Effects of trees on mean wind, turbulence and momentum exchange within and above a real urban environment

Effects of trees on mean wind, turbulence and momentum exchange within and above a real urban environment

Accepted Manuscript Effects of trees on mean wind, turbulence and momentum exchange within and above a real urban environment M.G. Giometto, A. Chris...

14MB Sizes 3 Downloads 57 Views

Accepted Manuscript

Effects of trees on mean wind, turbulence and momentum exchange within and above a real urban environment M.G. Giometto, A. Christen, P.E. Egli, M.F. Schmid, R.T. Tooke, N.C. Coops, M.B. Parlange PII: DOI: Reference:

S0309-1708(17)30648-6 10.1016/j.advwatres.2017.06.018 ADWR 2878

To appear in:

Advances in Water Resources

Received date: Revised date: Accepted date:

22 June 2016 14 June 2017 22 June 2017

Please cite this article as: M.G. Giometto, A. Christen, P.E. Egli, M.F. Schmid, R.T. Tooke, N.C. Coops, M.B. Parlange, Effects of trees on mean wind, turbulence and momentum exchange within and above a real urban environment, Advances in Water Resources (2017), doi: 10.1016/j.advwatres.2017.06.018

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

ACCEPTED MANUSCRIPT

Highlights • LES and tower measurements are used to quantify effect of urban trees

CR IP T

on flow • Roughness length and displacement height increase with increasing foliage density

• The ratio of turbulent to mean kinetic energy increases with increasing foliage

AC

CE

PT

ED

M

AN US

• Dense trees reduce downward turbulent transport of high-momentum fluid.

1

ACCEPTED MANUSCRIPT

Effects of trees on mean wind, turbulence and momentum exchange within and above a real urban environment

CR IP T

M. G. Giomettoa,∗, A. Christenb , P. E. Eglic , M. F. Schmida , R. T. Tooked , N. C. Coopsd , M. B. Parlangea a Department

AN US

of Civil Engineering, Faculty of Applied Science, University of British Columbia, Vancouver, BC, Canada b Geography / Atmospheric Science Program, University of British Columbia, Vancouver, BC, Canada c Institute of Earth Surface Dynamics, University of Lausanne, Lausanne, Switzerland d Department of Forestry, University of British Columbia, Vancouver, BC, Canada

Abstract

Large-eddy simulations (LES) are used to gain insight into the effects of trees on turbulence, aerodynamic parameters, and momentum transfer rates

M

characterizing the atmosphere within and above a real urban canopy. Several areas are considered that are part of a neighbourhood in the city of Vancouver, BC, Canada where a small fraction of trees are taller than buildings. In this

ED

area, eight years of continuous wind and turbulence measurements are available from a 30 m meteorological tower. Data from airborne light detection and

PT

ranging (LiDAR) are used to represent both buildings and vegetation at the LES resolution. In the LES algorithm, buildings are accounted through an immersed boundary method, whereas vegetation is parameterized via a location-

CE

specific leaf area density. LES are performed including and excluding vegetation from the considered urban areas, varying wind direction and leaf area density. Surface roughness lengths (z0 ) from both LES and tower measurements are

AC

sensitive to the 0 ≤ LAI/λbf < 3 parameter, where LAI is the leaf area index

and λbf is the frontal area fraction of buildings characterizing a given canopy. For instance, tower measurements predict a 19% seasonal increase in z0 , slightly ∗ Corresponding

author Email address: [email protected] (M. G. Giometto)

Preprint submitted to Advances in Water Resources

June 27, 2017

ACCEPTED MANUSCRIPT

lower than the 27% increase featured by LES for the most representative canopy (leaves-off LAI/λbf = 0.74, leaves-on LAI/λbf = 2.24). Removing vegetation from such a canopy would cause a dramatic drop of approximately 50% in z0

CR IP T

when compared to the reference summer value. The momentum displacement height (d) from LES also consistently increases as LAI/λbf increases, due in

large part to the disproportionate amount of drag that the (few) relatively

taller trees exert on the flow. LES and measurements both predict an increase

in the ratio of turbulent to mean kinetic energy (TKE/MKE) at the tower sampling height going from winter to summer, and LES also show how including

AN US

vegetation results in a more (positive) negatively skewed (horizontal) vertical

velocity distribution – reflecting a more intermittent velocity field which favors sweep motions when compared to ejections. Within the urban canopy, the effects of trees are twofold: on one hand, they act as a direct momentum sink for the mean flow; on the other, they reduce downward turbulent transport of high-momentum fluid, significantly reducing the wind intensity at the heights

M

where people live and buildings consume energy.

Keywords: aerodynamic roughness, roughness sublayer, trees, turbulence,

ED

urban canopy, urban forest, vegetation, wind.

PT

1. Introduction

The physical structure of cities controls and modifies the exchange of momentum, heat, water, and air pollutants between the surface and the atmo-

CE

sphere. The three-dimensional form of cities is intrinsically spatially variable,

5

and this influences the properties of wind, turbulence and turbulent exchange on

AC

street, neighbourhood and city scales (Belcher, 2005). Up to a height of several times the size of buildings – in the layer known as roughness sublayer (RSL) – flow statistics are strongly dependent on the actual shape and arrangement of buildings, and Monin-Obukhov similarity (MOS) scaling is generally inaccurate

10

(Roth, 2000). Field experiments and numerical studies have shown that in the surface layer above the RSL – the inertial sublayer (ISL) – the air flow over

3

ACCEPTED MANUSCRIPT

cities (and complex surfaces in general) becomes statistically homogeneous, and the neutral flow is completely determined by the friction velocity u∗ , aerodynamic roughness length z0 and zero-plane displacement d, and is obeying MOS scaling for diabatic situations (Parlange and Brutsaert, 1989; Brutsaert et al.,

CR IP T

15

1989; Parlange and Brutsaert, 1990; Brutsaert and Parlange., 1992; Katul and Parlange, 1992; Parlange et al., 1995; Feigenwinter et al., 1999; Bou-Zeid et al., 2004; Yang, 2016). Aerodynamic properties of cities control exchange rates of

momentum and scalars such as heat and humidity and therefore have impor20

tant implications for weather and hydrological modelling at urban and regional

AN US

scales (Brutsaert, 1982).

