with respect to wind
direction are also important when describing random urban roughness. There is still a long way
to achieve a rather complete parameterization of the morphometric factors to surface wind drag.
Also, deriving an accurate morphometric formula requires a high quality experimental data.
The geometric factors are intensively investigated by researchers, but the thermal
stratification effects are usually ignored, though there were few evidences in the literature as
reviewed by Zilitinkevich et al. [8]. Zilitinkevich et al. suggested that buoyancy effect is an
important factor to modify 𝑧0 and 𝑑 and proposed parameterization formulae based on eddy
viscosity theory and field measurement data. They hypothesized the relevant physical
mechanism is due to the development of convective updrafts and downdrafts developed

between warm and cold roughness elements. They suggested that the stability ratio 𝐿 , based on

roughness height, is a relevant parameter controlling 𝑧0 and 𝑑. However, their parameterization


is on the vegetative type roughness. Due to the much large physical dimension of urban

roughness elements compared to vegetation ones and thus potentially larger 𝐿 , the dependence

of stability effect can be significant. However, a systematic study concerning stratification


effect on urban roughness is still lacking.

30
2.3.3 Parameterization of roughness properties on mean temperature

Unlike aerodynamic roughness length 𝑧0 , the roughness length of temperature 𝑧ℎ is


much more difficult to be found through measurement and forming a parameterization model
is also difficult. There are several reasons for the difficulty.
Firstly, 𝑧ℎ is very sensitive to how the wall temperature 𝜃𝑤 in Equation (2.35) is defined,
as discussed by Kanda et al. [6]. Since the wall temperatures of building surfaces in a typical
urban area almost vary substantially from one to the other one and the temperature is usually
non-uniform over a building façade, which make a single uniform value of 𝜃𝑤 impossible. To
overcome the problem, other representative wall temperatures were defined for practicality to
replace 𝜃𝑤 , e.g. area averaged wall temperature, radiative surface temperature, screen
temperature, etc (see Kanda et al. [6] for the review). However, different definition of 𝜃𝑤 can
lead to quite a different value of 𝑧ℎ . This can be explained by the resistance law of heat transfer
from building walls to the outer space driven by the temperature difference across any pairs of
two points. Very different from the mechanism of momentum transfer that the pressure force
on the frontal walls of bluff-body dominates the momentum transfer, there is no such analogues
pressure transfer mechanism for heat transfer on bluff-body obstacles. The resulting heat
transfer thus ultimately relies on molecular diffusivity at the immediate vicinity of building
surfaces (Brutsaert [91] and Brutsaert [92]). This gives a much higher thermal resistance
compared to the flow-drag resistance on bluff obstacles. If the path for heat transfer is divided
into series of segments, it can be shown that the majority of thermal resistance is from the
segment nearest to the surface on which the molecular diffusion dominates. As a result, since
the roughness length of temperature 𝑧ℎ is actually a measure of thermal resistance of heat
transfer from the whole urban surface to the bulk flow above (larger 𝑧ℎ means lower resistance),
𝑧ℎ strongly depends on the exact location at which the temperature is taken as the wall
temperature 𝜃𝑤 in Equation (2.35) or how the temperature of different surfaces are averaged
if the temperature is non-uniform throughout the urban surface.
Secondly, 𝑧ℎ is Reynolds number dependent even if the Reynolds number is very high
as in real urban area, in contrary to the aerodynamic roughness length 𝑧0 that is Reynolds
number independent in fully rough regime. This can be similarly explained by the nature of
heat transfer by molecular diffusion on the building wall surfaces as discussed above. For
example, Brutsaert [93] suggested an empirical relationship of the ratio 𝑧0 /𝑧ℎ and the
roughness Reynolds number 𝑅𝑒 ∗ for rough surfaces:

31
𝑧0
ln ( ) = 𝑎(𝑅𝑒 ∗ )0.25 − 2.0 (2.46)
𝑧ℎ
where 𝑎 is an empirical constant. Brutsaert [93] suggested that 𝑎 = 2.46 while Kanda et al. [6]
showed that 𝑎 = 1.29 based on their large scale urban roughness experiments and suggested
that Equation (2.46) is a robust model for urban roughness.
Even if the definition of 𝜃𝑤 , the point(s) of measurement and the Reynolds number
effect are confirmed, a further difficulty is caused by the variation of surface temperature on
the surface in most situation due to the uneven heating of solar radiation, uneven composition
of building materials, anthropogenic heat and transpiration effect of plants and trees, etc. As a
result, many measurement points on surface may be needed in field measurements or a network
of a high number of heat flow paths may be taken into account to in numerical modelling.
Finally, similar to 𝑧0 , the buoyant stability effect on 𝑧ℎ is still not known. It is also
valuable to systematically study its effect based on idealized urban setting like the present LES
study.

2.4 Interaction of Near Wall Region and Outer Layer

The validity of MOST in ISL or SL requires critically that the turbulence is “local”,
which means neither the smaller scales like roughness properties nor the larger scales like
boundary layer depth 𝑧𝑖 interact with the local turbulence. As discussed in Section 2.2, given
that there is a sufficient separation of scale between the roughness length scale and 𝑧𝑖 , the
influence of roughness only confines within RSL and most studies suggest that the flow above
RSL is independent of any roughness length scales, which is the “wall similarity” hypothesis
proposed by Townsend [94]. In addition, Townsend suggested that the near-wall and outer layer
motions do not interact with each other, although the outer layer motions are present in the near
wall region. However, MOST is challenged by some studies in recent years that the near-wall
turbulent motions responsible for vertical momentum and heat transports shows certain degree
of dependence on 𝑧𝑖 . In this section, the reasoning of Townsend’s hypothesis and evidences on
the violation of the assumptions of MOST and Townsend’s hypothesis are reviewed.

2.4.1 Townsend’s Hypothesis

As a complementary theory of MOST, Townsend’s hypothesis (Townsend [3]) states

32
that the turbulent motions in the near wall region, which is also the inner layer of the boundary
layer (i.e. SL or combination of ISL and RSL), can be decomposed into active parts and inactive
parts. The active parts are the local scale motions that are responsible for vertical transports of
momentum, thermal energy and other scalars. The inactive parts are the much larger turbulent
motions of the scale of boundary layer depth 𝑧𝑖 that only behave like sloshing motions locally
and do not participate in vertical transports. Townsend’s hypothesis states that these two types
of turbulence do not interact with each other. Interpreting the mechanism by turbulence kinetic
energy (TKE) budget, Townsend proposed that, in the inner layer, the local rates of TKE
production and dissipation are so large that aspects of the turbulent motions concerned with
these processes are independent of conditions elsewhere in the flow. Therefore, Townsend’s
hypothesis forms an explanation of the central assumption of MOST that the statistical flow
quantities at the near wall region (inner layer) do not depend on any the outer layer parameter
𝑧𝑖 . If Townsend’s hypothesis is wrong such that the inner layer flow quantities depend on 𝑧𝑖 ,
then MOST will fail when it is applied on those quantities.
Townsend [94] proposed that three conditions to be satisfied in order for Townsend’s
hypothesis to be valid. The first condition is the local equilibrium of TKE budget, which states
that the TKE generation is balanced by the dissipation while the advection and transport terms
of TKE are negligible. This implies the turbulent processes are locally dominated such that the
turbulence is negligibly influenced by the larger scale motions. The second condition is the
layer that Townsend’s hypothesis is applied must be thin compared to the boundary layer depth
𝑧𝑖 . The third condition requires the variation of Reynolds stress within the layer must be small.
Observations show that the second and the third conditions are basically satisfied by SL or ISL
while the first condition “local equilibrium” requires further discussion.
As another complementary theory of MOST, Townsend [94] proposed the “attached-
eddy hypothesis”, which states that the turbulence in the inner layer is a superposition of
geometrically similar eddies that scales with the distance of the eddy centers from the wall (i.e.
attached to the wall). It is the local scale attached eddies, i.e. scaled with the local height 𝑧, that
contain the strongest cross-flow transporting motions while the eddies of scale larger than 𝑧
cannot contain vertical velocity due to the blocking of underlying surface and thus carries no
Reynolds stress (Jimenez [95]).This provides a justification of MOST that the local height z is
the most important parameter to characterize the near-wall flow statistics. It was at a later time
after the introduction of “attached-eddy hypothesis” that the “attached-eddies” were shown to
be hairpin-shaped vortices by Head and Bandyopadhyay [96], though the hairpin vortex had

33
been postulated much earlier by Theodorsen [97] (reviewed in the later section).

2.4.2 Evidences on violation of MOST and Townsend’s hypothesis

It has long been known that the second moments of horizontal velocity fluctuation in
SL is scaled by 𝑧𝑖 rather than the local height 𝑧 (Kaimal et al. [98] and Kaimal [99]), which
clearly violates the assumption of MOST and therefore MOST cannot be applied on the
horizontal velocity fluctuation. Kaimal [99] showed from their spectra analysis that it is the
low-frequency motions that are scaled on 𝑧𝑖 and dominated the second moments of horizontal
velocity fluctuation. According to Townsend’s hypothesis, these low-frequency motions refer
to the inactive parts (Bradshaw [100]) and represent the large eddies spanning the whole
boundary layer. Fortunately, the vertical velocity fluctuation and the resulting Reynolds stress
in SL seem to satisfy the assumptions of MOST. In contrast to the horizontal velocity
fluctuations, the vertical one shows that it is dominated by the motions scaled by 𝑧 while its
spectral energy decays at lower frequency. This suggests that the vertical velocity fluctuation
in SL represents the active part and is responsible for the production of Reynolds stress. The
spectral measurement of Kaimal et al. [98] also suggested that the temperature fluctuation in
SL is scaled by 𝑧 rather than 𝑧𝑖 .
However, the above reasoning is based on the premise that the 𝑧𝑖 -scaled inactive
motions do not interact with the local 𝑧-scaled active motions. If these two scale motions indeed
interact with each other, then MOST and Townsend’s hypothesis have to be refined. In recent
decades, due to the advancement of technology, including measurement techniques and
numerical computation, more and more evidences on the interaction of active and inactive
turbulences emerge. For example, the field measurement of Wilson [101] found that the neutral
limit of vertical velocity fluctuation deviates markedly from the generally accepted standard
value and they then concluded that the Monin-Obukhov constant is in fact not “universal” and
the MOST should never be regarded as “perfect”. The LES study by Khanna and Brasseur [102]
showed that the temperature field (i.e. mean temperature gradient and temperature fluctuation)
was MO-similar while the velocity field (i.e. mean streamwise velocity gradient and vertical
velocity fluctuation) was indirectly influenced by the 𝑧𝑖 -scaled eddies. The conclusions of
Khanna and Brasseur [102] were then supported by the field measurement of Johansson et al.
[34]. Considering the turbulent processes, Smedman et al. [9], McNaughton et al. [10] and
Laubach and McNaughton [11] provided evidences of how the outer layer turbulence

34
influenced the SL flow and temperature field. McNaughton [103] showed that the peak
positions of 𝑧𝑖 -cospectra reflect the dominance of large-scale wedge-like structures and the
positions of these peaks are insensitive to local instability. They collectively suggested a “top-
down mechanism”, which is a non-local process, that the outer layer eddies (not in attached to
the ground), which are resulted from the Richardson cascade of boundary layer scale eddies,
was carried by the larger scale eddies impinging onto the near surface. This phenomenon is not
consistent with the assumption of MOST that 𝑧𝑖 has no role. The dependence on 𝑧𝑖 , which
signifies the turbulent processes actually originated from the outer layer, is thus also a potential
factor of the scattering of field measurement data when scaled by MOST. The current
understanding of the interaction between the outer layer turbulence and the SL flow is
inadequate. A few publications concerning the inner-outer interaction include McNaughton
[103], McNaughton [104] and Laubach and McNaughton [11] for unstable atmospheric
boundary layer, Hunt and Morrison [105], Hunt and Carlotti [106], Hogstrom et al. [107],
Drobinski et al. [108] and Drobinski et al. [109] for near-neutral atmospheric boundary layer,
as well as Castillo et al. [110] and Inagaki et al. [111] for convective boundary layer over urban-
like roughness. Further studies are needed to investigate the physical processes in details so as
to parameterize those processes.

2.5 Mechanism of Turbulent Processes

To understand how the flow statistics behave and to further improve the parametrization
models, the investigation of mechanisms of turbulent processes including the turbulence
generation, dissipation and turbulent transports of fluid properties is inevitable. However, due
to the intractability of Navier-Stokes equations, the mechanism of turbulent processes can only
be studied through the flow realizations from observations and numerical simulations. The
study of mechanism requires more complete flow patterns captured in order for any insightful
arguments to be deduced. This is the reasons that the study of turbulent mechanisms is difficult
by field measurements while LES and DNS show advantages in this aspect. The studies of
turbulent mechanisms based on turbulence energetics (TKE budget) and patterns (coherent
structures) are reviewed here.

35
2.5.1 Turbulence kinetic energy (TKE) budget

TKE budget analysis is in fact a single-point ensemble statistical analysis, but it reveals
how and where the turbulent motions are generated, dissipated and transported by different
medium, including mean flow advection, viscous diffusion, turbulent diffusion and pressure
transport (Wyngaard [17]). Besides understanding the mechanisms of turbulent processes, the
study of TKE budget can also provide improved modelling when the TKE transport equation
is employed to represent the properties of turbulence for a flow, such as 𝑘 − 𝜀 model (Pope
[8]). The study of TKE budget can be divided based on the scale under study. The scales that
this study concerns include the roughness scale (a.k.a. microscale) and the local scale (i.e. ISL
and RSL over rough surface). The derivations of TKE budgets are showed in Appendix A.2.
There have already been many studies of TKE budget conducted on SL or ISL and these
studies treat each term of the TKE budget as a one-dimensional (vertical) profile due to
horizontally homogenous nature of SL. Classical theories like MOST require that the local
production by buoyancy and shear is balanced by the local dissipation (Wyngaard and Cote
[112]) and this implies the sum of transport terms by turbulent diffusion and pressure must be
zero. However, the review of field measurement data by Hogstrom [23] suggested that the sum
of transport terms is not zero even under neutral condition. Hogstrom found that the local
dissipation is about 25% larger than local the local production while the extra energy dissipated
is transported from above by pressure transport term, which is due to the inactive turbulence,
hypothesized by Hogstrom. Very different from Hogstrom[23], the field measurement of
Frenzen and Vogel [113] in ISL over vegetative roughness found that the local dissipation was
10 to 15% smaller than the total TKE production. However, due to the difficulty of measuring
pressure, the pressure transport term calculated by Hogstrom [23] and Frenzen and Vogel [113]
are treated as a remaining term of the budget equation fluctuation in field measurement, which
may introduce significant error to the calculation. In this aspect, numerical simulation offers
an advantage that it calculates a rather complete field of pressure. Under unstable condition,
Hogstrom suggested that the pressure transport term and the turbulent transport term become
increasingly important, but the local balance in unstable condition has not been sufficiently
explored until now.
Moving towards the urban roughness from the ISL, the one-dimensional approach of
TKE budget applicable in ISL has to be reformulated due to the horizontal inhomogeneity in
RSL. In this case, the horizontal average of TKE budget equation (Appendix A.3) proposed by

36
Wilson and Shaw [114] can be used here. Two new terms arise in the horizontal averaged TKE
budget equation: (i) wake production (or dispersive shear production) and (ii) dispersive
transport, which are due to the horizontal variation of mean velocity and Reynolds stress and
were reviewed by Raupach et al. [12] for neutral condition, but for vegetative roughness only.
The urban study by Christen et al. [115] found that the wake production was important inside
street cavity and at roof-level and it then decreased with height while TKE generated at roof-
level was transported down to the street cavity through pressure transport.
In order to gain a deeper understanding of the mechanism on the turbulent processes
around roughness elements, it is more insightful to investigate the spatial variation of different
terms of the original TKE budget equation without horizontal averaging. However, this type of
study is rare because it is difficult to plan a large array of measuring probes around a street
cavity. Numerical simulation here again offers an advantage to study the spatial variation. For
example, the direct numerical simulation (DNS) of flow over two-dimensional rib roughness
by Ashrafian and Andersson [116] showed that the local shear production attains its maximum
at the roughness cavity roof-level opening and it decreases at the upstream corner of roughness
crest. Unfortunately, there is almost no detailed study of the effect of buoyant instability.

2.5.2 Coherent structures

Turbulence was considered a stochastic phenomenon in the past. With the


advancement of research technology and technique, it is nowadays recognized that turbulent
fluctuations are not entirely random and they possess certain organizing properties. The
repetitive patterns of these organized motions are defined as coherent structures. Specific types
of coherent structure are repeatedly found at specific locations and some of them contribute
dominantly the turbulent transports of momentum, heat and other scalars (Robinson [15]).
Therefore, it is believed that understanding the coherent structures, including their topology,
how they are generated, dissipated, maintained and relocated and how coherent structures of
different types interact, would lead us to better understand the mechanism of turbulent
transports and thus promote more accurate turbulent flow modelling. For the research progress
of coherent structures of wall turbulence, Robinson [15] split its history into four eras (until the
year of publish): (i) discovery era (1932-1959); (ii) flow-visualization era (1960-1971); (iii)
conditional-sampling era (1972-1979) and (iv) computer simulation era (1980-1991). The trend
shows that computer simulation today is still a main research tool to investigate the coherent

37
structures, due to its rather complete simulated flow field that is very difficult or costly to obtain
from experiments.
Many experiments and numerical simulations have shown different forms of coherent
structures in high Reynolds number smooth wall turbulent flow, such as hairpin vortices, low-
speed streaks, streamwise vortices, ejections and sweeps (Kline et al. [117], Head and
Bandyopadhyay [96] and Monty et al. [118]) through different techniques such as visualization,
conditional averaging and filtering. Hairpin vortex, first proposed by Theodorsen [97], is now
widely recognized as the primary coherent structure responsible for turbulent transports of
momentum and scalars in the SL. It is found that hairpin vortices are generated by low-speed
streaks and streamwise vortices that form alternatingly in spanwise direction near the surface
and hairpin vortices evolve self-similarly in wall-normal direction (Jimenez [95]). Later,
Adrian [119] proposed that hairpin vortices tend to form collectively in streamwise direction
as coherent packets, which manifest themselves as low-momentum region, whose evidences
were given by Adrian et al. [120] and Tomkins and Adrian [121]. These features are also found
in the SL of neutral atmospheric boundary layer and Hutchins et al. [122] therefore proposed
to reconcile these structures in both atmospheric and laboratory flows.
With the well-established understanding of smooth wall coherent structures, there are
increasing demands to understand coherent structures of rough wall flows and how the SL or
ISL interact with these roughness induced structures. For example, Grass [123] pointed out the
sweeps (or inrushes) and ejections from roughness cavities are enhanced compared to those
over smooth wall. Both the study of open channel flow over glass-bead roughness by
Nakagawa and Nezu [124] and the study over field vegetation by Finnigan [125] showed that
sweep are dominant over ejection in RSL. By conditionally averaging the scalar and velocity
vector fields over a mixed deciduous forest, Gao et al. [126] showed that a strong sweep upwind
of scalar microfront and a weaker ejection ahead of mircofront are the typical structures in RSL.
For flow over cubical roughness, Coceal et al. [82] found that the coherent structures in ISL
are similar to those of smooth wall and atmospheric SL which supports the attached-eddy
hypothesis and outer-wall-layer (or wall) similarity. Within RSL, Coceal et al. found some
similarities of structures with those over vegetative roughness and they suggested this is due to
the similar mixing-layer nature caused by the inflectional velocity profile at the roughness
interface, which shares certain degree of similarity with vegetative roughness flows. The study
of coherent structure above a vegetation canopy by Finnigan et al. [127] showed that the
turbulence in RSL is more coherent and more efficient at transporting momentum and scalars.
They showed by composite averaging that the characteristic eddy consists of an upstream head-
38
down sweep-generating hairpin superimposed on a downstream head-up ejection-generating
hairpin.
Under unstable condition, large-scale motions induced by buoyancy are developed
over the entire boundary layer even the instability level is weak. Different types of large-scale
organized motions in ABL are reviewed by Young et al. [128]. The dominant motions in slightly
unstable condition with significant mean wind are the large-scale horizontal roll vortices, which
were simulated in LES by Moeng and Sullivan [129], while the ones in very strong or pure
convective boundary layer (CBL) are the cellular patterns similar to those in Rayleigh-Benard
convection, which were simulated in LES by Schmidt and Schumann [130]. These large-scale
structures are scaled by the boundary layer depth 𝑧𝑖 and are regarded as the inactive motions
according to Townsend’s hypothesis.
Unfortunately, the understanding of how instability modifies the near-wall coherent
structures is still insufficient and the related study is rare. Hommema and Adrian [131] showed
from their smoke visualization in unstable ABL that the orientation of hairpin vortices is lifted
from the wall compared to neutral condition. By studying the dissimilarity between the
turbulent transports of momentum and scalars in a neutral and an unstable ABLs, Li and Bou-
Zeid [132] found that the momentum and scalar transports become decorrelated by the presence
of thermal plumes in unstable condition. The LES study of flow over building array with heated
bottom by Park and Baik [133] also showed the plume-shaped structures, which cause stronger
ejections and more vertically expanded hairpin structures compared to no-heating case. These
plume-shaped structures are regarded as the active motions that are induced by local buoyancy
force, directly modify the local turbulent transports and are MOST like.
Recently, Laubach and McNaughton [11] challenged MOST and Townsend’s
hypothesis by proposing a top-down mechanism that active and inactive motions interact. In
their conceptual model of turbulent transports, the near wall region of CBL is separated into
two layers. The bottom layer, which is named shear SL by them, is dominated by shear-induced
attached hairpin like vortices, which are originally driven by the large outer eddies like
horizontal roll vortices or convective cells and grow self-similarly in wall-normal direction,
similar to the hairpin vortices of high-Reynolds number smooth wall flows. The upper layer on
the other hand is dominated by the impinging outer eddies, which are originally generated by
the Richardson cascade of the outer large eddies successively into smaller eddies, active in
local turbulent transports and scaled with the local height 𝑧 . Laubach and McNaughton
suggested that eddies of different kinds but similar size cannot exist in the same location at the
same instant. As a result, the shear-induced eddies are confined within the bottom shear SL
39
while the penetration of the upper impinging Richardson eddies to the shear SL is limited. The
characteristic of this upper layer is consistent to the conventional “local free convection layer”
under the reign of MOST that the turbulent statistics for active motions are scaled by 𝑧 and are
independent of shear (i.e. friction velocity 𝑢𝜏 ). Both conceptual models agree with the presence
of plume-like structures. However, there is one important distinction concerning the origin of
these plume-like structures: the “local free convection hypothesis” assumes the plume-like
motions are generated by local buoyancy force while the top-down mechanism by Laubach and
McNaughton [11] states that the plume-like motions are the consequence of the Richardson
cascade from large-scale outer eddies. This may explain why the profile forms of MOST
statistics are well scaled by the local stability parameter 𝑧/𝐿 (e.g. the power functional form of
vertical velocity fluctuation) but the multiplying constants are far from “universal”. This may
imply the multiply constants are indirectly governed by the nature of outer eddies, i.e.
modifying the outer eddies could modify the values of multiplying constants. They also
proposed that this mechanism leads to the dissimilarity between momentum and scalar
transports.
Apart from Laubach and McNaughton [11], Hogstrom et al. [107] and Hunt and
Morrison [105] also proposed a top-down mechanism that the outer layer motions in a high
Reynolds number neutral flow could interact with the near-wall motions: detached outer layer
eddies generated by shear away from wall, which do not present in low Reynolds number flows,
impinge to the near-wall region and strengthen the near-wall transports. This conceptual model
explains the abnormality of MOST statistics in the neutral or near-neutral ABL SL (Smedman
et al. [9] and Drobinski et al. [109]).
The above illustrates the importance of coherent structure study towards the
understanding of transport mechanism and the advantages of studying by numerical simulation.
The next step is to investigate how the buoyancy modifies or modulates the near-wall motions
if the underlying surface is physically rough.

40
Chapter 3

Numerical Model Configuration

A series of hypothetical convective urban boundary layers (CUBL) bounded atop by a


stable interfacial layer are simulated by LES. The urban surface is modeled by idealized two-
dimensional urban roughness elements. The wind above urban roughness is driven by a
background pressure gradient. The convective condition due to the heating of urban surface is
modeled by a constant and uniform temperature boundary condition. Subsequent analyses are
performed on the quasi-steady state where the flow statistics are sufficiently invariant with time.
This chapter describes the configurations of the LES models, including (i) the domain
construction, (ii) the fundamental transport equations with boundary conditions, (iii) large-
eddy simulation modeling, (iv) meshing of domain and numerical methods and (v) processing
and post-processing of models. The collections of LES models at various configurations,
including computational domain and boundary conditions, used in this study for different
purposes are summarized. Finally, the validation of the LES models is excised by comparing
the simulated flow statistics with those of a wind tunnel experiment.

3.1 Computational Domain

The domain (Figure 3.1) consists of a free-stream region and a number of street cavity
cavities separated by building elements of square cross-sectional shape and equal dimensions
at equal spacing. The whole domain is two-dimensional and therefore the building elements
are square ribs in geometry. The domain is composed of a number of periodic roughness units,
with each unit containing a street cavity and a vertical column above the cavity (Figure 3.2).
Each canyon cavity is bounded by a leeward wall, a windward wall and a ground surface. The
wind is driven by the background pressure gradient above the urban roughness in the

41
Figure 3.1: Computational domain of the standard case at roughness aspect ratio 1 (case A1).

streamwise direction, perpendicular to the street cavity axis, which is defined as the spanwise
direction. The Cartesian coordinate x-, y-, z-axes are defined as the streamwise, spanwise and
vertical directions, respectively. The ground surface of canyon cavity is defined as 𝑧 = 0 while
the coordinates 𝑥 and 𝑦 are described as relative positions depending on the type of analysis,
e.g. 𝑥 may be defined relative to domain inlet or the cavity center plane of certain periodic
roughness unit. The height of building elements ℎ , which is also the building width in x-
direction, is used as a reference length in describing the geometry and dimensions of the domain.
The height of the free-stream region 𝐻, extending from the building roof level to the domain
top, is set to 𝐻 = 7ℎ and is unchanged for all the cases. The overall domain height is thus 𝐿𝑧 =
𝐻 + ℎ = 8ℎ . The blockage ratio (domain height to building height) is 𝐿𝑧 /ℎ = 8 . The
roughness cavity width 𝑏 is varied to simulate different urban roughness conditions. Three
aspect ratios (building height to cavity width), 𝐴𝑅 = ℎ/𝑏 = 0.5, 1 and 2, are included in the
study and the corresponding models are denoted by A0.5, A1 and A2 with corresponding cavity
widths 𝑏 = 2ℎ, ℎ and 0.5ℎ, respectively. In addition, a special case of 𝐴𝑅 = 0.5 with all the
roughness dimensions halved (i.e. building height: 0.5ℎ, cavity width: ℎ and roof length 0.5ℎ)
is included to test the effect of roughness height on the flow above roughness and is denoted

42
Figure 3.2: A periodic roughness unit at roughness aspect ratio 1 (case A1).

by AH0.5. Cases of different domain streamwise lengths 𝐿𝑥 (from domain inlet to outlet) and
spanwise 𝐿𝑦 lengths are simulated. As a result, the number of periodic roughness unit depends
on the domain streamwise length. The case with domain streamwise length 𝐿𝑥 = 24ℎ (i.e. 12
periodic roughness units corresponding to the model A1) and spanwise length of 𝐿𝑦 = 5ℎ is
regarded as the standard one, in which the buoyancy effect on the flow statistics and structures
are investigated intensively. Additional cases with variant domain configurations are also
simulated for other specific analyses. A summary of all cases simulated is given in the last
section of this chapter.

3.2 Fundamental Equations and Boundary Conditions

In this section, the fundamental physics of the transport phenomena being simulated are
described. The flow velocity, pressure, temperature are aimed to be solved from the governing
equations and the boundary conditions. To enable efficient data processing and analysis, the
problem is idealized and simplified by some assumptions.
Dry air with ideal gas property and no composition change is considered in this study. The

43
effect due to planetary rotation is neglected. The fluid is assumed to be inert and thus no
chemical reaction is involved. The heating by radiation is modeled by boundary conditions
while the radiation effect on the fluid volume is ignored. Newtonian fluid is assumed. The
heating due to viscous dissipation is neglected. To avoid solving a variable fluid density field,
Boussinesq approximation is applied such that the governing equations are simplified to the
incompressible form, except there is an additional buoyancy force term, a linear function of
temperature, in the momentum equation. The justification of Boussinesq approximation for the
atmospheric boundary layer flow will be given after stating the adopted transport equations.

3.2.1 Basic transport equations

The equations governing flow and heat transfer based on the above assumptions include
the continuity equation, the momentum equation, and the energy equation (collectively termed
the Navier-Stoke equations), which are the mathematical expressions of the conservation
principles of mass, linear momentum and energy. The continuity equation is:
𝜕𝑢𝑖
=0 (3.1)
𝜕𝑥𝑖
The momentum equation is:
𝜕𝑢𝑖 𝜕𝑢𝑖 𝜕𝑝 𝜕 2 𝑢𝑖
+ 𝑢𝑗 =− +𝜈 + 𝛼𝑔(𝜃 − 𝜃0 )𝛿𝑖3 + 𝜆Π𝛿𝑖1 (3.2)
𝜕𝑡 𝜕𝑥𝑗 𝜕𝑥𝑖 𝜕𝑥𝑗 𝜕𝑥𝑗
The energy equation is:
𝜕𝜃 𝜕𝜃 𝜕 2𝜃
+ 𝑢𝑗 = 𝐷𝑇 (3.3)
𝜕𝑡 𝜕𝑥𝑗 𝜕𝑥𝑗 𝜕𝑥𝑗
The index notation, in which the suffix 𝑖 or 𝑗 = 1, 2, 3 represents coordinate axes 𝑥, 𝑦, 𝑧 , is
used. The summation convention, in which repeated suffix of a term denotes the summation
over three values of index 𝑖 = 1, 2, 3, is used as well. The index appeared only once within a
term is a free index, which denotes a vector component.
Cartesian coordinate system, in which 𝑥𝑖 and 𝑡 are coordinates and time, is used. 𝛿𝑖𝑗 is
Kronecker delta. The field variables 𝑢𝑖 ( 𝑢1 = 𝑢, 𝑢2 = 𝑣, 𝑢3 = 𝑤 ), 𝑝 and 𝜃 are velocity
component, kinematic pressure and potential temperature, respectively. 𝑔 is gravitational
acceleration. The fluid properties 𝜈 , 𝛼 and 𝐷𝑇 are kinematic viscosity, thermal expansion
coefficient at constant pressure and thermal diffusivity, respectively, which are assumed to be
constants. 𝜃0 is a reference temperature. Π is the background pressure gradient force that drives

44
the flow field (𝜆 = 0 inside all cavity cavities and 𝜆 = 1 above building roof level). The precise
natures of kinematic pressure 𝑝, potential temperature 𝜃, thermal expansivity 𝛼 and reference
temperature 𝜃0 will be given amidst the discussion of Boussinesq approximation.

3.2.2 Boussinesq approximation

In Boussinesq approximation, the fluid density has no effect on the flow except the
buoyancy force due to thermal expansion of fluid. It is commonly utilized to study the
atmospheric boundary layer problem in which the thermal convection is of interest. For
Boussinesq approximation to be accurate, the overall spatial and temporal variations of
temperature should be sufficiently small, as commonly satisfied by the atmospheric boundary
layer flows. It is widely known that the buoyancy effect is mainly due to the deviation of
temperature from the hydrostatic state, rather than the pressure counterpart. The effect of
density variation due to the deviation of pressure from the hydrostatic state, manifested in the
phenomenon of sound wave, is thus negligible. Boussinesq assumption enables the original
governing equations in compressible form to be reduced to the incompressible form. The
incompressible form of equations eliminates the need of solving the density field, reduces the
coupling of field properties and the non-linearity of equations. These help to reduce the time
of numerical iterations and the likelihood of solution divergence.
The derivation of Boussinesq approximation requires a hydrostatic background state,
dependent only on the elevation 𝑧, to be defined first. For an atmospheric boundary layer, it is
convenient to define the background state to be neutrally stratified, so that a fluid particle is at
the same density with the environment when it travels vertically under isentropic (frictionless
and adiabatic) expansion or compression process. This hydrostatic state is described by the
static equation:
𝑑𝑃𝑠
= −𝜌𝑠 𝑔 (3.4)
𝑑𝑧
and the condition derived from the thermodynamic property relation of isentropic process:
𝑑ℎ̂𝑠 𝑑𝑇𝑠 1 𝑑𝑃𝑠
= 𝑐𝑝 = (3.5)
𝑑𝑧 𝑑𝑧 𝜌𝑠 𝑑𝑧
in which the ideal gas properties 𝑃 = 𝜌𝑅0 𝑇 and 𝑑ℎ̂ = 𝑐𝑝 𝑑𝑇 are used. 𝜌, 𝑃, 𝑇, ℎ̂, 𝑐𝑝 and 𝑅0 are
fluid density, thermodynamic pressure, thermodynamic temperature, specific enthalpy, specific
heat capacity at constant pressure, specific gas constant, respectively. The subscript 𝑠 denotes

45
the hydrostatic state. The hydrostatic temperature gradient is then deduced as:
𝑑𝑇𝑠 𝑔
=− (3.6)
𝑑𝑧 𝑐𝑝
which is the adiabatic lapse rate. The hydrostatic temperature profile is thus a linear function:
𝑔
𝑇𝑠 (𝑧) = − (𝑧 − 𝑧𝑟𝑒𝑓 ) + 𝑇0 (3.7)
𝑐𝑝
where the subscript 0 denotes a reference state at a selected reference height 𝑧𝑟𝑒𝑓 , which can
be conveniently defined as the mid height of boundary layer. The hydrostatic pressure and
temperature can also be deduced similarly.
For fluid in motion, the thermodynamic properties can then be written in terms of
deviations, 𝛿𝑇, 𝛿𝑃, 𝛿𝜌, from the hydrostatic state:
𝑇(𝑥𝑖 , 𝑡) = 𝑇𝑠 (𝑧) + 𝛿𝑇(𝑥𝑖 , 𝑡)
𝑃(𝑥𝑖 , 𝑡) = 𝑃𝑠 (𝑧) + 𝛿𝑃(𝑥𝑖 , 𝑡) (3.8)
𝜌(𝑥𝑖 , 𝑡) = 𝜌𝑠 (𝑧) + 𝛿𝜌(𝑥𝑖 , 𝑡)
In the derivation of transport equations using Boussinesq approximation, the fluid density is
linearized about the reference state:
𝜌 = 𝜌0 [1 − 𝛼(𝑇 − 𝑇0 ) + 𝛼𝑃 (𝑃 − 𝑃0 )] (3.9)
1 𝜕𝜌 1
𝛼 = − 𝜌 (𝜕𝑇) = 𝑇 is the thermal expansion coefficient at constant pressure and 𝛼𝑃 =
𝑃
1 𝜕𝜌 1
( ) = 𝑃 is the isothermal compressibility. They are then approximated by neglecting the
𝜌 𝜕𝑃 𝑇
1 1
temperature and pressure dependences: 𝛼 = 𝑇 and 𝛼𝑃 = 𝑃 . Subsequently, the dynamic
0 0

viscosity 𝜇, the specific heat capacity at constant pressure 𝑐𝑝 , the thermal conductivity 𝑘𝑇 are
approximated as constants. The kinematic viscosity and the thermal conductivity are
𝜇 𝑘𝑇
approximated as well: 𝜈 = and 𝐷𝑇 = . To further simplifying the modeling equations,
𝜌0 𝜌0 𝑐𝑝

the temperature is replaced by the potential temperature defined by:


𝑃0 𝑅⁄𝑐𝑝
𝜃 = 𝑇( ) (3.10)
𝑃
which can be interpreted as the temperature of a fluid parcel when brought to the reference
state at pressure 𝑃0 under an isentropic process. The potential temperature at the hydrostatic
state is a constant by definition: 𝜃𝑠 = 𝜃0 = 𝑇0 . The buoyancy force per unit mass can be
approximated by neglecting the pressure deviation from the hydrostatic state:
𝛿𝜌 𝛿𝜌
− 𝑔 ≈ − 𝑔 ≈ 𝛼𝛿𝑇𝑔 ≈ 𝛼𝛿𝜃𝑔 = 𝛼(𝜃 − 𝜃0 )𝑔 (3.11)
𝜌 𝜌0

46
By neglecting the unimportant terms, the transport equations (3.1) - (3.3) are then
deduced (see chapter 8 of Wyngaard [17] for the procedure of Boussinesq approximation). In
(3.2), the kinematic pressure 𝑝 is related to the thermodynamic pressure 𝑃 by:
𝛿𝑃 1
𝑝≈ = (𝑃 − 𝑃𝑠 ) (3.12)
𝜌0 𝜌0
The approximation requires two dimensionless numbers to be small:
∆𝑇 ∆𝑃
𝜀1 = 𝛼∆𝑇 ≈ ≪ 1, 𝜀2 = 𝛼𝑃 ∆𝑃 ≈ ≪1 (3.13)
𝑇0 𝑃0
where ∆𝑇 and ∆𝑃 are the overall temporal and spatial temperature and pressure variations
throughout the domain. Assuming 𝑇0 = 285𝐾 and 𝑃0 = 100𝑘𝑃𝑎 , 𝜀1 < 0.1 implies ∆𝑇 <
28.5𝐾 , which is normally satisfied by the atmospheric boundary. ∆𝑃 is dominated by the
hydrostatic pressure rather than the dynamic pressure due to fluid motion. Therefore,
∆𝑃~𝜌0 𝑔𝛿~10𝑘𝑃𝑎 , where 𝛿 is the boundary layer height and 𝛿~1𝑘𝑚 for both neutral and
unstable atmospheric boundary layers. As a result, 𝜀2 ~0.1 and is small. Similar statements on
the requirements were also discussed extensively in the literature (e.g. Spiegel and Veronis [134]
and Ogura and Phillips [135]).
In the following text, the kinematic pressure 𝑝 and the potential temperature 𝜃 are
simply termed pressure and temperature, respectively.

3.2.3 Boundary and initial conditions

The boundary conditions for velocity and temperature are shown in Figure 3.3. The
urban surface is modeled by applying no-slip boundary conditions (𝑢 = 𝑣 = 𝑤 = 0) on all its
surfaces (ground, roof, leeward and windward walls). The stable interfacial layer over the urban
𝜕𝑢 𝜕𝑣
boundary layer is modeled by a free-slip boundary condition ( 𝜕𝑧 = 𝜕𝑦 = 𝑤 = 0) at the domain

top. Symmetry boundary condition is applied to pressure at the urban surface and the domain
𝜕𝑝
top such that the gradient normal to its boundary is zero (𝜕𝑛 = 0, n is the coordinate normal to

the corresponding surface). The thermal stratification is modeled by applying different constant
Δ𝜃
temperatures at the bottom no-slip surface (𝜃 = 𝜃𝑊 = 𝜃0 + ) and the domain top (𝜃 = 𝜃𝑇 =
2
Δ𝜃
𝜃0 − ). All the flow field variables 𝑝, 𝑢𝑖 and 𝜃 are periodic between the domain inlet and
2

outlet as well as between the two spanwise walls.


For the initial condition (𝑡 = 0), a uniform velocity field is set above the roof level (𝑢 =

47
Figure 3.3: Boundary conditions for velocity and temperature.

𝑈0 , 𝑣 = 𝑤 = 0 ) while a zero velocity field is set inside all canyon cavities. A uniform
temperature field (𝜃 = 𝜃0 ) and a zero pressure field are set throughout the whole domain.

3.2.4 Quasi-steady state and wall units

The effects of the initial conditions will die out gradually as the time 𝑡 advances and
the analyses in the subsequent sections are on the quasi-steady state, under which the ensemble
averaged statistics are invariant of time 𝑡. The overall urban surface shear stress 𝜏0 at the quasi-
steady state can be derived in advance. Considering the whole computational domain at the
quasi-steady state and in ensemble average, the only forces that remain are the surface stress
and the background pressure gradient force as there are no force or net momentum flow at the
domain top, the inlet and the outlet. Based on force and momentum balance, the relation
between the surface stress and the background pressure gradient can be written as:
𝜏0
𝜏0 = 𝜌Π𝐻 𝑜𝑟 Π = (3.14)
𝜌𝐻
Friction velocity 𝑢𝜏 and viscous length scale 𝑙𝜈 can then be defined based on the surface stress
𝜏0 and the kinematic velocity 𝜈 as:
48
1
𝜏0 2 1 𝜈
𝑢𝜏 = ( ) = (Π𝐻)2 , 𝑙𝜈 = (3.15)
𝜌 𝑢𝜏
Analogously, friction temperature 𝑢𝜏 can be defined based on the surface heat flux 𝑞0 as:
𝑞0
𝜃𝜏 = − (3.16)
𝜌𝑐𝑝 𝑢𝜏
𝜃𝜏 is negative when the surface heat flux is upwards (𝑞0 > 0) and vice versa. However, unlike
the surface stress, the surface heat flux at the quasi-steady state can only be known through the
numerical simulation
The wall units, denoted by superscript +, are then defined as the dimensionless velocity,
temperature and length scales using 𝑢𝜏 , 𝜃𝜏 and 𝑙𝜈 , respectively:
𝑢𝑖 (𝜃 − 𝜃𝑊 ) 𝑥𝑖
𝑢𝑖 + = , 𝜃+ = , 𝑥𝑖 + = (3.17)
𝑢𝜏 𝜃𝜏 𝑙𝜈
which are measures of the wall shear effect.

3.2.5 Non-dimensionalization

The numerical simulation is actually implemented on the dimensionless form of the transport
equations (3.1) - (3.3). For convenience, the initial streamwise velocity 𝑈0 , the overall potential
temperature difference ∆𝜃, the roughness element height ℎ are chosen as the reference scales
of velocity, temperature and length, respectively. Subsequently, 𝑢𝑖 , 𝑥𝑖 , 𝑡, 𝑝 and (𝜃 − 𝜃0 ) are
normalized by 𝑈0 , ℎ, ℎ/𝑈0 , 𝑈02 and ∆𝜃 respectively. The governing equations (3.1) – (3.3) are
then transformed to:
𝜕𝑢̂𝑖
=0 (3.18)
𝜕𝑥̂𝑖
𝜕𝑢̂𝑖 𝜕𝑢̂𝑖 𝜕𝑝̂ 1 𝜕 2 𝑢̂𝑖
+ 𝑢̂𝑗 =− + + 𝑅𝑖0 𝜃̂𝛿𝑖3 + 𝜆Π
̂ 𝛿𝑖1 (3.19)
𝜕𝑡̂ 𝜕𝑥̂𝑗 𝜕𝑥̂𝑖 𝑅𝑒0 𝜕𝑥̂𝑗 𝜕𝑥̂𝑗
𝜕𝜃̂ 𝜕𝜃̂ 1 𝜕 2 𝜃̂
+ 𝑢̂𝑗 = (3.20)
𝜕𝑡̂ 𝜕𝑥̂𝑗 𝑃𝑟 ∙ 𝑅𝑒0 𝜕𝑥̂𝑗 𝜕𝑥̂𝑗
where 𝑢̂𝑖 , 𝑥̂𝑖 , 𝑡̂ , 𝑝̂ and 𝜃̂ are now dimensionless variables and 𝜃̂ here means a temperature
𝑈0 ℎ 𝛼𝑔ℎΔ𝜃
difference, replacing ( 𝜃 − 𝜃0 ). 𝑅𝑒0 = is a Reynolds number, 𝑅𝑖0 = − is a
𝜈 𝑈0 2
𝜈
Richardson number and 𝑃𝑟 = 𝐷 is a Prandtl number. The background pressure gradient
𝑇

Πℎ
̂ = 2 . For all the simulated cases, 𝑅𝑒0 = 104 , 𝑃𝑟 = 0.7,
driving force Π is normalized to Π 𝑈 0

̂ = 5 × 10−4 while 𝑅𝑖0 is varied to simulate different levels of buoyant convection.


Π
49
The boundary condition and initial condition are normalized accordingly. The
normalized temperature is 𝜃̂ = 𝜃̂ ̂ ̂
𝑊 = 1/2 at the urban surface and 𝜃 = 𝜃𝑇 = −1/2 at the

domain top. At 𝑡̂ = 0, the normalized velocity above roof level is 𝑢̂ = 1, 𝑣̂ = 𝑤


̂ = 0 and the
normalized temperature is 𝜃̂ = 0.

3.3 Large-eddy Simulation (LES) Modeling

To avoid solving motions of exhaustive range of scales, the governing equations (3.1)
– (3.3)are filtered such that only the larger scale motions are explicitly solved while the smaller
scale motions, which are more isotropic and less influenced by the boundary conditions, are
modeled. In LES, a flow variable 𝜙 is decomposed into a resolved scale component 𝜙 𝑟 ,
representing the larger scale resolved motion, and a sub-grid scale (SGS) (a.k.a. residual scale)
component 𝜙 𝑠 , representing the smaller scale unresolved motion:
𝜙 = 𝜙𝑟 + 𝜙 𝑠 (3.21)
The resolved scale component is obtained by filtering operation on the physical space with a
filter function 𝐺(𝑥𝑖′ − 𝑥𝑖 ; Δ𝑖 ):

𝜙 𝑟 (𝑥𝑖 , 𝑡) = ∫ 𝜙(𝑥𝑖′ , 𝑡)𝐺(𝑥𝑖′ − 𝑥𝑖 ; Δ𝑖 ) 𝑑𝑉′ (3.22)


𝑉

where Δ𝑖 is the filter width at each direction, 𝑉 is the whole domain volume and 𝑑𝑉 ′ =
𝑑𝑥1′ 𝑑𝑥2′ 𝑑𝑥3′ . In this study, box filter (or top-hat filter) is used and the filter widths are equal to
the mesh element widths, which will be discussed on the next subsection. Filtering by box filter
is an averaging process over the specified box volume and its filter function is defined as:
1 Δ𝑖
𝑖𝑓 |𝑥𝑖′ − 𝑥𝑖 | ≤
𝐺(𝑥𝑖′ )
− 𝑥𝑖 ; Δ𝑖 = {Δ1 Δ2 Δ3 2 (3.23)
0 𝑜𝑡ℎ𝑒𝑟𝑤𝑖𝑠𝑒
To illustrate the physical meanings, the filtering operation is first applied to the
dimensional governing equations (3.1) – (3.3). The filtered equations are:
𝜕𝑢𝑖𝑟
=0 (3.24)
𝜕𝑥𝑖
𝜕𝑢𝑖𝑟 𝜕𝑢𝑖𝑟 𝜕𝑝𝑟 𝜕 2 𝑢𝑖𝑟 𝜕𝜏𝑖𝑗
+ 𝑢𝑗𝑟 =− +𝜈 − + 𝛼𝑔(𝜃 𝑟 − 𝜃0 )𝛿𝑖3 + 𝜆Π𝛿𝑖1 (3.25)
𝜕𝑡 𝜕𝑥𝑗 𝜕𝑥𝑖 𝜕𝑥𝑗 𝜕𝑥𝑗 𝜕𝑥𝑗
𝜕𝜃 𝑟 𝜕𝜃 𝑟 𝜕 2𝜃𝑟 𝜕𝑞𝑗
+ 𝑢𝑗𝑟 = 𝐷𝑇 − (3.26)
𝜕𝑡 𝜕𝑥𝑗 𝜕𝑥𝑗 𝜕𝑥𝑗 𝜕𝑥𝑗

50
where 𝜏𝑖𝑗 and 𝑞𝑗 are defined as:
𝜏𝑖𝑗 = (𝑢𝑖 𝑢𝑗 )𝑟 − 𝑢𝑖𝑟 𝑢𝑖𝑟 (3.27)
𝑞𝑗 = (𝜃𝑢𝑗 )𝑟 − 𝜃 𝑟 𝑢𝑖𝑟 (3.28)
are two new terms due to the filtering operation. 𝜏𝑖𝑗 and 𝑞𝑗 are the SGS stress and SGS heat
flux, respectively, due to the SGS motions, analogues to the Reynolds stress and heat flux in
the Reynolds-averaged Navier-Stokes (RANS) equations. Although the filtering operation and
the partial differentiation are not strictly commutative due to the presence of domain boundaries
and non-uniform filter widths, the error decreases with the decrease of filter widths (see the
discussion of Ghosal and Moin [136] for the commutation error) and the commutativity is
approximated in the above derivation.
To model the SGS stress, it is decomposed in to a deviatoric part and an isotropic part
as:
2
𝑑
𝜏𝑖𝑗 = 𝜏𝑖𝑗 + 𝑘𝑠𝑔𝑠 𝛿𝑖𝑗 (3.29)
3
1
such that the trace of the deviatoric stress 𝜏𝑖𝑖𝑑 = 0. 𝑘𝑠𝑔𝑠 = 2 𝜏𝑖𝑖 is the SGS turbulent kinetic

energy (SGS-TKE). Eddy-viscosity hypothesis (a.k.a. Boussinesq hypothesis) is then


employed to model the SGS stress:

𝑑 𝜕𝑢𝑖𝑟 𝜕𝑢𝑗𝑟
𝜏𝑖𝑗 = −2𝜈𝑠𝑔𝑠 𝑆𝑖𝑗𝑟 , 𝑆𝑖𝑗𝑟 = ( + ) (3.30)
𝜕𝑥𝑗 𝜕𝑥𝑖
Similarly, the gradient-diffusion hypothesis is employed to model the SGS heat flux:
𝜈𝑠𝑔𝑠 𝜕𝜃 𝑟
𝑞𝑗 = − ( ) (3.31)
𝑃𝑟𝑠𝑔𝑠 𝜕𝑥𝑗
𝜈𝑠𝑔𝑠 is the SGS turbulent viscosity and 𝑃𝑟𝑠𝑔𝑠 is the SGS turbulent Prandtl number, which is the
ratio of the SGS turbulent viscosity to the SGS turbulent diffusivity of heat. As discussed by
Tominaga and Stathopoulos [137], the solution is quite insensitive to the value of 𝑃𝑟𝑠𝑔𝑠 since
most of the motions are resolved. In this study, 𝑃𝑟𝑠𝑔𝑠 is simply taken as 0.7, which is within the
range of values used by different researchers.
The next step is to model the SGS turbulent viscosity 𝜈𝑠𝑔𝑠 , which is usually formulated
as a product of a characteristic length scale and a characteristic velocity scale of SGS motions.
In the current study, the geometric average of filter (or mesh element) sizes Δ = (Δ1 Δ2 Δ3 )1/3
1/2
and the square root of SGS-TKE 𝑘𝑠𝑔𝑠 are respectively assigned as the length scale and the
velocity scale of SGS motions, such that:

51
1/2
𝜈𝑠𝑔𝑠 = 𝐶𝑘 𝑘𝑠𝑔𝑠 Δ (3.32)
where 𝐶𝑘 = 0.07 is a modeling constant.
The final closure is provided by the one-equation SGS-TKE model proposed by
Schumann [138], which solve 𝑘𝑠𝑔𝑠 from its transport equation:
3/2
𝜕𝑘𝑠𝑔𝑠 𝜕𝑘𝑠𝑔𝑠 𝜕 𝜕𝑘𝑠𝑔𝑠 𝑘𝑠𝑔𝑠
+ 𝑢𝑗𝑟 = 2𝜈𝑠𝑔𝑠 𝑆𝑖𝑗𝑟 𝑆𝑖𝑗𝑟 + [(𝜈 + 𝜈𝑠𝑔𝑠 ) ] − 𝐶𝜀 (3.33)
𝜕𝑡 𝜕𝑥𝑗 𝜕𝑥𝑗 𝜕𝑥𝑗 Δ
where 𝐶𝜀 = 1.05 is another modeling constant.
The one-equation model provides the history and transport effects of SGS motions to
the resolved field through the eddy-viscosity model. This is an advantage over the traditional
Smagorinsky model (Smagorinsky [139]), which assumes the SGS motion is in equilibrium at
dissipating its energy. By including the non-equilibrium effects, the one-equation model is
superior in modeling free shear layers, flow separation and reattachment, wall boundary layers
and buoyancy induced flows, which are all present in the current study. Therefore, the cost of
solving an additional transport equation is justified by the additional modeled effects being
included. More details of filtering operation, derivation of filtered Navier-Stokes equations and
SGS modeling can be found in the book of Pope [21].
To complete the problem specification, initial condition and boundary condition for
𝑘𝑠𝑔𝑠 are needed. 𝑘𝑠𝑔𝑠 is set to zero throughout the domain at 𝑡 = 0 . Symmetry boundary
𝜕𝑘𝑠𝑔𝑠
condition is applied at the urban surface and the domain top ( = 0). Similar to other field
𝜕𝑛

variables, periodic boundary conditions for 𝑘𝑠𝑔𝑠 are set between the domain inlet and outlet as
well as between the two spanwise walls. The filtered equations (3.24) – (3.26), the equation for
𝑘𝑠𝑔𝑠 (3.33) and the respective boundary and initial conditions can be non-dimensionalized,
similar to procedures from subsection 3.2.4.

3.4 Discretization and Numerical Methods

To numerically solve the problem, the above problem statements have to be transformed
into ones that are solvable using computer. This include (i) discretizing the internal domain into
mesh elements; (ii) discretizing different terms in the governing equations in the form of partial
differential equations and the boundary conditions into linear expressions in terms of dependent
variables at mesh element centroids and boundary face centroids; and (iii) decoupling pressure
and velocity from the momentum and continuity equations and constructing iterative

52
Figure 3.4: Mesh inside canyon cavity of model A1, a) VeryCoarse, b) Coarse, c) Fine.

algorithms for solutions.


Construction of mesh, compilation of solver that contains the algorithms and the
solution procedures and processing of iterative solution are performed using OpenFOAM [140],
an open source code for CFD, running in Linux platform. Finite Volume method is adopted to
discretize the governing equations by first integrating over the volume of a mesh element.
Therefore, the internal domain has to be divided into non-overlapping elements through
meshing before solution procedures.

3.4.1 Domain meshing

Structured mesh composed of rectangular brick elements is used. Three classes of mesh
depending on mesh element density are defined, namely, VeryCoarse (VC), Coarse (C) and
Fine (F). The mesh in the lower part of a periodic roughness unit of model A1 for each mesh
class is shown in Figure 3.4. Compared to mesh C, the mesh element length in mesh VC is
double in each direction, while the one in mesh F is halved. As a result, the numbers of mesh
elements are eight times less in mesh VC and eight times more in mesh F.

53
Canyon height Canyon width Roof length Domain spanwise Free-stream height
Model
(nz) (nx) (nx) length (ny) (nz)

A0.5 2h (32)

A1 h (16) h (16) h (16)


Ly = 5h (80) H = 7h (140)
A2 0.5h (12)

AH0.5 0.5h (12) h (24) 0.5h (12)

Table 3.1: Mesh configurations for single periodic roughness units in Coarse (C) mesh.

Mesh no. in canyon Mesh no. above roof Mesh no. No. of Total domain
Model
(nx×ny×nz) (nx×ny×nz) per unit canyons mesh no.

A0.5 32×80×16 48×80×140 578,560 8 4,628,480

A1 16×80×16 32×80×140 378,880 12 4,546,560

A2 12×80×16 28×80×140 328,960 16 5,263,360

AH0.5 24×80×12 36×80×140 426,240 16 6,819,840

Table 3.2: Mesh element number for the standard cases where 𝐿𝑥 = 24ℎ in Coarse (C) mesh.

For brevity, only the configuration of mesh C is described (and the other meshes can be
scaled easily). For the models A0.5, A1 and A2, the canyon cavity is of height ℎ and contains
16 mesh elements in the vertical (z-) direction. The widths of cavity 𝑏 in x-direction are 2ℎ, ℎ
and 0.5ℎ and the number of mesh elements in x-direction are 32, 16 and 12 for the models A0.5,
A1 and A2, respectively. For the standard case in which the domain spanwise length is 𝐿𝑦 =
5ℎ , the y-direction mesh is uniform and the number is 80 throughout the whole domain,
including both the cavities and the free-stream region. In the free-stream region above roof
level, the x-direction mesh directly above canyon cavity follows the inside cavity one while the
mesh directly above building roof, of the same length ℎ for all aspect ratios, contains 16 mesh
elements (i.e. 8 in half roof length) in x-direction. There are 140 mesh elements in the vertical
(z-) direction between roof level and domain top. The mesh configurations of single periodic
roughness unit in Coarse mesh (C) for models of different aspect ratios are listed in Table 3.1.
The mesh element numbers in a single cavity unit and in the whole domain for the standard
case (𝐿𝑥 = 5ℎ and 𝐿𝑦 = 5ℎ) are listed in Table 3.2. The tables also include the model AH0.5.
The mesh is stretched away from the solid wall so that the mesh is denser near the wall.
Within the length of mesh stretching (e.g. from leeward wall to cavity mid-plane), the first-last

54
Canyon Free-stream
𝚫𝒚
Model 𝚫𝒙,𝒎𝒊𝒏 , 𝚫𝒙,𝒎𝒂𝒙 𝚫𝒛,𝒎𝒊𝒏 , 𝚫𝒛,𝒎𝒂𝒙 𝚫𝒙,𝒎𝒊𝒏 , 𝚫𝒙,𝒎𝒂𝒙 𝚫𝒛,𝒎𝒊𝒏 , 𝚫𝒛,𝒎𝒂𝒙
(×10-2h)
(×10-2h) (×10-2h) (×10-2h) (×10-2h)

A0.5 43.2 , 86.4 43.1 , 86.4

A1 43.1 , 86.2 43.1 , 86.2 43.1 , 86.2


62.5 27.4, 82.3
A2 28.7 , 57.3 28.7 , 86.2

AH0.5 28.8 , 57.6 28.7 , 57.3 28.7 , 57.3

Table 3.3: Mesh element sizes in Coarse (C) mesh.

Canyon Free-stream
Model 𝚫+𝒚
𝚫+𝒙,𝒎𝒊𝒏 , 𝚫+𝒙,𝒎𝒂𝒙 𝚫+𝒛,𝒎𝒊𝒏 , 𝚫+𝒛,𝒎𝒂𝒙 𝚫+𝒙,𝒎𝒊𝒏 , 𝚫+𝒙,𝒎𝒂𝒙 𝚫+𝒛,𝒎𝒊𝒏 , 𝚫+𝒛,𝒎𝒂𝒙

A0.5 25.6 , 51.1 25.5 , 51.1

A1 25.5 , 51.0 25.5 , 51.0 25.5 , 51.0


37.0 16.2 , 48.7
A2 17.0 , 33.9 17.0 , 51.0

AH0.5 17.0 , 34.0 17.0 , 34.0 17.0 , 34.0

Table 3.4: Mesh element sizes in wall unit in Coarse (C) mesh.

ratio is 2 everywhere, except the one from roof level to domain top where the first-last ratio is
3. As a result, the mesh is non-uniform in x- and z-directions. The maximum and minimum
mesh element sizes in each direction relative to the reference length ℎ are listed in Table 3.3
and in terms of wall units in Table 3.4. As the flow at the vicinity of wall is intended to be
resolved, no wall model is used.

3.4.2 Numerical methods

Using Finite Volume method, the field variables at mesh element volume centroids are
solved and stored for later post-processing and analyses. The temporal derivative term is
discretized by the implicit second-order-accurate backward differencing scheme. The gradient,
the diffusion and the convection terms are discretized by the second order Gaussian integration,
which in turn requires interpolation of mesh element face centroids from the volume centroids
of neighboring mesh elements. The second order linear interpolation (central difference)
scheme is used here. In order to prevent the occurrence of unphysical oscillations from the

55
convection terms, total variation diminishing (TVD) scheme is used in conjunction with the
central difference scheme. The coupling of the momentum and continuity equations is handled
by the pressure-implicit with splitting of operators (PISO) scheme, in which one predictor step
and two corrector step are used to solve for pressure and velocity vector. The preconditioned
conjugate gradient (PCG) method with the diagonal incomplete-Cholesky (DIC) maxtrix
preconditioner is used to solve for the pressure iteratively. The preconditioned bi-conjugate
gradient (PBiCG) method with the diagonal incomplete-LU preconditioner is used to solve for
the velocity vector, temperature and SGS-TKE. The details of the numerical methods used in
this study and OpenFOAM can be found in the book of Moukalled et al. [141].

3.5 Processing and Post-processing of Numerical Models

With the aforementioned model setting and methods employed, the field data is
calculated numerically and the quasi-steady state is saved. Different averaging operations,
depending on the analysis, are then performed on the saved data to produce meaningful
statistics. These statistics are categorized into (i) ensemble averages, calculated by averaging
in the homogenous directions, including time and y-direction, (ii) spatial (or horizontal)
averages, calculated by additionally averaging the ensemble averages in x-direction, and (iii)
conditional averages, calculated with certain criteria specified by the analysis.

3.5.1 Processing of numerical models

To achieve the quasi-steady state, each numerical model is first processed in the
VeryCoarse (VC) mesh class for a time period of or 6000ℎ/𝑈0 ( ~350ℎ/𝑢𝜏 ) with an
incremental time step Δ𝑡 = 0.04ℎ/𝑈0 (= 2.37 × 10−3 ℎ/𝑢𝜏 , where ℎ/𝑢𝜏 (~16.9ℎ/𝑈0 ) is the
eddy turnover time, which is a characteristic time scale of the turbulence shed from the
roughness). From the result of simulation, the quasi-steady mean velocity in free-stream region
is at the order of 𝑈0 . For the standard cases with domain streamwise length 𝐿𝑥 = 24ℎ, the time
period of 6000ℎ/𝑈0 corresponding to fluid passing through 250 domains. For the models in
which the data of mesh VC is needed, they are processed for a further time period of 2500ℎ/𝑈0
(~150ℎ/𝑢𝜏 ) and the field data during this period is saved at a time interval of 1ℎ/𝑈0 (=
0.0592ℎ/𝑢𝜏 ) so that there are 2500 sets of field data for post-processing.

56
The field data from mesh VC is then mapped to the Coarse (C) mesh class and then
processed for a time period of 1000ℎ/𝑈0 (~60ℎ/𝑢𝜏 ) with an incremental time step Δ𝑡 =
0.02ℎ/𝑈0 (= 1.18 × 10−3 ℎ/𝑢𝜏 ). During this period, the field data is saved at a time interval
of 1ℎ/𝑈0 (= 0.0592ℎ/𝑢𝜏 ) so that there are 1000 sets of field data for post-processing. If the
field data of Fine (F) mesh class is needed. The field data from mesh C is mapped to the mesh
F and then processed for a time period of 500ℎ/𝑈0 (~30ℎ/𝑢𝜏 ) with an incremental time step
Δ𝑡 = 0.01ℎ/𝑈0 (= 5.92 × 10−4 ℎ/𝑢𝜏 ). During this period, the field data is saved at a time
interval of 1ℎ/𝑈0 (= 0.0592ℎ/𝑢𝜏 ) so that there are 500 sets of field data for post-processing.

3.5.2 Post-processing of field data

The ensemble averages of different statistical quantities are computed from the captured
raw data and is denoted by overbar ∙̅. The ensemble averaging is performed on temporal and
spanwise domains as it is assumed that the field data used to compute the statistics are
statistically invariant in time and the spanwise direction is homogenous. The ensemble
averaging is additionally performed for each cavity unit as each one is identical. The mean of
a field variable 𝜙 is calculated by:
1 1 𝑡=𝑡2 𝑦=𝐿𝑦
𝜙̅(𝑥, 𝑧) = ∙ ∫ ∫ 𝜙(𝑥, 𝑦, 𝑧, 𝑡)𝑑𝑦 𝑑𝑡 (3.34)
𝑡2 − 𝑡1 𝐿𝑦 𝑡=𝑡1 𝑦=0
where 𝑡1 and 𝑡2 are the start time and the end time of the field data used in statistical
computation. The deviation from the mean, denoted by a prime, is defined as:
𝜙 ′ (𝑥, 𝑦, 𝑧, 𝑡) = 𝜙(𝑥, 𝑦, 𝑧, 𝑡) − 𝜙̅(𝑥, 𝑧) (3.35)
The second moment (or variance) of the field variable 𝜙 is calculated by:

̅̅̅̅̅̅ 1 1 𝑡=𝑡2 𝑦=𝐿𝑦


𝜙′𝜙′(𝑥, 𝑧) = ∙ ∫ ∫ [𝜙′(𝑥, 𝑦, 𝑧, 𝑡)]2 𝑑𝑦 𝑑𝑡 (3.36)
𝑡2 − 𝑡1 𝐿𝑦 𝑡=𝑡1 𝑦=0
Other ensemble statistics are computed similarly, e.g. covariance, third moment and two-point
correlation.
In some analyses of flow above roughness, the spatial (or horizontal) averages are
investigated. The averaging operation of spatial average is additionally performed on the
streamwise domain and is denoted by angle bracket 〈•〉. The spatial mean of an ensemble
averaged variable 𝜙̅, e.g. 𝑢̅, ̅̅̅̅̅
𝑢′𝑢′ and ̅̅̅̅̅̅
𝑢′𝑤′, is calculated by:
1 𝑥=𝐿𝑥
〈𝜙̅〉(𝑧) = ∫ 𝜙̅(𝑥, 𝑧)𝑑𝑥 (3.37)
𝐿𝑥 𝑥=0

57
The deviation from the spatial mean, denoted by a double prime, is defined as:
𝜙̅"(𝑥, 𝑧) = 𝜙̅(𝑥, 𝑧) − 〈𝜙̅〉(𝑧) (3.38)
The second moment of the ensemble variable 𝜙̅ is calculated by:
1 𝑥=𝐿𝑥
〈𝜙̅"𝜙̅"〉(𝑧) = ∫ [𝜙̅"(𝑥, 𝑧)]2 𝑑𝑥 (3.39)
𝐿𝑥 𝑥=0
The conditional average used to investigate the turbulent coherent structures is defined by the
averaging criteria. It will be discussed in the later sections.

3.6 Summary of numerical models

Different model configurations are used in different analyses. These include different
roughness configurations, different mesh densities and different streamwise or spanwise
domain sizes. Abbreviation is defined to label each model with the corresponding specification.
As discussed in previous subsections, the roughness configuration is labeled by A0.5, A1, A2
and AH0.5 (half roughness dimension of A0.5) while the mesh configuration is labeled by VC,
C and F. The domain streamwise extents used include 2ℎ, 4ℎ, 6ℎ, 12ℎ, 24ℎ and 48ℎ, labeled
by X2, X4, X6, X12, X24 and X48, respectively. The domain spanwise extents used include
5ℎ , 10ℎ and 20ℎ , labeled by Y5, Y10 and Y20. The list of model configurations with the
corresponding labels is shown in Table 3.5. The buoyancy effects in convective boundary layer
are intensively investigated in mesh C with domain horizontal extent X24-Y5 at different
roughness configurations (e.g., the case of aspect ratio 1 at Coarse mesh with domain
streamwise extent 24h and spanwise extent is labeled as A1-X24-Y5-C). The sensitivity of the
flow statistics to the domain horizontal extent is studied at different streamwise and spanwise
extents at the ranges mentioned above using the mesh VC and C and aspect ratio A1. The
sensitivity of flow statistics to the mesh densities are studied at domain horizontal extent X6-
Y5 for all the mesh classes (i.e. VC, C and F) at aspect ratio A1.
On the other hand, to complete the model specification, the Reynolds number and the
level of unstable stratification (or buoyancy) have to be specified. Considering the model as an
open-channel flow (as the top boundary is free-slip rather than no-slip) over roughness, it is
𝑢𝜏 𝐿𝑧
more physical to employ the friction Reynolds number 𝑅𝑒𝜏 = = 4,730 instead of the
𝜈
𝑈0 ℎ
previously mentioned one 𝑅𝑒0 = = 104 to specify the relative inertial effect of the flow.
𝜈
𝛼𝑔𝐿𝑧 Δ𝜃
The level of unstable stratification is specified by the Richardson number 𝑅𝑖τ = − 𝑢𝜏 2

58
Domain streamwise Domain spanwise
Aspect ratio Mesh density
extent extent
X2 (𝐿𝑥 = 2ℎ)
A0.5 (𝐴𝑅 = 0.5)
X4 (𝐿𝑥 = 4ℎ)
A1 (𝐴𝑅 = 1) VC (VeryCoarse) Y5 (𝐿𝑦 = 5ℎ)
X6 (𝐿𝑥 = 6ℎ)
A2 (𝐴𝑅 = 2) C (Coarse) Y10 (𝐿𝑦 = 10ℎ)
X12 (𝐿𝑥 = 12ℎ)
AH0.5 (𝐴𝑅 = 0.5, F (Fine) Y20 (𝐿𝑦 = 20ℎ)
X24 (𝐿𝑥 = 24ℎ)
half roughness size)
X48 (𝐿𝑥 = 48ℎ)

Table 3.5: Model configuration list.

ranging from 0 to -457. To summarize, the flow is completely characterized by 𝑅𝑒𝜏 = 4,730,
−457 ≤ 𝑅𝑖τ ≤ 0 and 𝑃𝑟 = 𝑃𝑟𝑠𝑔𝑠 = 0.7 with the SGS model constants 𝐶𝑘 = 0.07 and 𝐶𝜀 =
1.05, which are all prescribed before the processing the numerical models. Alternatively, the
level of unstable stratification can be characterized using the Monin-Obukhov length 𝐿 =
𝑢 3
𝜏
− 𝜅𝛽𝑞 (𝜅 = 0.4 in this study) in terms of ratio of two length scales, a.k.a. stability parameter,
0

e.g. 𝐿𝑧 /𝐿, ℎ/𝐿 or 𝑧/𝐿 depending on the analysis and the physics under study. These ratios are
intensively utilized in this study since it is directly related to the Monin-Obukhov theory
described in Chapter 2. However, 𝐿 can only be known after the processing of LES model since
the calculation of 𝐿 requires the surface heat flux 𝑞0 , which can only be extracted from the field
data. From the results of numerical models, the instability ratio 𝐿𝑧 /𝐿 ranges from 0 to -9.59,
which forms a substitute of the Richardson number 𝑅𝑖τ .

3.7 Model Validation by Wind Tunnel Experiment

From the author’s knowledge, the wind tunnel experiment of Uehara et al. [53] is the
only one reduced scale experiment to model the flow over roughness with surface heating.
Therefore, the LES model is compared to their result for validation, though the modeling
conditions, especially the roughness and heating configurations, are not exactly the same. This
section focuses on the flow statistics above the roof level of urban roughness.
A brief description about experiment of Uehara et al. [53] is given here. Their
experiment setup and measuring positions are shown in Figure 3.5, referring to their paper for
the details. The working section of the wind tunnel is 2𝑚 high, 3𝑚 wide and 24𝑚 long. To
model the urban roughness in wind tunnel, cubes (of length ℎ = 100𝑚𝑚, made of Styrofoam)
are regularly aligned in streamwise and spanwise directions. The cubes are separated by

59
Figure 3.5: Wind tunnel setting by Uehara et al. [53]: a) Arrangement of roughness elements and
model city blocks; b) Measuring points and the model setting.

100𝑚𝑚 in streamwise direction and 50𝑚𝑚 in spanwise direction, so the aspect ratio is 1. A
fetch including a 100𝑚𝑚 high tripping fence and shorter roughness elements of total length of
8.8𝑚 (regularly aligned Styrofoam cubes of dimension 100𝑚𝑚 × 100𝑚𝑚 × 50𝑚𝑚
separated by 100𝑚𝑚 on all sides) is placed before the urban roughness to generate a turbulent
boundary layer. The wind tunnel ground is heated to generate unstable stratification and the
walls of Styrofoam roughness elements can be approximated as thermally insulated. A laser
Doppler anemometer (LDA) and a cold wire are used to measure the flow field and temperature
inside and outside the urban roughness. Five rows of roughness elements are placed at the
upstream of the measurement locations while six rows are at its downstream. Measurements
are conducted along a vertical line from the ground to 700𝑚𝑚 at the mid position of the street
cavity between the upstream and downstream walls of roughness elements.
As the urban roughness is 2D (square ribs) in LES models but the roughness is 3D

60
Figure 3.6: Normalized vertical profiles of a) mean streamwise velocity; b) mean temperature; c)
streamwise velocity fluctuations S.D. (standard deviation); d) vertical velocity fluctuations S.D.; e)
temperature fluctuations S.D.. Symbols: wind tunnel data from Uehara et al. [53]; lines: current LES data.
Black diamond: neutral; blue cross: ground heated (𝑅𝑖2ℎ = −0.155); black solid line: neutral; blue dotted
line: ground heated (𝑅𝑖2ℎ = −0.208); blue dashed line: all walls heated (𝑅𝑖2ℎ = −0.132).

(cubes) in the experiment, the flow statistics at the vicinity of roughness elements are somewhat
different. But the influence to the flow above roof level is expected to be small since the
spanwise gap (= ℎ/2) between cubes are narrow compared to the streamwise cavity spacing
(= ℎ). The boundary layer heights in LES and experiment are also different. In view of these
differences and for better data comparison, the mean velocity and temperature at 𝑧 = 2ℎ are
used as the reference variables for data normalization since this height is sufficiently above the
roughness sublayer and it is far below the boundary top so that the influence due to boundary
layer height is minimal. The stability is described by the Richardson number, defined as:

61
̅ − 𝜃𝑤 )
𝛼𝑔(2ℎ)(𝜃2ℎ
𝑅𝑖2ℎ = (3.40)
𝑢̅2ℎ 2
where 𝜃𝑤 is the ground temperature in experiment. The definition and the magnitudes of
Richardson number used here are thus different from those described in the paper of Uehara et
al. [53] that used the values at roughness top (𝑧 = ℎ) instead. A comparison with the experiment
is made to the LES model with aspect ratio 1 (model setting: A1-X24-Y5-C) with all bottom
surface heated as discussed in previous sections. In addition to the LES models with all surface
being heated, we additionally simulate a LES model of same configuration except only grounds
are heated but other walls are insulated (gradient of temperature perpendicular to wall is set
zero). The mean velocity, temperature and second moment statistics along the vertical line from
the ground to 7ℎ at the mid position of the street cavity are compared between LES models and
experiment under both neutral and unstable stratification with similar Richardson number 𝑅𝑖2ℎ ,
as shown in Figure 3.6. For the LES results presented here, only the resolved scale variables
are used in calculations and the statistics denoted by overbar are ensemble averaged in time
domain, in spanwise domain and for all cavity units, as described in sub-section 3.5.2. The

standard deviations of fluctuation in Figure 3.6 is defined as 𝜎𝑢 = √̅̅̅̅̅ ̅̅̅̅̅̅ and


𝑢′𝑢′, 𝜎𝑤 = √𝑤′𝑤′

𝜎𝜃 = √̅̅̅̅̅
𝜃′𝜃′.
The mean velocity (Figure 3.6a) in LES matches the wind tunnel one well inside and
above cavity for both neutral and unstable conditions, except near roof level where velocity
gradient is the greatest and near ground where the effect due to flow entrainment of street
intersection is large. Both show that the wind flow above roughness is more vertically uniform
above roughness when the ground is heated, which agrees with MOST. It is remarkable that
the difference in the mean flow between the two LES cases of ground heated and all walls
heated is slight and essentially the same below 𝑧 = 3ℎ. The mean temperature (Figure 3.6b) of
the LES model of ground heated case agrees well with the wind tunnel one above roof level.
This implies the influence of the dissimilar roughness types on the mean fields above roof level
is practically small. The velocity fluctuations 𝜎𝑢 and 𝜎𝑤 normalized by the reference velocity
(Figure 3.6c and d) above roughness in the LES cases are underestimated compared to the
experimental ones. The larger turbulence intensity above roughness in wind tunnel cases may
arise from the upstream condition (the tripping fence as well as a sudden change of roughness
height, from 50𝑚𝑚 to 100𝑚𝑚, at the first building elements). The turbulence intensity inside
cavity in LES models is also much less than that in experiment, which is due to the stronger
sheltering effect in 2D roughness and stronger turbulence generation by the entrainment from

62
street intersections. Both results show that the 𝜎𝑢 attains maximum at roof level and then
decreases with height. This implies there is certain similarity of turbulence generation
mechanism at the roughness top, which is hypothesized to be the plane mixing layer type
turbulence. The experiment does not show the increasing trend of 𝜎𝑤 from the roof level to 𝑧 =
3ℎ in the LES unstable cases but rather shows a relatively uniform profile. This may be due to
the much stronger turbulence around the 3D roughness elements. As described by MOST, the
vertical velocity fluctuation in unstable stratification should increase with height above RSL
and given that the roughness is much smaller than the boundary layer height. This condition
may not be satisfied in the experiment, which may be due to the 3D roughness nature and its
larger lockage ratio (i.e. roughness element is not sufficiently small compared to boundary
layer thickness) that leads to a thick RSL with no or insufficient thickness of ISL. Nevertheless,
both LES and experiment show that the turbulence intensity generally increases when the
ground is heated and is even stronger in the LES model of all surfaces being heated. Lastly,
both show that the temperature fluctuation (Figure 3.6e) is quite uniform with similar
magnitude when the ground is heated, though the peak at roof level is not present in the
experiment, which is again possibly due to the 3D roughness nature.