The term “urban canopy” refers to the assemblage of buildings, built objects and vegetation that make up the three dimensional interface between the land surface and the atmosphere in cities. In many low- to medium-density urban 25

areas, vegetation, including tall trees, forms an important component of the urban canopy (Oke et al., 2017). Trees can significantly modify meteorological and

M

hydrological processes in the RSL and also affect the flow in the ISL above (Oke, 1989). Furthermore, trees are an important design tool to manage building and

30

ED

street climates (e.g. Oke, 1988; Shashua-Bar and Hoffman, 2000; Johansson and Emmanuel, 2006; Bowler et al., 2010). At the coarser scales, urban vegetation is relevant for managing heat stress and air quality at neighborhood to city

PT

scales. For example, model simulations by Li and Bou-Zeid (2013) suggest that urban forestry is a critical tool to prevent exacerbated urban-rural temperature

CE

differences during heat waves. To ensure safety and health of urban dwellers, and to efficiently manage

35

energy and water in urban systems, we need numerical models that represent

AC

weather, climate and hydrology under current conditions and future climate and planning scenarios (Chen et al., 2012). The presence of urban canopies on the atmosphere is generally represented in hydrology and mesoscale weather models

40

using “urban land surface models” (ULSMs). ULSMs describe the energetics, hydrology and dynamics of the urban surface, by either implying certain bulk values or simulating the behaviour of different tiles such as roofs, walls, roads 4

ACCEPTED MANUSCRIPT

and urban vegetation. More complex ULSMs targeted to resolve the climate of the RSL and inside 45

the urban canopy are called urban canopy parameterizations (UCPs). UCPs

CR IP T

represent the three-dimensional geometry of an urban surface, and account for interactions between tiles, the atmosphere, the building interior and subsurface, often applying a multi-layer approach (Masson, 2005).

Recent assessments of our capabilities to accurately reproduce land-atmosphere 50

exchange in urban environments using ULSMs have shown that the lack of, or a

crude and improper representation of urban vegetation, in particular the direct

AN US

interaction of trees with buildings and vice versa, is a major weakness. The

international urban energy balance model comparison concluded that “taking vegetation cover into account (or not) significantly impacts model performance” 55

in particular related to energy partitioning and evapotranspiration (Grimmond et al., 2010).

Trees interact in a three-dimensional way with buildings radiatively (e.g. shading

M

and radiation trapping), energetically (e.g. transpiration), and dynamically (e.g. wind sheltering). However, most of UCPs to date do not include trees as three dimensional objects. Recently, there have been efforts to represent the three-

ED

60

dimensional radiative and energetic effects in urban modelling systems (Lee and Park, 2008; Krayenhoff et al., 2014; Ryu et al., 2015; Lee et al., 2016).

PT

The dynamical impact of urban trees, however, is not included in operational ULSMs and UCPs.

The dynamical effect is not only referring to the sheltering effect and change

CE

65

of scaling parameters z0 and d, but also the fact that trees modify turbulence

and consequently turbulent length scales that affect transfer of scalars and dis-

AC

sipation rates in the RSL. Krayenhoff et al. (2015) use CFD simulations over generic urban geometries to inform parameterizations of the dynamical effect of

70

urban trees for given densities and clumping. The influence of trees is also timedependent. Deciduous trees in mid- and high-latitude cities change their leaf area density (a in the following) with season. Although some ULSMs for hydrology applications prescribe changing leaf area densities over the year (e.g. J¨ arvi 5

ACCEPTED MANUSCRIPT

et al., 2011), no operational USLM considers the effect of seasonally changing a 75

on mean wind and turbulent exchange. The overall goal of this study is to assess the role of urban trees and changing

CR IP T

leaf densities on the flow dynamics within and above a realistic urban canopy. This helps to inform the next-generation ULSMs when, where and how incorporating or not the effects of urban vegetation at the scale of neighbourhoods 80

or cities is appropriate and possible. The specific research questions are:

• To determine the effect of trees and changing leaf density on the aerody-

AN US

namic roughness length and zero-plane displacement in the urban environment.

• To test the sensitivity of integral turbulence statistics in the ISL above the urban canopy to the characteristics of the urban vegetation.

85

• To characterize the effect of trees and changing a on vertical transfer of

M

momentum and mean kinetic energy in the urban RSL.

A combination of detailed large-eddy simulations (LES) and long-term mea-

90

ED

surements from a tower over an urban canopy in Vancouver, Canada, is used to answer these questions. A Digital Surface Model (DSM) derived from airborne Light Detection and Ranging (LiDAR) data is used to characterize the

PT

surface and represent both buildings and vegetation at high spatial resolution in the LES. The LES is run under the same forcing with the DSM including

CE

vegetation at various densities and once without vegetation (reference case). 95

The latter condition is of course not possible to reproduce on the tower measurements. Measurements from the tower are instead conditionally sampled

AC

discerning between leaves-on (Summer) and leaves-off (Winter) seasons to investigate the effect of varying leaf area density in the real atmosphere.

100

To date, numerical simulations of the flow and turbulence over cities with

buildings resolved (as opposed to modelled) have primarily represented cities as idealized rectangular patches of different surface roughness and arrays of cubes or cuboids with varying spatial organization (Kanda et al., 2004; Kanda, 6

ACCEPTED MANUSCRIPT

2006; Xie and Castro, 2006; Coceal et al., 2007b; Xie et al., 2008; Inagaki et al., 2012; Anderson et al., 2015; Li et al., 2016b; Yang et al., 2016; Yang, 2016; 105

Sadique et al., 2016; Anderson, 2016; Zhu et al., 2016). In addition, several

CR IP T

studies modelled the aerodynamic effect of urban trees using simple, generic street geometries to investigate the effect of trees on pollutant dispersion (Balcz´ o et al., 2009; Salim et al., 2011; Moonen et al., 2013; Gromke and Blocken, 2015;

Vranckx et al., 2015). Such idealized studies show consistently that trees reduce 110

the vertical exchange of scalars and deteriorate air quality when sources are below the canopy, a conclusion also supported by wind tunnel simulations (e.g.