63
Chapter 4

Examination of Monin-Obukhov Similarity Theory


(MOST) by Large-eddy Simulation (LES)

In this chapter, the validity and the applicability of MOST are tested using LES on three
configurations of idealized urban-type roughness as introduced in Chapter 3. Three roughness
configurations specified by the aspect ratios, 0.5, 1 and 2 (namely, model A0.5, A1 and A2) are
studied. Due to a thicker RSL for the case of aspect ratio 0.5 that may mask the ISL, a halved
roughness size while keeping the same roughness configuration and the domain dimension
(model AH0.5) is also included so that the ISL is thicker for testing. The single-point statistics,
including gradients of mean velocity and temperature and second moments of vertical velocity
and temperature fluctuations, in the forms of MOST functions (𝜙𝑚 , 𝜙ℎ , 𝜙𝑤 and 𝜙𝜃 defined in
Section 2.1) are analyzed. A functional form of MOST functions of mean velocity and
temperature gradients, which is a pair of piecewise functions – linear function at near neutral
state and power function at strong unstable state, is proposed and tested against the LES results.
Given the proposed MOST function of mean velocity gradient 𝜙𝑚 , the mixing length under the
Prandtl’s mixing length hypothesis is derived and compared to the LES calculated results.
Finally, the sensitivity of the MOST statistics to the LES domain streamwise and spanwise
extents are tested.

4.1 MOST Statistics in ISL

The statistics are obtained from the LES model X24-Y5-C (with fixed domain
streamwise length 𝐿𝑥 = 24ℎ and spanwise length 𝐿𝑦 = 5ℎ, processed at Coarse mesh). The
one-dimensional statistics (vertical profile) are then calculated by averaging in temporal,

64
spanwise and streamwise domains, as introduced in subsection 3.5.2. The cases at neutral state
are served as the reference case to compare with the unstable cases. Starting from the neutral
state, 12 instability states with approximately equal intervals are simulated for each model of
A0.5, A1 and A2 while 7 instability states are simulated for the model AH0.5. The adjusted
local height 𝑧𝑒𝑓𝑓 = 𝑧 − 𝑑 is used. The von Karman constant 𝜅 = 0.4 and the turbulent Prandtl
number 𝑃𝑟𝑡 = 0.95 are found to be fit nicely to the data and will be used throughout. The
method to obtain displacement height d will be discussed later. Two instability parameters are
𝑧𝑖
used to describe the instability level: (i) describes the instability level of the whole boundary
𝐿
𝑧𝑒𝑓𝑓
layer (UBL if neutral and CUBL if convective) while (ii) 𝜁 = describes the instability level
𝐿

at local height, where 𝑧𝑖 is the effective boundary layer depth, i.e. total domain height minus
displacement height, 𝑧𝑖 = 𝐿𝑧 − 𝑑 , and can also be approximated by 𝑧𝑖 ≈ 𝐿𝑧 − ℎ since the
building height ℎ usually does not differ too much from 𝑑. The maximum instability levels of
𝑧𝑖
the simulated cases are = −8.4, −8.3, −7.8 and −7.6, corresponding to the models A0.5, A1,
𝐿

A2 and AH0.5, respectively.


The normalized statistics 𝜙𝑚 , 𝜙ℎ , 𝜙𝑤 and 𝜙𝜃 are plotted against the local instability
level 𝜁 in Figures 4.1 – 4.4. Since the wall is rough, 𝑧𝑒𝑓𝑓 is adopted as the length scale in the
normalization of the mean gradients. It is found that the plots of mean gradients 𝜙𝑚 and 𝜙ℎ are
more sensitive to the value of d than the fluctuation moments 𝜙𝑤 and 𝜙𝜃 since the definitions
of 𝜙𝑚 and 𝜙ℎ include the displacement 𝑑 while 𝜙𝑤 and 𝜙𝜃 do not (i.e. 𝜙𝑤 and 𝜙𝜃 just shift in
the x-axis of the plot as a whole when 𝑑 is varied). Therefore, the displacement height 𝑑 is then
obtained by fitting the mean gradient of velocity 𝜙𝑚 onto a curve within 1.3ℎ < 𝑧 < 1.7ℎ for
each aspect ratio of A0.5, A1 and A2 while 0.8ℎ < 𝑧 < 1.2ℎ for AH0.5, such that the lower
bound is 0.3ℎ above roughness top and the upper bound is 0.7ℎ above roughness top (i.e. one-
tenth of the effective boundary layer depth 𝑧𝑖 ) and the stated ranges are assumed to be within
the ISL. The convergence of data (collapse onto a line within the assumed ISL) for A0.5 is not
satisfactory compared to A1 and A2 since the extent of RSL is relatively high in A0.5 because
of the increased inhomogeneity caused by increased roughness separation so that the ISL is too
thin or may not exist. Therefore, the model of aspect ratio 0.5 with half roughness dimension
(AH0.5) is employed to further test the applicability of MOST since its true ISL extent is larger
than that of A0.5, where the improvement of convergence is clearly shown by fluctuation
statistics 𝜙𝑤 and 𝜙𝜃 . To ensure an appropriate value of 𝑑 for the model A0.5, the value of 𝑑
obtained from AH0.5 replaces that of A0.5 (multiplying by 2 due to the double roughness

65
dimension of A0.5). Finally, the values of 𝑑/ℎ obtained are 0.9, 1, 1 and 0.45 for the model
A0.5, A1, A2 and AH0.5 and they are all found to be almost independent of the instability for
the entire range of instability considered. It is also found that the value of 𝑑 is not very sensitive
to data at different height 𝑧 (within in the assumed ISL, 1.3ℎ < 𝑧 < 1.7ℎ) employed in the
fitting process, which justifies the use of the above stated height ranges.

4.1.1 Mean Streamwise Velocity Gradient 𝝓𝒎

Figure 4.1a shows 𝜙𝑚 at each instability level plotted from the building top to the
domain top for the model A1. It clearly shows that a portion of data for each instability level
collapses onto a curve when plotted against the local instability level 𝜁 . It implies 𝜙𝑚 is a
𝑧𝑖 𝑧𝑖
function of 𝜁 only but not since is different for each model, which means MOST is
𝐿 𝐿

satisfied. By definition, the range where the data collapses onto a curve for each instability case
is the true ISL. To better show the true range of ISL, the same data of A1 within 1.15ℎ < 𝑧 <
1.7ℎ is shown in a separate plot in Figure 4.1b. The plot shows that the data collapses onto a
single curve very well within this range, except the most near-neutral case and near the upper
end (𝑧~1.7ℎ) of some unstable cases. It confirms that the local free convection hypothesis is
satisfied by the LES data of model A1 below the maximum instability level 𝜁𝑚𝑎𝑥 ~ − 0.7, i.e.
the data can be described by a power form with a power index -1/3 (red dashed-dotted line). At
near-neutral condition, the linear equation first proposed by Webb [28] describes the LES data
quite well (blue dashed line). The transition instability from the near-neutral state to the local
free convection state occurs at 𝜁𝑡𝑟,𝑚 ~ − 0.05. As a result, a pair of functions, under the realm
of MOST, is able to describe the LES data of 𝜙𝑚 below the maximum attainable instability
level:
1 + 5.5𝜁 −0.05 < 𝜁 < 0 (4.1)
𝜙𝑚 (𝜁) = { −1/3
0.27(−𝜁) −0.7 < 𝜁 < −0.05
However, 𝜙𝑚 calculated by LES is of lower magnitude compared to the one fitted from field
measurements (BD relation with the empirical constants fitted by Hogstrom [23], green dotted
line) and the deviation of magnitude increases with instability, at a value of approximately 0.2
lower at the local free convection state. This problem was also noted from the LES calculation
by Khanna and Brasseur [102] with also a value of 0.2 lower than the field measurement one,
but the reason is still not known. Nevertheless, given that the field measurement data used in
the regression of equation is highly scattered, especially under high instability level, it is

66
1.5
a) ϕm = 0.27(-zeff/L)-1/3

ϕm = 1
1
ϕm

0.5

Aspect Ratio = 1
(A1) ϕm = 1 + 5.5(zeff/L)
0
0.001 0.01 0.1 1 10
-zeff /L
b) 1.2 ϕm = 0.27(-zeff/L)-1/3 c) 1.2
ϕm = 0.27(-zeff/L)-1/3
1 ϕm = 1 1 ϕm = 1
ϕm = ϕm =
0.8 [1+19(-zeff/L)]-1/4 0.8 [1+19(-zeff/L)]-1/4
ϕm

ϕm

0.6 0.6

0.4 0.4
Aspect Aspect
Ratio = 1 Ratio = 2
0.2 0.2
(A1) (A2)
ϕm = 1 + 5.5(zeff/L) ϕm = 1 + 5.5(zeff/L)
0 0
0.002 0.02 0.2 2 0.002 0.02 0.2 2
-zeff /L -zeff /L
d) 1.2 ϕm = 0.27(-zeff/L)-1/3 e) 1.2
ϕm = 0.27(-zeff/L)-1/3
1 ϕm = 1 1 ϕm = 1
ϕm = ϕm =
0.8 [1+19(-zeff/L)]-1/4 0.8 [1+19(-zeff/L)]-1/4
ϕm

ϕm

0.6 0.6

0.4 0.4
Aspect
Aspect Ratio = 0.5,
Ratio = 0.5 Half Height
0.2 (A0.5) 0.2
(AH0.5)
ϕm = 1 + 5.5(zeff/L) ϕm = 1 + 5.5(zeff/L)
0 0
0.002 0.02 0.2 2 0.002 0.02 0.2 2
-zeff /L -zeff /L

Figure 4.1: Normalized mean streamwise velocity gradient 𝜙𝑚 against local stability level
−𝑧𝑒𝑓𝑓 /𝐿. From roughness top to domain top: a) aspect ratio 1 (A1). From 0.15ℎ to 0.7ℎ above the
roughness top: b) aspect ratio 1 (A1), c) aspect ratio 2 (A2), d) aspect ratio 0.5 (A0.5) and e) aspect ratio
0.5 and half roughness height (AH0.5). Black solid line: neutral state. Blue dashed line: weakly unstable
state, linear function. Red dashed-dotted line: local free-convection state, power function. Green dotted line:
Businger-Dyer relation with the coefficients by Hogstrom [23].

67
thought that the deviation of LES result from the field measurement is attributed to the
overestimation of MOST function from the scatted field data. The influence of the whole
boundary layer (characterized by 𝑧𝑖 ) does not appear significantly in 𝜙𝑚 of the current LES
𝑧𝑖
results at the ISL region given the instability level varies across the LES models. However,
𝐿
𝑧𝑖
we do not deny this possibility at real situation since (i) the global instability level in the
𝐿

current LES study is much lower than those in field measurement studies and (ii) some
processes like entrainment from the interfacial layer and transient growth of boundary layer are
neglected in the idealized LES models, although the influence is believed to be small if it exists.
𝜙𝑚 at other aspect ratios shows similar trends with the pair of functions fitted to the
model A1, although their quality of data collapse is not as good as the model A1. It is observed
that 𝜙𝑚 near wall region (i.e. maybe within RSL) for the model A2 shown in Figure 4.1c is
overestimated compared to the fitted MOST functions. This may be due to the weaker
mechanical turbulence generated by the roughness of the model A2 compared to model A1 and
thus there is a larger proportion of drag force coming from the molecular diffusivity and subgrid
scale turbulence which may not be fully resolved by the current LES as no wall model is applied
to the no-slip walls. Nevertheless, except very near the wall, 𝜙𝑚 of the model A2 generally
compares well with the parametric equations. On the other hand, the data alignment of the
model A0.5 shown in Figure 4.1d is not as good as the model A1, which may be due to the
higher RSL height when the roughness separation is larger. The model AH0.5 in Figure 4.1e,
with a comparatively thicker ISL due to smaller roughness size, shows an improvement of data
alignment. However, both the models A0.5 and AH0.5 show that their magnitudes of 𝜙𝑚 are
higher than the parametric MOST functions at the intermediate instability, but the difference is
small.
The above results show that MOST of 𝜙𝑚 is generally applicable for the urban-type
roughness at the local instability levels (−0.7 < 𝜁 < 0) and the aspect ratios (0.5, 1 and 2)
tested, although the data alignment at the very near neutral cases is comparatively less
satisfactory. The parametric MOST functions with the fitted coefficients extracted from model
A1 are also applicable for all tested models of different aspect ratios.

4.1.2 Mean temperature gradient 𝝓𝒉

The mean temperature gradient 𝜙ℎ of aspect ratio 1 is shown in Figure 4.2a. Similar to
the mean velocity gradient 𝜙𝑚 , a portion of data near wall for each instability level collapses
68
1.5
a)
ϕh = 0.105(-zeff/L)-2/3

ϕh = 1
1
ϕh

0.5

Aspect Ratio = 1
(A1) ϕh = 1 + 5.5(zeff/L)
0
0.001 0.01 0.1 1 10
-zeff /L
b) 1.2 ϕh = 0.105(-zeff/L)-2/3 c) 1.2
ϕh = 0.105(-zeff/L)-2/3
1 ϕh = 1 1 ϕh = 1

0.8
ϕh = 0.8 ϕh =
[1+11.6(-zeff/L)]-1/2 [1+11.6(-zeff/L)]-1/2
ϕh

ϕh

0.6 0.6
0.27(-zeff /L)-1/3
0.4 0.4
Aspect
Aspect Ratio = 2
0.2 Ratio = 1 0.2 (A2)
(A1)
ϕh = 1 + 5.5(zeff/L) ϕh = 1 + 5.5(zeff/L)
0 0
0.002 0.02 0.2 2 0.002 0.02 0.2 2
-zeff /L -zeff /L
d) 1.2 ϕh = 0.105(-zeff/L)-2/3 e) 1.2
ϕh = 0.105(-zeff/L)-2/3
1 ϕh = 1 1 ϕh = 1

ϕh = ϕh =
0.8 0.8
[1+11.6(-zeff/L)]-1/2 [1+11.6(-zeff/L)]-1/2
ϕh

ϕh

0.6 0.6

Aspect
0.4 0.4
Aspect Ratio = 0.5,
Ratio = 0.5 Half Height
0.2 0.2
(A0.5) (AH0.5)
ϕh = 1 + 5.5(zeff/L) ϕh = 1 + 5.5(zeff/L)
0 0
0.002 0.02 0.2 2 0.002 0.02 0.2 2
-zeff /L -zeff /L

Figure 4.2: Normalized mean temperature gradient 𝜙ℎ against local stability level −𝑧𝑒𝑓𝑓 /𝐿 .
Others are similar to Figure 4.1.

69
onto a curve when plotted against the local instability level 𝜁. It is also found that 𝜙ℎ at the
local free convection state satisfies the MOST well until 1/5 boundary layer depth, while 𝜙𝑚
only satisfies the MOST well at most below 1/10 boundary layer depth. Figure 5b shows only
a portion of data near wall 1.15ℎ < 𝑧 < 1.7ℎ, in which the alignment of data is quite good
except the most near neutral one. The fitted MOST function of ϕh from the model A1 is:
1 + 5.5𝜁 −0.05 < 𝜁 < 0 (4.2)
𝜙ℎ (𝜁) = { −2/3
0.105(−𝜁) −0.7 < 𝜁 < −0.05
The linear functional form, the same as 𝜙𝑚 , describes the near neutral data well, which implies
the Reynolds analogy is applicable at the near neutral state. After the transitional point 𝜁𝑡𝑟,ℎ ~ −
0.05, 𝜙ℎ is better described by the power form with the power index -2/3 rather than the power
index -1/3 deduced from the local free convection hypothesis, where the difference is clearly
by the two power function of different indices (the red dashed-dotted line for index -2/3 and
the orange solid line for index -1/3) in Figure 4.2b. Therefore, the Reynolds analogy is not
satisfied, although the MOST is well satisfied. The power index difference is also observed on
the Businger-Dyer relations of 𝜙𝑚 and 𝜙ℎ and their ratio of power indexes of 𝜙𝑚 and 𝜙ℎ is the
same as the current LES result.
𝜙ℎ at other aspect ratios are shown in Figure 4.2c-e and its behaviors are very similar
to those of 𝜙𝑚 , e.g. deviation at near-neutral state and mesh effect near wall for narrow
roughness cavities. The LES calculated 𝜙ℎ is also at a magnitude 0.2 lower than the field
measurement one (green dotted line). The data alignment of 𝜙ℎ for both the model A0.5 and
AH0.5 is good and the above fitted functions describe them well. Therefore, the RSL effect on
𝜙ℎ is comparatively weaker than 𝜙𝑚 .

4.1.3 Second moment of vertical velocity fluctuation 𝝓𝒘

The vertical velocity fluctuation 𝜙𝑤 of aspect ratio 1 is shown in Figure 4.3a including
for the whole boundary layer and in Figure 4.3b within 1.15ℎ < 𝑧 < 1.7ℎ. The fitted functions
are:
1.2 −0.1 < 𝜁 < 0 (4.3)
𝜙𝑤 (𝜁) = { 2/3
4.5(−𝜁) −0.7 < 𝜁 < −0.1
It shows that 𝜙𝑤 at the near neutral state converges to the neutral MOST value, 𝜙𝑤 =
1.2, around the upper limit of ISL 𝑧 = 1.7ℎ, that is a higher elevation for the MOST to be
satisfied when compared to 𝜙𝑚 and 𝜙ℎ . This problem is believed to be due to mesh resolution
It is also found that the neutral MOST value is satisfied by the near neutral cases well above

70
10
a)
Aspect Ratio = 1 ϕw = 4.5(-zeff/L)2/3
(A1)

ϕw = 1.2
ϕw

0.1
0.001 0.01 0.1 1 10
-zeff /L
4 4
b) c)
Aspect Aspect
Ratio = 1 Ratio = 2
3 (A1) 3 (A2)

ϕw = 1.69(1-3zeff/L)2/3 ϕw = 1.69(1-3zeff/L)2/3
ϕw
ϕw

2 2

1 ϕw = 1.2 1 ϕw = 1.2

ϕw = 4.5(-zeff/L)2/3 ϕw = 4.5(-zeff/L)2/3
0 0
0.002 0.02 0.2 2 0.002 0.02 0.2 2
-zeff /L -zeff /L
4 4
d) e)
Aspect Aspect
Ratio = 0.5 Ratio = 0.5,
3 (A0.5) 3 Half Height
(AH0.5)

ϕw = 1.69(1-3zeff/L)2/3 ϕw = 1.69(1-3zeff/L)2/3
ϕw
ϕw

2 2

1 ϕw = 1.2 1 ϕw = 1.2

ϕw = 4.5(-zeff/L)2/3 ϕw = 4.5(-zeff/L)2/3
0 0
0.002 0.02 0.2 2 0.002 0.02 0.2 2
-zeff /L -zeff /L

Figure 4.3: Normalized second moment of vertical velocity fluctuation 𝜙𝑤 against local stability
level −𝑧𝑒𝑓𝑓 /𝐿. From roughness top to domain top: a) aspect ratio 1 (A1). From 0.15ℎ to 0.7ℎ above the
roughness top: b) aspect ratio 1 (A1), c) aspect ratio 2 (A2), d) aspect ratio 0.5 (A0.5) and e) aspect ratio
0.5 and half roughness height (AH0.5). Black solid line: neutral state. Red dashed-dotted line: local free-
convection state, power function. Green dotted line: Panofsky et al. [36].

71
𝑧 = 1.7ℎ until 1/4 of the boundary layer depth. On the other hand, the power function of index
2/3 deduced from the local free convection hypothesis describes the data well at moderate
instability, starting at the transition instability 𝜁𝑡𝑟,ℎ ~ − 0.1 . The LES result shows that the
power function describes the data well down to the very low level near the roughness top.
Therefore, the flow dynamics near roughness is well consistent to the MOST and the local free
convection hypothesis when horizontally averaged, even the flow is highly inhomogeneous
there. Although the transition between the constant function and the power function is not
smooth, it shows that the deviation between LES data and the functions around the transition
point is small. Thus, the proposed pair of functions are sufficient to describe the relation of 𝜙𝑤
and 𝜁. Compared to the relation proposed by Panofsky et al. [36] fitted to the data of field
measurement, the LES one shows a smaller magnitude at near neutral state. This was also
observed in the LES of Khanna and Brasseur [102]. When the flow is more unstable, both the
LES data and the relation of Panofsky et al. [36] approach to a same trend.
The model A2 in Figure 4.3c shows a very similar result. Therefore, the mesh resolution
problem noted in 𝜙𝑚 and 𝜙ℎ does not appear in 𝜙𝑤 . However, the data of 𝜙𝑤 in model A0.5
fails to align on a single curve while the model AH0.5 shows a satisfactory alignment. This
implies the RSL effect of wider roughness separation strongly manifests on 𝜙𝑤 . In addition,
for the wider roughness separation cases, A0.5 and AH0.5, the near neutral cases show a higher
MOST value. The reason is still not known. The overall results support that the proposed
MOST functions of 𝜙𝑤 are applicable to all the aspect ratios tested.

4.1.4 Second moment of temperature fluctuation 𝝓𝜽

The temperature fluctuation 𝜙𝜃 of aspect ratio 1 is shown in Figure 4.4a for the whole
boundary layer and in Figure 4.4b within 1.15h < z < 1.7h. The fitted functions are:
3.5 −0.1 < 𝜁 < 0 (4.4)
𝜙𝜃 (𝜁) = { −2/3
0.9(−𝜁) −0.7 < 𝜁 < −0.1
It is found that the data alignment at near neutral state is even less satisfactory than 𝜙𝑤 .
It still does not converge to a curve at the assumed ISL upper limit 𝑧 = 1.7ℎ, but it appears to
converge at higher elevation. This is again believed to be resulted from the mesh resolution
effect. At stronger local instability, the LES data converges to the power function deduced from
the local free convection hypothesis, except the most near wall data, which suggests the RSL
effect on 𝜙𝜃 is stronger than 𝜙𝑤 . The LES data at the local free convection state agrees with
the MOST relation extracted from field measurement data by De Druin et al. [37] but not the
72
30
a)
ϕθ = 0.9(-zeff/L)-2/3
ϕθ

3 ϕθ = 3.5

Aspect Ratio = 1
(A1)
0.3
0.001 0.01 0.1 1 10
-zeff /L
8 8
b) c) ϕθ = 0.9(-zeff/L)-2/3
ϕθ = 0.9(-zeff/L)-2/3
ϕθ = 8.41(1-28.4zeff/L)-2/3
6 ϕθ 6
ϕθ

4 ϕθ = 3.5 4 ϕθ = 3.5

ϕθ = 8.41(1-28.4zeff/L)-2/3
2 2 Aspect
Aspect
Ratio = 1 Ratio = 2
(A1) (A2)
0 0
0.002 0.02 0.2 2 0.002 0.02 0.2 2
-zeff /L -zeff /L
8 8
d) ϕθ = 0.9(-zeff/L)-2/3
e) ϕθ = 0.9(-zeff/L)-2/3

ϕθ = 8.41(1-28.4zeff/L)-2/3 ϕθ = 8.41(1-28.4zeff/L)-2/3
6 6
ϕθ

ϕθ = 3.5
ϕθ

4 4 ϕθ = 3.5

Aspect
2 2 Ratio = 0.5,
Aspect
Ratio = 0.5 Half Height
(A0.5) (AH0.5)
0 0
0.002 0.02 0.2 2 0.002 0.02 0.2 2
-zeff /L -zeff /L

Figure 4.4: Normalized second moment of temperature fluctuation 𝜙𝜃 against local stability level
−𝑧𝑒𝑓𝑓 /L. Others are similar to Figure 4.3. Green dotted line: De Druin et al. [37].

73
near neutral state. The field measurement result shows a larger value of 𝜙𝜃 at near neutral state
mainly because the temperature scalar source in real surface is usually inhomogeneous while a
uniform and constant temperature boundary condition is prescribed to the LES.
At local free convection state, the data of the model A2 shows a better alignment than
the model A1 while the model A0.5 shows a worse alignment and the model AH0.5 improves
slightly. 𝜙𝜃 of all aspect ratios agree to the power function at the local free convection state.
At the near neutral state, the model A0.5 and AH0.5 show a smaller value of 𝜙𝜃 and better
agree to the MOST value 𝜙𝜃 = 3.5, which may be due to the stronger turbulent mixing by a
wider roughness separation.

4.2 Mixing Length in ISL Derived from MOST

As an approach to model the turbulent transport in ISL, the Prandtl’s mixing length
hypothesis states that the Reynolds stress is related to the mean velocity gradient through the
mixing length 𝑙𝑚 :
𝑑𝑈 2 (4.5)
̅̅̅̅̅̅
−〈𝑢 ′ 𝑤 ′ 〉 = (𝑙
𝑚 )
𝑑𝑧
In ISL, the Reynolds stress can be approximated by the surface stress such that:
𝜏
′ 𝑤 ′ 〉 ≈ 0 = 𝑢2
̅̅̅̅̅̅
−〈𝑢 𝜏
(4.6)
𝜌
Combined with the definition of MOST function 𝜙𝑚 , the mixing length 𝑙𝑚 can be expressed
as:
𝜅𝑧𝑒𝑓𝑓
𝑙𝑚 ≈ (4.7)
𝜙𝑚
which agrees with the mixing length 𝑙𝑚 = 𝜅𝑧𝑒𝑓𝑓 at neutral state in which 𝜙𝑚 = 1.
In Figure 4.5, the mixing length 𝑙𝑚 calculated by LES result (denoted by symbols) is
compared to 𝑙𝑚 calculated using Equation (4.7) and the proposed MOST function 𝑙𝑚 from
Equation (4.1) (denoted by solid lines). Their deviation is also shown in Figure 4.6. Figure 4.5
shows that the 𝑙𝑚 calculated using the proposed MOST function follows the LES results well
within the ISL region (black dotted line) at different instability levels for the cases A1, A2 and
AH0.5 while the deviation is comparatively larger for the case A0.5. This can be seen from
Figure 4.6 that their deviations are less than 20% for the cases A1, A2 and AH0.5 but the overall
deviation is larger for the case A0.5. This may be because of the thicker RSL for the case A0.5
and the MOST function fails to describe the mean velocity gradient well in RSL. The modeled
𝑙𝑚 by MOST function tends to be overestimated for all cases. Figure 4.6 also show that the

74
Figure 4.5: Mixing length (x-axis) against the effective height (y-axis), both normalized by the
reference length ℎ; a) aspect ratio 1 (A1), b) aspect ratio 2 (A2), c) aspect ratio 0.5 (A0.5), d) aspect ratio
0.5 of half roughness size (AH0.5). Line: Equation (4.7). Symbol: LES result.

modeled one and the LES one tends to deviate from each other as the ISL upper limit or as the
surface roughness is approached. This implies that the proposed MOST function 𝜙𝑚 may not
describe the LES results well over the entire ISL region, especially near the ISL upper limit
and the surface roughness. The deviation as the ISL upper limit is approached may be due to
three reasons: (i) inaccuracy in LES modeling (both processing and post-processing); (ii)
inaccuracy of the proposed MOST function of 𝜙𝑚 given that MOST is valid; and (iii) MOST
is not strictly valid and other parameter like the boundary layer thickness 𝑧𝑖 is able to influence
𝜙𝑚 in ISL, though the influence is not too large (as the deviation is less than 20% even at the
ISL upper limit). The reason (i) seems unlikely since the trend of deviation near ISL upper limit
is consistent for all cases of different aspect ratios (except A0.5) and instabilities. From Figure

75
Figure 4.6: Deviation of mixing lengths between LES results and MOST functions (x-axis) against
the effective height normalized by the reference length ℎ (y-axis); a) aspect ratio 1 (A1), b) aspect ratio 2
(A2), c) aspect ratio 0.5 (A0.5), d) aspect ratio 0.5 of half roughness size (AH0.5). Different colors mean
different instability levels, ranging from neutral (black) to moderately unstable (red).

4.1 in subsection 4.1.2, it does show that the LES calculated 𝜙𝑚 near the ISL upper limit at
different instabilities does not fall onto the “universal” curve. Therefore, at this current stage,
it is concluded that the reason (iii) is more probable.

4.3 Sensitivity Study on the Effect of Horizontal Domain


Dimensions on MOST Quantities at Local Free Convection State

The previous sections studied the MOST at fixed domain dimensions (X24-Y5), but the
effect of horizontal domain dimensions on the SL statistics is still not known. In view of the

76
1 1 1
a) b) c)
<θ'θ'>/θ*2 =
0.8 0.8 <w'w'>/w*2 = 0.8 1.66(zeff/δeff)-2/3
2.44(zeff/δeff)2/3

0.6 0.6 0.6


zeff /zi

0.4 0.4 0.4

0.2 0.2 0.2

0 0 0
0 1 2 3 0 0.5 1 1.5 2 0 4 8 12 16 20
<u'u'>/w*2 <w'w'>/w*2 <θ'θ'>/θ*2

Figure 4.7: Second moment versus the dimensionless height 𝑧𝑒𝑓𝑓 /𝑧𝑖 : a) streamwise velocity, b)
vertical velocity and c) temperature fluctuations, normalized by Deardorff scale 𝑤∗ or 𝜃∗, of the model A1
at Coarse mesh and 12 roughness cavity (A1-X24-Y5-C), from the near neutral state (dark blue) to the local
free convection state (red). Red dashed-dotted line: SL scaling reformulated from the local free convection
hypothesis.

unclearness, a sensitivity study of domain sizes is conducted: streamwise extent 𝐿𝑥 =


2ℎ, 4ℎ, 6ℎ, 12ℎ, 24ℎ, 48ℎ, denoted by X2, X4, X6, X12, X24, X48; and spanwise extent 𝐿𝑦 =
5ℎ, 10ℎ, 20ℎ, denoted by Y5, Y10, Y20;. It is well known that eddies at the length scale of
boundary layer depth 𝑧𝑖 dominates the turbulent motions at the core of the boundary layer
under strong instability. Given that other large scale factors are unchanged, the statistics should
be scaled by the Deardorff scales (Deardorff [65]) under free convective condition in which
shear is of minor effect. In Deardorff scaling (introduced in Section 2.2.3), the boundary layer
depth 𝑧𝑖 is the primary length scale, the convective velocity 𝑤∗ the primary velocity scale and
𝜃∗ is the corresponding temperature scale.
To demonstrate the Deardorff scaling by LES, figure 4.7 shows that the fluctuation
statistics within the whole boundary layer for the model A1-X24-Y5-C, which was used in the
study of the previous section, converges to single curves as instability increases. It means the
Deardorff scaling is satisfied. The local free convection hypothesis of the SL (or ISL) vertical
velocity and temperature fluctuations under MOST scaling can be reformulated in terms of
Deardorff scaling, as shown in Figure 4.7b and c (red dotted-dashed line), since both the shear
effect and the boundary layer depth effect, which are denoted by 𝑢𝜏 and 𝑧𝑖 , do not appear on
these statistics at the SL. However, the profiles of velocity fluctuation calculated by LES is
larger above ISL when compared to the atmospheric measurement, e.g. Deardorff and Willis
[142], and symmetric in shape while atmospheric measurement ones are asymmetric. It is

77
believed to be resulted from the difference in heating configuration such that a quasi-steady
and vertically uniform heat flux is resulted in current LES rather than a realistic growing
convective boundary with vertically diminishing heat flux is resolved. Nevertheless, the
temperature fluctuation calculated by LES is similar to the field measurement data below
0.8𝑧𝑒𝑓𝑓 /𝑧𝑖 and its MOST relation is applicable until the mid-boundary layer.
Before investigating how the horizontal domain dimensions modifies the MOST
quantities in ISL, it is instructive to observe how the horizontal domain dimension modifies the
turbulence quantities of the whole boundary layer. A series of models at the free convective
condition in which Deardorff scaling is applicable, with different streamwise and spanwise
domain extents are run at VeryCoarse (VC) mesh. The detailed model configurations are listed
in Section 3.4.1. Figure 4.8 shows that different 𝐿𝑥 and 𝐿𝑦 are able to modify the strengths of
boundary layer scaled eddies at the mid-boundary layer. When 𝐿𝑥 increases, both streamwise
and vertical fluctuations strengthen. When 𝐿𝑦 increases, streamwise fluctuation weakens while
vertical fluctuation weakens initially at Y10 but it then strengthens at Y20. This implies the
strength and the topology of the boundary layer scaled eddies depend on the LES domain
horizontal dimensions. These problems are specific to numerical simulations when periodic
boundary conditions are applied to the domain side walls, which leads to the quantization of
the boundary layer scaled eddies is resulted as discussed in the introduction. Although the
fluctuations of velocity components are modified by horizontal domain extents, the
temperature fluctuation seems unaffected, except the case of shortest streamwise domain (X2-
Y5).
In order to view the effect of horizontal domain dimensions on the MOST quantities
within ISL, the MOST quantities within 1.15ℎ < 𝑧 < 1.7ℎ are plotted against the local
instability level together with the parametric equations of the local free convection from the
previous section in Figure 4.9. As shown in Figure 4.9a(i), the mean velocity gradient 𝜙𝑚
shows a lower magnitude when 𝐿𝑥 is short, but it converges to a single curve as the 𝐿𝑥 increases
to 24ℎ. The converged data is comparable to the parametric curve except it deviates from the
parametric equation at the lower end, which is believed to be due to the RSL effect and/or the
mesh effect since a coarser mesh is used in this sensitivity test. The smaller magnitudes in the
cases where 𝐿𝑥 is short (i.e. the models X2, X4 and X6) are the indication that the turbulent
motions of larger scale (of scale of 𝐿𝑥 and > 𝑧𝑒𝑓𝑓 ) is able to modify the local flow that has
long been believed to be dominated by the local scale motions of scale 𝑧𝑒𝑓𝑓 . On the other hand,
when 𝐿𝑦 increases from 5ℎ to 20ℎ as shown in Figure 4.9a(ii), 𝜙𝑚 at the upper end of ISL

78
1 1
a) i) ii)

0.8 0.8

X2_Y5
0.6 X4_Y5 0.6 X24_Y5 X48_Y5
zeff /zi

zeff /zi
X6_Y5 X24_Y10 X48_Y10
X12_Y5
X24_Y20 X48_Y20
0.4 X24_Y5 0.4
X48_Y5

0.2 0.2

0 0
0 1 2 3 0 1 2 3
<u'u'>/w* 2 <u'u'>/w*2
1 1
b) i) ii)

0.8 0.8
X2_Y5
X4_Y5
0.6 0.6 X24_Y5 X48_Y5
zeff /zi

X6_Y5
zeff /zi
X24_Y10 X48_Y10
X12_Y5
X24_Y20 X48_Y20
0.4 X24_Y5 0.4
X48_Y5

0.2 <w'w'>/w*2 0.2 <w'w'>/w*2


= 2.44(zeff/zi)2/3 = 2.44(zeff/zi)2/3

0 0
0 0.5 1 1.5 0 0.5 1 1.5
<w'w'>/w*2 <w'w'>/w*2
1 1
c) i) ii)
<θ'θ'>/θ*2 <θ'θ'>/θ*2
0.8 = 1.66(zeff/zi)-2/3 0.8 = 1.66(zeff/zi)-2/3
X2_Y5
X4_Y5
0.6 0.6
X6_Y5
zeff /zi

zeff /zi

X12_Y5
X24_Y5 X48_Y5
X24_Y5
0.4 0.4
X48_Y5 X24_Y10 X48_Y10

X24_Y20 X48_Y20
0.2 0.2

0 0
0 4 8 12 16 20 0 4 8 12 16 20
<θ'θ'>/θ*2 <θ'θ'>/θ*2

Figure 4.8: Second moment versus the dimensionless height 𝑧𝑒𝑓𝑓 /𝑧𝑖 : a) streamwise velocity, b)
vertical velocity and c) temperature fluctuations, normalized by Deardorff scale 𝑤∗ or 𝜃∗, of the model A1
at VeryCoarse mesh and at a local free convection state, at variable domain i) streamwise (X) and ii)
spanwise (Y) extents (the meanings of symbol and color follow the legends of the plots). Red dashed-dotted
line: SL scaling reformulated from the local free convection hypothesis.

79
0.8 0.8
a) i) ii)
0.6 0.6
ϕm

ϕm
0.4 ϕ = 0.27(-z /L)-1/3 0.4
m eff ϕm = 0.27(-zeff/L)-1/3
X2_Y5
X4_Y5
0.2 X6_Y5
0.2 X24_Y5 X48_Y5
X12_Y5 X24_Y10 X48_Y10
X24_Y5
X48_Y5 X24_Y20 X48_Y20
0 0
0.1 1 0.1 1
-zeff /L -zeff /L
0.8 0.8
b) i)
X2_Y5
X4_Y5 ii) X24_Y5 X48_Y5
X6_Y5 X24_Y10 X48_Y10
X12_Y5 X24_Y20 X48_Y20
0.6 X24_Y5 0.6
X48_Y5
ϕh

ϕh
0.4 0.4

0.2 ϕh = 0.105(-zeff/L)
-2/3
0.2 ϕh = 0.105(-zeff/L)-2/3

0 0
0.1 1 0.1 1
-zeff /L -zeff /L
4 4
c) i) ii)
ϕw = 4.5(-zeff/L)2/3 ϕw = 4.5(-zeff/L)2/3
3 3
ϕw
ϕw

2 2
X2_Y5
X4_Y5
1 X6_Y5 1 X24_Y5 X48_Y5
X12_Y5 X24_Y10 X48_Y10
X24_Y5
X48_Y5 X24_Y20 X48_Y20
0 0
0.1 1 0.1 1
-zeff /L -zeff /L
4 4
d) i) ϕθ = 0.9(-zeff/L)-2/3
X2_Y5
ii)
X4_Y5
X6_Y5
ϕθ = 0.9(-zeff/L)-2/3
3 X12_Y5 3
X24_Y5
X48_Y5
ϕθ
ϕθ

2 2

1 1 X24_Y5 X48_Y5
X24_Y10 X48_Y10
X24_Y20 X48_Y20
0 0
0.1 1 0.1 1
-zeff /L -zeff /L

Figure 4.9: MOST statistics: Dimensionless gradient of mean a) streamwise velocity 𝜙𝑚 , b)


temperature 𝜙ℎ , dimensionless second moment of c) vertical velocity fluctuation 𝜙𝑤 , d) temperature
fluctuation 𝜙𝜃 , of the model A1 at VeryCoarse mesh (VC) and at the local free convection state within
1.15ℎ < 𝑧 < 1.7ℎ, at variable domain i) streamwise (X) and ii) spanwise (Y) extents (the meanings of
symbol and color follow the legends of the plots).

80
decreases in magnitude but it is unaffected at the lower end. The mean temperature gradient
𝜙ℎ (Figure 4.9b) on the other hand converges to a single curve as 𝐿𝑦 increases to 24ℎ (Figure
4.9b(i)) and it is not significantly affected by 𝐿𝑦 (Figure 4.9b(ii)). 𝜙ℎ is also larger in
magnitude compared to the parametric equation due to the mesh effect. The vertical velocity
fluctuation 𝜙𝑤 (Figure 4.9c) is significantly influenced by both the 𝐿𝑥 and 𝐿𝑦 . It deviates from
the parametric equation as 𝐿𝑦 increases. The temperature fluctuation 𝜙𝜃 (Figure 4.9d)
converges to the parametric equation when 𝐿𝑥 or 𝐿𝑦 increases.
All the results of X24 and X48 are similar in both the ISL and the whole boundary layer
and thus X24 is sufficient. It also shows that the temperature statistics 𝜙ℎ and 𝜙𝜃 in ISL are
less dependent on 𝐿𝑥 and 𝐿𝑦 compared to the velocity statistics 𝜙𝑚 and 𝜙𝑤 . This is in line with
the conclusion of Khanna and Brasseur [102]. 𝜙𝑤 depends on the domain configuration
significantly, which means 𝜙𝑤 is actually influenced by the larger scale eddies, despite the local
free convection functional form is still hold (i.e. the power index is the same, but the empirical
constant is different). This sensitivity test demonstrates that the boundary layer scale motion,
of scale of 𝐿𝑥 and 𝐿𝑦 , indeed modulates the local scale flow statistics in ISL or SL. From the
above evidence, MOST seems not strictly valid and the MOST statistics are not exactly
“universal”. This is therefore one of the reason for the scattering of field measurement data.

4.4 Concluding Remarks

In this chapter, the LES results show that the MOST is basically satisfied within ISL as
the data of each MOST function of 𝜙𝑚 , 𝜙ℎ , 𝜙𝑤 and 𝜙𝜃 collapses onto a single curve within
ISL for each aspect ratio (except A0.5 where the RSL is too thick for a significant extent of ISL
to exist) and each instability. However, the LES results show that the data collapse at slightly
unstable state is not good for 𝜙𝑚 , 𝜙ℎ , 𝜙𝑤 and 𝜙𝜃 while the LES calculated 𝜙𝑚 deviates from
the “universal” curve as the ISL upper limit is approached. It is possible that the boundary layer
height 𝑧𝑖 , which is excluded from the assumption of MOST, is a factor that causes the deviation
from the “universal” curve. Nevertheless, the deviation is not large (less than 20% within the
ISL).
Parameterized functional forms of the MOST functions of 𝜙𝑚 , 𝜙ℎ , 𝜙𝑤 and 𝜙𝜃 are
proposed to fit the “universal” curves concatenating the ISL data under a range of boundary
layer instability levels, characterized by 𝑧𝑖 /𝐿. The proposed functional forms are piecewise

81
functions. For the mean gradients 𝜙𝑚 and 𝜙ℎ , the piecewise function contains a linear function
for the near-neutral range and a power function for the moderately unstable range. For the
fluctuation moments 𝜙𝑤 and 𝜙𝜃 , the piecewise function contains a constant function for the
near-neutral range and a power function for the moderately unstable range. At moderately
unstable state, it is found that the power indexes of 𝜙𝑚 (-1/3), 𝜙𝑤 (2/3) and 𝜙𝜃 (-2/3) predicted
by the local free-convection hypothesis are agreed by the LES results. However, the power
index of mean temperature gradient 𝜙ℎ is found to be -2/3 from the LES result, which conforms
to the ratio of power indexes of 𝜙𝑚 and 𝜙ℎ in Businger-Dyer relations but does not agree with
the value -1/3 predicted by local free-convection hypothesis. The reason is not known here
while Li et al. [33] suggested the dissimilarity of power indexes of 𝜙𝑚 and 𝜙ℎ is due to a
“scale-resonance” between turnover eddies and excursions in the instantaneous temperature
profile. Based on the proposed function of 𝜙𝑚 , the mixing length 𝑙𝑚 in ISL based on the
Prandtl’s mixing length hypothesis is derived as well and compared well to the LES results.
From the sensitivity study of LES domain streamwise and spanwise extents at
moderately unstable state, the flow fluctuation statistics show that the domain horizontal
dimension is able to modify the boundary layer scale eddies in both strength and pattern, whose
statistics are consistent to the Deardorff’s scaling. The results show that the domain streamwise
extent 24ℎ (X24) used to investigate the MOST statistics in Section 4.1 is sufficiently long for
MOST statistics to converge in ISL. It also shows that the temperature statistics 𝜙ℎ and 𝜙𝜃 in
ISL are basically unaffected by the spanwise extent but the velocity statistics 𝜙𝑚 and 𝜙𝑤 do
not, especially near the upper limit of ISL. The result implies the larger scale eddies of different
intensities and patterns, determined by the domain horizontal extents, are able to modulate the
ISL flow statistics, though the influence is not very significant. This provides an evidence that
MOST is not strictly valid and the MOST statistics are not “universal”.

82
Chapter 5

Properties of Roughness Sublayer (RSL) at Different


Instability Conditions and Roughness Configurations

RSL is defined as a layer in which the local flow properties show the influence of
roughness. In this chapter, the statistical flow properties in RSL at different instability
conditions and roughness aspect ratios are investigated. Two approaches are used to study the
influence of roughness on RSL flow properties and RSL height 𝑧∗ : (i) deviation of horizontally
averaged profiles of flow statistics from the MOST profiles that are applicable in ISL and (ii)
the horizontal inhomogeneity of flow statistics due to the inhomogeneous nature of roughness
morphology. The advantages and disadvantages of these two approaches to calculate 𝑧∗ are
discussed. As RSL is near to the surface, the near wall resolution may cause problems in
resolving the turbulent motions there. Therefore, the impact of mesh resolution is tested by
including simulation of models of aspect ratio 1 at three mesh resolution levels, namely,
VeryCoarse (VC), Coarse (C) and Fine (F), with a smaller domain size (A1-X6-Y5).

5.1 Deviation from MOST Statistics in RSL and the Effect of


Mesh Resolution at Roughness Aspect Ratio 1

This section illustrates how the urban type roughness modifies the horizontal averaged
mean flow at region influenced by roughness, i.e. RSL, compared to the standard SL or ISL
under horizontal averaging. The near-wall region, including the cavities and a layer just above
the roughness top, is more sensitive to the mesh resolution compared to the outer layer since
the turbulent transports in the near wall region are dominated by smaller scale motions, which
may not be fully resolved by LES. In order to study the MOST in ISL and deviation of

83
Figure 5.1: Vertical profiles or MOST statistics at near-wall region: a) Dimensionless gradient of
mean (time + horizontal averaged) streamwise velocity 𝜙𝑚 , b) dimensionless gradient of mean temperature
𝜙ℎ , c) dimensionless second moment of vertical velocity fluctuation 𝜙𝑤 , d) dimensionless second moment
of temperature fluctuation 𝜙𝜃 . i) Neutral stratification, ii) unstable stratification. Dashed line and cicle:
VeryCoarse mesh (A1-X6-Y5-VC), dashed-dotted line and triangle: Coarse mesh (A1-X6-Y5-C), dotted
line and square: Fine mesh (A1-X6-Y5-F), thin solid line and cross: 12 cavities at Coarse mesh (A1-X24-
Y5-C), thick black solid line without symbol: modelled MOST functions fitted to LES results.

84
MOST in RSL by the current LES model with stronger confidence, the resolved flow statistics
at the near-wall region above roughness top are compared at different mesh resolutions,
including both the neutral state and an unstable state at which the local free-convection
condition, as studied in Chapter 4, is satisfied within ISL. Only the roughness configuration of
aspect ratio 1 is considered in this section. The model setting A1-X6-Y5 (i.e. a shorter domain
streamwise extent compared to the one used in previous chapter) is employed in this
comparison study at the three mesh resolutions: VeryCoarse (VC), Coarse (C) and Fine (F).
The model A1-X24-Y5-C that was used in the previous chapter to investigate the MOST
functions is also included for comparison.
The dimensionless mean gradients of streamwise velocity 𝜙𝑚 and temperature 𝜙ℎ as
well as the variances of vertical velocity 𝜙𝑤 and temperature 𝜙𝜃 above the roughness roof
level 𝑧 = ℎ and below the upper bound of ISL 𝑧 = 1.7ℎ (defined as one-tenth of boundary
layer depth), are shown in Figure 5.1. The calculated statistics take the resolved scale
contributions only and are horizontally averaged (in temporal, streamwise and spanwise
domains). In order to demonstrate the impact of roughness on RSL, the roughness height ℎ
instead of the Obukhov length 𝐿 is used to normalize the vertical coordinate in Figure. 5.1. It
is noted that the length scale used in the normalization of 𝜙𝑚 and 𝜙ℎ is the effective height
𝑧𝑒𝑓𝑓 = 𝑧 − 𝑑, which is the same as the roof-level adjusted height 𝑧 − ℎ for the case of aspect
ratio 1 since the displacement height 𝑑 is almost equal to the roughness height ℎ under the
entire instability range simulated, as discussed in Section 4.1. The MOST functions proposed
in the previous chapter are also included in Figure 5.1 (black solid line) to illustrate the
deviation of RSL flow from the ISL flow as well as the RSL height.

5.1.1 Neutral stratification

From Figure 5.1a(i), mean velocity gradient 𝜙𝑚 increases from the roof level and then
develops a local maximum peak. When the mesh is coarser, the peak is at a higher location and
its magnitude is larger. This indicates a coarser mesh underestimates the mixing efficiency such
that a steeper mean velocity gradient (i.e. higher 𝜙𝑚 ) is required to sustain the same amount of
momentum flux (which is unchanged for all models because of the same prescribed background
pressure gradient). It is thought to be resulted from a portion of unresolved smaller scale
motions that are partially responsible for turbulent mixing. 𝜙𝑚 at different mesh resolutions
approaches the value 1, consistent to MOST at neutral state. It shows that the impact of mesh

85
resolution on 𝜙𝑚 is weaker away from the wall than near to the wall, which implies that the
unresolved motions near the wall does not influence 𝜙𝑚 at higher elevation. The error due to
mesh resolution does not pass to a higher elevation, which means that the turbulent mechanism,
e.g. generation and destruction, in ISL under neutral stratification is highly local. However, 𝜙𝑚
of the model VC converges to the neutral MOST value slower than the models C and F.
Considering the model F, 𝜙𝑚 converges to the neutral MOST value at 𝑧~1.5ℎ , though 𝜙𝑚
within 1.05ℎ < 𝑧 < 1.5ℎ does not differ from the MOST value (𝜙𝑚 = 1) too much.
From the review of Finnigan [47] on the flows over vegetation canopy, the plane mixing
layer type turbulence induced at the roughness top enhances the eddy viscosities of momentum
and heat compared to standard SL and thus 𝜙𝑚 is less than unity as predicted by MOST. On
the other hand, the study of urban-type roughness by Letzel et al. [143] showed that the plane
mixing layer type turbulence also appeared due to the dynamic instability of the air stream
velocity mismatch at the interface between the roughness cavities and the outside unobstructed
flow. As illustrated by the F mesh result, although there is a local maximum, 𝜙𝑚 is less than 1
within ℎ < 𝑧 < 1.3ℎ (RSL), which implies the local mixing is enhanced compared to MOST.
It is believed the plane mixing layer type coherent structures are responsible for the enhanced
transport at roughness top. However, the local maximum of 𝜙𝑚 just above urban-type
roughness does not appear in those of horizontally homogenous vegetation canopies, as
demonstrated by the observation and model of flow over forest by de Ridder [62] and Harman
and Finnigan [61]. Therefore, the local maximum is hypothesized to be specific to this 2D
urban-type roughness. The difference is thought to be resulted from the different characteristics
of motions between urban type and vegetation canopy roughness as well as the consequence of
horizontal averaging. In urban-type roughness, the motions are highly inhomogeneous in the
horizontal direction in RSL, especially near roughness top where the mean wind gradient is
very strong due to the no-slip boundary condition. This behavior is expected to vary with
different urban roughness configurations (i.e. different aspect ratio or 2D/3D geometry). A
homogenous vegetation canopy on the other hand shows no similar behavior since the motions
are horizontally homogenous in RSL.
The mean gradient of temperature 𝜙ℎ at neutral state shown in Figure 5.1b(i) is very
similar to 𝜙𝑚 for 𝑧 < 1.3ℎ, which suggests the similar effect of mesh resolution and similar
transport mechanism of momentum and temperature. It also suggests the Reynolds analogy is
well satisfied under neutral condition at the near wall region where the direct roughness effect
is strong. However, a gradual but small deviation between 𝜙𝑚 and 𝜙ℎ at higher elevation is

86
believed to be due to the different driving mechanisms of momentum and heat transports in the
LES models where the momentum flux decreases linearly with height while the heat flux is
uniform with height.
The vertical velocity fluctuation 𝜙𝑤 shown in Figure 5.1c(i) increases smoothly with
height in RSL for VC and C mesh models but a mild peak appears in F model. For the VC
model, 𝜙𝑤 is underestimated, due to the failure to resolve all dominant motions. The
underestimation of 𝜙𝑤 in VC model persists until the limit of ISL 𝑧 = 1.7ℎ while 𝜙𝑤 in C
model has already approached the neutral MOST value 𝜙𝑤 = 1.2 for 𝑧 > 1.5ℎ and F mesh
models for 𝑧 > 1.3ℎ. All of these suggest C mesh model is sufficient to resolve the ISL motions
and also a considerable portion of motions in RSL. However, the reason of the slightly lower
magnitude of 𝜙𝑤 in ISL for F model is unknown.
For the temperature fluctuation 𝜙𝜃 , a peak also appears just above roughness top as
shown in Figure 5.1d(i). 𝜙𝜃 then decreases with height until it reaches the neutral MOST value
𝜙𝜃 = 3.5 (but the F mesh indicates the neutral MOST value may be around 4). The C mesh
model shows that 𝜙𝜃 has still not approached the neutral MOST value at the ISL upper limit
𝑧 = 1.7ℎ, while 𝜙𝜃 in the F mesh model has already converged to a value ~4, but different
from the converged value 3.5 at C model. The VC model significantly overestimates 𝜙𝜃 below
𝑧 = 1.7ℎ.
The results of the neutral cases suggest that the VC mesh is insufficient for the current
LES to resolve all the motions for 𝑧 > 1.3ℎ, while the C mesh seems sufficient. The mesh
effect is stronger in the fluctuation statistics 𝜙𝑤 and 𝜙𝜃 than in the mean gradient statistics 𝜙𝑚
and 𝜙ℎ . The result shows that the mesh resolution effect and roughness effect in RSL are
intertwined, so it is difficult to obtain the RSL height 𝑧∗ to high accuracy from the current LES
results. Also, the RSL height 𝑧∗ defined by different statistics are different. From the F model,
which is regarded as the most accurate one, the RSL height 𝑧∗ defined by the convergence of
profile is estimated to be 1.3ℎ from 𝜙𝑤 and 𝜙𝜃 and 1.5ℎ from 𝜙𝑚 , while 𝜙ℎ crosses the
MOST value at 𝑧 = 1.4ℎ but it keeps increasing with height. A very good similarity between
the model of domain streamwise length 6ℎ (X6) and the model of domain streamwise length
24ℎ (X24) at the same mesh for all statistics is noted, which suggests that the active turbulence
is sufficiently calculated by the model X6 and the current domain size is of little impact on the
near-wall region under neutral condition.

87
5.1.2 Unstable stratification

For the unstable cases of which the ISL region is under local-free-convective condition,

the instability levels at roughness scale for models of different meshes are − 𝐿 ~ − 0.87 (X6-

VC), −0.84 (X6-C), −0.93 (X6-F) and −0.88 (X24-C). The instability of these models are
slightly different even though the prescribed parameters (e.g. background pressure gradient)
and the boundary conditions are kept unchanged because the quasi-steady horizontal averaged
upward heat flux, which determines the Obukhov length 𝐿 , is calculated from the LES
calculated results rather than prescribed in advance like the momentum flux so that 𝐿 may be
influenced by the near wall mesh resolution and/or the domain size. Nevertheless, the
instabilities of different models are approximately equal.
As shown in Figure 5.1a(ii), the LES result of models X6 shows that 𝜙𝑚 at different
mesh resolutions similarly approaches the same trend when 𝑧 > 1.3ℎ, but different from the
model X24-C, which has a longer domain streamwise length. It has been demonstrated in the
Chapter 4 that the horizontal domain extent, which modifies the outer layer eddies, is able to
modify the motions and turbulent transports in ISL. For 𝑧 < 1.3ℎ , it is surprising that the
modeled MOST function of 𝜙𝑚 follows the LES data down to very near roughness top at
𝑧~1.05ℎ, which can be regarded as the RSL height of the mean gradient or the mean profile of
velocity. Nearer to the surface roughness, a local peak appears, similar to the neutral cases.
This local peak may be the result of the roughness shear layer induced turbulence rather than
buoyancy induced turbulence, at least at the current instability level considered. Nevertheless,
buoyancy still exhibits certain degree of modulation near roughness so that the magnitude of
the local maximum is not exactly the same as the neutral cases. Similar to the neutral cases, the
position of the local peak still significantly depends on mesh resolution, but the magnitude at
the near wall (around the local peak) in unstable state is less influenced by the mesh resolution.
The possible reason is that the turbulent motions there contain certain amount of buoyancy
induced turbulence (e.g. buoyancy generated turbulence or smaller scale turbulence transported
down by the outer layer larger eddies) which is stronger in magnitude and larger in scale than
the purely shear turbulence and are able to be captured by LES.
As shown in Figure 5.1b(ii), 𝜙ℎ at different mesh resolutions also approaches the same
trend with small amount of deviation when 𝑧 > 1.4ℎ and, unlike 𝜙𝑚 , the trend is similar to the
modeled 𝜙ℎ fitted to the LES result of model X24. This is consistent to the result of previous
chapter that the temperature statistics are less influenced by the domain dimensions. 𝜙ℎ also

88
develops a peak just above the roughness roof level, though with a smaller magnitude compared
to 𝜙𝑚 . Therefore, the Reynolds analogy, which states that 𝜙𝑚 ≈ 𝜙ℎ , is approximately satisfied
at the near wall region around the local peak. However, as discussed in the previous section
and shown in the later section, 𝜙ℎ is not the same as 𝜙𝑚 in ISL, although it is not clearly
observed in Figure 5.1a(ii) and 5.1b(ii). This has been reported by many studies and it is
consistent to the fact that the functional forms of 𝜙ℎ and 𝜙𝑚 are different in ISL as
demonstrated in the previous chapter.
As shown in Figure 5.1c(ii), the vertical velocity fluctuation 𝜙𝑤 at different mesh
resolutions shows similar increasing trends with height, comparable to the MOST function at
local free convection (black solid in the figure) in ISL. It is notable that the LES calculated
profile follow the local free-convection functional form down to 𝑧~1.05ℎ, very near the roof-
level, similar to the mean gradients 𝜙𝑚 and 𝜙ℎ , However, there is deviation between 𝜙𝑤 at
different mesh resolutions, starting from near the roughness top 𝑧~1.05ℎ, and the deviation
persists even above 𝑧 = 1.3ℎ to the ISL top with uniform offsets. This implies, when
parameterized by the MOST function of power form with power index -1/3, the empirical
constant does not consist at different mesh resolutions. On the contrary, 𝜙𝑤 in neutral cases
converges at higher elevation at different mesh resolutions. This problem has also been
addressed by the LES study of a purely convective boundary layer by Sullivan and Patton [144].
They also found that the second moment of resolved vertical velocity fluctuation was
underestimated throughout the whole boundary layer if the mesh was not fine enough and the
underestimation is largest at roughly the mid-boundary layer where the magnitude of vertical
velocity fluctuation is at maximum. They suggested there should be adequate scale separation
between energy-containing eddies and those near filter-cutoff in order for the resolved solution
to converge. The top-down mechanism proposed by Laubach and McNaughton [11] suggests
that the outer layer motions can transport the smaller scale turbulence, which is originated in
the outer layer by the Richardson cascade, down to the near wall region (i.e. 𝜙𝑤 at SL may
possess some outer layer flow properties). Therefore, sufficiently dense mesh is required to
resolve these small scale turbulence that are important to the near surface statistics throughout
the whole boundary layer. It is hypothesized that 𝜙𝑤 is underestimated near wall if the motions
are not fully captured by the LES in the core part of the boundary layer. This problem is
prominent in the current study since gradient mesh resolution is employed in the LES models
of which the mesh in the core region is coarser than the near wall.
Finally, as shown in Figure 5.1d(ii), the value of 𝜙𝜃 in the unstable state is generally

89
less than that in neutral state and decreases with height, following the MOST function of local
free convection. 𝜙𝜃 just above the roughness top is similar to the neutral case that a local peak
is also developed there, which supports that the flow there is mainly controlled by shear
dynamics and the buoyancy only provides a modulating effect. However, it is not known why
there is a persistent deviation above the local maximum in the F model but not in the VC and
C models.
Overall, it shows that the motions are sufficiently resolved by C mesh model in ISL and
above ( 𝑧 > 1.4ℎ ) for the unstable cases, although there are still slight deviations in the
fluctuation moments 𝜙𝑤 and 𝜙𝜃 throughout the whole SL (ℎ < 𝑧 < 1.7ℎ), possibly, due to the
top-down mechanism. The modeled MOST functions follow the data points down to the near
roughness top 𝑧~1.05ℎ for 𝜙𝑚 , 𝜙ℎ and 𝜙𝑤 and 𝑧~1.3ℎ for 𝜙𝜃 , which implies the
corresponding RSL heights 𝑧∗ . It is hypothesized that the ISL profile agrees to the data down
to such a low height is due to the top-down mechanism proposed by Laubach and McNaughton
[11]. The flows are similar at both the neutral and the unstable cases when nearer to the rough
wall (around the peak at ~1.05ℎ) where the dynamics of near wall motions are shear dominated
and this region is strongly influenced by mesh resolution. Unstable buoyancy effect is relatively
weaker near wall than at higher elevations since the locally generated turbulence by shear
outweighs the Richardson cascade turbulence transported down from the core of boundary
layer. As a rule of thumb, the RSL and ISL flows can be accurately modeled in neutral
stratification if the mesh is sufficiently dense in RSL and ISL, while the RSL and ISL flows
can be accurately modeled in moderately unstable stratification if the mesh is sufficiently dense
throughout the whole boundary layer, which is more computationally demanding.

5.2 Deviation from MOST Functions in RSL at Different


Instability Levels and Roughness Aspect Ratios

In this section, the deviations of horizontally averaged flow statistics from the MOST
functions fitted to the LES data within ISL in Chapter 4 are compared in the near wall region
at different instability levels and roughness aspect ratios. The mean velocity gradient 𝜙𝑚 and
the vertical velocity fluctuation 𝜙𝑤 are studied here, while the mean temperature gradient 𝜙ℎ
and the temperature fluctuation 𝜙𝜃 do not since 𝜙ℎ is very similar to 𝜙𝑚 in the near wall region
while 𝜙𝜃 in RSL is too complicated to draw any conclusion. These horizontal averaged

90
Figure 5.2: Ratio of LES calculated dimensionless mean velocity gradient 𝜙𝑚 and MOST function
from Equation (4.1) at various instability levels from neutral (black line) to moderately unstable (red line)
and a) aspect ratio 1 (A1), b) aspect ratio 2 (A2), c) aspect ratio 0.5 (A0.5), d) aspect ratio 0.5 and half
roughness size (AH0.5).

statistics are plotted in a ratio, e.g. 𝜙𝑚 /𝜙𝑚_𝑀𝑂𝑆𝑇 , against the normalized height 𝑧/ℎ starting
from the building roof level until the ISL upper limit, where 𝜙𝑚 is the statistics directly
calculated by LES and 𝜙𝑚_𝑀𝑂𝑆𝑇 is the MOST function fitted to the ISL data in Chapter 4.
𝜙𝑚_𝑀𝑂𝑆𝑇 is from the Equation (4.1) and 𝜙𝑤_𝑀𝑂𝑆𝑇 is from Equation (4.3). The models used in
this study are the same as those used in the study of MOST in Chapter 4 (i.e. model setting of
X24-Y5-C).

5.2.1 Mean velocity gradient 𝝓𝒎

As shown in Figure 5.2, it is surprising that the ratio 𝜙𝑚 /𝜙𝑚_𝑀𝑂𝑆𝑇 is independent of


the instability condition for each the roughness configuration (A1, A2 and AH0.5) except the
model A0.5, whose profiles of the ratio at different unstable levels do not collapse onto one
curve. It is possible that the problem in A0.5 is due to the displacement height 𝑑 which is taken
to be constant with respect to instability in this study while 𝜙𝑚 is sensitive to 𝑑 as explained in
Section 4.1. Therefore, it is possible that 𝑑 indeed depends on instability for aspect ratio 0.5.

91
However, it is very difficult to obtain an accuracy value of 𝑑 since 𝑑 is not a well-defined
parameter, which can only be obtained through fitting of data in ISL to the known MOST
function and it requires a sufficient depth of ISL. Nevertheless, the profiles under a range of
instability in model A0.5 do not differ too much from each other. Therefore, the results from
A1, A2 and AH0.5 suggest that the deviation of 𝜙𝑚 from the MOST function is independent
of instability for the range of instability levels simulated in this study and therefore the RSL
height 𝑧∗ defined based on the deviation from MOST is independent of the instability.
For the models of roughness aspect ratio 1 (A1), the roughness effect on 𝜙𝑚 is confined
within 𝑧 < 1.15ℎ . It shows there is a local peak for all instability levels around 𝑧 = 1.1ℎ ,
which has been discussed in Section 5.1. For the models of aspect ratio 2 (A2), the roughness
effect on 𝜙𝑚 is confined within 𝑧 < 1.3ℎ where the region is of larger extent compared to A1
and the local peak is larger in magnitude. This larger extent of stronger peak are hypothesized
to be due to the insufficient mesh resolution for the model A2. As the width of separation
between building elements is halved for the model A2 compared to the model A1, a portion of
the smaller scale motions may not be resolved and the mean gradient may therefore be
overestimated which leads to a stronger peak in 𝜙𝑚 . For the models of roughness aspect ratio
0.5 of half roughness size (AH0.5), the roughness effect on 𝜙𝑚 is confined within 𝑧 < 0.7ℎ
(i.e. 0.2 above roof level). The deviation increases in magnitude as the roughness is approached
and there is no local peak in the plot as it is located under the building roof level since the mean
flow diverges down a small depth under the roof level to the roughness cavity behind the
roughness corner and the displacement height 𝑑 is therefore under the roof level (different from
A1 and A2 that their displacement heights are approximately equal to their building roof levels
due to their skimming flow behavior).

5.2.2 Second moment of vertical velocity fluctuation 𝝓𝒘

As shown in Figure 5.3, the models A1 and A2 show that the deviation of 𝜙𝑤 from
the MOST function decreases with instability level. This means that the RSL height 𝑧∗ based
on this definition decreases with height as instability level enhances. This is different from the
mean velocity gradient 𝜙𝑚 that the deviation from MOST in RSL is independent of instability.
It is hypothesized the decreased deviation of 𝜙𝑤 from MOST profile in the near wall region
when instability levels increases is because a certain amount of turbulence generated above,
e.g. ISL or outer layer, transports down to the near wall region by the large eddies and the

92
Figure 5.3: Ratio of LES calculated dimensionless second moment of vertical velocity fluctuation
𝜙𝑤 and MOST function from Equation (4.3) at various instability levels from neutral (black line) to
moderately unstable (red line) and a) aspect ratio 1 (A1), b) aspect ratio 2 (A2), c) aspect ratio 0.5 (A0.5),
d) aspect ratio 0.5 and half roughness size (AH0.5).

advected turbulence bears similar characteristics of ISL turbulence which is well described by
the MOST function. This explanation is in line with the top-down mechanism proposed by
Laubach and McNaughton [11]. Therefore, the turbulence motions that 𝜙𝑤 represents contain
certain amount of turbulence that is not locally generated and destroyed. On the other hand, the
turbulent transporting motions in both ISL and RSL represented by the mean gradient 𝜙𝑚
contain only the active motions that account for the momentum transports. This implies that a
portion of turbulence represented by 𝜙𝑤 is inactive to momentum transport and contains
information of larger scale motions, e.g. buoyancy induced eddies of boundary layer scale.
However, the models of aspect ratio 0.5 (A0.5 and AH0.5) show a more complicated
pattern. For this case, 𝜙𝑤 tends to be underestimated from MOST function in the near wall
region under neutral condition while it tends to be overestimated there under strong unstable
condition. This complicated behavior may be again due to the displacement height 𝑑 which is
assigned a constant value with respect to the instability and fitted using the mean velocity
gradient 𝜙𝑚 rather than the vertical velocity fluctuation 𝜙𝑤 . Apart from the small dependence
of 𝑑 to instability, this may imply an additional problem that the fitted value of 𝑑 may be
different if different flow statistics are taken. It has been tested (the result not shown here) that

93
the profiles of the ratio 𝜙𝑚 /𝜙𝑚_𝑀𝑂𝑆𝑇 in A0.5 are similar to those of models A1 and A2 if a
lower value of 𝑑 (while 𝑑~0.9ℎ based on mean gradient 𝜙𝑚 ) is used. The result of AH0.5 is
similar to that of A0.5 but with smaller deviation at the near wall region, which is due to the
smaller roughness size so that the instability effect at roughness scale (parameterized by ℎ𝑟 /𝐿,
ℎ𝑟 is the height of building) is smaller in model AH0.5.
The results of 𝜙𝑚 and 𝜙𝑤 in the near wall region suggest that the extent of RSL 𝑧∗
based on the deviation from the MOST function require an accurate pre-determined value of
displacement height 𝑑, which is different to obtain accurately through fitting of profile, and it
is even harder in field measurement studies. Secondly, the obtained value of 𝑧∗ is also different
if different flow statistics are considered.

5.3 Horizontal Inhomogeneity of Flow Statistics in Near Wall


Region at Different Instability Levels and Roughness Aspect Ratios

In this section, the inhomogeneity of ensemble averaged flow statistics (i.e. averaged
in temporal and spanwise domains and for all cavity units) in streamwise (x-) direction is
investigated under different instability levels and roughness aspect ratios. As proposed by
Florens et al. [14], the RSL height 𝑧∗ based on this inhomogeneous properties is defined by a
certain percentage of variability of a given set of ensemble statistics. The horizontal variability

is defined as 𝐷(𝜙̅)/〈𝜙̅〉 , where 𝐷(𝜙̅) = √(𝜙̅ − 〈𝜙̅〉)2 is the dispersion of 𝜙̅ from the
horizontal mean 〈𝜙̅〉, and 〈𝜙̅〉 is the streamwise averge of 𝜙̅. This idea is tested on simulated
2D urban type roughness flows with unstable stratification in this section. The models A0.5,
A1 and A2 are used to investigate how both the instability levels and roughness aspect ratios
modify the inhomogeneous properties at the near wall region. The instability level concerning
RSL flow is better described by the ratio of building height to Obukhov length, ℎ/𝐿. ℎ is used
as a reference length scale here as it is the most relevant length scale to describe the flow
properties in RSL that is under the strong influence of physical roughness.

94
5.3.1 Streamwise variation of vertical profiles of mean streamwise velocity
̅̅̅̅̅̅
̅ and Reynolds stress 𝒖′𝒘′
𝒖

̅̅̅̅̅̅ are investigated here to


The mean streamwise velocity 𝑢̅ and Reynolds stress 𝑢′𝑤′
illustrate the meaningful characteristics of the inhomogeneous properties while the overall
behaviors of other single-point statistics are discussed collectively in the later subsection. The
𝑧
streamwise variations 𝑢̅(𝑧) including both inside roughness cavity (0 < ℎ < 1) and above roof
𝑧
level (ℎ > 1) are shown in Figure 5.4. Only the models A0.5 and A1, but not the model A2, are

taken here to illustrate their significant difference caused by roughness geometrical


configuration since the result of model A2 are very similar to that of model A1. The results of
both models A1 and A0.5 show a strong spatial variation, both streamwise and vertical, inside
the roughness cavity at different instability levels. This spatial variation is actually caused by
the flow recirculation inside the cavity, which behaves similarly for all cases due to the
skimming flow characteristics at aspect ratios 1 and 0.5. When it is interpreted by the mean
flow state, the mean flow recirculation is confined in the cavity and is basically decoupled from
the flow above roof-level. However, the interpretation on the mean flow does not imply the
turbulence fluctuating flow fields as we will subsequently demonstrate it on the Reynolds stress.
As a result, the streamwise variability decreases sharply just above roof-level and the profiles
of 𝑢̅ start to converge onto a single profile. The elevation levels at which 5% horizontal
variability is attained are also shown in the figures. For model A1 (Figure 5.4 (a)), it indicates
that 𝑢̅ quickly attains the level of 5% horizontal variability (𝑧~ℎ ) just near roof level at
different instability levels whatever the instability level is. For model A0.5, the horizontal
inhomogeneity at 5% is also quickly attained just slightly above the roof level, but its level
raises to 𝑧~1.3ℎ − 1.4ℎ for the other two unstable cases. Therefore, when the separation of
building roughness increases, the horizontal inhomogeneity of 𝑢̅ is more sensitive to the
buoyancy effect.
Similarly, the streamwise variation of vertical profiles of Reynolds stress ̅̅̅̅̅̅
𝑢′𝑤′ are
shown in Figure 5.5. The models at different instability levels and roughness aspects show that
̅̅̅̅̅̅
𝑢′𝑤′ is strongest and most inhomogeneous around the roof level. ̅̅̅̅̅̅
𝑢′𝑤′ is much weaker inside
the roughness cavity, though it is relatively stronger for model A0.5 compared to model A1 and
when it is under stronger instability level. Above roof level, the horizontal inhomogeneity
enhances with both wider separation of building elements and stronger instability level, which

95
Figure 5.4: Profiles of ensemble mean streamwise velocity 𝑢̅ at different streamwise (x-) positions
(within a roughness unit) at a) aspect ratio 1 (A1) and b) aspect ratio 2 (A2); under three levels of instability
condition i), ii) and iii). The dotted line represents the level at which 5% horizontal variability of 𝑢̅ is
attained.

96
Figure 5.5: Profiles of ensemble averaged Reynolds stress ̅̅̅̅̅̅ 𝑢′𝑤′ at different streamwise (x-)
positions (within a roughness unit) at a) aspect ratio 1 (A1) and b) aspect ratio 2 (A2); under three levels of
instability condition i), ii) and iii). The dotted line represents the level at which 5% horizontal variability of
̅̅̅̅̅̅
𝑢′𝑤′ is attained.

97
Figure 5.6: 2D contours of a) Reynolds stress ̅̅̅̅̅̅
𝑢′𝑤′, b) momentum flux by mean velocity advection
𝑢̅ ∙ 𝑤 ̅̅̅̅̅̅
̅ and c) total momentum flux (𝑢 ′𝑤 ′ + 𝑢 ̅), normalized by 𝑢𝜏2 at i) neutral and ii) moderately unstable
̅∙𝑤

conditions ( ~ − 1.2) for model A1.
𝐿

can be observed from the level of 5% horizontal variability for each case in Figure 5.5. It is
also observed that the inhomogeneity due to instability is much more significant than the effect
on aspect ratios. Figure 5.5(b)(iii) shows that the 5% horizontal variability is at a much higher
level, even above ISL. This implies that the inhomogeneity property of Reynolds stress profiles
̅̅̅̅̅̅
𝑢′𝑤′ can be protrude to the outer layer above ISL when the instability is strong and this effect
on ̅̅̅̅̅̅
𝑢′𝑤′ is much stronger than on the mean velocity profiles 𝑢̅.
To explain the reason of the strong inhomogeneity of Reynolds stress under unstable
condition, the contour of ̅̅̅̅̅̅
𝑢′ 𝑤 ′ on x-z plane is shown in Figure 5.6(a) for the model A1, and its

98
Figure 5.7: Vertical profiles of horizontal variability of a) mean streamwise velocity 𝑢̅ and b)
Reynolds stress ̅̅̅̅̅̅
𝑢′𝑤′ at i) aspect ratio 0.5 (A0.5), ii) aspect ratio 1 (A1) and iii) aspect ratio 2 (A2). Symbol
color: from neutral condition (black) to moderately unstable condition, ℎ/𝐿 ~ − 1.2 (red).

99
pattern are similar to the model A0.5 (and thus not shown here). When it is under neutral
condition, ̅̅̅̅̅̅
𝑢′ 𝑤 ′ of peak magnitude is located near the top corner of downstream building wall
(𝑥 = 1.5ℎ, 𝑦 = ℎ). In contrast, when it is under unstable condition, the local peak is located
near the top corner of upstream building wall (𝑥 = 0.5ℎ, 𝑦 = ℎ) and the region of strong ̅̅̅̅̅̅
𝑢′ 𝑤 ′
spreads to a much higher elevation compared to the neutral case. To understand the mechanism
of momentum transfer from free-stream region to the roughness elements, the vertical
momentum flux by the mean flow field, 𝑢̅ ∙ 𝑤
̅, is illustrated in Figure 5.6(b). The neutral case
(Figure 5.6(b)(i)) shows that 𝑢̅ ∙ 𝑤
̅ is concentrated at the downstream corner and 𝑢̅ ∙ 𝑤
̅ is
negligible away from the roof-level. On the other hand, the unstable case (Figure 5.6(b)(i))
shows that the flux is downward (negative 𝑢̅ ∙ 𝑤
̅ ) near the downstream corner but upward
(positive 𝑢̅ ∙ 𝑤
̅) near the upstream corner and the region with significant magnitude of 𝑢̅ ∙ 𝑤
̅
extends far away from the roof-level. This implies that the mean recirculation flow than
containing a significant proportion of momentum, protrudes above the roof-level of the cavity
opening under sufficiently strong unstable condition and it is believed to be the reason of the
high horizontal variability of Reynolds stress profiles ̅̅̅̅̅̅
𝑢′ 𝑤 ′ . Therefore, the mean flow velocity
is not strictly horizontally inhomogeneous above roof-level, although its variability is small in
percentage and its effect is large on 𝑢̅ ∙ 𝑤
̅ given such a small horizontal variability of 𝑢̅.
When the Reynolds stress (flux due to flow fluctuations) and the flux due to mean flow
̅̅̅̅̅̅
advection are summed up as a total momentum flux (𝑢 ′𝑤 ′ + 𝑢 ̅) as shown in Figure 5.6(c),
̅ ∙𝑤
it can be seen that the momentum is actually transferred down to the roughness elements only
through the region near the downstream corner no matter it is under neutral or unstable
condition. The momentum is subsequently transferred to the downstream building wall as
pressure force (form drag). It also shows that the Reynolds stress ̅̅̅̅̅̅
𝑢′ 𝑤 ′ is counterbalanced by
the momentum flux due to mean advection 𝑢̅ ∙ 𝑤
̅ at the upstream corner under unstable
condition so that no momentum is in fact transported there. This completes the origin why the
Reynolds stress profiles ̅̅̅̅̅̅
𝑢′ 𝑤 ′ is strongly inhomogeneous under unstable condition.
The vertical profiles of horizontal variabilities of 𝑢̅ and ̅̅̅̅̅̅
𝑢′ 𝑤 ′ , i.e. 𝐷(𝑢̅)/〈𝑢̅〉 and
̅̅̅̅̅̅
𝐷(𝑢 ̅̅̅̅̅̅
′ 𝑤 ′ )/〈𝑢 ′ 𝑤 ′ 〉, at aspect ratios 0.5, 1 and 2 under different levels of instability are shown in

Figure 5.7. It clearly indicates that their horizontal variabilities strongly depends on both the
roughness aspect ratios and instability levels. It seems that the horizontal variability is a non-
linear function of roughness aspect ratios and instability levels, e.g. widening the cavity
separation (i.e. decreasing aspect ratio) can magnify the instability effect on inhomogeneity. It
also demonstrated that 𝑢̅ and ̅̅̅̅̅̅
𝑢′ 𝑤 ′ can behave very differently in terms of horizontal variability.