AN US

Gromke and Ruck, 2008).

With increasing parallel computing power and the availability of Digital Surface Models (DSMs) it is now feasible to simulate subsets of a real cities (e.g. 115

Xie and Castro, 2009; Kanda et al., 2013; Giometto et al., 2016; Hertwig et al., 2016a,b). However, to the best of our knowledge, no LES study in the literature has simulated a real urban neighborhood including the effect of individually

M

resolved, clumped trees. In the absence of buildings, the effect of a plant canopy on airflow and turbulence and, inversely, the effect of flow on trees, has been studied using LES assuming uniform a distributions (e.g. Shaw and Schumann,

ED

120

1992a; Kanda and Hino, 1994; Dwyer et al., 1997; Finnigan et al., 2009a) and more recently with explicitly resolved plants and realistically clumped a (Yue

PT

et al., 2007; Huang et al., 2009; Bailey and Stoll, 2013; Patton et al., 2016). Likewise, observational studies in cities have provided limited insight on the role of urban trees. Most studies quantified turbulence generated within and

CE

125

above urban canopies using multiple measurements located at different heights on towers in the RSL and/or ISL (e.g. Rotach, 1993; Feigenwinter et al., 1999;

AC

Christen et al., 2007; Nelson et al., 2007; Christen et al., 2009; Ramamurthy and Pardyjak, 2011, 2015). However, observational studies with simultaneous

130

measurements within and above urban canopies have targeted urban canopies without or with very limited trees, and no observational study has been able to isolate the effect of trees on mean wind and turbulence from the effect of buildings. 7

CE

PT

ED

M

AN US

CR IP T

ACCEPTED MANUSCRIPT

Figure 1: Filled color contour of buildings and vegetation height (hb and hv respectively) in a 974 × 620 m2 area surrounding the 28.8 m “Vancouver-Sunset” meteorological tower in

AC

Vancouver, BC, Canada. Surface data are based on LiDAR measurements. Selected subsets for use in large-eddy simulation (LES) are highlighted (S1, S2, S3, and S4).

8

ACCEPTED MANUSCRIPT

2. Methods 135

2.1. Study area

CR IP T

Figure 1 depicts a 974×620 m2 area in south central Vancouver, BC, Canada, centered on the micrometeorological research tower “Vancouver-Sunset” (Fluxnet ID “Ca-VSu”, 123.0784◦ W, 49.2261◦ N, 78 m a.s.l.). The area is characterized by a relatively homogeneous mix of streets and detached one to two storey 140

houses (12.8 Bldg. ha−1 ), and is representative of the majority of Vancouver’s neighbourhoods and North American cities (Local Climate Zone 6, according to

AN US

Stewart and Oke (2012)). The area has been the focus of a number of observa-

tional campaigns related to flow and turbulence characteristics over suburban surfaces (Schmid et al., 1991; Roth and Oke, 1993; Roth, 1993; Roth and Oke, 145

1995; Grimmond et al., 1998; Grimmond and Oke, 2002)

Vegetation in the area (circle of 400 m radius around tower) has a density of 19.2 Stems ha−1 and comprises of about 23% evergreen trees and 77% deciduous

M

trees (Liss et al., 2013). The large fraction of deciduous trees is resulting in a strong seasonal variation of the Leaf Area Index (LAI). Liss et al. (2013) estimates that the LAI of trees is 0.39 m2 m−2 in summer. The average tree

ED

150

height is estimated 8.4 m for broadleaf trees (n = 742) and 11.1 m (n = 223) for coniferous trees.

PT

The entire surface (S4) displayed in Fig. 1 and three smaller surfaces (S1, S2 and S3) comprised within S4 are selected for use in LES. The reason to 155

separately consider a number of sub-surfaces is to provide a broader range of

CE

realistic canopies, each characterized by its peculiar distribution of buildings and trees, to further support conclusions and to study the dependence of flow

AC

statistics on relevant surface parameters (such as LAI and the frontal-area index of buildings). For the considered wind approaching direction (see Sect. 2.4.2), all

160

surfaces (S1, S2, S3 and S4) are mostly within the 70%, and entirely within the 80% long-term cumulative footprint of the tower. The 80% long-term cumulative footprint climatology covers an area of 1.04 km2 and is shown in Fig. 1 of

Crawford and Christen (2015a).

9

ACCEPTED MANUSCRIPT

2.2. Light detection and ranging data The detailed building heights and vegetation information used in the LES

165

are derived from a LiDAR point cloud dataset, mapped to a two-dimensional

CR IP T

horizontal Cartesian grid with stencil of 0.5 m in both directions. The LiDAR dataset was acquired in February 2014, during leaves-off conditions, which al-

lowed for good penetration of the laser through the vegetation canopies. Ex170

tracted buildings information comprise the building heights (hb ), function of the horizontal coordinates x and y, whereas vegetation elements are represented as a spatially distributed leaf area density (a) coefficient. The a coefficient has been

AN US

computed as a function of the foliage probability distribution (Pf ) following Tooke et al. (2011):

a(x, λ, k) = − ln [1 − Pf (x, λ, k)] , 175

(1)

where the functions λ, k : (x, y) → R+ and where Pf is itself approximated as a two parameters Weibull distribution

k  z (k−1) exp [−(z/λ)k ] . λ λ

M

Pf =

(2)

ED

The λ(x, y) > 0 and k(x, y) > 0 are the scale and shape parameters, and are evaluated based on a non-linear fit to the LiDAR beam interception probability, representing probability of branches and evergreen foliage at a given height. A random spatial orientation of leaves was assumed. To calculate summer a, the

PT

180

overall LAI resulting from the LiDAR dataset has been rescaled to match the

CE

LAI = 0.39 value of Liss et al. (2013), representative of the area in the summer season.