100
It is interesting to observe that a local peak of ̅̅̅̅̅̅
𝑢′ 𝑤 ′ horizontal variability appears at 𝑧~1.5ℎ
when under strong unstable condition for the model A0.5, but the reason is still unknown.

5.3.2 Horizontal variability of single-point ensemble statistics and the


corresponding RSL height 𝒛∗

Some researchers, e.g. Florens et al. [14], proposed to define the RSL height 𝑧∗ by
certain percentage of horizontal variability of certain single-point statistics. However, the
problem of this definition is the arbitrariness of the assigned percentage and the choice of
ensemble statistics that qualifies as the candidates. In reduced scale experiment or even field
measurement, this problem may be crucial since the measurement cost is high and there is
usually a certain degree of variability due to measurement inaccuracy, which may cause this
definition difficult to apply. Nevertheless, the present LES study offers an opportunity to test
this idea in an easier and more systematic manner.
To give a comprehensive view on the inhomogeneous properties of different single-
point statistics, the RSL heights defined by horizontal variabilities of mean velocity
components 𝑢̅ and 𝑤 ̅̅̅̅̅̅ , turbulent heat flux 𝜃′𝑤′
̅ , Reynolds stress 𝑢′𝑤′ ̅̅̅̅̅̅ as well as second

moments of fluctuating quantities ̅̅̅̅̅


𝑢′𝑢′, ̅̅̅̅̅ ̅̅̅̅̅̅ and ̅̅̅̅̅
𝑣′𝑣′, 𝑤′𝑤′ 𝜃′𝜃′ are presented collectively in Figure
5.8, which includes the models of three roughness aspect ratios and all the instability levels
simulated. The RSL heights 𝑧∗ based on 5% and 10% horizontal variabilities are shown in
Figure 5.8(a) and (b), respectively, which show that 𝑧∗ defined by different variabilities can be
significantly different. This means there is no sharp-cut value of variability percentage that is
proper to define 𝑧∗ . The plots clearly show that the horizontal inhomogeneous properties of
different statistics are very different. For example, the turbulent flux ̅̅̅̅̅̅
𝑢′𝑤′ and ̅̅̅̅̅̅
𝜃′𝑤′ are the most
inhomogeneous ones while the mean velocities components 𝑢̅ and 𝑤
̅ are the least
inhomogeneous ones, so that the RSL height 𝑧∗ based on a turbulent flux can be much higher
than 𝑧∗ based on a mean velocity component given the same variability percentage. The results
also show that 𝑧∗ increases non-linearly with the instability, with the increase being much
stronger in stronger unstable condition. Therefore, the results shown in this section suggest that
the method to define the RSL height 𝑧∗ by the horizontal inhomogeneous properties of single-
point statistics can causes problems, especially under strongly unstable condition.

101
Figure 5.8: Ratio of LES calculated dimensionless second moment of vertical velocity fluctuation
𝜙𝑤 and MOST function from Equation (4.3) at various instability levels from neutral (black line) to
moderately unstable (red line) and a) aspect ratio 1 (A1), b) aspect ratio 2 (A2), c) aspect ratio 0.5 (A0.5),
d) aspect ratio 0.5 and half roughness size (AH0.5).

102
5.4 Concluding Remarks

In this chapter, the statistical flow properties in RSL in terms of horizontally averaged
profiles and horizontal inhomogeneity as well as the mesh resolution effect are investigated. In
addition, two methods to define the RSL height 𝑧∗ are tested under a range of unstable levels:
(i) deviation of horizontally averaged profiles of flow statistics from the MOST profiles that
are applicable in ISL and (ii) the horizontal inhomogeneity of flow statistics due to the
inhomogeneous nature of roughness morphology.
Firstly, the horizontally averaged flow statistics are compared to the corresponding
MOST functions at the near wall region for the model of aspect ratio 1 (A1) on three mesh
resolutions VeryCoarse (VC), Coarse (C) and Fine (F). At neutral condition, the results suggest
that the mean gradients 𝜙𝑚 and 𝜙ℎ are less than the MOST ones near the roughness, which is
similar to vegetation canopies. It is observed that there is a local peak just above the building
roof-level for both 𝜙𝑚 and 𝜙ℎ for any mesh resolution. This is thought to be a specific flow
characteristic over urban type roughness. The results of 𝜙𝑚 and 𝜙ℎ together with fluctuation
statistics 𝜙𝑤 and 𝜙𝜃 suggest that the near wall region is crucially dependent on the mesh
resolution. Insufficient mesh resolution tends to overestimate 𝜙𝑚 and 𝜙ℎ due to the subgrid-
scale turbulence model, but the mesh resolution effect diminishes with height and the ISL is
sufficiently modeled by the Coarse (C) mesh model. This implies the turbulent mechanism is
local or top-down in both RSL and ISL so that the RSL properties and effect of insufficient
mesh resolution in RSL will not transfer up to the ISL. At unstable condition, the result shows
that the flow at the vicinity of roughness is dominated by shear rather than buoyancy such that
𝜙𝑚 and 𝜙ℎ are very similar to those under neutral condition, even though the ISL region is
under local free-convection condition, while buoyancy only provides a secondary effect. The
result of vertical velocity fluctuation 𝜙𝑤 suggests that the boundary layer scaled motions can
modify the turbulence statistics in both ISL and RSL in the form of inactive turbulence.
Therefore, insufficient mesh resolution throughout the domain will influence the near wall
motions.
Secondly, considering the Coarse (C) mesh for different instability levels and different
aspect ratios, it is found that the deviation of mean gradient 𝜙𝑚 from its MOST function in
RSL is generally independent of the instability levels. This finding may be able to provide
certain simplification when modeling the RSL mean gradient 𝜙𝑚 or mean velocity profile
under unstable stratification. On the other hand, the result for fluctuation 𝜙𝑤 shows that it

103
depends on both roughness aspect ratio and instability level.
For the study of horizontal inhomogeneity near roughness, the unstable results show
that the turbulent fluxes, e.g. Reynolds stress ̅̅̅̅̅̅
𝑢′𝑤′ and heat flux ̅̅̅̅̅̅
𝜃′𝑤′, is more inhomogeneous
̅. The reason for strong inhomogeneity on ̅̅̅̅̅̅
than the mean flow quantities, e.g. 𝑢̅ and 𝑤 𝑢′𝑤′ is
found to be due to the difference in momentum transport mechanism under neutral and unstable
conditions. It is found that the momentum transport from the mean flow advection 𝑢̅ ∙ 𝑤
̅ above
building roof-level is significant under unstable condition but not under neutral condition and
̅ subsequently causes ̅̅̅̅̅̅
the non-negligible field of 𝑢̅ ∙ 𝑤 𝑢′𝑤′ to be highly inhomogeneous above
roof-level. The result also shows unstable condition can magnify the inhomogeneous properties
at wider separation of building elements, which is a non-linear effect.
Finally, the tests on the two approaches (i) and (ii) to define and calculate the RSL
height 𝑧∗ suggest that the result of 𝑧∗ can be very different based on different approach and
each approach has its own disadvantages. For example, 𝑧∗ based on the method (i) is quite
sensitive to the displacement height 𝑑, whose accurate value is difficult to be found, especially
in field measurement. The method (ii) may strongly exaggerate 𝑧∗ in unstable condition and
the threshold of dispersion level of inhomogeneity is arbitrary. Both the two methods show that
𝑧∗ is different depending on which single-point statistic is used, while the method (ii) is the
more sensitive one. Based on the above comparison, it is suggested that, in terms of RSL flow
modelling, the method (i) is the more practical one and preferable one.

104
Chapter 6

Parameterization of Surface Roughness on Inertial


Sublayer (ISL) Flow under Unstable Condition

In this chapter, the ISL mean velocity and temperature profiles, 𝑈(𝑧) and 𝛩(𝑧) ,
calculated by LES are investigated. The mean profile equations are derived analytically based
on the MOST functions proposed in Chapter 4. The analytical equations are compared to the
LES calculated results. Then the applicability and validity of these profile equations are
discussed. The aerodynamic roughness length 𝑧0 and the temperature roughness length 𝑧ℎ are
fitted by matching the derived profile equations and the LES results. Based on the derived mean
profile equation, the effects of local buoyancy at ISL and the effect of roughness on the mean
profile can be decoupled and their physical properties are discussed. The aerodynamic
roughness length 𝑧0 , which has long been regarded as a property of roughness geometrical
configuration alone, is tested against the instability level to study whether it is also a property
of buoyant instability. Additionally, the Reynolds number effect on the temperature roughness
length 𝑧ℎ is also studied. The detailed flow properties around roughness elements and their
relationships with 𝑧0 and 𝑧ℎ are investigated. Based on the LES results, the mechanisms of how
the different types of turbulent motions, e.g. buoyancy-induced updrafts and downdrafts and
eddies generated by Kelvin-Helmholtz instability, influences 𝑧0 and 𝑧ℎ are proposed.

6.1 Analytical Derivation of Mean Profiles Based on the


Proposed MOST Functions

From Chapter 4, the proposed MOST functions of mean gradients 𝜙𝑚 and 𝜙ℎ that
satisfied the LES results at three roughness configurations and within a range of unstable

105
𝒊=𝒎 𝒊=𝒉
𝒂𝒊 5.5 5.5

𝑪𝒊 0.27 0.105

𝒏𝒊 , -1/3 -2/3

𝜻𝒕𝒓,𝒊 -0.05 -0.05

𝜻𝒎𝒂𝒙,𝒊 -0.7 -0.7

Table 6.1: Values of numerical constants in Equation (6.1)

stratification is a pair of piece functions – a linear function at near neutral state and a power
function at moderately unstable state:
𝜙𝑖,1 (𝜁) = 1 + 𝑎𝑖 𝜁 𝜁𝑡𝑟,𝑖 < 𝜁 < 0 (6.1)
𝜙𝑖 (𝜁) = { 𝑛 , 𝑖 = 𝑚 𝑜𝑟 ℎ
𝜙𝑖,2 (𝜁) = 𝐶𝑖 (−𝜁) 𝑖 𝜁𝑚𝑎𝑥,𝑖 < 𝜁 < 𝜁𝑡𝑟,𝑖
where 𝑎𝑖 , 𝐶𝑖 and 𝑛𝑖 , are empirical constants fitted the LES data. 𝜁𝑡𝑟,𝑖 are the transition stability
from linear function to power function and 𝜁𝑚𝑎𝑥,𝑖 is the maximum stability that is valid for the
present LES results. The values of these constants are shown in Table 6.1. 𝜁𝑡𝑟,𝑖 is dependent on
the empirical constants 𝑎𝑖 , 𝐶𝑖 and 𝑛𝑖 . Their relationship is due the smooth transition between
𝑑𝜙𝑖,1 𝑑𝜙𝑖,2
the two functions at the transition point 𝜁𝑡𝑟,𝑖 , such that = when 𝜁 = 𝜁𝑡𝑟,𝑖 . The
𝑑𝜁 𝑑𝜁

relationship is then derived as:


𝜙𝑖,1 (𝜁) = 1 + 𝑎𝑖 𝜁 𝜁𝑡𝑟,𝑖 < 𝜁 < 0 (6.2)
𝜙𝑖 (𝜁) = { 𝑛 , 𝑖 = 𝑚 𝑜𝑟 ℎ
𝜙𝑖,2 (𝜁) = 𝐶𝑖 (−𝜁) 𝑖 𝜁𝑚𝑎𝑥,𝑖 < 𝜁 < 𝜁𝑡𝑟,𝑖
It is advantageous that this pair of piecewise functions provides a smooth transition at
the two stability ranges and it contains only elementary functions that permit analytical
integration to derive the mean velocity and temperature profiles applicable within ISL. A mean
profile here means a horizontally averaged vertical profile, i.e. 𝑈 =< 𝑢̅ >. As discussed in
Section 2.3.1, the mean velocity and temperature profiles after mathematical integration are:
𝑈(𝑧) 1 𝑧𝑒𝑓𝑓 (6.3)
= [ln ( ) − Ψ𝑚 (𝜁) + Ψ𝑚 (𝜁0 )]
𝑢𝜏 𝜅 𝑧0
𝛩(𝑧) − 𝜃𝑤 𝑃𝑟𝑡 𝑧𝑒𝑓𝑓 (6.4)
= [ln ( ) − Ψℎ (𝜁) + Ψℎ (𝜁ℎ )]
𝜃𝜏 𝜅 𝑧ℎ
𝑧 𝑧
where 𝜁0 = 𝐿0 and 𝜁ℎ = 𝐿ℎ and the stability correction function is:
𝜁
1 − 𝜙𝑖 (𝜁) (6.5)
Ψ𝑖 (𝜁) = ∫ 𝑑𝜁 , 𝑖 = 𝑚 𝑜𝑟 ℎ
0 𝜁
These profile equations will be used to obtain the parameters of surface roughness 𝑧0 and 𝑧ℎ
from the LES calculated results. The mathematical integration of Equation (6.1) yields:

106
−𝑎𝑖 𝜁 𝜁𝑡𝑟,𝑖 < 𝜁 < 0
Ψ𝑖 (𝜁) = { 𝜁 𝐶𝑖 (6.6)
ln ( ) − [(−𝜁)𝑛𝑖 − (−𝜁𝑡𝑟 )𝑛𝑖 ] − 𝑎𝑖 𝜁𝑡𝑟,𝑖 𝜁𝑚𝑎𝑥,𝑖 < 𝜁 < 𝜁𝑡𝑟,𝑖
𝜁𝑡𝑟,𝑖 𝑛𝑖
and

Ψ𝑖 (𝜁0,ℎ ) = −𝑎𝑖 𝜁0,ℎ (6.7)


given that 𝜁0 < 𝜁𝑡𝑟,𝑚 and 𝜁ℎ < 𝜁𝑡𝑟,ℎ , which are satisfied in most situations and in this study as
well. 𝜁0,ℎ means 𝜁0 or 𝜁ℎ . The detailed integration procedures are shown in Appendix A1.

6.2 Mean Streamwise Velocity and Temperature Profiles

The mean velocity and temperature profiles 𝑈(𝑧) and 𝛩(𝑧) calculated by LES of model
A1 are compared to the analytically derived profiles Equations (6.3) and (6.4) with the fitted
𝑧0 and 𝑧ℎ at all instability levels simulated in Figure 6.1. The results of other aspect ratios are
similar to A1 and are thus no shown here. The values of 𝑧0 and 𝑧ℎ are fitted within the assumed
ISL (1.3ℎ < 𝑧 < 1.7ℎ , denoted by two solid black lines in the plots) such that the LES
calculated profiles and the derived profiles are best-fit (the characteristics of 𝑧0 and 𝑧ℎ will be
discussed in the next section).
As expected, the mean profiles calculated by LES follow the derived profiles very well
within the assumed ISL. Both the plots show that the derived profiles also describe the LES
data well down to roughness top (𝑧 = ℎ). This is because the mean gradients 𝜙𝑚 and 𝜙ℎ do
not deviate significantly from the MOST ones in the RSL above building roof-level for the
model A1 (as discussed in Chapter 5), while the deviation for model A0.5 are comparative
larger (results not shown here). The mean velocity profile 𝑈(𝑧) calculated by LES begins to
deviate from the derived profile at all instabilities above the upper limit of the assumed ISL
(𝑧 = 1.7ℎ ). The mean temperature 𝛩(𝑧) above the ISL upper limit also deviates from the
derived profile at the near neutral state, but it keeps following the derived profile at strong
unstable level. This is because the mean temperature gradient 𝜙ℎ also follows its MOST
function well above the ISL upper limit under strong unstable condition. Therefore, the LES
results demonstrates that applicability of the integrated stability correction functions, Equations
(6.3) - (6.7), within ISL as well as the potential applicability outside ISL, e.g. RSL above roof-
level and mixed layer under sufficient instability level.
𝑈(𝑧)
In Equation (6.3), it is observed that the normalized mean velocity profile is
𝑢𝜏

107
Figure 6.1: Profiles of normalized a) mean streamwise velocity, b) mean temperature. The models
are of aspect ratio 1 (A1-X24-Y5). Line: integrated profile function with roughness length 𝑧0 or 𝑧ℎ fitted to
LES data. Symbol: LES data. The two vertical black solid lines represent the lower limit (0.3h above
roughness top) and the upper limit (0.7h above roughness top or 1/10 of the boundary layer depth 𝑧𝑖 ) of
ISL.

108
Figure 6.2: Profiles of normalized mean streamwise velocity corrected by the stability correction
1
function, − Ψ𝑚 (𝜁). Others are the same as Figure 6.1.
𝜅

1 𝑧𝑒𝑓𝑓
composed of three terms: (i) a logarithmic term, 𝜅 ln ( ) , that resembles the log-law at
𝑧0
1
neutral condition, (ii) a stability correction term, − 𝜅 Ψ𝑚 (𝜁) , and (iii) another stability
1
correction term for 𝑧0 , 𝜅 Ψ𝑚 (𝜁0 ). The term (iii) can be combined to 𝑧0 in term (i) to form a new
1 𝑧 1
logarithmic term, 𝜅 ln ( 𝑧𝑒𝑓𝑓′ ), where 𝑧0′ = 𝑧0 exp[− 𝜅 Ψ𝑚 (𝜁0 )], or it can be neglected since it’s
0

the deviation of 𝑧0′ from 𝑧0 is normally very small, less than 5% for the most unstable case of
the present LES study. Therefore, the mean velocity profile can be viewed as contributed
separately by the terms (i) and (ii) and each term represent a different physical mechanism. The
term (i) containing the parameter 𝑧0 represents the dynamical effect of roughness on the ISL
mean flow, which influences the bulk mean profile rather than modifies the local flow
properties. The term (ii), as discussed in Section 6.1, is from the MOST function and therefore
it represents the local effect of how the buoyancy induced turbulence modifies the local flux-
gradient transport properties in ISL and independent of any roughness effect.
To take a deeper look on the contribution of each term in Equation (6.3), the LES results
𝑈(𝑧) 1
of the mean velocity profile adjusted by the stability correction term (ii), + − 𝜅 Ψ𝑚 (𝜁), is
𝑢𝜏

compared to the parametric equation without the stability term, such that the sum of term (i)
1 𝑧𝑒𝑓𝑓 1 1 𝑧
and term (iii), 𝜅 ln ( ) + 𝜅 Ψ𝑚 (𝜁0 ) , or 𝜅 ln ( 𝑧𝑒𝑓𝑓′ ) in Figure 6.2 under different instability
𝑧0 0

109
levels. It shows that there are offsets among each curve. The offset is due to the variation of the
parameter 𝑧0 with the instability level, which causes a uniform offset for the mean profile as a
whole. Within the range of instability level simulated, the offset can be as high as ~1 (or the
deviation in mean velocity can be as high as one unit of friction velocity 𝑢𝜏 ). Therefore, the
effect of the instability dependence of 𝑧0 can be significant on the mean velocity profile,
especially when nearer to the wall since the magnitude of the stability correction term (ii)
decreases when nearer to the wall.
The above analysis can be similarly applied on the mean temperature profile and thus
not presented here for simplicity. The dependence of 𝑧0 of buoyant instability will be studied
further in the next section.

6.3 Aerodynamic and Temperature Roughness Lengths

The roughness lengths 𝑧0 and 𝑧ℎ appears as integration constants in the integration of


MOST mean gradient functions 𝜙𝑚 and 𝜙ℎ and are found here by fitting of derived profile
equations to the LES data within the range 1.3ℎ < 𝑧 < 1.7ℎ . The roughness length 𝑧0
extracted from LES of different aspect ratios is shown in Figure 6.3a, in which the physical
height of roughness ℎ𝑟 (= ℎ for models A0.5, A1 and A2, = 0.5ℎ for model AH0.5) is the
length scale for normalization.
The LES results of aspect ratios 1 and 2 show that 𝑧0 slightly decreases with instability
at near neutral state and it then increases clearly with instability level. For example, the model
A2 shows that 𝑧0 /ℎ𝑟 ~0.02 at ℎ𝑟 /𝐿 = 0 and 𝑧0 /ℎ𝑟 ~0.04 at ℎ𝑟 /𝐿 = −1 , which is
approximately double of the neutral value of 𝑧0 . The instability effect on 𝑧0 is shown to be
comparatively stronger for narrower street cavity (model A2) than for wider roughness cavity
(model A1), given the same instability level parameterized by ℎ𝑟 /𝐿. However, it is unsure why
𝑧0 decreases slightly under slightly unstable condition. It is guessed to be related to the
boundary layer scaled eddies generated by weak buoyancy force field as it is the only difference
between the neutral and slightly unstable flow fields. The near wall flow field of slightly
unstable condition should be the same as the neutral one if there is no influence from the outer
layer because of its small scale compared to the outer scale and thus there is negligibly small
locally buoyant effect. If the modulation effect on 𝑧0 by outers scale motions is true, then it
implies that eddies spanning the whole boundary layer are able to influence the near wall flow
properties that governs 𝑧0 .
110
Figure 6.3: a) Aerodynamic roughness length 𝑧0, b) temperature roughness length 𝑧ℎ , normalized
by building height ℎ𝑟 (= ℎ for models A0.5, A1 and A2, (= 0.5ℎ for model AH0.5), against instability
parameter −ℎ𝑟 /𝐿. Blue solid square: aspect ratio 1 (A1), red solid triangle: aspect ratio 2 (A2), green solid
circle: aspect ratio 0.5 (A0.5) and green open circle: aspect ratio 0.5 with half roughness dimension (AH0.5).

111
The results of 𝑧0 for A0.5 and AH0.5 compare well with each other, which means the
mesh resolution effect near roughness (e.g. number of meshes per roughness cavity in AH0.5
is less than in A0.5) and the effect of the blockage ratio (ratio of roughness dimension to
boundary layer depth, e.g. larger for A0.5) are not significant on 𝑧0 predicted by the LES. This
also implies that the Reynolds number effect on model A0.5 is negligible and thus it is in fully
rough regime. Similar to models A1 and A2, 𝑧0 in models A0.5 and AH0.5 decreases with
instability initially, but its decrease is stronger than those of models A1 and A2. It then stays
constant approximately afterward within the range of instability level of the present study and
it does not show any increasing trend further like the models A1 and A2. This will be explained
by the properties of flow structure around roughness elements next.
The above result shows that 𝑧0 not just depends on the geometrical configuration and
dimension of roughness, but also the buoyant instability. This dependence was also
demonstrated by the model and field measurement by Zilitinkevich et al. [8]. They suggested
the physical mechanism: the convective updrafts developing at side walls of roughness extend
upwards and provide extra resistances to the mean flow, so that a larger 𝑧0 is resulted. This is
hypothesized to be the cause of increase in 𝑧0 at sufficiently strong instability level (e.g. 𝑧0 in
ℎ𝑟 ℎ𝑟
model A1 starts to increase at ~ − 0.25 while 𝑧0 in model A2 starts to increase at ~−
𝐿 𝐿
ℎ𝑟
0.15). For model A0.5, the relatively constant trend of 𝑧0 with is believed to be due to its
𝐿

much stronger mechanical turbulence generated by the direct impingement of flow on the
building walls than the buoyantly generated convective updrafts. Therefore, it is hypothesized
that the instability level at which 𝑧0 starts to increase depends on the roughness geometrical
properties, e.g. aspect ratio for 2D urban roughness. Although both the present LES study and
the measurement study of Zilitinkevich et al. shows that 𝑧0 is a function of buoyant instability,
the behavior suggested by present study seems very different from the model proposed by
Zilitinkevich et al.. Their model suggested that 𝑧0 increases sharply with instability at the
neutral condition and the increasing trend decreases further. This cannot be observed from the
present study and rather the LES result suggests that 𝑧0 decreases slightly with instability at
neutral condition, then reaches a minimum and then increases thereafter at certain instability
level, which depends on the roughness geometrical configuration like aspect ratio, 𝐴𝑅. Finally,
the conclusion is that 𝑧0 /ℎ𝑟 can be formulated as a function:
𝑧0 ℎ𝑟 ℎ1 ℎ2 (6.8)
= 𝐹𝑚 ( , 𝐴𝑅, , , … )
ℎ𝑟 𝐿 ℎ𝑟 ℎ𝑟
where ℎ1 , ℎ2 , … are the length scales specifying the geometry of roughness.

112
Figure 6.3b shows that the LES result of temperature roughness length 𝑧ℎ . All models
of different aspect ratios show that 𝑧ℎ increases with instability initially, which is different from
the behavior of 𝑧0 . Depending on the aspect ratios, 𝑧ℎ then decreases slightly with different
amount and then stays constant with instability level. It is thought that the initial increase and
the sequent decreases in 𝑧ℎ is similarly governing by the large boundary layer eddies as
discussed in the previous paragraph. This also shows that the effect of large eddies in slightly
unstable condition is different on 𝑧0 and 𝑧ℎ , but the detailed mechanism is still unknown. For
model A1 and A2, 𝑧ℎ increases after certain instability level. This behavior is very similar to
that of 𝑧0 and it is thought that both 𝑧0 and 𝑧ℎ are governed by similar mechanism at
sufficiently strong instability level, which is the contribution of convective updrafts developed
on the building walls. Hence, 𝑧ℎ indeed depends on buoyant instability level.
On the other hand, the dissimilarity of 𝑧ℎ /ℎ𝑟 from models A0.5 and AH0.5 suggests
that the magnitude of 𝑧ℎ /ℎ𝑟 is different when the roughness physical dimension is different,
even though the roughness geometrical is exactly the same. This can be explained by the
Reynolds number effect, as discussed in Subsection 2.3.2, which states that heat is ultimately
transported by molecular diffusion at the very vicinity of roughness wall and thus depends on
molecular diffusivity, in contrary to the momentum transport that is dominated by pressure
force on roughness wall that is Reynolds number independent. Since the roughness elements
of model A0.5 are double in size relative to model AH0.5, the Reynolds number of flow around
roughness elements of model A0.5 is also double. From Figure 6.3b, it shows that 𝑧ℎ /ℎ𝑟 for
model AH0.5 is larger than that of A0.5, which means the efficiency of heat transport is stronger
when the Reynolds number is lower. This is analogous to the momentum transport of smooth
surface where momentum is ultimately transported through viscous diffusivity without
pressure (form) drag. For example, the friction factor (i.e. transport efficiency) of smooth
surface pipe flow decreases continuously with Reynolds number whatever the Reynolds
number.
The conclusion here is that 𝑧ℎ /ℎ𝑟 depends on roughness geometry, Reynolds number
and also buoyant instability. Since heat transfer depends on molecular diffusivity, the molecular
Prandtl number 𝑃𝑟 = 𝜈/𝐷𝑇 should also be a factor, which is, however, not studied by the
present study. Therefore, the function of 𝑧ℎ /ℎ𝑟 can be written as:
𝑧ℎ ℎ𝑟 ℎ𝑟 𝑢𝜏 ℎ1 ℎ2 (6.9)
= 𝐹ℎ ( , 𝐴𝑅, , 𝑃𝑟, , , … )
ℎ𝑟 𝐿 𝜐 ℎ𝑟 ℎ𝑟
Practically, 𝑃𝑟 can be taken out from the function since 𝑃𝑟 of air is a constant in most
applications.

113
A note to mention is about the Brutsaert’s relationship between 𝑧ℎ and Reynolds
number as discussed Subsection 2.3.2 in (Brutsaert [93]):
𝑧0
ln ( ) = 𝑎(𝑅𝑒 ∗ )0.25 − 2.0 (6.10)
𝑧ℎ
𝑧 𝑢
The Reynolds number defined by Brutsaert is 𝑅𝑒 ∗ = 0𝜈 𝜏. If we use the LES results to find the

empirical constant 𝑎, it is found that for both the models A0.5 and AH0.5 𝑎~1.4 under neutral
condition and 𝑎~1.5 under unstable condition at ℎ𝑟 /𝐿~ − 0.6. The value is quite similar to the
value found by Kanda et al. [6] (𝑎 = 1.29). Therefore, Brutsaert’s relationship is probably a
valid model for 𝑧ℎ and the empirical constant 𝑎 may be slightly depends on instability level.
Finally, to deduce the properties of the dependence between roughness lengths and
buoyant instability at moderately unstable level, both the results of 𝑧0 and 𝑧ℎ suggest that they
can be strengthened by the instability once the physical roughness length scale, e.g. ℎ𝑟 , is
comparable to the Obukhov length 𝐿 . The explanation is that the turbulent transport by
mechanical turbulence is characterized by a physical roughness length while the turbulent
transport by buoyant induced turbulence is characterized by 𝐿. As the mechanical turbulence
for wider roughness cavity is much stronger (e.g. A0.5 and AH0.5), it makes sense that the
instability level at which the roughness lengths start to increase is larger (and this may be the
reason that 𝑧0 and 𝑧ℎ for model A0.5 still stay constant even at the strongest unstable level
simulated by LES). The influence of roughness length by buoyant induced turbulence is further
investigated in the next section.

6.4 Relationships of Roughness Lengths with Buoyancy Induced


Turbulence at Roughness Scale

To study how buoyancy influences the roughness lengths, the vertical velocity
fluctuations 𝑤′ at building roof-level and inside roughness cavity are investigated. The vertical
velocity fluctuation rather than the horizontal ones is studied here since the vertical velocity
fluctuation near wall is less influenced by the outer boundary layer motions and are directly
induced by local buoyancy force while the streamwise and spanwise counterparts, which are
generated by both the shear-induced motions and the pressure redistribution from the vertical
fluctuation, are less distinct from effect of buoyancy.
The aerodynamic and temperature roughness lengths, 𝑧0 and 𝑧ℎ , together with the

114
Figure 6.4: a) Aerodynamic roughness length 𝑧0 , denoted by red open triangle Δ; temperature
roughness length 𝑧ℎ , denoted by black open square □;streamwise averaged standard deviation of vertical
̅̅̅̅̅̅̅
velocity at roof-level normalized by friction velocity 〈𝑤 ′ 𝑤 ′ 〉1/2 /𝑢 , scaled by 0.01, denoted by green cross
𝜏
×, plotted against stability parameter −ℎ𝑟 /𝐿; b) streamwise profiles of standard deviation of vertical
1/2
velocity, ̅̅̅̅̅̅̅
𝑤 ′ 𝑤 ′ /𝑢𝜏 at roof-level under different instabilities (block solid line: neutral, red dotted line:
most unstable, ℎ𝑟 /𝐿 = −1.19), 𝑥/𝑏 = −0.5 is the leeward wall corner and 𝑥/𝑏 = 0.5 is the windward wall
corner, where 𝑏 is the roughness cavity width.

115
̅̅̅̅̅̅〉1/2 at building
horizontally averaged vertical velocity fluctuation standard deviation 〈𝑤′𝑤′
roof-level (𝑧 = ℎ𝑟 ) are plotted against instability level −ℎ𝑟 /𝐿 in Figure. 6.4a. It shows that
̅̅̅̅̅̅〉1/2 basically stays constant with instability level when −ℎ𝑟 /𝐿 < 0.4 and then
roof-level 〈𝑤′𝑤′
̅̅̅̅̅̅〉1/2 against −ℎ𝑟 /𝐿 and those of
increases afterwards. It is noteworthy that the trends of 〈𝑤′𝑤′
and the roughness lengths, 𝑧0 and 𝑧ℎ , are comparable: they all start to increases with instability
̅̅̅̅̅̅〉1/2 at
level when −ℎ𝑟 /𝐿 > 0.4 . This finding implies that 𝑧0 and 𝑧ℎ are correlated to 〈𝑤′𝑤′
roof-level through some turbulent processes in unstable condition.
̅̅̅̅̅̅ along the roof-level at
The streamwise profiles of vertical velocity fluctuation 𝑤′𝑤′
different unstable stratification are plotted in Figure 6.4b. It shows that the entire streamwise
̅̅̅̅̅̅ keeps unchanged at slightly unstable condition but it increases with instability
profile of 𝑤′𝑤′
level at all point along the roof-level level when the instability level is sufficiently strong (e.g.
̅̅̅̅̅̅ increases linearly from
−ℎ𝑟 /𝐿 > 0.4 for model A1). Under neutral or near-neutral state, 𝑤′𝑤′
the leeward corner (𝑥/𝑏 = −0.5) to the windward corner (𝑥/𝑏 = 0.5). It is hypothesized that
̅̅̅̅̅̅ under neutral condition is mainly generated by the Kelvin-Helmholtz
the roof-level 𝑤′𝑤′
instability. The mechanism has been suggested by Letzel et al. [143]: turbulence is generated
by the Kelvin-Helmholtz instability locally due to the intense shear layer formed by the velocity
mismatch between the roughness cavity and the free-stream at the roof-level. However, when
̅̅̅̅̅̅ strengthens
the buoyant instability is strong enough, the entire streamwise profile of 𝑤′𝑤′
̅̅̅̅̅̅ appear near both the leeward and windward corners. The peaks at these
while peaks of 𝑤′𝑤′
two corners, which do not appear in neutral and near-neutral cases, are probably due to the
buoyancy-driven updraft developed near the leeward and windward walls which are at a higher
temperature.
The contours of instantaneous vertical velocity fluctuation 𝑤′ under neutral and strong
unstable conditions (ℎ𝑟 /𝐿 = −1.19) are shown in Figure 6.5. Under neutral condition (Figure
6.5a), the turbulent motions in roughness cavities and at the roof-level of cavity opening are
sporadic and small in scale compared to the turbulence above roof-level. The can be explained
by the difference in turbulence generation mechanisms at different locations: (i) turbulence in
the form of hairpin-like vortices are generated by streamwise streaky structures in ISL, which
is similar for both flows over rough and smooth surfaces (Coceal et al. [82]); (ii) turbulence is
generated by Kelvin-Helmholtz instability at the interface (roof-level) between the roughness
cavity and the outer free-stream layer, whose mechanism is similar to that of free shear layer
(Letzel et al. [143]) and (iii) turbulence generated at roof-level acts as Reynolds stress to drive
the recirculation inside cavity and the turbulence is transported down to the cavity core by
116
Figure 6.5: Contours of instantaneous vertical velocity fluctuation 𝑤′ normalized by 𝑢𝜏 on the
vertical x-z mid-plane of domain (𝑦 = 𝐿𝑦 /2) under a) neutral; and b) unstable condition (ℎ𝑟 /𝐿 = −1.19).
Both plots use the same scale and positive magnitude (red color) denotes upward velocity.

advection of cavity flow recirculation (Liu and Barth [145] and Liu et al. [146]).
Under unstable condition (Figure 6.5b), both the intensity and size of turbulent eddies
in the roughness cavities are larger than those under neutral condition. It is observed that the
instantaneous fluctuations 𝑤′ tend to extend from the cavity bottom to the outer free-stream
and they tend to develop near the upstream and downstream walls. These results are consistent
̅̅̅̅̅̅ in Figure 6.4b. These vertically aligned and
to the two peaks in the streamwise profile of 𝑤′𝑤′
elongated structures are the manifestation of the updrafts generated by buoyancy and the
downdrafts to replace the flow driven out by updrafts (continuity). The updrafts near heated
wall is also observed in some studies, e.g. the street cavity experiment by Louka et al. [147]
observed that a thin thermal layer developed locally within a few centimeters from the heated
wall and they visualized that convective flow close to windward wall carries air masses from
the street level upwards.
It is hypothesized that these vertically elongated buoyant structures promote the
efficiency of vertical exchanges of momentum and heat exchanges. Buoyant updrafts and
downdrafts both occur in ISL and RSL. The buoyant mixing processes in ISL have already
been accounted for by the stability correction functions Ψ𝑖 (𝜁). The roughness lengths 𝑧0 and
𝑧ℎ represents the absorption efficiency of momentum and release of thermal energy from the

117
bulk flow by the roughness. Therefore, it is hypothesized that 𝑧0 and 𝑧ℎ are governed by both
the shear-induced and buoyant-induced turbulent structures around roughness elements, as also
suggested by Zilitinkevich et al. [8], by modifying the transport efficiency through the
aforementioned mechanisms. It is also thought that the strength of updrafts is proportional to
the height of roughness since a deeper street cavity allows a longer time development of
updrafts while the mechanical turbulence is weaker due to a comparatively narrower cavity
opening width. This explains why roughness lengths of narrower cavity (e.g. model A2) tend
to be more sensitive to the buoyant instability. Therefore, ℎ𝑟 /𝐿 is an appropriate parameter to
describe the buoyant effect on the roughness lengths 𝑧0 and 𝑧ℎ .

6.5 Concluding Remarks

In this chapter, the equations of mean velocity and temperature profiles, 𝑈(𝑧) and 𝛩(𝑧),
are derived analytically by mathematically integrating the MOST functions proposed in
𝑈(𝑧)
Chapter 4. The normalized mean velocity profile (similarly for 𝛩(𝑧)) is composed of three
𝑢𝜏
1 𝑧𝑒𝑓𝑓
terms: (i) a logarithmic term, 𝜅 ln ( ), that resembles the log-law at neutral condition, (ii) a
𝑧0
1
stability correction term, − 𝜅 Ψ𝑚 (𝜁) , that represents the local effect of buoyancy-induced
1
turbulent mixing in ISL and (iii) another stability correction term for 𝑧0 , 𝜅 Ψ𝑚 (𝜁0 ), which can

be absorbed into 𝑧0 and it is of minor contribution (less than 5% on value of 𝑧0 ). The values of
aerodynamic and temperature roughness lengths 𝑧0 and 𝑧ℎ are fitted by matching the LES data
to the derived equations within ISL. The result shows that 𝑈(𝑧) and 𝛩(𝑧) calculated by LES
match the derived profile equations very well in ISL and also satisfactorily in RSL above
building roof-level. It also shows that 𝑧0 and 𝑧ℎ are actually depends on buoyant instability and,
if the variation of 𝑧0 with instability level is ignored, the error on mean profile can be
significant (deviation on 𝑈(𝑧) can attain one unit of friction velocity 𝑢𝜏 ). The error is
especially significant at the location nearer to the wall, where the magnitudes of mean profiles
are small and the effect of the stability correction term (ii) is also weak.
To test the variation of 𝑧0 and 𝑧ℎ with instability level, the ratio of building height to
Obukhov length, ℎ𝑟 /𝐿, is used and it is thought to be an appropriate parameter to describe to
how strong the buoyancy effect at roughness scale. It is found that 𝑧0 decreases slightly and 𝑧ℎ
increases slightly from their respective neutral values under slightly unstable condition and

118
they then stays constant further. It is hypothesized that the slight change is due to modulation
of roughness scale flow by the outer boundary layer eddies induced by the slight buoyancy
force rather than the direct modification of turbulent flow by local buoyancy force at roughness
scale. It is observed that both 𝑧0 and 𝑧ℎ start to increase under sufficiently strong instability
level (e.g. ℎ𝑟 /𝐿~ − 1 for aspect ratio 1). It is found that the critical instability level at which
the roughness lengths start to increase depends on the aspect ratio, e.g. 𝑧0 and 𝑧ℎ of narrower
roughness cavity tend to increase at weaker instability level (i.e. 𝑧0 of model A2 starts to
increase earlier than that of A1 while 𝑧0 of A0.5 still does no increase even at the strongest
unstable condition being simulated) and the increases of 𝑧0 and 𝑧ℎ are more significant with
ℎ𝑟 /𝐿 for narrower roughness cavity. The LES result of 𝑧ℎ from models A0.5 and AH0.5 also
confirms its dependence of Reynolds number and it satisfies the Brutsaert’s relationship of the
ratio 𝑧0 /𝑧ℎ and Reynolds number.
A closer look at the turbulence at roughness scale suggests that it is the updrafts
developed near the upstream and downstream building walls that strengthens the turbulent
mixing efficiency between the cavity and the above free-stream, which causes larger roughness
lengths 𝑧0 and 𝑧ℎ under strong unstable condition. This is reflected on the plot of roof-level
̅̅̅̅̅̅ , which shows two peaks at the
streamwise profile of vertical velocity fluctuation 𝑤′𝑤′
upstream and downstream corners, as well as the plot of instantaneous vertical velocity
fluctuation 𝑤′ around roughness elements on the x-z plane, which shows that the turbulent
motions are elongated in vertical direction and tend to developed next to the building walls.