AC

2.3. Tower-based turbulence measurements

185

The 30 m triangular lattice tower that hosted turbulence instrumentation

is situated 5 m from the south-eastern corner of a power substation (Fig. 1). Continuous long-term measurements were made from 1 May, 2008 to 30 April, 2016 at a single height, using a CSAT-3 three-axis ultrasonic anemometerthermometer (Campbell Scientific, Logan, UT, USA) operated at 28.4 m above 10

ACCEPTED MANUSCRIPT

190

local ground facing an azimuth of 180◦ . Because the base of the tower is recessed by 4 m from the surrounding terrain, an effective measurement height of z=24.4 m is used for the measurements in further calculations (Crawford and Christen,

CR IP T

2015b). Measured quantities include the instantaneous three components of the wind vector (u,v,w, all rotated into the mean wind using a double-rotation) sam195

pled at 60 Hz and stored at 20 Hz. Block averages over 30 min were calculated and filtered following the procedures outlined in Crawford et al. (2012).

Conditional sampling of the tower data has been performed, isolating 30 minute blocks with a neutral dynamic stability following the criteria of −0.075 < 200

AN US

(z − d)/L < 0.075, where d is the zero-plane displacement (set to d = 5.46 m, see Sect. 4.2), and L is the Obukhov-Length measured on the tower as: L=

T u∗ κ g θ∗

(3)

where κ is the von K´ arm´ an constant (we assumed κ = 0.4), T the absolute 2

2

temperature, u∗ the friction velocity calculated as u∗ = (u0 w0 + v 0 w0 )0.25 and

M

θ∗ the temperature scale calculated as θ∗ = −w0 T 0 /u∗ . To ensure consistent wind direction over the 30 minute block, the RMSE of the cyclic wind direction (α) calculated for 6 subintervals of 5 minutes each, was set to not exceed 15◦ in

ED

205

each block. Only data with a wind speed of > 2 m s−1 and wind direction 150◦ ≤

α ≤ 300◦ was considered, hence approaching between SSE and WNW (excluding

PT

sectors where potential flow distortion is found by the instrumentation and the wake of a nearby school building). This resulted in 4700 datasets further considered (3.4% of the period of 8 years, or 3.2 months of data). To avoid effects

CE

210

of seasonally changing wind direction to affect results, conditional averages were formed as follows: Averages and statistics were calculated separately for 30◦

AC

wind direction sectors and each month of the year. To calculate summer values, conditional averages from all 30◦ sectors between 150◦ and 300◦ and all months

215

between May and September were selected and equally weighted.

11

ACCEPTED MANUSCRIPT

2.4. Numerical algorithm The LES algorithm adopted in this study was originally developed in Albertson and Parlange (1999a,b) and updated to develop linear and non-linear

220

CR IP T

LES closure models (Meneveau et al., 1996; Port´e-Agel et al., 2000; Bou-Zeid

et al., 2005; Lu and Port´e-Agel, 2010), wall models (Anderson and Meneveau,

2010, 2011; Anderson et al., 2012; Hultmark et al., 2013), and has been extensively used in wind energy applications (Calaf et al., 2010, 2011; Sharma et al., 2016), urban applications (Tseng et al., 2006; Bou-Zeid et al., 2009; Cheng and

Port´e-Agel, 2013, 2015; Anderson et al., 2015; Giometto et al., 2016) and in studies of flow over vegetation canopies (Chester and Meneveau, 2007; Chester

AN US

225

et al., 2007; Yue et al., 2007, 2008; Bailey and Stoll, 2013).

It integrates the spatially filtered Navier-Stokes equations in their rotation form

M

∂u ˜i ∂u ˜i ∂u ˜j ∂π ˜ ∂τij +u ˜j ( − )=− − + Fi + f˜ib + f˜iv ∂t ∂xj ∂xi ∂xi ∂xj ∂u ˜i = 0, ∂xi

(5)

ED

Ω × [0, T ] where Ω denotes a regular domain, T is the total integration

in 230

(4)

time, u ˜i are the space filtered velocity components in the x, y and z Cartesian coordinate directions, π ˜ is a modified filtered kinematic pressure (˜ π = +

1 ˜i u ˜i ), 2u

+

ρ is a reference constant density, τij denotes the subgrid-scale

PT

1 3 τii

1 ˜ ρp

(SGS) kinematic stress tensor, and Fi ≡ (u2∗ /Lz , 0, 0) is a constant kinematic pressure gradient forcing that is introduced to drive the flow in the x coordinate direction. In addition, f˜ib is an immersed-boundary method (IBM) forcing term

CE 235

that is used to impose the desired boundary condition at the building interface

AC

and in the below-interface locations (it has a finite value at and inside the buildings interface and is zero elsewhere), and f˜iv is a volumetric force that is introduced to account for drag from vegetation elements (see Sect. 2.4.1).

240

The equations are solved on a regular domain Ω using a mixed pseudospec-

tral - finite differences algorithm (Orszag, 1969, 1970), time advancement is performed using an explicit second-order accurate Adam Bashforth scheme, and 12

ACCEPTED MANUSCRIPT

the pressure field is evaluated through a fractional step method (Chorin, 1968). A free-lid boundary condition is imposed at the top of the domain (H), periodic 245

boundary conditions apply in the horizontal directions, due to the pseudospec-

CR IP T

tral approach, and zero velocity is enforced at Γb and below via the f˜ib forcing term (this approach is known in literature as the discrete forcing approach IBM

method (Mittal and Iaccarino, 2005)). In addition, tangential stresses are applied in a band centered around the buildings interface Γb via a logarithmic equi250

librium wall-law. Specifically, the band is defined as −1.1∆ ≤ Φ(x, y, z) ≤ 1.1∆, where Φ denotes the distance from Γb and ∆ = (∆x×∆y×∆z)1/3 , where ∆x, ∆y

AN US

and ∆z are the grid stencils in the three coordinate directions. The wall-law reads τw = −



κ(k˜ ut k2 ) ln (1 + ∆/z0 )

2

in

Γb × [0, T ] ,

(6)

where u ˜t = u ˜ − (˜ u · n)n is the tangential-to-surface velocity vector, n is the 255

surface normal vector, and z0 is the aerodynamic roughness length parameter.