119
Chapter 7

Mechanism of Turbulent Transports at Near-Wall


Region

To understand the mechanism of how momentum and temperature are transported by


turbulence from the rough surface to the bulk flow and gain a deeper insight of the
corresponding flow statistics at the near-wall region of unstable UBL, the LES-simulated flow
fields will be investigated by considering their turbulent kinetic energy (TKE) budget and
coherent structures. The TKE budget study includes a one-dimensional (vertical) analysis
above the building roof-level and a two-dimensional analysis at the roughness scale. The study
of coherent structures includes two-point correlation, instantaneous flow visualization and
conditional average. Many recent studies in the literature suggest that MOST and Townsend’s
hypothesis are in fact not straightly valid. The present LES study would allow us to test whether
those hypotheses are correct or not and subsequently reveal the reason behind.

7.1 Turbulent Kinetic Energy (TKE) Budget at Roughness Scale

The TKE budget equation applicable to the present LES study at roughness scale is:
𝜕𝑘̅ 𝜕𝑘̅ ̅̅̅̅̅ 𝜕𝑢̅ − ̅̅̅̅̅̅ 𝜕𝑢̅ ̅̅̅̅̅̅ 𝜕𝑤̅ ̅̅̅̅̅̅ 𝜕𝑤
̅
𝜕𝑘̅ [−𝑢̅ −𝑤̅ ] [−𝑢′𝑢′ 𝑢′𝑤′ − 𝑢′𝑤′ − 𝑤′𝑤′ ]
= ⏟ 𝜕𝑥 𝜕𝑧 + ⏟ 𝜕𝑥 𝜕𝑧 𝜕𝑥 𝜕𝑧
𝜕𝑡
𝐴 𝑃𝑠
𝜕 ̅̅̅̅ 𝜕 ̅̅̅̅̅ 𝜕 ̅̅̅̅̅ 𝜕 ̅̅̅̅̅̅
[− (𝑘𝑢′ ) − (𝑘𝑤′ )] [− (𝑝′𝑢′ ) − (𝑝′𝑤′ )] (7.1)
+ ⏟ 𝜕𝑥 𝜕𝑧 + ⏟ 𝜕𝑥 𝜕𝑧
𝑇𝑡 𝑇𝑝
̅̅̅̅̅̅̅̅̅̅

𝜕𝑢 𝜕𝑢 ′
0 𝜈 𝑖 𝑖 ⏟
⏟ ̅̅̅̅̅̅
𝛼𝑔𝑤 ′𝜃′
+ − ⏟𝜕𝑥𝑗 𝜕𝑥𝑗 + =0
𝑇𝑣 𝐵
𝜀

120
1
where 𝑘 = 2 𝑢𝑖′ 𝑢𝑖′ is the instantaneous TKE, 𝐴 is the mean advection term, 𝑃𝑠 is the mean-

gradient (shear) production term, 𝑇𝑡 is the turbulent transport term, 𝑇𝑝 is the pressure transport
term, 𝑇𝑣 is the viscous diffusion term, 𝜀 is the viscous dissipation term and 𝐵 is the buoyant
production term. The derivation of Equation (7.1) is shown in Appendix A.2. Assumptions
based on the conditions of the LES configurations have been applied to the derivation of
̅
𝜕𝑘
Equation (7.1): quasi-steadiness (i.e. = 0), spanwise homogeneity (i.e. two-dimensionality)
𝜕𝑡

and negligible viscous diffusion (i.e. 𝑇𝑣 = 0). In this LES study, the terms of TKE budget are
calculated by resolved scale flow variables only while the SGS parts are ignored. In analyzing
the LES results, an alternative budget equation is used:

0 = 𝐴 + 𝑃𝑠 + 𝑇𝑡 + 𝑇𝑝 + 𝐵 + 𝑅 (7.2)
where 𝑅 is the residual term accounting for the viscous dissipation and errors due to neglects
of temporal term, spanwise variation, viscous diffusion and SGS contributions which are small
but in fact exist due to the imperfect LES solution and finite samples of flow field realization.
The 2D contours of different budget terms, normalized by characteristic velocity scale 𝑢𝜏 and
length scale ℎ, are shown in Figure 7.1 and Figure 7.2 for the aspect ratio 1 (model A1) and
aspect ratio 0.5 (model A0.5), respectively. The figures include a neutral state and an unstable
state (ℎ𝑟 /𝐿 = −1.19 for A1 and ℎ𝑟 /𝐿 = −1.20 for A0.5).
Considering the model A1 (Figure 7.1) first, the shear production 𝑃𝑠 (Figure 7.1b)
dominates the TKE budget for both the neutral and the unstable states at the building roof-level
interface at which the mean streamwise velocity profile is inflected. Comparing the neutral and
unstable states, the neutral state shows that 𝑃𝑠 develops gradually from the upstream building
corner to the downstream corner in both strength and vertical extent while the unstable state
shows that there are two peaks, one near the upstream corner and one near the downstream
corner. The dissimilarity implies their modes of turbulence production at the interface are
difference, but their amount of production are similar. At the unstable state, another part of
turbulence production is from the buoyancy, indicated by the buoyant production term 𝐵
(Figure 7.1f). It shows that 𝐵 is strongest along the downstream building wall but not at the
remaining part of cavity. 𝐵 is also significant at the roof-level opening interface and it tends to
be more horizontally homogenous with the increase of height. However, the buoyant
production 𝐵 is much weaker than the shear production 𝑃𝑠 , at about an order of magnitude, at
the locations of strongest magnitude. This indicates the shear production is the main generation
mechanism at roughness level even at such a strong unstable level (ℎ𝑟 /𝐿 = −1.19). From the

121
Figure 7.1: See next page for caption. (cont.)

122
Figure 7.1: 2D contours of TKE budget terms of model A1 at roughness scale: a) mean advection
term 𝐴; b) mean gradient (shear) production term 𝑃𝑠 ; c) turbulent transport term 𝑇𝑡 ; d) pressure transport
term 𝑇𝑝 ; e) residual term 𝑅; f) buoyant production term 𝐵, normalized by 𝑢𝜏3 /ℎ, under i) neutral condition
and i) unstable condition, (ℎ𝑟 /𝐿 = −1.19).

123
previous chapter, however, it shows that the vertical velocity fluctuation is much stronger at
the roof level under the unstable state (Figure 6.4b). This may suggest that the extra turbulence
at roughness scale is actually from the upper part of the boundary layer and therefore the top-
down mechanism, e.g. the one proposed by Laubach and McNaughton [11], seems to be more
supported than the local mechanism.
The TKE budget also unravels how the turbulence is transported from one location to
the others by mean advection 𝐴 (Figure 7.1a), turbulent transport 𝑇𝑡 (Figure 7.1c) and pressure
transport 𝑇𝑝 (Figure 7.1d). At neutral state, 𝐴 is negative at the upstream corner and extending
to the downstream along roof-level while 𝐴 is positive around the downstream corner, which
represents the turbulence generated at upstream is advected to the downstream by mean flow.
Similar pattern with stronger magnitude is observed at roof-level under unstable condition.
However, it shows that 𝐴 is very strong above roof-level and its influence extends to the higher
ISL. It shows a strong TKE gain above upstream corner while a strong but relatively weaker
TKE loss above the downstream corner by mean advection. This can be explained by the mean
velocity flow field that the mean flow recirculation, which is confined within the cavity below
roof-level under neutral condition, penetrates the roof-level opening to the above under
unstable condition, as discussed in Section 5.3. The turbulent transport 𝑇𝑡 (Figure 7.1c) shows
that it merely transports TKE generated by shear to the downstream corner at neutral state. It
is very different under unstable state, which shows that TKE is transported by 𝑇𝑡 from upper
level to the roof-level, which may explain the strong vertical velocity fluctuation at roof-level,
though more evidences are needed to conform this argument. The 2D pattern of pressure
transport 𝑇𝑝 (Figure 7.1d) is similar to both neutral and unstable states. 𝑇𝑝 is only active around
the downstream corner, which is thought to be due to the unsteady pressure force by the
unsteady flow impingement on the upper downstream building wall. A notable difference is the
much stronger gain by 𝑇𝑝 at the right side of the downstream corner and a non-negligible gain
along the building roof at unstable state. This may be due to the top-down turbulent
impingement on the building roof, but this argument is not ensured. Finally, the TKE
dissipation is represented by the residual term 𝑅 (Figure 7.1e), which is approximate rather
than exact as it is calculated by residual of other budget terms and based on the assumptions
mentioned previously. It shows that the dissipation is mainly confined at the building roof-level
and the upper downstream building wall. It is active starting at the downstream corner
following the building roof and then it becomes weaker at the cavity opening. The pattern of
dissipation is quite similar for both neutral and unstable states, although it is comparative

124
Figure 7.2: See next page for caption. (cont.)

125
Figure 7.2: 2D contours of TKE budget terms of model A0.5 at roughness scale: a) mean advection
term 𝐴; b) mean gradient (shear) production term 𝑃𝑠 ; c) turbulent transport term 𝑇𝑡 ; d) pressure transport
term 𝑇𝑝 ; e) residual term 𝑅; f) buoyant production term 𝐵, normalized by 𝑢𝜏3 /ℎ, under i) neutral condition
and ii) unstable condition, (ℎ𝑟 /𝐿 = −1.19).

stronger in unstable state, which may be due to the additional buoyant production.
To investigate how the physical geometry of roughness affects the turbulent mechanism,
the TKE budget terms for aspect ratio 0.5 are also analyzed, as shown in Figure 7.2. At neutral
state, each budget term for aspect ratio 0.5 are not much different from those of aspect ratio 1,
except the shear production is mainly significant near upstream corner. This implies the flows
at both aspect ratios are governed by the similar mechanism around roof-level: the Kelvin-
Helmholtz instability by the inflected velocity profile at the roof-level interface (Letzel et al.
[143]). At unstable state, however, the budget terms are quite different between these two
aspect ratios at the upstream corner, where mean advection 𝐴, shear production 𝑃𝑠 , turbulent
transport 𝑇𝑡 and pressure transport 𝑇𝑝 are much stronger and their regions of influence extend
to a very high location for model A0.5 compared to model A1. This is thought to be due to the
unsteady buoyant motions at the upstream corner of model A0.5 that causes strong spatial, both
streamwise and vertical, variation of the turbulent fluctuation fields, which are more uniform

126
for A1 because of its relatively skimming nature of roof-level flow. This is indicated by the
buoyant production 𝐵 (Figure 7.2f), which shows a stronger peak that is more vertical extended
at the upstream corner compared to model A1. Apart from the behavior at the upstream corner,
others are quite similar between these two roughness configurations.

7.2 Horizontally Averaged Turbulent Kinetic Energy (TKE)


Budget above Building Roof-level

To study how turbulence is generated, dissipated and transported in the vertical


direction above urban roughness, including RSL and ISL, the horizontally averaged TKE
budget equation is employed:


0
𝜕
− 〈𝑘̅"𝑤 ̅̅̅̅̅̅〉 𝜕〈𝑢̅
̅"〉 −〈𝑢′𝑤′
+ ⏟ 𝜕𝑧 +⏟ 𝜕𝑧
𝐴′
𝑇𝑑 𝑃𝑠 ′

̅̅̅̅̅" 𝜕𝑢̅"〉 − 〈𝑢′𝑤′


[− 〈𝑢′𝑢′ ̅̅̅̅̅̅" 𝜕𝑢̅"〉 − 〈𝑢′𝑤′
̅̅̅̅̅̅" 𝜕𝑤̅" ̅̅̅̅̅̅" 𝜕𝑤
〉 − 〈𝑤′𝑤′
̅"
〉]
+⏟ 𝜕𝑥 𝜕𝑧 𝜕𝑥 𝜕𝑧 (7.3)
𝑃𝑤
𝜕 ̅̅̅̅̅ 𝜕 ̅̅̅̅̅̅ ̅̅̅̅̅̅̅̅̅̅
𝜕𝑢𝑖′ 𝜕𝑢𝑖′
− 〈𝑘𝑤′ 〉 − (𝑝′𝑤′ ) ⏟ 0 𝜈 〈 ⏟ ̅̅̅̅̅̅
〉 𝛼𝑔〈𝑤 ′𝜃′〉
+ ⏟ 𝜕𝑧 + ⏟ 𝜕𝑧 + − ⏟ 𝜕𝑥𝑗 𝜕𝑥𝑗 + =0
𝑇𝑣 ′ 𝐵 ′
𝑇𝑡 ′ 𝑇𝑝 ′ 𝜀′
which is derived in Appendix A.3. Two new terms, dispersive transport 𝑇𝑑 and wake production
𝑃𝑤 , arise due to horizontal averaging, while others (denoted by prime) are similar to the ones
of the two-dimensional version. 𝑇𝑑 and 𝑃𝑤 are expected to be significant within RSL only since
these two terms are due to the horizontal (i.e. streamwise for this study) inhomogeneity of
ensemble averaged mean flow and turbulent fields. For the wake production 𝑃𝑤 , it has been
̅̅̅̅̅̅" 𝜕𝑢̅"〉 is the dominant component while the other three are about two orders
tested that − 〈𝑢′𝑤′ 𝜕𝑧

of magnitude less (result not shown here). In analyzing the LES results, the Equation (7.3) is
alternatively written as the simplified version:

0 = 𝑇𝑑 + 𝑃𝑠′ + 𝑃𝑤 + 𝑇𝑡′ + 𝑇𝑝′ + 𝐵 ′ + 𝑅′ (7.4)


where 𝑅′ is the residual of the other terms. The vertical profiles of the budget terms in Equation
(7.4) are shown in Figure 7.3 for aspect ratio 1 (model A1) and Figure 7.4 for aspect ratio 0.5
(model A0.5), respectively, both including a neutral case and a strong unstable case (ℎ𝑟 /𝐿~ −
1.20).
127
Figure 7.3: Vertical profiles of horizontally averaged TKE budget terms for model A1 above
roughness roof-level: a) neutral condition and b) unstable condition, (𝑧𝑖 /𝐿 = −8.3).

The neutral cases of model A1 (Figure 7.3a) and model A0.5 (Figure 7.4a) show that
the TKE budget equation reduces to a balance between shear production 𝑃𝑠′ and dissipation
represented by 𝑅′ above 𝑧 = 1.3ℎ for model A1 and 𝑧 = 1.4ℎ for model A0.5. Therefore, from
the perspective of TKE budget, the turbulent behavior in ISL under neutral condition is “local”,
i.e. turbulence is generated and dissipated at the same local height, and there is a limit of the

128
Figure 7.4: Vertical profiles of horizontally averaged TKE budget terms for model A0.5 above
roughness roof-level: a) neutral condition and b) unstable condition, (𝑧𝑖 /𝐿 = −8.4).

influence of roughness, so that the wall similarity hypothesis is supported, at least for the
roughness configurations (A1 and A0.5) being studied. Within the region where roughness
effect is significant, turbulent transport 𝑇𝑡′ and wake production 𝑃𝑤 are the secondary dominant
terms next to 𝑃𝑠′ and 𝑅′ . Both 𝑇𝑡′ and 𝑃𝑤 are negative within the region, which means they
represent loss of TKE there. TKE loss through 𝑇𝑡′ implies TKE is transferred from this region

129
to other region(s), which is possibly inside the cavity. On the other hand, the loss of TKE
through wake production 𝑃𝑤 is peculiar since most previous field studies suggested that 𝑃𝑤 is
a gain term rather than a loss term. 𝑃𝑤 is physically interpreted as the turbulence created in the
wakes of buildings due to the horizontal variation of Reynolds stress and velocity gradient
(Raupach et al. [12] and Christen et al. [115]). However, those studies did not explicitly
calculate 𝑃𝑤 since a large horizontal array of probes is needed in order for explicit calculation.
Rather, those studies usually “approximate” 𝑃𝑤 by interpreting it as the rate of working of mean
̅̅̅̅̅̅̅〉
𝜕〈𝑢′𝑤′
flow against drag, i.e. 𝑃𝑤 = −〈𝑈〉 by Raupach et al. [148]. This “approximation” method
𝜕𝑧

leads to positive values of 𝑃𝑤 for the present LES result, which is contradictory to the explicit
calculation method that leads to negative values. Therefore, the “approximate” method seems
to be wrong and 𝑃𝑤 seems to be a loss of TKE in most of the RSL. Comparatively, pressure
transport 𝑇𝑝′ and dispersive transport 𝑇𝑑 are much smaller in magnitude.
Under unstable condition, both the model A1 (Figure 7.3b) and the model A0.5 (Figure
7.4b), similar behaviors in budget terms in ISL (approximately above 𝑧 = 1.2ℎ for both
models): the sum of shear production 𝑃𝑠′ and buoyant production 𝐵′ balance the dissipation 𝑅′
while turbulent transport 𝑇𝑡′ (a loss term) balances pressure transport 𝑇𝑝′ (a gain term),
consistent to most field studies (e.g. Wyngaard and Cote [112]) while dispersive transport 𝑇𝑑
and wake production 𝑃𝑤 are negligible. Within RSL, the behavior of TKE budget is more
complicated and all the terms are significant. 𝑃𝑠′ and 𝑅′ are weaker for A0.5 compared to A1.
𝑇𝑝′ is a gain term and it is much stronger in A0.5. 𝑇𝑡′ is a loss term in the upper part of RSL but
a gain term in the lower part for both A0.5 and A1. 𝑃𝑤 is a loss term for A1 while it is a loss in
the lower part of RSL but a gain term in the upper part for A0.5. 𝑇𝑑 is basically a loss term.
These complicated patterns are thought to be due to the strong horizontal inhomogeneity of
ensemble mean flow and turbulent fields caused by the buoyant instability.
The above results for RSL show that (i) the turbulence generated under strong unstable
condition is not locally dissipated, but transported vertically downward towards the cavity
region since turbulent transport 𝑇𝑡′ , pressure transport 𝑇𝑝′ and dispersive transport 𝑇𝑑 are gain
terms near roof-level; (ii) both wake production 𝑃𝑤 and dispersive transport 𝑇𝑑 can be
significant under unstable condition; and (iii) wake production 𝑃𝑤 is a loss term rather than a
gain term and the method of Raupach et al. [148] to calculate 𝑃𝑤 seems to be not feasible.

130
7.3 Visualization of Instantaneous Turbulent Fields

The understanding of the mechanism of turbulent processes and coherent structures can
be gained through visualizing single realizations of instantaneous fields of flow fluctuations
(e.g. velocity 𝑢′ , 𝑤 ′ and temperature 𝜃 ′ ) and fluxes (e.g. 𝑢′ 𝑤′ and 𝜃 ′ 𝑤′ ). From this
instantaneous visualization study, the size, mode and distribution of different types of coherent
structures can be observed qualitatively. This also allows us to explain the corresponding
ensemble statistics. Figure 7.5 and Figure 7.6 show the x-z fields of 𝑢′ , 𝑤′ and 𝜃 ′ under a
neutral state and a strong unstable state, respectively, for the whole domain.
Comparing the two instability levels shows that streamwise fluctuation 𝑢′ (Figure 7.5a
and Figure 7.6a) above roughness is dominated by small scale structures under neutral state but
large scale structures (at the scale of boundary layer depth 𝑧𝑖 ) under unstable state. It seems
that the size of 𝑢′ in neutral state increases with height while the probability (or spatial density)
of occurrence decreases with height. 𝑢′ in unstable state seems to form in bulk shape and
occupy half the boundary layer for each bulk structure, which occurs in alternatingly (i.e.
positive and negative 𝑢′ ) in the streamwise direction. This suggests that 𝑢′ under unstable
condition is controlled by the boundary layer scaled eddies that occur periodically in the
horizontal directions (the figure only shows the streamwise one).
Under neutral condition, the vertical velocity fluctuation 𝑤 ′ (Figure 7.5b) is smaller in
size and occurs in less frequency compared to 𝑢′ while the shapes of 𝑤 ′ are spot-like while the
shapes of 𝑢′ is more elongated in streamwise direction. Under unstable condition, 𝑤 ′ (Figure
7.6b) in the core of boundary layer is vertical extended, scaled by 𝑧𝑖 and occurs periodically in
streamwise direction as 𝑤 ′ is controlled by the large-scale buoyant convection at the core
region. Unlike 𝑢′ , 𝑤 ′ at the near wall-region is dominated by the small scale structures and it
is vertically elongated rather than spot-like in shape, consistent to the local scaling (i.e. MOST,
scaled by 𝑧) and 𝑤 ′ is dominated by the active turbulence at the near-surface.
Comparing the temperature fluctuations 𝜃 ′ under neutral (Figure 7.5c) and unstable
conditions (Figure 7.6c), it shows that the near-wall region of both instabilities is dominated
by small scale structures elongated in streamwise direction and tilted upwards, with stronger
tilting at unstable state. This suggests that 𝜃 ′ at near wall is also dominated by act motions at
both instabilities. The near-wall 𝜃 ′ at neutral state is quite similar to 𝑢′ , which is consistent to
the Reynolds analogy that momentum and heat transports are similar. However, the near-wall

131
Figure 7.5: Instantaneous fluctuation flow fields in x-z plane at neutral condition for model A1:
(a) streamwise velocity fluctuation 𝑢′; (b) vertical velocity fluctuation 𝑤′ and (c) temperature fluctuation
𝜃′, normalized by friction velocity 𝑢𝜏 and temperature 𝜃𝜏 .

132
Figure 7.6: Instantaneous fluctuation flow fields in x-z plane at unstable condition (𝑧𝑖 /𝐿 = −8.3)
for model A1: (a) streamwise velocity fluctuation 𝑢′; (b) vertical velocity fluctuation 𝑤′ and (c) temperature
fluctuation 𝜃′, normalized by friction velocity 𝑢𝜏 and temperature 𝜃𝜏 .

133
Figure 7.7: Instantaneous vertical velocity fluctuation 𝑤′ fields in y-z plane at neutral condition for
model A1 at different heights: (a) 𝑧 = 0.9ℎ; (b) . 𝑧 = 1.01ℎ; (c) . 𝑧 = 1.1ℎ; (d) . 𝑧 = 1.5ℎ; (e) . 𝑧 = 2ℎ;
(f) . 𝑧 = 3.5ℎ.

134
Figure 7.8: Instantaneous vertical velocity fluctuation 𝑤′ fields in y-z plane at unstable condition
(𝑧𝑖 /𝐿 = −8.3) for model A1 at different heights: (a) 𝑧 = 0.9ℎ; (b) . 𝑧 = 1.01ℎ; (c) . 𝑧 = 1.1ℎ; (d) . 𝑧 =
1.5ℎ; (e) . 𝑧 = 2ℎ; (f) . 𝑧 = 3.5ℎ.

135
𝜃 ′ at unstable state is very different to 𝑢′ in scale. At the core region of boundary layer, 𝜃 ′
continues to be tilted and increases in size with height under neutral condition while the
structure of 𝜃 ′ is sparse and vertically extended through the whole boundary layer depth,
roughly following the pattern of 𝑤 ′ . The above results suggest that 𝑢′ and 𝜃 ′ are highly
correlated in neutral condition (satisfying the Reynolds analogy) while 𝑤 ′ and 𝜃 ′ are highly
correlated in unstable condition, at which the large scale buoyant convection force controls
both the vertical velocity and temperature structures at the cores.
To observe the spatial structures of fluctuation in a deeper way, 𝑤 ′ on the horizontal
plan at different heights is shown in Figure 7.6 and Figure 7.7 for the neutral and unstable
conditions, respectively. 𝑤 ′ is focused here since it is the essential links of vertical transports
of momentum and heat. Under neutral condition (Figure 7.6), it clearly shows that the 𝑤 ′
structure increases in size (spanwise and streamwise) from the roughness cavity to the
boundary layer core: (i) the structure is scaled by roughness size in cavity (𝑧 = 0.9ℎ) and at
roof-level interface (𝑧 = 1.01ℎ); (ii) the structure is scaled by local height 𝑧 in the near wall
region (𝑧 = 1.1ℎ, 𝑧 = 1.5ℎ and 𝑧 = 2ℎ), but (iii) the structure ceases to increase in size above
( 𝑧 = 3.5ℎ ). These structures are fairly uniform at all heights above roughness. The
development of the spatial structure of 𝑤 ′ in vertical direction actually signifies the
development of hairpin vortices in self-similar fashion at the near-wall region. Under unstable
condition (Figure 7.7), it shows that 𝑤 ′ is much stronger and larger in size at the regions of
cavity (𝑧 = 0.9ℎ) and roof-level (𝑧 = 1.01ℎ), which are also discussed in Section 6.4. At 𝑧 =
1.1ℎ, the shapes of 𝑤 ′ is not quite uniform as in neutral state: some are narrow in shape similar
to the neutral ones and some are bulge-like. It is thought that different structures are due to
different mechanisms: the narrow ones are shear-induced and the bulge-like ones are buoyancy-
induced. Further away from the surface, the non-uniformity of 𝑤 ′ structure strengthens, which
is due to the gradual dominance of boundary layer eddying motions with increase of height.
To investigate how momentum and heat are transported by turbulence from the wall to
free-stream, the instantaneous momentum flux 𝑢′ 𝑤′ and heat flux 𝜃 ′ 𝑤 ′ in vertical (x-z) plane
are investigated in Figure 7.9 and Figure 7.10 for neutral and unstable states, respectively. In
order to understand the nature of transport mechanism, the quadrant analysis, by which 𝑢′ 𝑤 ′
and 𝜃 ′ 𝑤 ′ are divided into their respective four quadrant parts according to the sign of 𝑢′ , 𝑤 ′
and 𝜃 ′ , is employed. The quadrants of 𝑢′ 𝑤′ are defined as:
𝑢′𝑤′|𝑞1 : 𝑢′ > 0, 𝑤 ′ > 0 (outward interaction)
𝑢′𝑤′|𝑞2 : 𝑢′ < 0, 𝑤 ′ > 0 (ejection) (7.5)
𝑢′𝑤′|𝑞3 : 𝑢′ < 0, 𝑤 ′ < 0 (inward interaction)

136
𝑢′𝑤′|𝑞4 : 𝑢′ > 0, 𝑤 ′ < 0 (sweep)
and the quadrants of 𝜃 ′ 𝑤 ′ are defined as:
𝜃′𝑤′|𝑞1 : 𝜃 ′ > 0, 𝑤 ′ > 0 (hot updraft)
𝜃′𝑤′|𝑞2 : 𝜃 ′ < 0, 𝑤 ′ > 0 (outward interaction) (7.6)
𝜃′𝑤′|𝑞3 : 𝜃 ′ < 0, 𝑤 ′ < 0 (cold downdraft)
𝜃′𝑤′|𝑞4 : 𝜃 ′ > 0, 𝑤 ′ < 0 (inward interaction)
The ejection 𝑢′𝑤′|𝑞2 and the sweep 𝑢′𝑤′|𝑞4 contribute to the downward momentum flux while
the outward interaction 𝑢′𝑤′|𝑞1 : 𝑢′ and the inward interaction 𝑢′𝑤′|𝑞4 contribute to the upward
momentum flux, whose magnitude on average is much less than the downward flux. Therefore,
this study focuses on the ejection 𝑢′𝑤′|𝑞2 and the sweep 𝑢′𝑤′|𝑞4 . 𝑢′𝑤′|𝑞2 is represented by
positive value (red) while 𝑢′𝑤′|𝑞4 is represented by negative value (blue) in Figure 7.9a and
Figure 7.10a. Similarly, the hot updraft 𝜃′𝑤′|𝑞1 and the cold downdraft 𝜃′𝑤′|𝑞3 contribute to
the upward heat flux. 𝜃′𝑤′|𝑞1 is represented by positive value (red) while 𝜃′𝑤′|𝑞3 is
represented by negative value (blue) in Figure 7.9b and Figure 7.10b.
In neutral state, Figure 7.9 shows that the near-wall transporting (active) eddies,
including ejection and sweep for momentum flux as well as hot updraft and cold downdraft for
heat flux, are dominated by small scale structures. It also shows that the near-wall structures of
momentum flux and heat flux are very similar in shape and they tend to form at the same
locations. This implies both near-wall momentum flux and heat flux under neutral condition
are governed by similar mechanism, which is the ejection and inrush induced by hairpin
vortices. This argument is consistent to the Reynolds analogy specifying the similarity between
momentum and heat transfers. Away from the wall, the transporting eddies detach from the
wall and their sizes increase with height. It seems that the detached eddies are distributed in
random locations. Overall, it seems that the occurrence of upward transporting motions (i.e.
ejection and hot updraft) is in similar proportion to the downward transporting motions (i.e.
sweep and cold downdraft). It is noted that the transporting motions of heat flux is stronger
than the momentum flux at the upper region of boundary layer, which is due to the different
momentum and heat settings in LES: the flow momentum is driven by background pressure
gradient, which leads to a linearly decreasing shear stress with height on average, while the
heat flux is driven by the heat temperature at bottom wall, which leads to a vertically uniform
heat flux on average.
Under unstable condition, both the momentum and heat transporting eddies tend to
aggregate into larger structures that occupy a significant region of boundary layer, which
implies these transporting elements are governed by boundary layer scaled motions. It also

137
Figure 7.9: Instantaneous fields of fluxes in x-z plane at neutral condition for model A1: (a)
momentum flux 𝑢′𝑤′, where (i) ejection motion (𝑢′ < 0, 𝑤 ′ > 0) is positive (red); (ii) sweep (𝑢′ > 0, 𝑤 ′ <
0) is negative (blue); (b) heat flux 𝜃′𝑤′, where (i) hot updraft motion (𝜃 ′ > 0, 𝑤 ′ > 0) is positive (red); (ii)
cold downdraft (𝜃 ′ < 0, 𝑤 ′ < 0) is negative (blue). Normalization by friction velocity 𝑢𝜏 and temperature
𝜃𝜏 is used.

shows that much smaller scale motions accompany the large-scale motions. At the near-wall
region, both small- and large-scale transporting motions are found. These small scale motions
may be the Richardson cascade eddies generated from the large-scale eddies, whose process is
proposed by Laubach and McNaughton [11]. A notable difference from the neutral state is that
significant amount of transporting eddies are vertically extended from the wall-region to the
boundary layer core. The momentum and heat fluxes under unstable condition are found to be
rather dissimilar, especially at the boundary layer core region, in which the large-scale heat flux
carrying motions tend to attach to the wall and form at vertically elongated shape while the

138
Figure 7.10: Instantaneous fields of fluxes in x-z plane at unstable condition (𝑧𝑖 /𝐿 = −8.3) for
model A1: (a) momentum flux 𝑢′𝑤′, where (i) ejection motion (𝑢′ < 0, 𝑤 ′ > 0) is positive (red); (ii) sweep
(𝑢′ > 0, 𝑤 ′ < 0) is negative (blue); (b) heat flux 𝜃′𝑤′, where (i) hot updraft motion (𝜃 ′ > 0, 𝑤 ′ > 0) is
positive (red); (ii) cold downdraft (𝜃 ′ < 0, 𝑤 ′ < 0) is negative (blue). Normalization by friction velocity
𝑢𝜏 and temperature 𝜃𝜏 is used.

large-scale momentum carrying motions are detached from the wall and comparatively bulge-
shaped.
The above results concerning the instantaneous flow and temperature structures suggest
that, under sufficiently strong unstable condition, (i) the near-wall turbulent transports are
multiscale phenomena, containing both local scale and boundary layer scale transporting eddies,
which implies the Townsend’s hypotheses (i.e. large scale eddies does not participate the
turbulent transports of momentum and heat) is not strictly valid; and (ii) momentum and heat
transporting eddies become dissimilar in shape, size and occurring locations, which means

139
momentum and heat transports are governed by different transporting eddies and thus the
Reynolds analogy is not applicable.
Finally, to understand more about the role of large scale motions at the near-wall region,
the instantaneous fluctuation and flux fields are low-pass filtered (or moving-averaged) in
streamwise direction, whose concepts and functions are discussed by Castillo et al. [110].
Spatial filtering allows us to the influence of the large scale eddies on the local scale eddies at
the near-wall region as well as the interaction of motions at multi-scale. The streamwise filter
size used is 10ℎ , which is sufficient to remove most structures of scale smaller than the
boundary layer depth 𝑧𝑖 . The filtered instantaneous structures at the near-wall region are shown
in horizontal (x-y) plane at two heights, (i) 𝑧 = 1.01ℎ (roof-level, within RSL) and (ii) 𝑧 =
1.5ℎ (mid-ISL), in Figure 7.11 and 7.12 at neutral state and in Figure 7.13 and 7.14 at unstable
state.
Under neutral condition, Figure 7.11 shows that the roof-level (𝑧 = 1.01ℎ ) filtered
fluctuation structures, 𝑢′, 𝑤′ and 𝜃′, are streaky and narrow in shape. These streamwise streaks
represent there is a certain organizing mechanism for the near-wall turbulence. This mechanism
is thought to be the packet structures proposed by Adrian et al. [120], who suggests that hairpin
vortices tend to from successively in the streamwise between these streamwise streaks. For the
momentum and heat transporting motions (Figure 7.11d and e), the streaky structures are also
easily observed, but they are thicker and less alternating in spanwise direction. These suggest
that the turbulent transports are organized at certain region rather than distributed evenly on the
x-z plane. At the mid-ISL (𝑧 = 1.5ℎ) in Figure 7.12, the streaky structures of 𝑢′ and 𝜃′ are
much thicker than the roof-level ones, but the streaky structures of 𝑤′ are relatively narrow.
This is consistent to the laboratory finding that the hairpin-vortices grow in size vertically and
scaled with the local height 𝑧. In contrast, the transporting motions (Figure 7.12d and e) are
much thicker in shape, which represents their spatial organizing power is stronger than the mere
flow and temperature fluctuations and thus it is more likely that the turbulent transports are
driven by organized motions (or coherent structures). Figure 7.11 and Figure 7.12 also show
that 𝑢′ and 𝜃′ are highly correlated while 𝑢′𝑤′ and 𝜃′𝑤′ are highly correlated since each pair
tends to form at the same locations, which again implies the validity of Reynolds analogy.
Under unstable condition, Figure 7.13 and Figure 7.14 show that the near-wall filtered
structures are shorter in streamwise length and larger in spanwise width at both roof-level (𝑧 =
1.01ℎ) and mid-ISL (𝑧 = 1.5ℎ). It suggests that the boundary layer eddies prevent the long
streamwise streaks to form at near-wall region. The two figures also show that the structures

140
Figure 7.11: Streamwise filtered instantaneous fields in y-z plane at neutral condition for model A1
at roof-level (𝑧 = 1.01ℎ): (a) streamwise velocity fluctuation 𝑢′; (b) vertical velocity fluctuation 𝑤′; (c)
temperature fluctuation 𝜃′; (d) momentum flux 𝑢′𝑤′, where (i) ejection motion (𝑢′ < 0, 𝑤 ′ > 0) is positive
(red); (ii) sweep (𝑢′ > 0, 𝑤 ′ < 0) is negative (blue); (e) heat flux 𝜃′𝑤′, where (i) hot updraft motion (𝜃 ′ >
0, 𝑤 ′ > 0) is positive (red); (ii) cold downdraft (𝜃 ′ < 0, 𝑤 ′ < 0) is negative (blue). Normalization by
friction velocity 𝑢𝜏 and temperature 𝜃𝜏 is used. Filtering unit is 10ℎ.

141
Figure 7.12: Streamwise filtered instantaneous fields in y-z plane at neutral condition for model A1
at mid-ISL (𝑧 = 1.5ℎ). Others are similar to Figure 7.11.

142
Figure 7.13: Streamwise filtered instantaneous fields in y-z plane at unstable condition (𝑧𝑖 /𝐿 =
−8.3) for model A1 at roof-level (𝑧 = 1.01ℎ). Others are similar to Figure 7.11.

143
Figure 7.14: Streamwise filtered instantaneous fields in y-z plane at unstable condition (𝑧𝑖 /𝐿 =
−8.3) for model A1 at mid-ISL (𝑧 = 1.5ℎ). Others are similar to Figure 7.11.

144
are highly correlated at both heights, which means the structures at these two heights are from
the same large scale motions, though the fluctuations and fluxes due to the boundary layer
eddies are stronger at mid-ISL compared to roof-level.
From the results of filtered structures of transporting eddies, it is observed that the
boundary layer eddies are in fact not entirely inactive at the SL and the influence of these large
eddies on the fluxes increases with height. The influence of boundary layer eddies on near-wall
turbulent transports are also strengthened by the buoyant instability. This further illustrates why
Townsend’s hypothesis is invalid, especially under strongly unstable condition.