M

τw is the kinematic surface stress defined in the previously introduced band and 0 0 it corresponds to τ13 = τ31 in a local Cartesian coordinate system centered at

ED

the given collocation node and defined by e01 = u ˜ t /k˜ uk, e02 = (˜ ut × n)/k˜ uk,

e03 = n. After evaluation τw is rotated back in the {e1 , e2 , e3 } canonical co-

260

ordinate system to provide τij at Γb . Gibbs oscillations that arise when IBM

PT

methods are implemented within spectral-like discretizations are here reduced via a Laplacian smoothing approach, as described in Chester et al. (2007). Al-

CE

ternative approaches to perform such a task have also been proposed recently (Fang et al., 2011; Li et al., 2016a) A dynamic Smagorinsky LES model with planar (x, y) averaged coefficient

265

AC

is adopted to evaluate τij from spatially-resolved quantities, thus closing the system of equations 4 and 5 in the bulk of the flow (Germano et al., 1991; Lilly, 1992). The dynamic model is based on the mixing length concept, and evaluates the SGS kinematic stresses as a function of the local resolved strain rate tensor: ˜ 2 S˜ij , τij = −2νt S˜ij = −2(cs,∆ ∆)2 kSk 13

(7)

ACCEPTED MANUSCRIPT

270

where cs,∆ is the Smagorinsky coefficient at grid scale ∆ (Scotti et al., 1993). Unlike the original Smagorinsky model (Smagorinsky, 1963), which requires an ad-hoc specification of cs,∆ , the dynamic model computes cs,∆ dynamically

CR IP T

by taking advantage of information contained in the smallest resolved scales (the scales between ∆ and a test-filter scale e.g. 2∆). More sophisticated 275

Smagorinsky-type SGS models have been proposed during the past two decades,

see for instance the works of Meneveau et al. (1996); Port´e-Agel (2004); Bou-Zeid et al. (2005), and a plethora of other approaches exist to close the space-filtered

Navier-Stokes equations (Meneveau and Katz, 2000). However, when it comes

280

AN US

to LES of flow over bluff elements such as urban canopies, pressure (or form)

drag is the main sink of momentum and SGS contributions are – on average – only a minor percentage of the overall momentum flux (see e.g., Giometto et al., 2016). Because of this, the velocity field is known to be poorly dependent on the specific SGS approach when compared to flow over streamlined elements, and simple and efficient closure models yield accurate results. Note that these conclusions do not apply to convective scalar transfer within these same envi-

M

285

ronments, as scalar transfer occurs via molecular diffusion independently of the

ED

shape of the obstacles (see e.g., Li et al., 2016b). Note also that using Eq. 6 for the specification of surface stresses, in conjunction with Eq. 7, results in a system of equations that does not depend on the molecular viscosity parameter, hence guaranteeing scaling similarity of solutions (i.e., normalized solutions are

PT

290

independent of the imposed friction velocity or outer velocity U∞ ). A more de-

CE

tailed description of the LES algorithm and specifics on the IBM implementation can be found in Chester et al. (2007).

AC

2.4.1. Representation of vegetation in the LES

295

The term f˜iv is included in the momentum conservation equation (Eq. 4) to

account for the kinematic drag from vegetation elements, and is defined as f˜iv (x, t) = −CV a ˜(x, λ, k)˜ ui k˜ uk2 ,

14

(8)

ACCEPTED MANUSCRIPT

where CV is a constant sectional drag coefficient and a ˜ is the spatially-filtered (filter width = ∆) leaf area density, derived from high-resolution LiDAR measurements. CV = 0.2 is assumed, which is the optimized value determined by Katul and Albertson (1998) and used in Krayenhoff et al. (2015).

CR IP T

300

Vegetation drag is set proportional to a velocity squared, in line with previous studies of flow over vegetation canopies (Shaw and Schumann, 1992b; Finni-

gan et al., 2009b). This implies that leaves elements are assumed to be in fully rough regime (i.e., the total drag from vegetation elements is form drag), which 305

is an acceptable approximation within the goal of the current study (Thom,

AN US

1968; Brunet et al., 1994).

In addition, it is known that the turbulent wakes behind vegetation elements, resulting from the action of pressure and viscous drag on branches and leaves, yield a transfer of energy from the mean flow and from larger-scale turbulence 310

to smaller-scale turbulence, thus short circuiting the inertial energy cascade (e.g., Wilson, 1988; Poggi et al., 2004; Cava and Katul, 2008). Wake turbulence

M

dissipates quickly, but it is a non-negligible contribution to the SGS kinetic energy budget (Shaw and Patton, 2003), and should therefore be accounted for

315

ED

in 1.5-order closure schemes, in which SGS TKE is predicted with a prognostic energy equation.

First-order algebraic LES closures – such as the dynamic Smagorinsky model

PT

used herein – are relatively simple and do not carry information on SGS energetics. As such, they are not amenable to modifications to account for the

CE

aforementioned effects induced by wake motions on SGS dynamics. This might 320

limit their applicability in flows over vegetation when the canopy occupies a significant fraction of the airspace, thus constantly sustaining small-scale wake

AC

turbulence and affecting SGS dissipation rates. However, the behavior of Smagorinsky models is expected to improve when

used in flows within explicitly resolved sparse canopies with sub-tree grid reso-

325

lution as in the current settings. Under such conditions the energetic range of wake motions, which are about the size of full vegetation elements (Cava and Katul, 2008), is explicitly resolved by LES, whereas SGS wake turbulence is 15

ACCEPTED MANUSCRIPT

expected to survive over a limited portion of the overall airspace. This is due to the sparsity of trees and to the rapid dissipation rates of SGS wake energy, 330

which support the working conjecture that SGS wake energy can be neglected

CR IP T

with little loss in accuracy in the predicted velocity field.

ED

M

AN US

2.4.2. Setup of LES simulations

Figure 2: Pseudocolor plot of the instantaneous streamwise velocity field corresponding to

PT

simulation S4SV W (see Table 1).

To quantify the effect of vegetation on mean flow and turbulence over and

CE

within the selected urban canopies (S1, S2, S3 and S4 in Fig. 1), three vegetative configurations are considered for each surface, namely

AC

335

• Urban surfaces with no vegetation (NV); • Urban surfaces with wintertime vegetation (WV); • Urban surfaces with summertime vegetation (SV).

Table 2 displays the overall LAI for the S1, S2, S3 and S4 canopies in the sum340

mertime, which are based on LiDAR data. The LiDAR-derived a ˜(x) has been 16

ACCEPTED MANUSCRIPT

rescaled so that for a 1 × 1 km2 surface centered around the measurement tower the LAI matches the one reported in Liss et al. (2013) for the same area. To simulate wintertime vegetative conditions the LAI value of each surface is reduced

345

CR IP T

by an estimated factor of 3, based on separate LAI determined for broadleaf and coniferous trees in the study area by Liss et al. (2013), assuming broadleaf trees are all deciduous and considering the effect of trunks and branches to the total projected area in winter. Table 1 summarizes the setup of all simulations.

Table 1: Summary of simulation parameters. Surfaces correspond to those displayed in Fig. 1.

Lx , Ly , Lz and Nx , Ny , Nz denote the domain size and number of collocation nodes in the

AN US

x, y, z coordinate directions, and α denotes the approaching wind direction (α = 180◦ → wind approaching from south, and α = 270◦ → wind approaching from west). Notation: S1α β, α = NV (no vegetation) or α = WV (winter vegetation) or α = SV (summer vegetation); β = W (wind from west) or β = S (wind from south).

Label

surface

Lx × Ly × Lz (m)

Nx × Ny × Nz

α (deg)

LAI

NV S1NV W , S1S

410 × 210 × 96

320 × 160 × 192

180, 270

0

S1WV S

S1

320 × 160 × 192

0.094

S1SV S

410 × 210 × 96

180, 270

S1SV W ,

S1

320 × 160 × 192

180, 270

0.282

NV S2NV W , S2S

410 × 210 × 96

S2

410 × 210 × 96

320 × 160 × 192

180, 270

0

S2

410 × 210 × 96

320 × 160 × 192

180, 270

0.082

S2

410 × 210 × 96

320 × 160 × 192

180, 270

0.246

S3

380 × 410 × 96

320 × 320 × 192

180, 270

0

S3

320 × 320 × 192

180, 270

0.152

SV S3SV W , S3S

380 × 410 × 96

S3

320 × 320 × 192

180, 270

0.458

NV S4NV W , S4S

380 × 410 × 96

S4

974 × 620 × 192

512 × 320 × 384

180, 270

0

S4

974 × 620 × 192

512 × 320 × 384

180, 270

0.117

S4

974 × 620 × 192

512 × 320 × 384

180, 270

0.351

S2WV S

S2SV W ,

S2SV S

PT

NV S3NV W , S3S

ED

S2WV W ,

S3WV S

CE

S3WV W ,

AC

S4WV W ,

S4WV S

SV S4SV W , S4S

M

S1

S1WV W ,

Two approaching wind directions are considered for each surface, namely

west and south, corresponding to situations excluding the sectors influenced by

350

sensor mounting and the unusually tall school-building to the East (see Fig. 1). The Cartesian reference system in each case is chosen so that the x axis points 17

ACCEPTED MANUSCRIPT

in the streamwise direction, and the y axis points in the spanwise direction (the reference system displayed in Fig. 1 corresponds to that used in simulations with west approaching wind angle). To evaluate z0 in equation 6 the approach of Giometto et al. (2016) is adopted, which results in z0 = ∆/15 for the considered

CR IP T

355

runs. Simulations are initialized imposing a velocity field obtained from LES at

twice as coarse resolution for each case, corresponding to a condition of dynamic equilibrium of the system (Monin, 1977). LES are then run for an additional

T ≡ h/u∗ = 400 and temporal averaging is carried out over the last T = 150. This yields well converged statistics throughout the boundary layer.

AN US

360

3. Averaging

The intrinsic double averaging (DA) approach is used herein (Nikora et al., 2007). Within the DA framework, a generic variable θ(x, y, z, t) can be decom-

(9)

and a fluctuation component (θ0 ≡ θ − hθi). Specifically, the intrinsic spatial

ED

365

M

posed into a space and time average component Z Z Z Z 1 1 hθi(z) ≡ θ(r, τ )dV dτ , T T VF (z) VF (z)

averaging is here performed first, over horizontal slabs of total volume VF (z) = Lx × Ly × ∆z in the fluid domain only, and the time averaging is performed

PT

afterwards.

CE

4. Results and discussion Using data from the airborne LiDAR, from LES runs and from turbulence

370

instruments on the tower, we identify, isolate and discuss the importance of the

AC

urban tree canopy on flow and turbulence above and within the urban canopy. Section 4.1 summarizes the characteristics of the urban building and vege-

tation layers extracted from LiDAR. In section 4.2 we then present the effect

375

of trees on integral scaling parameters z0 and d. Next, we quantify turbulence characteristics and turbulent exchange in the ISL (Section 4.3) and provide a

18

ACCEPTED MANUSCRIPT

detailed analysis of the vertical exchange of momentum and energy in the RSL (Section 4.4).