7.4 Comparison of Momentum and Heat Transports

This section investigates how the momentum and heat transfer modes in RSL and ISL
above roof-level change with buoyant instability and how buoyant instability modifies their
dissimilarity. These are studied through quadrant analysis and correlation coefficient of
momentum and heat fluxes. The quadrants of momentum and heat fluxes have already been
defined in Equations (7.5) and (7.6) and the quadrants are averaged in temporal, spanwise and
̅̅̅̅̅̅̅̅̅
streamwise domains in this section. The ratio of ejection 〈𝑢′𝑤′| ̅̅̅̅̅̅̅̅̅
𝑞2 〉 and sweep 〈𝑢′𝑤′|𝑞4 〉 as well

̅̅̅̅̅̅̅̅̅
as the ratio of hot updraft 〈𝜃′𝑤′| ̅̅̅̅̅̅̅̅̅
𝑞1 〉 and cold downdraft 〈𝜃′𝑤′|𝑞3 〉 are studied here in Figure

7.15. The ratio can be viewed as the relative dominance of upward and downwards motions on
the turbulent fluxes. The correlation coefficient between momentum and heat fluxes (also
averaged in streamwise direction) is defined as:
̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
(𝑢 ′ 𝑤 ′ − ̅̅̅̅̅̅
𝑢′𝑤′)(𝜃 ′ 𝑤 ′ − ̅̅̅̅̅̅
𝜃′𝑤′)
𝑅𝑢′ 𝑤′ ,𝜃′ 𝑤′ = 〈 1 1
〉 (7.7)
̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
2 2 ̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
2 2
[(𝑢′ 𝑤 ′ − ̅̅̅̅̅̅
𝑢′𝑤′) ] [(𝜃 ′ 𝑤 ′ − ̅̅̅̅̅̅
𝜃′𝑤′) ]
which is plotted in Figure 7.16.
̅̅̅̅̅̅̅̅̅
Figure 7.15a shows that, at neutral state, the ejection motions 〈𝑢′𝑤′| 𝑞2 〉 dominate the

momentum flux above roughness except at the vicinity of the roof-level where the sweep
̅̅̅̅̅̅̅̅̅
motions 〈𝑢′𝑤′| 𝑞4 〉 dominate. The dominance of ejection attains the maximum around 𝑧 = 1.2ℎ,

which is within RSL, and it then decreases slightly in ISL and keeps uniform with height further.
Sweep is usually quoted as a dominant momentum carrying motions within RSL of vegetative
roughness (e.g. Raupach et al. [12]) and it is usually explained that the dominance of sweep is
due to the plane mixing-layer at the interface of roughness and free-stream flow (e.g. Finnigan
[47]). However, the LES result suggests that the dominance of sweep motions only confined

145
within a very thin layer at the interface between roughness and free-stream flow for urban-type
roughness, but not the whole RSL, within which ejection dominates the flux at the upper part.
As buoyant instability level increases, the ejection motions around the peak ( 𝑧 = 1.2ℎ )
increases at slightly unstable state but decreases after certain strong unstable state. When the
instability is sufficiently strong (i.e. at local free-convection state), there seems to be no local
maximum of the ratio and the ratio increases smoothly, which implies ejection motions become
increasingly dominated with height. Throughout the range of instability level reported, the
momentum carrying motions at the vicinity of roof-level are kept dominated by sweeps, which
is kept unchanged with instability.
Figure 7.15b shows that the mode of heat transport is very similar to that of momentum
transport. However, the behavior of heat transporting motions is quite different from the
momentum transporting motions when buoyant instability is active. It shows that the ratio of
̅̅̅̅̅̅̅̅̅
hot updraft 〈𝜃′𝑤′| ̅̅̅̅̅̅̅̅̅
𝑞1 〉 and cold downdraft 〈𝜃′𝑤′|𝑞3 〉 attains a local maximum around 𝑧 = 1.2ℎ,

similar to the momentum flux at neutral and slightly unstable state, but the magnitude of the
peak increases with instability level. Unlike momentum flux, the ratio decreases with height in
ISL and further upwards, while the respective ratio in momentum transport increases with
height. This implies the modes of momentum and heat transports are different from each other.
At the vicinity of roof-level, the heat transport is dominated by cold downdraft, which is from
the downward velocity fluctuations, similar to the dominance of sweep motion in momentum
transport. Therefore, the momentum and heat transports around the interface of urban
roughness and free-stream flow due to similar mechanism, which is shear-induced rather than
buoyancy-induced.
Figure 7.16 shows that the overall correlation between momentum and heat fluxes,
𝑅𝑢′ 𝑤′ ,𝜃′ 𝑤′ , is high at neutral state and its magnitude decreases with instability, which means
the momentum and heat are transported by similar organized motions at neutral state but
dissimilar organized motions at unstable state. At neutral state, the correlation is around -0.8 at
the near-wall region and it then decreases mildly with height. With increasing instability levels,
the near-wall correlation decreases and the decreasing trend with height is also enhanced. At
sufficiently strong unstable level, 𝑅𝑢′ 𝑤′ ,𝜃′ 𝑤′ settles down to around -0.3 and it is uniform with
height.
The results above consistently suggest that momentum and heat are transported by
different organized motions within ISL at strong unstable level, consistent to the field
measurement study by Li and Bou-Zeid [132]. However, there seems to be a conflicting result

146
Figure 7.15: Vertical profiles of flux quadrant ratios above roughness roof-level of model A1 at
̅̅̅̅̅̅̅̅̅
different instability levels: a) ratio of ejection 〈𝑢′𝑤′| ̅̅̅̅̅̅̅̅̅
𝑞2 〉 and sweep 〈𝑢′𝑤′|𝑞4 〉 and b) ratio of hot updraft
̅̅̅̅̅̅̅̅̅ ̅̅̅̅̅̅̅̅̅
〈𝜃′𝑤′|𝑞1〉 and cold downdraft 〈𝜃′𝑤′|𝑞3〉. Inability levels from neutral state (black) to 𝑧𝑖 /𝐿 = −8.3 (red).

147
Figure 7.16: Vertical profiles of horizontally averaged correlation coefficient between momentum
flux 𝑢′𝑤′ and heat flux 𝜃′𝑤′ above roughness roof-level for model A1 at different instability levels. Inability
levels from neutral state (black) to 𝑧𝑖 /𝐿 = −8.3 (red).

concerning the turbulent transports at the vicinity of roof-level at strong unstable state:
𝑅𝑢′ 𝑤′ ,𝜃′ 𝑤′ at the vicinity of roof-level is much lower (𝑅𝑢′ 𝑤′ ,𝜃′ 𝑤′ ~ − 0.3) than the neutral value
(𝑅𝑢′ 𝑤′ ,𝜃′ 𝑤′ ~ − 0.85), while some aforementioned results (e.g. quadrant analysis and MOST
statistics), suggested that those momentum and heat results are similar there, for example, the
mean velocity gradient 𝜙𝑚 and mean temperature gradient 𝜙ℎ are similar near roughness while
̅̅̅̅̅̅̅̅̅
the ratio 〈𝑢′𝑤′| ̅̅̅̅̅̅̅̅̅ ̅̅̅̅̅̅̅̅̅ ̅̅̅̅̅̅̅̅̅
𝑞2 〉/〈𝑢′𝑤′|𝑞4 〉 and the ratio 〈𝜃′𝑤′|𝑞1 〉/〈𝜃′𝑤′|𝑞3 〉 are also quite similar (i.e.

downward motions dominated) near roughness. Therefore, it seems that momentum and heat
at strong unstable state are transported by different organized motions around roughness
elements, at different time instants and different locations, but their carrying motions lead to
similar effects in the mean. These motions are possibly from the same class of organized motion
but due to different origins. The instantaneous visualization in Section 7.3 have shown that the
large-scale eddies produce different effects on the large-scale 𝑢′ and 𝜃′ fields and these large-
scale structures are able to extend down to near roughness. This may imply that the dissimilarity
of momentum and heat transports are due to these large-scale eddies. However, the detailed
mechanism for this dissimilarity is still not figured out.

148
7.5 Two-point Correlation and Integral Length Scale

To study to size and shape of coherent structures in average, the two-point correlation
in x-z plane is employed here, defined as:
̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
𝜙′(𝑥, 𝑦, 𝑧, 𝑡)𝜙′(𝑥𝑟 , 𝑦, 𝑧𝑟 , 𝑡)
𝑟𝜙𝜙 (𝑥, 𝑧; 𝑥𝑟 , 𝑧𝑟 ) = 1 1 (7.8)
̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
2 2 ̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
[𝜙′(𝑥, 𝑦, 𝑧, 𝑡) ] [𝜙′(𝑥𝑟 , 𝑦, 𝑧𝑟 , 𝑡) ]2
2

where 𝜙 is any flow variable, (𝑥𝑟 , 𝑧𝑟 ) is the reference position and averaging in temporal and
spanwise domains is used (denoted by overbar).
At neutral state, Figure 7.17 shows that the two-point correlations of streamwise
velocity 𝑟𝑢𝑢 and temperature 𝑟𝜃𝜃 are somehow similar in size and shape, which are tilted
upwards, while 𝑟𝑤𝑤 is smaller in size and circular in shape. The tilting angles of 𝑟𝑢𝑢 and 𝑟𝜃𝜃
are due to the shearing motions in ISL and represent successive hairpin vortices that carry 𝑢′
and 𝜃′ are developing in both streamwise and vertical directions. The small size and circular
shape of 𝑟𝑤𝑤 mean the vertical motions of fluid are localized, small-scale structures and
independent of other large-scale structures. The patterns are basically similar to those over
vegetative roughness (Raupach et al. [12] and Finnigan [47]).
At unstable state, Figure 7.18 shows that 𝑟𝑢𝑢 is cell-shape like, two-layered and periodic
in streamwise direction throughout the whole boundary layer, which suggests that 𝑢′ in ISL is
actually dominated by the highly organized large-scale flow recirculation within boundary
layer, consistent to the many field measurement results suggesting that 𝑢′ is dominated by
inactive and scaled by boundary layer depth 𝑧𝑖 . The feature of 𝑟𝑢𝑢 is not seen in 𝑟𝑤𝑤 and 𝑟𝜃𝜃 .
𝑟𝑤𝑤 is circular in shape, similar to that in neutral condition, but with a larger size, while 𝑟𝜃𝜃
spreads through a larger region compared to 𝑟𝑤𝑤 . Therefore, the vertical velocity fluctuation
𝑤 ′ responsible for the flux carrying motions is more susceptible to large-scale motion under
unstable condition than under neutral condition, which may imply the flux carrying motions in
ISL are not completely independent of large-scale motions, contrary to Townsend’s hypothesis.
As 𝑤′ in fact represents the flux carrying motions, 𝑟𝑤𝑤 at three levels in the near-wall
region, including mid-ISL, roof-level and mid cavity, are compared in Figure 7.19. The neutral
state results show that 𝑟𝑤𝑤 are confined within small regions at these three positions, with
radius smaller than the roughness size ℎ. 𝑤′ inside cavity is basically decoupled from the free-
stream flow, but there are small positive values of 𝑟𝑤𝑤 across different cavity centers, which
implies certain large-scale motions above roughness provide an organizing effect for roughness

149
Figure 7.17: Two-point correlations in x-z plane at neutral condition of reference position at mid-
ISL (𝑥𝑟 , 𝑧𝑟 = 1.5ℎ, 1.5ℎ): (a) streamwise velocity fluctuation 𝑢′, (b) vertical velocity fluctuation 𝑤′ and (c)
temperature fluctuation 𝜃′.

150
Figure 7.18: Two-point correlations in x-z plane at unstable condition (𝑧𝑖 /𝐿 = −8.3) of reference
position at mid-ISL ( 𝑥𝑟 , 𝑧𝑟 = 13ℎ, 1.5ℎ): (a) streamwise velocity fluctuation 𝑢′, (b) vertical velocity
fluctuation 𝑤′ and (c) temperature fluctuation 𝜃′.

151
Figure 7.19: Two-point correlations of vertical velocity fluctuation 𝑤′ in x-z plane at different
reference positions: (a) mid-ISL (𝑥𝑟 , 𝑧𝑟 = 13ℎ, 1.5ℎ), (b) roof-level (𝑥𝑟 , 𝑧𝑟 = 13ℎ, 1.01ℎ) and (c) mid-
cavity (𝑥𝑟 , 𝑧𝑟 = 13ℎ, 0.5ℎ); under (i) neutral condition and (ii) unstable condition (𝑧𝑖 /𝐿 = −8.3).

scale flows. This is also studied by Alvarez and Christensen [149], who discovered that some
meandering superstructures, with scale even larger than boundary layer depth, were present in
the RSL over irregular roughness, but these superstructures are not the focus of this study. At
unstable condition, the region of significant 𝑟𝑤𝑤 in ISL is much broader than that in roof-level
and 𝑟𝑤𝑤 at cavity center is confined within the roof-level. This again suggests that 𝑤′ inside
cavity is decoupled from the free-stream flow. Similar to neutral condition, there are a small
positive values across different cavity centers, which shows the indirect influence of some large
scale motions.

152
Figure 7.20: Vertical profiles of integral length scales of vertical velocity fluctuation at streamwise
direction 𝐿𝑤𝑤,𝑥 (𝑧𝑟 ) and vertical direction 𝐿𝑤𝑤,𝑧 (𝑧𝑟 ) for neutral and unstable condition (𝑧𝑖 /𝐿 = −8.3).

Finally, the spatial size of dominant turbulent motion can be quantitatively calculated
by integrating the two-point correlation at specific direction, resulting in integral length scale.
Vertical velocity fluctuation 𝑤 ′ is again considered here due to its importance in turbulent
transports. The integral length scales of in streamwise direction, 𝐿𝑤𝑤,𝑥 , and vertical direction,
𝐿𝑤𝑤,𝑧 , are defined respectively as:

𝐿𝑤𝑤,𝑥 (𝑥𝑟 , 𝑧𝑟 ) = ∫ 𝑟𝑤𝑤 (𝑥, 𝑧𝑟 ; 𝑥𝑟 , 𝑧𝑟 )𝑑𝑥 (7.9)


𝐷

𝐿𝑤𝑤,𝑧 (𝑥𝑟 , 𝑧𝑟 ) = ∫ 𝑟𝑤𝑤 (𝑥𝑟 , 𝑧; 𝑥𝑟 , 𝑧𝑟 )𝑑𝑧 (7.10)


𝐷
where D represents the range within domain boundaries or solid walls. The reference position
𝑥𝑟 is fixed at cavity center plane (i.e. between upstream and downstream building walls).
Figure 7.20 shows that 𝐿𝑤𝑤,𝑥 and 𝐿𝑤𝑤,𝑧 are of similar magnitude at both neutral and unstable
conditions above roof-level (𝑧𝑟 = ℎ), so the structures of 𝑤 ′ are of circular shape in average at
both instability levels, while the radius of mean structure at unstable state is larger. Above ISL

153
(𝑧𝑟 > 1.7ℎ), the structure become vertically elongated at both instability levels and the degree
of vertical elongation is stronger in unstable state. At roof-level, the mean 𝑤 ′ structures at both
instability levels are circular (i.e. 𝐿𝑤𝑤,𝑥 ~𝐿𝑤𝑤,𝑧 ) and of similar size. Therefore, 𝑤 ′ at roof-level
are governed by similar mechanisms under different instability levels. Inside cavity, the two-
point correlation 𝑟𝑤𝑤 in Figure 7.19 showed that the mean 𝑤 ′ structures are slightly vertically
elongated and similar in size for both conditions. This is consistent to the neutral result of
integral length scale, but it is peculiar that 𝐿𝑤𝑤,𝑧 is negative inside cavity under unstable
condition. The negative value of 𝐿𝑤𝑤,𝑧 is thought to be due to the non-zero value of 𝑟𝑤𝑤 in the
upper boundary layer since 𝐿𝑤𝑤,𝑧 is calculated by integrating 𝑟𝑤𝑤 until the boundary layer top.
Hence, the negative value of 𝐿𝑤𝑤,𝑧 inside cavity may indicate there is certain influence on the
cavity turbulence from some large-scale motions.

7.6 Conditional Average

Conditional average requires a condition as a selection criteria on captured flow fields,


e.g. Coceal et al. [82] showed a low-momentum region above cube roughness from conditional
averaging given a point event of negative fluctuating streamwise velocity, while Finnigan et al.
[127] revealed that the characteristic eddy consists of an upstream head-down sweep-
generating hairpin vortex superimposed on a downstream hear-up ejection-generating hairpin
by using a local maxima of pressure at the vegetative canopy top as the selective condition.
This section investigates whether the pattern of coherent structure over two-dimensional urban
roughness is similar to that over vegetative roughness revealed by Finnigan et al. [127] and
how the presence of unstable stratification modifies the structures. The patterns of coherent
structure near wall given strong updraft and strong downdraft under unstable condition are
studied.
Figure 7.21, shows that conditional averaged velocity field (arrows) and temperature
field (color contour) given 1% raw data of the greatest pressure fluctuation for all pressure
records at roof-level or mid-ISL. Under neutral condition (Figure 7.21a), a temperature
microfront is clearly shown at the point of pressure maximum in both conditional averages at
roof-level and mid-ISL. A cold downdraft and a sweep are at the upstream of the temperature
microfront while a hot updraft and an ejection are at the downstream part. This form a typical
eddy for the momentum and heat transports at both RSL and ISL. The result also strongly

154
Figure 7.21: Conditional averages of flow velocity vector (arrows) and temperature (color contour)
on x-z plane under (a) neutral condition, (b) unstable condition (𝑧𝑖 /𝐿 = −8.3), given local pressure maxima
at (i) roof-level (𝑧 = 1.01ℎ), (ii) mid-ISL (𝑧 = 1.5ℎ).

suggests that momentum and heat are transported by the same organized motion at the same
time and position. Comparing the coherent structures at roof-level and ISL, the one at roof-
level shows that the upstream cold downdraft and sweep are stronger than the downstream hot
updraft and ejection and vice versa at ISL. Therefore, the typical eddy (i.e. stronger upstream
sweep superimposed on a weaker downstream ejection) found in the conditional averaged field
in vegetative roughness by Finnigan et al. [127] is also found at the roof-level of urban
roughness, which means the same mechanism shared by flows in vegetative roughness and
urban roughness and this mechanism is thought to be the Kelvin-Helmholtz instability due to
the inflected velocity profile at the roughness interface.

155
Under unstable condition, Figure 7.21b shows that the conditional averaged flow and
temperature fields given local pressure maxima are dominated by the strong upstream
downdraft and this pattern is similar for both pressure triggers at roof-level and mid-ISL. For
the roof-level one, the conditional averaged field shows that there is only a very week
downstream ejection at roof-level. There seems to be no ejection motion for the mid-ISL
pressure trigger. It also shows that the strong downdraft from the free-stream drives the
recirculation flow inside the roughness cavity. These results may suggest that the pressure
maximum event is actually generated by the strong downdraft from the boundary layer in
unstable condition, rather than the pairing of upstream sweep and downstream ejection in
neutral condition.
The three-dimensional coherent structures educed by local pressure maxima at roof-
level are investigated in a series of cross-sectional views in y-z plane in Figure 7.22 for neutral
condition and Figure 7.23 for unstable condition. In Figure 7.22, the streamwise variation of
coherent structure in neutral condition shows that a spanwise diverging flow accompanying the
sweep motion upstream and a pair of counter-rotating streamwise vortices accompanying the
ejection motion downstream. These streamwise vortices increase in height with in streamwise
direction. This implies these streamwise vortices are slightly tilted upward. The three-
dimensional structures can be explained by the conceptual model of dual-hairpin eddy from
Finnigan et al. [127] that the vortex legs are strained by the mean shear. The counter-rotating
streamwise vortices calculated from present LES are about 2ℎ apart, so that the structure is
scaled by roughness size.
At unstable state, the streamwise variation of coherent structure in Figure 7.23 shows
that the strong downdraft and spanwise diverging flows are dominated above roof-level. A
downward flow in the upstream part of cavity and an upward flow in the downstream part are
found, which represent the recirculation flow inside the cavity. A pair of counter-rotating
streamwise vortices are also found at the downstream position and are similarly about 2ℎ apart,
but the ejection is much weaker than in neutral condition. This suggests that the coherent
structures at roof-level in neutral condition are also found in unstable condition. However, the
recirculation flow driven by strong downdraft motion in boundary layer in unstable condition
is not present in neutral condition. This implies the conditional averaged fields in unstable
condition contains superimposed events: (i) pairing of upstream sweep and downstream
ejection as well as (ii) the strong downdraft penetrating the roughness interface.
Finally, the conditional averaged fields in unstable condition given the strong updraft
and downdraft are investigated in Figure 7.24. The updraft and downdraft conditional averaged
156
Figure 7.22: Conditional averages of flow velocity vector (arrows) and temperature (color contour)
on y-z plane under neutral condition given local pressure maxima at roof-level (𝑧 = 1.01ℎ): cross-sectional
views at: (a) 𝑥 = −0.6ℎ, (b) 𝑥 = −0.4ℎ, (c) 𝑥 = −0.2ℎ, (d) 𝑥 = 0, (e) 𝑥 = 0.2ℎ, (f) 𝑥 = 0.4ℎ, (g) 𝑥 =
0.6ℎ.

157
Figure 7.23: Conditional averages of flow velocity vector (arrows) and temperature (color contour)
on y-z plane under unstable condition (𝑧𝑖 /𝐿 = −8.3) given local pressure maxima at roof-level (𝑧 = 1.01ℎ):
cross-sectional views at: (a) 𝑥 = −0.6ℎ, (b) 𝑥 = −0.4ℎ, (c) 𝑥 = −0.2ℎ, (d) 𝑥 = 0, (e) 𝑥 = 0.2ℎ, (f) 𝑥 =
0.4ℎ, (g) 𝑥 = 0.6ℎ.

158
Figure 7.24: Conditional averages of flow velocity vector (arrows) and temperature (color contour)
on x-z plane under unstable condition (𝑧𝑖 /𝐿 = −8.3), based on (a) local maximum of vertical velocity, (b)
local minimum of vertical velocity, at roof-level (𝑧 = 1.01ℎ).

flows are from the 1% records of strongest upward velocity fluctuation and strongest downward
velocity fluctuation, respectively, at roof-level. The roof-level updraft in Figure 7.24a shows
that the upward motion, accompanied by strong positive temperature fluctuation, is a large-
scale structure connecting the roughness cavity and the boundary layer and the flow is even in
reversed direction above roof-level. The roof-level downdraft in Figure 7.24b shows the
downward motion, accompanied by strong negative temperature fluctuation, drives the flow
from the boundary layer to the roughness cavity that leads to a recirculating flow in the cavity.
The conditional averaged updraft motion drives a strong negative (backward) streamwise
velocity fluctuation upwards, which means the updraft motion assists in downward momentum
transport. However, the conditional averaged downdraft motion does not drive a similar amount
of positive (forward) streamwise velocity fluctuation downwards, which implies the downward
momentum transports by downdraft is much weaker than the updraft and the momentum
transports by downward motions seem to be carried by other turbulent motions.

159
7.7 Concluding Remarks

The mechanisms of momentum and heat transported by turbulence in both neutral and
unstable conditions and the difference between these two conditions are investigated through
the TKE budget and coherent structures. TKE budget contains the information about the
turbulent energetics and is studied in (i) two-dimensional contour at roughness scale and (ii)
one-dimensional profile above urban roughness top in this section. Coherent structures, which
represent the patterns of organized motions that were found to be important to turbulent
transports and occur repeated at certain locations, are studied through (i) visualization of
instantaneous turbulent fields with and without spatial filter, (ii) quadrant analysis of
momentum and heat fluxes and correlation between momentum and heat fluxes, (iii) two-point
correlation and integral length scale and (iv) conditional averaging.
The two-dimensional TKE budget analysis finds that shear production 𝑃𝑠 at roof-level
is the dominant production mechanism under both neutral and unstable conditions. The buoyant
production 𝐵 is even much weaker than the shear production 𝑃𝑠 at about an order of magnitude
under unstable condition, which implies that the enhanced turbulent fluctuations and fluxes
around roughness are not due to the locally buoyancy-induced turbulence. The transport terms,
including mean advection 𝐴 , turbulent transport 𝑇𝑡 and pressure transport 𝑇𝑝 , are similar
between neutral and unstable conditions, except that these transport terms are very active at the
upstream upper corner of cavity under unstable condition. The TKE budget results between
aspect ratio 1 and 0.5 are very similar at both instability levels, which suggests they are
governed by similar mechanism, Kelvin-Helmholtz instability, at roof-level. The one-
dimensional TKE budget analysis above roughness reveals that the wake production 𝑃𝑤 can be
significant near roughness and it is calculated to be negative by LES, rather than positive
suggested by some field measurement studies. The results also suggest that the turbulence
generated under unstable condition is not locally dissipated, but transported vertically
downward towards roughness cavity. The dispersive transport 𝑇𝑑 is negligible in neutral
condition, but it can be significant under unstable condition.
The study of organized motions by instantaneous visualization shows that the velocity
and temperature fluctuations as well as momentum and heat fluxes at near-wall region are
small-scaled (scaled by local height) and distributed evenly under neutral condition while the
fluctuations and fluxes at near-wall region are both small- and large-scaled (scaled by boundary
layer depth) and the small-scaled structures tend to occur collectively with the large-scaled

160
ones. By filtering the instantaneous fluctuation fields in streamwise direction, the streaky
structures elongated in streamwise direction and scaled with local height in the near-wall region
are revealed in neutral condition. Under unstable condition, these streaky structures are
shortened in streamwise direction and get wider in spanwise direction, which show that
influence of boundary layer eddies near the surface. The modes of momentum and heat
transports are compared by using quadrant analysis and correlation coefficient between fluxes.
Quadrant analysis suggests that both momentum and heat transports are dominated by
downward motions near roughness at both neutral and unstable conditions while the result of
correlation between fluxes suggests that momentum and heat fluxes are transported by similar
flux carrying motions in neutral condition but dissimilar motions in unstable state. The study
by two-point correlation and integral length scale shows that the influence of vertical velocity
fluctuation 𝑤 ′ , which was thought to be independent of motions larger than local scale, by
large-scale motions is much stronger in unstable condition. This implies there is certain
mechanism controlling the local scale motions by the large-scale motions. Finally, the
conditional averaging reveals the pair of upstream head-down sweep and downstream hear-up
ejection above urban roughness under neutral condition by LES using the local pressure
maxima as the trigger, similar to the structure found above vegetative roughness. However, the
conditional averaged structure under unstable condition is actually a superimposition of the
sweep-ejection pair eddy and large-scale downdraft.
It is concluded from all the above results that, unlike the neutral condition, the turbulent
transports in near-wall region under unstable condition are multi-scale processes and governed
by top-down mechanism while momentum and heat are transported by different organized.
These imply Townsend’s hypothesis is not strictly valid and the Reynolds analogy for
momentum and heat transfers is also not valid under unstable condition.

161
Chapter 8

Conclusion

In this study, LES is employed to investigate the turbulent flow over idealized two-
dimensional urban roughness at different roughness aspect ratios (0.5, 1 and 2) under a range
of buoyant instability, starting from the neutral state to an unstable state at which local free-
convective condition is satisfied at the ISL. In LES models, flow is driven by background
pressure gradient force and the unstable stratification is driven by a higher temperature on the
walls. Boussinesq approximation is used to model the buoyancy force field. One equation SGS-
TKE model is used to model SGS motions. The LES model is validated by the wind tunnel
result by Uehara et al. [53] on the mean velocity, mean temperature and velocity fluctuation
variances above urban roughness under neutral and unstable condition.
In Chapter 4, the LES calculated single-point statistics, including gradients of mean
velocity and temperature (𝜙𝑚 and 𝜙ℎ ) as well as second moments of vertical velocity and
temperature fluctuations (𝜙𝑤 and 𝜙𝜃 ), in the forms of MOST functions are studied in ISL.
These MOST statistics show good collapses of data onto a single curve, except at slightly
unstable state and ISL upper limit, which may be due to the influence of large boundary layer
eddies. Therefore, the applicability of MOST over idealized urban roughness is basically
supported by the LES results. Parameterized functional forms that fit the LES data are proposed,
which are piece functions: (i) a linear function for 𝜙𝑚 and 𝜙ℎ and a constant function for 𝜙𝑤
and 𝜙𝜃 at near-neutral state and (ii) a power function at strong unstable state. It is found that
the power function exponent of 𝜙ℎ is -2/3, different from the prediction of local free-
convective hypothesis. Sensitivity study of LES domain dimensions shows that the temperature
statistics 𝜙ℎ and 𝜙𝜃 in ISL are basically unaffected by the spanwise extent but the velocity
statistics 𝜙𝑚 and 𝜙𝑤 do not, which implies that larger scale eddies are able to modulate the ISL
flow statistics.

162
In Chapter 5, two approaches are used to study the influence of roughness on RSL flow
statistics and RSL height 𝑧∗ : (i) deviation of horizontally averaged profiles of flow statistics
from the prediction of MOST functions and (ii) the horizontal inhomogeneity of flow statistics.
The deviation from MOST functions are studied by LES models of three different mesh
resolutions. The result in (i) shows that 𝜙𝑚 and 𝜙ℎ are less than the MOST ones near roughness,
similar to vegetative roughness, but a local peak appears just above roof-level, which is
different from vegetative one. It is found that mesh resolution effect is important at the vicinity
of roughness, but its effect diminishes with height and the ISL flow is independent of mesh
resolution. This implies mesh resolution effect does not travel upwards and the turbulent
processes are thought to be local or top-down. For the study in (ii), the unstable result shows
that Reynolds stress ̅̅̅̅̅̅
𝑢′𝑤′ and heat flux ̅̅̅̅̅̅
𝜃′𝑤′ are more inhomogeneous than other mean and
fluctuation statistics. The strong inhomogeneity of ̅̅̅̅̅̅
𝑢′𝑤′ is found to be due to the strong
momentum transport by mean advection 𝑢̅ ∙ 𝑤
̅ above roof-level. The results suggest that
unstable stratification can magnify the inhomogeneous properties additional to roughness
morphology, which is non-linear with the effect of aspect ratio.
In Chapter 6, the mean velocity and temperature profile equations, 𝑈(𝑧) and 𝛩(𝑧), are
derived analytically based on the MOST functions proposed in Chapter 4 and are compared
well to the LES results, which supports the applicability of the proposed MOST functions. The
aerodynamic and temperature roughness lengths, 𝑧0 and 𝑧ℎ , obtained by matching the derived
profiles with the LES results are found to increase with buoyant instability level after
sufficiently strong unstable state, contrary to the traditional view that 𝑧0 and 𝑧ℎ are not
functions of instability. It is also found from the LES result that 𝑧0 and 𝑧ℎ decreases suddenly
from the neutral values at slightly unstable state, which may be due to the influence of large
scale turbulence. Additionally, the dependence of 𝑧ℎ with Reynolds number is found to satisfy
the Brutsaert’s relationship. A closer look at the behaviors of turbulence at roughness scale
suggests that the increased values of 𝑧0 and 𝑧ℎ (or the enhanced efficiency of turbulent
transports at roughness scale) under strongly unstable state are correlated to the strengthened
turbulence fluctuation and the more vertically elongated turbulent structures extending from
roughness cavity to the above free-stream.
In Chapter 7, the mechanisms of momentum and heat transported by turbulence are
investigated by TKE budget in (i) two-dimensional contour at roughness scale and (ii) one-
dimensional profile above urban roughness top in this section; and patterns of coherent
structures from (i) visualization of instantaneous turbulent fields with and without spatial filter,

163
(ii) quadrant analysis of momentum and heat fluxes and correlation between momentum and
heat fluxes, (iii) two-point correlation and integral length scale and (iv) conditional averaging.
It is found from the TKE budget analysis that the buoyant production 𝐵 is even much weaker
than the shear production 𝑃𝑠 at about an order of magnitude under unstable condition, which
implies that the enhanced turbulent fluctuations and fluxes near wall are not due to the locally
buoyancy-induced turbulence. The transports terms are much stronger near roughness under
unstable condition. The wake production 𝑃𝑤 in one-dimensional TKE budget is significant near
roughness and is calculated to be negative by LES, rather than positive suggested by some field
studies. The instantaneous visualization shows that the near-wall structures of fluctuations and
fluxes are small-scaled and disturbed evenly under neutral condition, but they are multi-scaled
and the small-scaled ones tend to occur collectively with the large-scaled ones. After filtering
in streawmise direction, streaky structures elongated in streamwise direction are found in the
neutral state while the streaky structures are shortened and wider under unstable state, which
shows the influence of boundary layer eddies to the near-wall region. The momentum and heat
fluxes are found to become uncorrelated with increasing instability, but their modes of
transports near roughness are similar in quadrant analysis. It is therefore hypothesized they are
transported by similar but different organized motions. The study by two-point correlation and
integral length scale shows that the vertical velocity fluctuation 𝑤 ′ is localized in neutral
condition but its spatial range of influence is much larger under unstable condition. The
conditional averaged fields reveals the pair of upstream head-down sweep and downstream
hear-up ejection above urban roughness under neutral condition, similar to one above the
vegetative roughness, while the conditional averaged structure under unstable condition is
actually a superimposition of the sweep-ejection pair eddy and large-scale downdraft.
The above conclusions from the LES results suggest that, unlike the neutral condition,
the turbulent transports in near-wall region under unstable condition are in fact multi-scale
processes and governed by top-down mechanism. This provides evidences that the assumption
of MOST, which states that the turbulence is scaled by local height 𝑧 but not boundary layer
depth 𝑧𝑖 , and the Townsend’s hypothesis, which stats that the large scale motions are inactive
to near-wall transports, are in fact not slightly true under unstable condition. However, MOST
can still approximately describe the ISL flow properties well given the instability level is not
too high, which supports the applicability of MOST. The LES results also suggest that
momentum and heat transports are different under unstable condition, which mean the
Reynolds analogy for the similarity between momentum and heat transfers is in fact not valid.

164
Appendix

A.1 Steps of Mathematical Integration of the Proposed MOST


functions

Starting from the proposed MOST functional forms in Chapter 4,


𝜙1,𝑖 (𝜁) = 1 + 𝑎𝑖 𝜁 𝜁𝑡𝑟,𝑖 < 𝜁 < 0 (A.1)
𝜙𝑖 (𝜁) = { 𝑛 , 𝑖 = 𝑚 𝑜𝑟 ℎ
𝜙2,𝑖 (𝜁) = 𝐶𝑖 (−𝜁) 𝑖 𝜁𝑚𝑎𝑥,𝑖 < 𝜁 < 𝜁𝑡𝑟,𝑖
And from the definitions of integrated stability correction function:
𝜁
1 − 𝜙𝑖 (𝜁) (A.2)
Ψ𝑖 (𝜁) = ∫ 𝑑𝜁 , 𝑖 = 𝑚 𝑜𝑟 ℎ
0 𝜁
As 𝜙𝑖 (𝜁) is a piecewise function, the mathematical integration has to be performed in
piecewise steps:
𝜁
1 − 𝜙1,𝑖 (𝜁)
∫ 𝑑𝜁 𝜁𝑡𝑟,𝑖 < 𝜁 < 0
0 𝜁 (A.3)
Ψ𝑖 (𝜁) = 𝜁 𝜁
1 − 𝜙1,𝑖 (𝜁) 1 − 𝜙2,𝑖 (𝜁)
∫ 𝑑𝜁 + ∫ 𝑑𝜁 𝜁𝑚𝑎𝑥,𝑖 < 𝜁 < 𝜁𝑡𝑟,𝑖
{ 0 𝜁 0 𝜁
The analytical integration of the linear function 𝜙1,𝑖 (𝜁) is:
1 − 𝜙1,𝑖 (𝜁) 1 − (1 + 𝑎𝑖 𝜁) (A.4)
∫ 𝑑𝜁 = ∫ 𝑑𝜁 = ∫ −𝑎𝑖 𝑑𝜁 = −𝑎𝑖 𝜁 + 𝑐𝑜𝑛𝑠𝑡𝑎𝑛𝑡
𝜁 𝜁
The analytical integration of the power function 𝜙2,𝑖 (𝜁) is:
1 − 𝜙2,𝑖 (𝜁) 1 − 𝐶𝑖 (−𝜁)𝑛𝑖 1
∫ 𝑑𝜁 = ∫ 𝑑𝜁 = ∫ [ + 𝐶𝑖 (−𝜁)𝑛𝑖 −1 ] 𝑑𝜁
𝜁 𝜁 𝜁 (A.5)
𝐶𝑖
= 𝑙𝑛𝜁 − (−𝜁)𝑛𝑖 + 𝑐𝑜𝑛𝑠𝑡𝑎𝑛𝑡
𝑛𝑖
Finally, the integrated stability correction function is derived as:
−𝑎𝑖 𝜁 𝜁𝑡𝑟,𝑖 < 𝜁 < 0
Ψ𝑖 (𝜁) = { 𝜁 𝐶𝑖 (A.6)
ln ( ) − [(−𝜁)𝑛𝑖 − (−𝜁𝑡𝑟 )𝑛𝑖 ] − 𝑎𝑖 𝜁𝑡𝑟,𝑖 𝜁𝑚𝑎𝑥,𝑖 < 𝜁 < 𝜁𝑡𝑟,𝑖
𝜁𝑡𝑟,𝑖 𝑛𝑖

165
A.2 Derivation of Turbulent Kinetic (TKE) Budget Equation

The basic procedures of derivation are documented in detail by Wyngaard (2010). The key
steps that lead to the form of TKE budget equation used by the present study are shown here.
Starting from the Navier-Stokes Equation (3.2) that describes the fundamental physics of flow:
𝜕𝑢𝑖 𝜕𝑢𝑖 𝜕𝑝 𝜕 2 𝑢𝑖 (A.7)
+ 𝑢𝑗 =− +𝜈 + 𝛼𝑔(𝜃 − 𝜃0 )𝛿𝑖3 + 𝜆Π𝛿𝑖1
𝜕𝑡 𝜕𝑥𝑗 𝜕𝑥𝑖 𝜕𝑥𝑗 𝜕𝑥𝑗
Ensemble averaging leads to
𝜕𝑢̅𝑖 𝜕𝑢̅𝑖 𝜕𝑝̅ 𝜕 2 𝑢̅𝑖 ̅̅̅̅̅̅
𝜕𝑢 ′ ′
𝑖 𝑢𝑗
+ 𝑢̅𝑗 =− +𝜈 ̅
+ 𝛼𝑔(𝜃 − 𝜃0 )𝛿𝑖3 + 𝜆Π𝛿𝑖1 − (A.8)
𝜕𝑡 𝜕𝑥𝑗 𝜕𝑥𝑖 𝜕𝑥𝑗 𝜕𝑥𝑗 𝜕𝑥𝑗
where the notations of overbar and single prime refer to Section 3.5.2.
Subtracting Equation (A.7) from Equation (A.8) produces the equation of velocity fluctuation:
𝜕𝑢𝑖′ 𝜕𝑢𝑖′ 𝜕𝑢̅𝑖 𝜕 𝜕𝑝′ 𝜕 2 𝑢𝑖′ (A.9)
+ 𝑢̅𝑗 + 𝑢𝑗′ + (𝑢𝑖′ 𝑢𝑗′ − ̅̅̅̅̅̅
𝑢𝑖′ 𝑢𝑗′ ) = − +𝜈 + 𝛼𝑔𝜃′𝛿𝑖3
𝜕𝑡 𝜕𝑥𝑗 𝜕𝑥𝑗 𝜕𝑥𝑗 𝜕𝑥𝑖 𝜕𝑥𝑗 𝜕𝑥𝑗
By dotting Equation (A.9), ensemble averaging and dividing it by 2, the TKE budget equation
is derived as
𝜕𝑘̅ 𝜕𝑘̅ 𝜕𝑢̅𝑖 𝜕 𝜕 ̅̅̅̅̅′ 𝜕 2 𝑘̅
= −𝑢̅𝑗 − ̅̅̅̅̅̅
𝑢𝑖′ 𝑢𝑗′ − ̅̅̅̅̅𝑗′ ) −
(𝑘𝑢 (𝑝′𝑢𝑖 ) + 𝜈
𝜕𝑡 𝜕𝑥𝑗 𝜕𝑥𝑗 𝜕𝑥𝑗 𝜕𝑥𝑖 𝜕𝑥𝑗 𝜕𝑥𝑗 (A.10)
̅̅̅̅̅̅̅̅̅̅

𝜕𝑢𝑖 𝜕𝑢𝑖 ′
−𝜈 ̅̅̅̅̅̅
+ 𝛼𝑔𝑤′𝜃′
𝜕𝑥𝑗 𝜕𝑥𝑗
1
where 𝑘 = 2 𝑢𝑖′ 𝑢𝑖′ is the instantaneous TKE and the continuity equation is used in the derivation.