CR IP T

4.1. LiDAR derived characteristics of buildings and vegetation Relevant geometric statistics for the subsets S1, S2, S3 and S4 extracted from

380

the LiDAR point cloud are reported in Tab. 2. The range of plan area fractions of buildings (λbp ≡ Abp /A0 ) and frontal area fractions of buildings (λbf ≡ Abf /A0 )

in absence of vegetation, classify the selected surfaces as moderately dense with a wake-interference type flow regime (Oke, 1988; Britter and Hanna, 2003; Barlow,

2014). Here, Abp (m2 ) is the total plan area of buildings, Abf (m2 ) is the total

AN US

385

frontal (silhouette) area of buildings, and A0 ≡ Lx × Ly (m2 ) is the total planar area of the considered subset. The tree canopy cover (plan area fraction of trees, λvp ≡ Avp /A0 ranges between 9 and 13%, and matches well the reported 11% for a

1.9 × 1.9 km box surrounding the tower (Kellett et al., 2013). λvp is substantially 390

lower that the average tree canopy cover for the City of Vancouver with 18%

M

(City of Vancouver) and lower than the tree canopy coverage reported for US North-Western urban forests of 33% in (Nowak and Crane, 2002). The four

ED

considered surfaces S1, S2, S3 and S4 are relatively similar in terms of geometry of building and vegetation characteristics, S3 has a slightly higher amount of 395

taller trees and a higher LAI, whereas S2 is the most sparsely vegetated subset.

PT

As is typical of realistic urban canopies, and in contrast to uniform-height, synthetically-generated cube-array surfaces, the building plan and frontal area fractions are a function of z throughout the RSL. The vertical structure of

CE

the buildings plan and frontal area fractions, as well as the vegetation plan

400

area fraction and vertical profiles of the horizontally-averaged (fluid only) leaf

AC

area density (h˜ ai), important parameters from a aerodynamic perspective (Di

Sabatino et al., 2008), are shown in Fig. 3. Figure 4 depicts the probability density function (pdf ) of building and veg-

etation height for the S4 surface (qualitatively similar to those of the other

405

canopies, which are not shown). The building height distribution is characterized by three principal modes at z ≈ 1 m, z ≈ 3 m and z ≈ 6 m, corresponding 19

ACCEPTED MANUSCRIPT

CR IP T

Table 2: Relevant statistics of buildings and vegetation of the modeled urban canopy subsets in simulations S1, S2 and S3 and S4. S4 corresponds to the full surface displayed in Fig. 1. hb

and hv denote the height of an individual building or vegetation element (tree, bushes, etc.),

respectively. h ≡ mean (hb ) + σhb , where σhb is the standard deviation of buildings height. v b b λb p and λp are the plan area fraction of buildings and vegetation, respectively. λf,west , λf,south

denote the frontal area fraction of buildings with respect to west and south approaching wind

Unit

S1

S2

S3

S4

4.5

4.6

4.4

4.5

8.3

8.9

9.0

9.0∗

6.6

6.7

6.3

6.7

0.340

0.337

0.309

0.282



0.199

0.201

0.158

0.161



0.182

0.186

0.155

0.152

Buildings m

max (hb )

m

h

m

λbp



ED

λbf,south

M

mean (hb )

λbf,west

AN US

directions. LAI denotes the summer leaf area index of the canopy.

Vegetation

m

3.5

3.0

4.6

3.84

max (hv )

m

25.2

20.7

26.2

26.2

λvp



0.108

0.088

0.128

0.109

LAI



0.282

0.278

0.430

0.35

CE

PT

mean (hv )

* the unusually tall school building has been artificially reduced to 9 m height,

AC

to homogenize the surface and reduce undesired propagation of wake motions from such a structure in the inlet boundary (due to the prescribed periodic boundary conditions).

20

M

AN US

CR IP T

ACCEPTED MANUSCRIPT

ED

Figure 3: Top left: vertical structure of the building plan area fraction Ab p /AT . Top right: frontal area fraction dAf /dV as seen by west (black lines) and south (red lines) approaching wind. Bottom left: vertical structure of the vegetation plan area fraction Avp /AT . Bottom right: vertical structure of the horizontally averaged leaf area density h˜ ai(z) in the summer

PT

season. Different line-styles correspond to different subsets (see legend).

CE

to short surface elements (e.g., fences, landscaping), to one and to two storey houses, respectively. Conversely, the vegetation height distribution decreases rapidly from its surface value, indicative of a large number of small plants and shrubs.

AC

410

4.2. Vegetation effects on integral scaling parameters z0 and d in the ISL The aerodynamic roughness length (z0 ) and the zero-plane displacement (d)

are key parameters to properly simulate flow and turbulence over urban areas at city to regional scale in weather prediction, air quality dispersion, hydrology 21

CR IP T

ACCEPTED MANUSCRIPT

Figure 4: Binned p.d.f. of the roofs height (left) and of the vegetation height (right) for the

415

AN US

S4 canopy (see Fig. 1).

and climate models (Stensrud, 2007). Current morphometric methods employed to estimate z0 and d from urban form do not consider urban vegetation at all (Grimmond et al., 1998), despite its ubiquity. To address this issue, the dependence of z0 and d on urban vegetation is explored in this section.

420

M

Table 3 compares the LES modeled and corresponding tower measured d and z0 parameters. In both cases, values for each surface in different vegetation

ED

conditions are proposed (no vegetation, leaves-off and leaves-on). z0 is evaluated from the LES runs via logarithmic fit of the LES h˜ ui i(z) velocity in the range p −1 3 < z/h < 5, with fixed u∗ ≈ Lz kFi k = 1 (m s ), assuming κ = 0.4, and

425

PT

relating the displacement height d to the center of action of the drag force from buildings and vegetation (Jackson, 1981). The fitting interval was chosen based

CE

on the observed velocity profiles, which display a logarithmic behavior down to z − d ≈ h in most cases (see e.g. mean velocity profiles for cases over the S4 subset displayed in Fig. 5). Logarithmic velocity profiles well within the UCL

AC

(z ≤ max (hb , hv )) were previously reported in the wind tunnel measurements

430

of Cheng and Castro (2002) for flow over urban-like surfaces, in LES of flow over surface mounted cubes of variable height in Yang et al. (2016), and in LES over surface mounted rectangular prisms with high aspect ratio in Sadique et al. (2016). Unlike LES, direct tower measurements cannot be used to determine the

22

AN US

CR IP T

ACCEPTED MANUSCRIPT

Figure 5: Left: DA (intrinsic) streamwise LES velocity over the S4 canopy for west (black lines) and south (gray lines) wind approaching directions. Symbols: solid lines, no vegetation cases; dashed lines: winter vegetation cases; dot-dashed lines: summer vegetation cases. A reference logarithmic profile is shown in red. Center: % increase in z0 as a function of LAI/λb f. Right: % increase in d as a function of LAI/λb f.