Equation (A.10) is symbolized as


𝜕𝑘̅ (A.11)
= 𝐴 + 𝑃𝑠 + 𝑇𝑡 + 𝑇𝑝 + 𝑇𝑣 − 𝜀 + 𝐵
𝜕𝑡
Each term in Equation (A.11) is defined as:
̅
𝜕𝑘
(i) 𝐴 = −𝑢̅𝑗 𝜕𝑥 : mean advection term;
𝑗

′ ′ 𝜕𝑢̅̅̅𝑖
(ii) ̅̅̅̅̅̅
𝑃𝑠 = −𝑢 𝑖 𝑢𝑗 𝜕𝑥 : mean-gradient (shear) production term;
𝑗

𝜕
(iii) ̅̅̅̅̅𝑗′ ) : turbulent transport term;
𝑇𝑡 = − 𝜕𝑥 (𝑘𝑢
𝑗

𝜕 ̅̅̅̅̅′
(iv) 𝑇𝑝 = − 𝜕𝑥 (𝑝′𝑢 𝑖 ) : pressure transport term;
𝑖

𝜕2 𝑘
̅
(v) 𝑇𝑣 = 𝜈 𝜕𝑥 : viscous diffusion term;
𝑗 𝜕𝑥𝑗

̅̅̅̅̅̅̅̅
𝜕𝑢′ 𝜕𝑢′
(vi) 𝜀 = 𝜈 𝜕𝑥 𝑖 𝜕𝑥 𝑖 : viscous dissipation term; and
𝑗 𝑗

(vii) ̅̅̅̅̅̅ : buoyant production term.


𝐵 = 𝛼𝑔𝑤′𝜃′

166
For the present LES study, Equation (A.10) can be simplified by conditions of quasi-steadiness,
spanwise homogeneity and negligible viscous diffusion:
𝜕𝑘̅ 𝜕𝑘̅ ̅̅̅̅̅ 𝜕𝑢̅ − ̅̅̅̅̅̅ 𝜕𝑢̅ ̅̅̅̅̅̅ 𝜕𝑤
̅ ̅̅̅̅̅̅ 𝜕𝑤̅
[−𝑢̅ −𝑤̅ ] [−𝑢′𝑢′ 𝑢′𝑤′ − 𝑢′𝑤′ − 𝑤′𝑤′ ]
⏟ 𝜕𝑥 𝜕𝑧 + ⏟ 𝜕𝑥 𝜕𝑧 𝜕𝑥 𝜕𝑧
𝐴 𝑃𝑠
𝜕 ̅̅̅̅ 𝜕 ̅̅̅̅̅ 𝜕 ̅̅̅̅̅ 𝜕 ̅̅̅̅̅̅
[− (𝑘𝑢′ ) − (𝑘𝑤′ )] [− (𝑝′𝑢′ ) − (𝑝′𝑤′ )] (A.12)
+ ⏟ 𝜕𝑥 𝜕𝑧 + ⏟ 𝜕𝑥 𝜕𝑧
𝑇𝑡 𝑇𝑝
̅̅̅̅̅̅̅̅̅̅

𝜕𝑢 𝜕𝑢 ′
0 𝜈 𝑖 𝑖 𝛼𝑔𝑤
⏟ ⏟ ̅̅̅̅̅̅
′𝜃′
+ − ⏟𝜕𝑥𝑗 𝜕𝑥𝑗 + =0
𝑇𝑣 𝐵
𝜀
In analyzing the LES results, the symbolized Equation (A.11) can be alternatively written as:

0 = 𝐴 + 𝑃𝑠 + 𝑇𝑡 + 𝑇𝑝 + 𝐵 + 𝑅 (A.13)
where 𝑅 is the residual term accounting for the viscous dissipation and errors due to neglects
of temporal term, spanwise variation, viscous diffusion and subgrid scale (SGS) contributions
which are small but in fact exist due to the imperfect LES solution.

A.3 Derivation of Horizontally Averaged Turbulent Kinetic


(TKE) Budget Equation

Horizontal averaging the ensemble averaged TKE budget Equation (A.10) leads to
𝜕〈𝑘̅〉 𝜕〈𝑘̅〉 𝜕 𝜕〈𝑢̅𝑖 〉 𝜕𝑢̅𝑖 "
= −〈𝑢̅𝑗 〉 − 〈𝑘̅"𝑢̅𝑗 "〉 − 〈𝑢 ̅̅̅̅̅̅
′ ′
𝑖 𝑢𝑗 〉
̅̅̅̅̅̅
− 〈𝑢 ′ ′
𝑖 𝑢𝑗 " 〉
𝜕𝑡 𝜕𝑥𝑗 𝜕𝑥𝑗 𝜕𝑥𝑗 𝜕𝑥𝑗 (A.14)
𝜕 ̅̅̅̅̅′ 𝜕 2 〈𝑘̅〉 ̅̅̅̅̅̅̅̅̅̅
𝜕𝑢𝑖′ 𝜕𝑢𝑖′ ̅̅̅̅̅̅〉
− 〈𝑝′𝑢𝑖 〉 + 𝜈 −𝜈〈 〉 + 𝛼𝑔〈𝑤′𝜃′
𝜕𝑥𝑖 𝜕𝑥𝑗 𝜕𝑥𝑗 𝜕𝑥𝑗 𝜕𝑥𝑗
where the notations of angle bracket and double prime refer to Section 3.5.2.
Equation (A.14) is symbolized as
𝜕〈𝑘̅ 〉 (A.15)
= 𝐴′ + 𝑇𝑑 + 𝑃𝑠 ′ + 𝑃𝑤 + 𝑇𝑡 ′ + 𝑇𝑝 ′ + 𝑇𝑣 ′ − 𝜀′ + 𝐵′
𝜕𝑡
Each term in Equation (A.15) is defined as:
̅〉
𝜕〈𝑘
(i) 𝐴′ = −〈𝑢̅𝑗 〉 𝜕𝑥 : mean advection term;
𝑗

𝜕
(ii) 𝑇𝑑 = − 𝜕𝑥 〈𝑘̅"𝑢̅𝑗 "〉 : dispersive transport term;
𝑗

′ ′ 𝜕〈𝑢 ̅̅̅〉
(iii) ̅̅̅̅̅̅
𝑃𝑠 ′ = −〈𝑢 𝑖
𝑖 𝑢𝑗 〉 𝜕𝑥 : mean-gradient (shear) production term;
𝑗

′ ′ 𝜕𝑢 ̅̅̅"
(iv) ̅̅̅̅̅̅
𝑃𝑤 = − 〈𝑢 𝑖
𝑖 𝑢𝑗 " 𝜕𝑥 〉 :wake production term;
𝑗

167
𝜕
(v) ̅̅̅̅̅𝑗′ 〉 : turbulent transport term;
𝑇𝑡 ′ = − 𝜕𝑥 〈𝑘𝑢
𝑗

𝜕 ̅̅̅̅̅′
(vi) 𝑇𝑝 ′ = − 𝜕𝑥 〈𝑝′𝑢 𝑖 〉 : pressure transport term;
𝑖

𝜕2 〈𝑘
̅〉
(vii) 𝑇𝑣 ′ = 𝜈 𝜕𝑥 : viscous diffusion term;
𝑗 𝜕𝑥𝑗

̅̅̅̅̅̅̅̅
𝜕𝑢′ 𝜕𝑢′
(viii) 𝜀′ = 𝜈 〈𝜕𝑥 𝑖 𝜕𝑥 𝑖 〉 : viscous dissipation term; and
𝑗 𝑗

(ix) ̅̅̅̅̅̅〉 : buoyant production term.


𝐵′ = 𝛼𝑔〈𝑤′𝜃′
Compared to Equation (A.11), two new terms, (ii) dispersive transport 𝑇𝑑 and (iv) wake
production 𝑃𝑤 , arise due to the horizontal averaging, while other terms are similar to those in
Equation (A.11). Similarly, Equation (A.14) can be simplified by conditions of quasi-steadiness,
spanwise homogeneity and negligible viscous diffusion for the present LES study:


0
𝜕
− 〈𝑘̅"𝑤 ̅"〉 −〈𝑢′𝑤′ ̅̅̅̅̅̅〉 𝜕〈𝑢̅〉
+ ⏟ 𝜕𝑧 +⏟ 𝜕𝑧
𝐴′
𝑇𝑑 𝑃𝑠 ′
𝜕 ̅̅̅̅̅
̅̅̅̅̅" 𝜕𝑢̅"〉 − 〈𝑢′𝑤′
[− 〈𝑢′𝑢′ ̅̅̅̅̅̅" 𝜕𝑢̅"〉 − 〈𝑢′𝑤′̅̅̅̅̅̅" 𝜕𝑤̅" ̅̅̅̅̅̅" 𝜕𝑤
〉 − 〈𝑤′𝑤′
̅"
〉] − 〈𝑘𝑤′ 〉
+⏟ 𝜕𝑥 𝜕𝑧 𝜕𝑥 𝜕𝑧 + ⏟ 𝜕𝑧 (A.16)
𝑃𝑤 𝑇𝑡 ′
𝜕 ̅̅̅̅̅̅ ̅̅̅̅̅̅̅̅̅̅

𝜕𝑢𝑖 𝜕𝑢𝑖 ′
− (𝑝′𝑤′ ) ⏟ 0 𝜈 〈 ⏟ ̅̅̅̅̅̅
〉 𝛼𝑔〈𝑤 ′𝜃′〉
+ ⏟ 𝜕𝑧 + − ⏟ 𝜕𝑥𝑗 𝜕𝑥𝑗 + =0
𝑇𝑣 ′ 𝐵 ′
𝑇𝑝 ′ 𝜀′
In analyzing the LES results, the symbolized Equation (A.15) can be alternatively written as:

0 = 𝑇𝑑 + 𝑃𝑠′ + 𝑃𝑤 + 𝑇𝑡′ + 𝑇𝑝′ + 𝐵 ′ + 𝑅′ (A.17)


where 𝑅′ is the residual term accounting for the viscous dissipation and errors due to neglects
of temporal term, spanwise variation, viscous diffusion and subgrid scale (SGS) contributions
which are small but in fact exist due to the imperfect LES solution.

168
References

[1] Monin, A. S., and AMf Obukhov. "Basic laws of turbulent mixing in the SL of the
atmosphere." Contrib. Geophys. Inst. Acad. Sci. USSR 151 (1954): 163-187.
[2] Obukhov, A. M. "Turbulence in an atmosphere with a non-uniform temperature."
Boundary-layer meteorology 2.1 (1971): 7-29.
[3] Townsend, A. A. "Equilibrium layers and wall turbulence." Journal of Fluid Mechanics
11.01 (1961): 97-120.
[4] Izumi, Yutaka. Kansas 1968 Field Program Data Report. No. AFCRL-72-0041. AIR
FORCE CAMBRIDGE RESEARCH LABS LG HANSCOM FIELD MASS, 1971.
[5] Foken, Thomas. "50 years of the Monin–Obukhov similarity theory." Boundary-Layer
Meteorology 119.3 (2006): 431-447.
[6] Kanda, M., et al. "Roughness lengths for momentum and heat derived from outdoor
urban scale models." Journal of Applied Meteorology and Climatology 46.7 (2007):
1067-1079.
[7] Kramm, Gerhard, et al. "Hans A. Panofsky’s Integral Similarity Function—At Fifty."
Atmospheric and Climate Sciences 3.04 (2013): 581.
[8] Zilitinkevich, Sergej S., et al. "The effect of stratification on the aerodynamic roughness
length and displacement height." Boundary-layer meteorology 129.2 (2008): 179-190.
[9] Smedman, Ann‐Sofi, et al. "Heat/mass transfer in the slightly unstable atmospheric SL."
Quarterly Journal of the Royal Meteorological Society 133.622 (2007): 37-51.
[10] McNaughton, K. G., R. J. Clement, and J. B. Moncrieff. "Scaling properties of velocity
and temperature spectra above the surface friction layer in a convective boundary
layer."
[11] Laubach, Johannes, and Keith G. McNaughton. "Scaling properties of temperature
spectra and heat-flux cospectra in the surface friction layer beneath an unstable outer
layer." Boundary-layer meteorology 133.2 (2009): 219-252.
[12] Raupach, M. R., R. A. Antonia, and S. Rajagopalan. "Rough-wall turbulent boundary
layers." Applied Mechanics Reviews 44.1 (1991): 1-25.

169
[13] Rotach, Mathias W. "On the influence of the urban roughness sublayer on turbulence
and dispersion." Atmospheric Environment 33.24 (1999): 4001-4008.
[14] Florens, Emma, Olivier Eiff, and Frédéric Moulin. "Defining the roughness sublayer
and its turbulence statistics." Experiments in fluids 54.4 (2013): 1-15.
[15] Robinson, Stephen K. "Coherent motions in the turbulent boundary layer." Annual
Review of Fluid Mechanics 23.1 (1991): 601-639.
[16] Khanna, Samir, and James G. Brasseur. "Three-dimensional buoyancy-and shear-
induced local structure of the atmospheric boundary layer." Journal of the atmospheric
sciences 55.5 (1998): 710-743.
[17] Wyngaard, John C. Turbulence in the Atmosphere. Cambridge University Press, 2010.
[18] Wyngaard, John C. "Atmospheric turbulence." Annual Review of Fluid Mechanics 24.1
(1992): 205-234.
[19] Wiernga, Jon. "Representative roughness parameters for homogeneous
terrain." Boundary-Layer Meteorology 63.4 (1993): 323-363.
[20] Kanda, M. "Progress in the scale modeling of urban climate: Review." Theoretical and
Applied Climatology 84.1-3 (2006): 23-33.
[21] Pope, Stephen B. "Turbulent flows." (2001).
[22] Obukhov, A. M. “Turbulentnost’ v temperaturnoj – neodnorodnoj atmosfere
(Turbulence in an Atmosphere with a Non-uniform Temperature)”, Trudy Inst. Theor.
Geofiz. AN SSSR 1 (1946): 95–115.
[23] Högström, U. L. F. "Review of some basic characteristics of the atmospheric SL."
Boundary-Layer Meteorology 25th Anniversary Volume, 1970–1995. Springer
Netherlands, 1996. 215-246.
[24] Kader, B. A., and A. M. Yaglom. "Mean fields and fluctuation moments in unstably
stratified turbulent boundary layers." Journal of Fluid Mechanics 212 (1990): 637-662.
[25] Businger, Joost A., et al. "Flux-profile relationships in the atmospheric SL." Journal of
the atmospheric Sciences 28.2 (1971): 181-189.
[26] Dyer, A. J., and B. B. Hicks. "Flux‐gradient relationships in the constant flux layer."
Quarterly Journal of the Royal Meteorological Society 96.410 (1970): 715-721.
[27] Wilson, D. Keith. "An alternative function for the wind and temperature gradients in
unstable SLs." Boundary-layer meteorology 99.1 (2001): 151-158.
[28] Webb, Eric K. "Profile relationships: The log‐linear range, and extension to strong
stability." Quarterly Journal of the Royal Meteorological Society 96.407 (1970): 67-90.
[29] Dyer, A. J. "A review of flux-profile relationships." Boundary-Layer Meteorology 7.3
170
(1974): 363-372.
[30] Kaimal, J. C., and J. C. Wyngaard. "The kansas and minnesota experiments." Boundary-
Layer Meteorology 50.1-4 (1990): 31-47.
[31] Carl, Douglas M., Terry C. Tarbell, and Hans A. Panofsky. "Profiles of wind and
temperature from towers over homogeneous terrain." Journal of the Atmospheric
Sciences 30.5 (1973): 788-794.
[32] Katul, Gabriel G., Alexandra G. Konings, and Amilcare Porporato. "Mean velocity
profile in a sheared and thermally stratified atmospheric boundary layer." Physical
review letters 107.26 (2011): 268502.
[33] Li, Dan, Gabriel G. Katul, and Elie Bou-Zeid. "Mean velocity and temperature profiles
in a sheared diabatic turbulent boundary layer." Physics of Fluids (1994-present) 24.10
(2012): 105105.
[34] Johansson, Cecilia, et al. "Critical test of the validity of Monin-Obukhov similarity
during convective conditions." Journal of the atmospheric sciences 58.12 (2001): 1549-
1566.
[35] Wyngaard, J. C., O. R. Coté, and Y. Izumi. "Local free convection, similarity, and the
budgets of shear stress and heat flux." Journal of the Atmospheric Sciences 28.7 (1971):
1171-1182.
[36] Panofsky, H. A., et al. "The characteristics of turbulent velocity components in the SL
under convective conditions." Boundary-Layer Meteorology 11.3 (1977): 355-361.
[37] De Bruin, H. A. R., W. Kohsiek, and B. J. J. M. Van den Hurk. "A verification of some
methods to determine the fluxes of momentum, sensible heat, and water vapour using
standard deviation and structure parameter of scalar meteorological quantities."
Boundary-Layer Meteorology 63.3 (1993): 231-257.
[38] Roth, Matthias. "Review of atmospheric turbulence over cities." Quarterly Journal of
the Royal Meteorological Society 126.564 (2000): 941-990.
[39] Jiménez, Javier. "Turbulent flows over rough walls." Annu. Rev. Fluid Mech. 36 (2004):
173-196.
[40] Krogstad, P-Å., R. A. Antonia, and L. W. B. Browne. "Comparison between rough-and
smooth-wall turbulent boundary layers." Journal of Fluid Mechanics 245 (1992): 599-
617.
[41] Krogstadt, P-Å., and R. A. Antonia. "Surface roughness effects in turbulent boundary
layers." Experiments in Fluids 27.5 (1999): 450-460.
[42] Leonardi, S., et al. "Direct numerical simulations of turbulent channel flow with
171
transverse square bars on one wall." Journal of Fluid Mechanics 491 (2003): 229-238.
[43] Bhaganagar, Kiran, John Kim, and Gary Coleman. "Effect of roughness on wall-
bounded turbulence." Flow, turbulence and combustion 72.2-4 (2004): 463-492.
[44] Moriwaki, Ryo, and Manabu Kanda. "Flux-gradient profiles for momentum and heat
over an urban surface." Theoretical and applied climatology 84.1-3 (2006): 127-135.
[45] Wood, C. R., et al. "Turbulent flow at 190 m height above London during 2006–2008:
a climatology and the applicability of similarity theory." Boundary-layer meteorology
137.1 (2010): 77-96.
[46] Al-Jiboori, M. H., Yumao Xu, and Yongfu Qian. "Local similarity relationships in the
urban boundary layer." Boundary-layer meteorology 102.1 (2002): 63-82.
[47] Finnigan, John. "Turbulence in plant canopies." Annual Review of Fluid Mechanics
32.1 (2000): 519-571.
[48] Reynolds, Ryan T., and Ian P. Castro. "Measurements in an urban-type boundary layer."
Experiments in Fluids 45.1 (2008): 141-156.
[49] Cheng, Hong, and Ian P. Castro. "Near wall flow over urban-like roughness."
Boundary-Layer Meteorology 104.2 (2002): 229-259.
[50] Rotach, M. W. "Turbulence close to a rough urban surface part I: Reynolds stress."
Boundary-Layer Meteorology 65.1-2 (1993): 1-28.
[51] Coceal, O., et al. "Mean flow and turbulence statistics over groups of urban-like cubical
obstacles." Boundary-Layer Meteorology 121.3 (2006): 491-519.
[52] Castro, Ian P., Hong Cheng, and Ryan Reynolds. "Turbulence over urban-type
roughness: deductions from wind-tunnel measurements." Boundary-Layer
Meteorology 118.1 (2006): 109-131.
[53] Uehara, Kiyoshi, et al. "Wind tunnel experiments on how thermal stratification affects
flow in and above urban street cavities." Atmospheric Environment 34.10 (2000): 1553-
1562.
[54] Li, Xian-Xiang, Rex E. Britter, and Leslie K. Norford. "Transport processes in and
above two-dimensional urban street cavities under different stratification conditions:
results from numerical simulation." Environmental Fluid Mechanics 15.2 (2015): 399-
417.
[55] Cai, Xiaoming. "Effects of differential wall heating in street cavities on dispersion and
ventilation characteristics of a passive scalar." Atmospheric environment 51 (2012):
268-277.
[56] Cheng, W. C., and Chun-Ho Liu. "Large-eddy simulation of turbulent transports in
172
urban street cavities in different thermal stabilities." Journal of Wind Engineering and
Industrial Aerodynamics 99.4 (2011): 434-442.
[57] Yusup, Yusri, and Jing‐Fen Lim. "Turbulence variances in the convective urban
roughness sublayer: an application of similarity theory using local scales."
Meteorological Applications 21.2 (2014): 149-160.
[58] Amir, Mohammad, and Ian P. Castro. "Turbulence in rough-wall boundary layers:
universality issues." Experiments in fluids 51.2 (2011): 313-326.
[59] Belcher, S. E., N. Jerram, and J. C. R. Hunt. "Adjustment of a turbulent boundary layer
to a canopy of roughness elements." Journal of Fluid Mechanics 488 (2003): 369-398.
[60] Coceal, O., and S. E. Belcher. "A canopy model of mean winds through urban areas."
Quarterly Journal of the Royal Meteorological Society 130.599 (2004): 1349-1372.
[61] Harman, Ian N., and John J. Finnigan. "A simple unified theory for flow in the canopy
and roughness sublayer." Boundary-layer meteorology 123.2 (2007): 339-363.
[62] De Ridder, Koen. "Bulk transfer relations for the roughness sublayer." Boundary-layer
meteorology 134.2 (2010): 257-267.
[63] Arnqvist, Johan, and Hans Bergström. "Flux‐profile relation with roughness sublayer
correction." Quarterly Journal of the Royal Meteorological Society 141.689 (2015):
1191-1197.
[64] Seibert, Petra, et al. "Review and intercomparison of operational methods for the
determination of the mixing height." Atmospheric environment 34.7 (2000): 1001-1027.
[65] Deardorff, James W. "Convective velocity and temperature scales for the unstable
planetary boundary layer and for Rayleigh convection." Journal of the atmospheric
sciences 27.8 (1970): 1211-1213.
[66] Deardorff, James W. "Preliminary results from numerical integrations of the unstable
planetary boundary layer." Journal of the Atmospheric Sciences 27.8 (1970): 1209-1211.
[67] Deardorff, James W. "Numerical investigation of neutral and unstable planetary
boundary layers." Journal of the Atmospheric Sciences 29.1 (1972): 91-115.
[68] Willis, G. E., and J. W. Deardorff. "A laboratory model of the unstable planetary
boundary layer." Journal of the Atmospheric Sciences 31.5 (1974): 1297-1307.
[69] Lenschow, D. H., J. Co Wyngaard, and Wo To Pennell. "Mean-field and second-
moment budgets in a baroclinic, convective boundary layer." Journal of the
Atmospheric Sciences 37.6 (1980): 1313-1326.
[70] Hibberd, M. F., and B. L. Sawford. "A saline laboratory model of the planetary
convective boundary layer." Boundary-layer meteorology 67.3 (1994): 229-250.
173
[71] Zilitinkevich, Sergej S., et al. "The influence of large convective eddies on the surface‐
layer turbulence." Quarterly Journal of the Royal Meteorological Society 132.618
(2006): 1426-1456.
[72] Yaglom, A. M. "Similarity laws for constant-pressure and pressure-gradient turbulent
wall flows." Annual Review of Fluid Mechanics 11.1 (1979): 505-540.
[73] Paulson, Ca A. "The mathematical representation of wind speed and temperature
profiles in the unstable atmospheric SL." Journal of Applied Meteorology 9.6 (1970):
857-861.
[74] Kanda, Manabu, and Takanobu Moriizumi. "Momentum and heat transfer over urban-
like surfaces." Boundary-layer meteorology 131.3 (2009): 385-401.
[75] Nikuradse, Johann. "Gesetzmäßigkeiten der turbulenten Strömung in glatten Rohren
(Nachtrag)." Forschung im Ingenieurwesen 4.1 (1933): 44-44.
[76] Uehara, Kiyoshi, Shinji Wakamatsu, and Ryozo Ooka. "Studies on critical Reynolds
number indices for wind-tunnel experiments on flow within urban areas." Boundary-
layer meteorology 107.2 (2003): 353-370.
[77] Snyder, William H., and Ian P. Castro. "The critical Reynolds number for rough-wall
boundary layers." Journal of Wind Engineering and Industrial Aerodynamics 90.1
(2002): 41-54.
[78] Britter, R. E., and S. R. Hanna. "Flow and dispersion in urban areas." Annual Review of
Fluid Mechanics 35.1 (2003): 469-496.
[79] Bauer, Bernard O., Douglas J. Sherman, and John F. Wolcott. "Sources of Uncertainty
in Shear Stress and Roughness Length Estimates Derived from Velocity Profiles*." The
Professional Geographer 44.4 (1992): 453-464.
[80] Andreas, Edgar L., et al. "Evaluations of the von Kármán constant in the atmospheric
SL." Journal of Fluid Mechanics 559 (2006): 117-149.
[81] Jackson, P. S. "On the displacement height in the logarithmic velocity profile." Journal
of Fluid Mechanics 111 (1981): 15-25.
[82] Coceal, O., et al. "Structure of turbulent flow over regular arrays of cubical roughness."
Journal of Fluid Mechanics 589 (2007): 375-409.
[83] Leonardi, Stefano, and Ian P. Castro. "Channel flow over large cube roughness: a direct
numerical simulation study." Journal of Fluid Mechanics 651 (2010): 519-539.
[84] Nagib, Hassan M., and Kapil A. Chauhan. "Variations of von Kármán coefficient in
canonical flows." Physics of Fluids 20.10 (2008): 1518.
[85] Oke, T. R. "Street design and urban canopy layer climate." Energy and buildings 11.1
174
(1988): 103-113.
[86] Perry, Anthony Edward, William H. Schofield, and Peter N. Joubert. "Rough wall
turbulent boundary layers." Journal of Fluid Mechanics 37.02 (1969): 383-413.
[87] Grimmond, C. S. B., and Timothy R. Oke. "Aerodynamic properties of urban areas
derived from analysis of surface form." Journal of applied meteorology 38.9 (1999):
1262-1292.
[88] Bottema, Marcel. "Urban roughness modelling in relation to pollutant dispersion."
Atmospheric Environment 31.18 (1997): 3059-3075.
[89] Raupach, M. R. "Simplified expressions for vegetation roughness length and zero-plane
displacement as functions of canopy height and area index." Boundary-Layer
Meteorology 71.1-2 (1994): 211-216.
[90] Hanna, Steven, Rex Britter, and Pasquale Franzese. "Simple screening models for urban
dispersion." Proc. Eighth Conf. Harmonisation within Atmos. Disp. Mod. for Reg.
Purposes. 2002.
[91] Brutsaert, Wilfried. "A model for evaporation as a molecular diffusion process into a
turbulent atmosphere." Journal of Geophysical Research 70.20 (1965): 5017-5024.
[92] Brutsaert, Wilfried. "A theory for local evaporation (or heat transfer) from rough and
smooth surfaces at ground level." Water resources research 11.4 (1975): 543-550.
[93] Brutsaert, Wilfried. "Some exact solutions for nonlinear desorptive diffusion."
Zeitschrift für angewandte Mathematik und Physik ZAMP 33.4 (1982): 540-546.
[94] Townsend, Albert A. The structure of turbulent shear flow. Cambridge university press,
1976.
[95] Jiménez, Javier. "Cascades in wall-bounded turbulence." Annual Review of Fluid
Mechanics 44.1 (2011): 27.
[96] Head, M. R., and P. Bandyopadhyay. "New aspects of turbulent boundary-layer
structure." Journal of Fluid Mechanics 107 (1981): 297-338.
[97] Theodorsen T. “Mechanism of Turbulence”, in Proc 2nd Midwestern Conf of Fluid
Mechanics, Ohio State University, 1952.
[98] Kaimal, J. C., et al. "Spectral characteristics of surface‐layer turbulence." Quarterly
Journal of the Royal Meteorological Society 98.417 (1972): 563-589.
[99] Kaimal, J. C. "Horizontal velocity spectra in an unstable SL." Journal of the
Atmospheric Sciences 35.1 (1978): 18-24.
[100] Bradshaw, P. "‘Inactive’motion and pressure fluctuations in turbulent boundary layers."
Journal of Fluid Mechanics 30.02 (1967): 241-258.
175
[101] Wilson, J. D. "Monin-Obukhov functions for standard deviations of velocity."
Boundary-layer meteorology 129.3 (2008): 353-369.
[102] Khanna, S. A. M. I. R., and James G. Brasseur. "Analysis of Monin-Obukhov similarity
from large-eddy simulation." Journal of Fluid Mechanics 345 (1997): 251-286.
[103] McNaughton, K. G. "Attached eddies and production spectra in the atmospheric
logarithmic layer." Boundary-layer meteorology 111.1 (2004): 1-18.
[104] McNaughton, K. G. "On the kinetic energy budget of the unstable atmospheric SL."
Boundary-layer meteorology 118.1 (2006): 83-107.
[105] Hunt, Julian CR, and Jonathan F. Morrison. "Eddy structure in turbulent boundary
layers." European Journal of Mechanics-B/Fluids 19.5 (2000): 673-694.
[106] Hunt, J. C. R., and Pierre Carlotti. "Statistical structure at the wall of the high Reynolds
number turbulent boundary layer." Flow, Turbulence and Combustion 66.4 (2001): 453-
475.
[107] Högström, Ulf, J. C. R. Hunt, and Ann-Sofi Smedman. "Theory and measurements for
turbulence spectra and variances in the atmospheric neutral SL." Boundary-Layer
Meteorology 103.1 (2002): 101-124.
[108] Drobinski, Philippe, et al. "The structure of the near-neutral atmospheric SL." Journal
of the atmospheric sciences 61.6 (2004): 699-714.
[109] Drobinski, Philippe, et al. "Numerical and experimental investigation of the neutral
atmospheric SL." Journal of the atmospheric sciences 64.1 (2007): 137-156.
[110] Castillo, Marieta Cristina, Atsushi Inagaki, and Manabu Kanda. "The effects of inner-
and outer-layer turbulence in a convective boundary layer on the near-neutral inertial
sublayer over an urban-like surface." Boundary-layer meteorology 140.3 (2011): 453-
469.
[111] Inagaki, Atsushi, et al. "Large-eddy simulation of coherent flow structures within a
cubical canopy." Boundary-layer meteorology 142.2 (2012): 207-222.
[112] Wyngaard, J. C., and O. R. Coté. "The budgets of turbulent kinetic energy and
temperature variance in the atmospheric SL." Journal of the Atmospheric Sciences 28.2
(1971): 190-201.
[113] Frenzen, Paul, and Christoph A. Vogel. "Further studies of atmospheric turbulence in
layers near the surface: scaling the TKE budget above the roughness sublayer."
Boundary-layer meteorology 99.2 (2001): 173-206.
[114] Wilson, N. Robert, and Roger H. Shaw. "A higher order closure model for canopy flow."
Journal of Applied Meteorology 16.11 (1977): 1197-1205.
176
[115] Christen, Andreas, Mathias W. Rotach, and Roland Vogt. "The budget of turbulent
kinetic energy in the urban roughness sublayer." Boundary-layer meteorology 131.2
(2009): 193-222.
[116] Ashrafian, Alireza, and Helge I. Andersson. "The structure of turbulence in a rod-
roughened channel." International journal of heat and fluid flow 27.1 (2006): 65-79.
[117] Kline, S. J., et al. "The structure of turbulent boundary layers." J. Fluid Mech30.4
(1967): 741-773.
[118] Monty, J. P., et al. "Large-scale features in turbulent pipe and channel flows. "Journal
of Fluid Mechanics 589 (2007): 147-156.
[119] Adrian, Ronald J. "Hairpin vortex organization in wall turbulence." Physics of Fluids
(1994-present) 19.4 (2007): 041301.
[120] Adrian, R. J., C. D. Meinhart, and C. D. Tomkins. "Vortex organization in the outer
region of the turbulent boundary layer." Journal of Fluid Mechanics 422 (2000): 1-54.
[121] Tomkins, Christopher D., and Ronald J. Adrian. "Spanwise structure and scale growth
in turbulent boundary layers." Journal of Fluid Mechanics 490 (2003): 37-74.
[122] Hutchins, Nicholas, et al. "Towards reconciling the large-scale structure of turbulent
boundary layers in the atmosphere and laboratory." Boundary-layer meteorology 145.2
(2012): 273-306.
[123] Grass, A. J. "Structural features of turbulent flow over smooth and rough boundaries."
Journal of fluid Mechanics 50.02 (1971): 233-255.
[124] Nakagawa, Hiroji, and Iehisa Nezu. "Prediction of the contributions to the Reynolds
stress from bursting events in open-channel flows." Journal of fluid mechanics 80.1
(1977): 99-128.
[125] Finnigan, J. J. "Turbulence in waving wheat." Boundary-Layer Meteorology 16.3
(1979): 181-211.
[126] Gao, W., and R. H. Shaw. "Observation of organized structure in turbulent flow within
and above a forest canopy." Boundary Layer Studies and Applications. Springer
Netherlands, 1989. 349-377.
[127] Finnigan, JOHN J., ROGER H. Shaw, and Edward G. Patton. "Turbulence structure
above a vegetation canopy." J Fluid Mech 637 (2009): 387-424.
[128] Young, George S., et al. "Rolls, streets, waves, and more: A review of quasi-two-
dimensional structures in the atmospheric boundary layer." Bulletin of the American
Meteorological Society 83.7 (2002): 997-1001.
[129] Moeng, Chin-Hoh, and Peter P. Sullivan. "A comparison of shear-and buoyancy-driven
177
planetary boundary layer flows." Journal of the Atmospheric Sciences 51.7 (1994): 999-
1022.
[130] Schmidt, Helmut, and Ulrich Schumann. "Coherent structure of the convective
boundary layer derived from large-eddy simulations." Journal of Fluid Mechanics 200
(1989): 511-562.
[131] Hommema, Scott E., and Ronald J. Adrian. "Packet structure of surface eddies in the
atmospheric boundary layer." Boundary-layer meteorology 106.1 (2003): 147-170.
[132] Li, Dan, and Elie Bou-Zeid. "Coherent structures and the dissimilarity of turbulent
transport of momentum and scalars in the unstable atmospheric SL." Boundary-layer
meteorology 140.2 (2011): 243-262.
[133] Park, Seung-Bu, and Jong-Jin Baik. "A large-eddy simulation study of thermal effects
on turbulence coherent structures in and above a building array." Journal of Applied
Meteorology and Climatology 52.6 (2013): 1348-1365.
[134] Spiegel, E. A., and G. Veronis. "On the Boussinesq approximation for a compressible
fluid." The Astrophysical Journal 131 (1960): 442.
[135] Ogura, Yoshimitsu, and Norman A. Phillips. "Scale analysis of deep and shallow
convection in the atmosphere." Journal of the atmospheric sciences 19.2 (1962): 173-
179.
[136] Ghosal, Sandip, and Parviz Moin. "The basic equations for the large eddy simulation of
turbulent flows in complex geometry." Journal of Computational Physics 118.1 (1995):
24-37.
[137] Tominaga, Yoshihide, and Ted Stathopoulos. "Turbulent Schmidt numbers for CFD
analysis with various types of flowfield." Atmospheric Environment 41.37 (2007):
8091-8099.
[138] Schumann, Ulrich. "Subgrid scale model for finite difference simulations of turbulent
flows in plane channels and annuli." Journal of computational physics 18.4 (1975):
376-404.
[139] Smagorinsky, Joseph. "General circulation experiments with the primitive equations: I.
the basic experiment*." Monthly weather review 91.3 (1963): 99-164.
[140] OpenFOAM, ‘The open source cfd toolbox’, Website. http://www.openfoam.org/.
[141] Moukalled, F., L. Mangani, and M. Darwish. The Finite Volume Method in
Computational Fluid Dynamics. Springer, 2016.
[142] Deardorff, J. W., and G. E. Willis. "Further results from a laboratory model of the
convective planetary boundary layer." Boundary-Layer Meteorology 32.3 (1985): 205-
178
236.
[143] Letzel, Marcus Oliver, Martina Krane, and Siegfried Raasch. "High resolution urban
large-eddy simulation studies from street cavity to neighbourhood scale." Atmospheric
Environment 42.38 (2008): 8770-8784.
[144] Sullivan, Peter P., and Edward G. Patton. "The effect of mesh resolution on convective
boundary layer statistics and structures generated by large-eddy simulation." Journal of
the Atmospheric Sciences 68.10 (2011): 2395-2415.
[145] Liu, Chun-Ho, and Mary C. Barth. "Large-eddy simulation of flow and scalar transport
in a modeled street cavity." Journal of Applied Meteorology 41.6 (2002): 660-673.
[146] Liu, Chun-Ho, Mary C. Barth, and Dennis YC Leung. "Large-eddy simulation of flow
and pollutant transport in street cavities of different building-height-to-street-width
ratios." Journal of Applied Meteorology 43.10 (2004): 1410-1424.
[147] Louka, P., et al. "Thermal effects on the airflow in a street cavity–Nantes' 99
experimental results and model simulations." Water, air and soil pollution: Focus 2.5-
6 (2002): 351-364.
[148] Raupach, M. R., P. A. Coppin, and B. J. Legg. "Experiments on scalar dispersion within
a model plant canopy part I: The turbulence structure." Boundary-Layer Meteorology
35.1-2 (1986): 21-52.
[149] Mejia-Alvarez, R., Y. Wu, and K. T. Christensen. "Observations of meandering
superstructures in the roughness sublayer of a turbulent boundary layer." International
Journal of Heat and Fluid Flow 48 (2014): 43-51.

179

Вам также может понравиться