435

M

flow characteristics in complete absence of trees. However, tower measurements allow one to study the effect of varying LAI on mean wind and momentum

ED

transfer by conditionally sampling and comparing leaves-off (between November to February), and leaves-on season (between May to September). Using a SV WV SV fixed d/h = 0.81 (average over the S4WV W , S4W , S4S , S4S LES runs), z0 is

440

PT

calculated based on tower measurements for each block satisfying the (neutral) stability and wind-direction requirements, inverting the logarithmic wind profile

CE

equation, following Grimmond et al. (1998):   ku z0 ≡ (z − d) exp , u∗

(10)

AC

with mean horizontal wind u and friction velocity u∗ directly measured by the eddy-covariance system (see Sect. 2). Note that since ln (zt ) ≈ ln (zt − d)

variations in d results in minor changes in the resulting z0 , hence the choice

445

of using a fixed d. This choice also ensures that the two approaches (LES and tower measurements) remain independent. Computing d as center of action of the surface drag forces predicts a signif23

ACCEPTED MANUSCRIPT

Table 3: Aerodynamic roughness lengths and displacement heights for the considered cases. For the LES runs the displacement height d is computed as center of action of the sum of drag forces from the building and vegetation canopies (hf˜ib i(z) and hf˜iv i(z) respectively), whereas z0 is obtain by linear fit of the LES h˜ ui i(z) velocity in the range 3 < z/h < 5. The

CR IP T

displacement height for tower measurements is computed as average from the LES runs S4WV W , WV , S4SV and roughness lengths are evaluated as z ≡ (z − d) exp (−uκ/u ), where S4SV t ∗ 0 W , S4S S

zt denotes the measurements height and u and u∗ are the time-averaged velocity and frictionvelocity at the tower sampling location. Tower data from all wind sectors SE to NW have

been equally weighted into 15◦ bins to remove directional effects. Near-Neutral conditions (−0.075 < (z − d)/L < 0.075) and wind > 2 ms−1 , RMSE(α) < 15◦ , and 150◦ < α < 300◦ . Summer is day of year (DOY) 121 to 274. Winter is DOY 305 to 59.

Wind

No veg. d/h

West

0.54

S1

South

0.61

S2

West

0.54

S2

South

0.62

S3

West

S3

South

S4

d/h

z0 /h

d/h

z0 /h

0.056

0.77

0.078

0.89

0.094

0.041

0.84

0.069

0.89

0.087

0.055

0.66

0.069

0.69

0.075

0.044

0.73

0.056

0.77

0.068

0.049

0.97

0.096

1.06

0.130

0.53

0.083

1.01

0.088

1.11

0.124

West

0.48

0.056

0.75

0.086

0.84

0.106

South

0.52

0.048

0.79

0.077

0.88

0.102

150 < α < 300





0.81

0.119

0.81

0.142

PT

ED

0.50

S4 Tower

Summer veg.

z0 /h

M

S1

Winter veg.

AN US

Surface

icant increase in both d and z0 as a function of LAI/λbf , as displayed in Fig. 5

CE

and tabulated in Tab. 3, highlighting the important role of vegetation elements

450

in affecting the aerodynamics of the flow over the considered realistic urban

AC

settings.

LES runs predict variations of d from the leaves-off and leaves-on cases be-

tween +5% and +15%, due to significant contributions to the overall drag force that taller trees are exerting on the flow, despite their modest plan area fraction

455

Avp /AT (see Fig. 3). Corresponding variations in z0 are between +10% and +40%. It is also apparent that in complete absence of trees z0 would dramat-

24

ACCEPTED MANUSCRIPT

0.30

2.0

95% 75% Mean

1.5

0.20 0.15 0.10 615

660

1035

967

504

537

226

318

452

514 844

549

0.05 0.00

J

F

M

A

M

J

J

A

S

O

N

Month of the year

D

CR IP T

1.0

50% 25% 5%

z0 (m)

z0 / hb

0.25

0.5 0.0

Leaves-on Transition

Leaves-off

AN US

Figure 6: Aerodynamic roughness length (z0 ) measured on the tower at a height of zt = 24.4

m and sorted by month of the year (see Sect. 2.3 for additional details). The numbers under each category refer to the number of half hourly measurements represented in the statistics.

ically decrease, down to more than half of their reference summer values for the 2 / LAI/λbf / 3 cases. In complete absence of trees d is also significantly

M

reduced.

Specifically, z0 and d both display a sub-linear dependence on LAI/λbf , high-

460

ED

lighting how increasing LAI results in additional sheltering induced by vegetation on buildings and on other vegetation elements. Sheltering between roughness elements is known to be an important dynamical mechanism, responsible

465

PT

to reduce the overall drag that a surface exerts on the flow in the high λbf or LAI range (Macdonald et al., 1998; Raupach, 1992; Yang et al., 2016; Sadique et al., 2016). Conversely, the strong non-linear dependence of d on LAI/λbf can

d∼

LAIHv + λbf Hb , LAI + λbf

(11)

AC

CE

be justified by considering that

where Hv and Hb are reference heights for the vegetation and buildings elements.

Equation 11 is a simplified version of the formula for the evaluation of the center

470

of action (barycenter) of buildings and vegetation forces (Jackson, 1981). Note that with opportunely chosen Hv and Hb , Eq. 11 well approximates the d dependence on LAI/λbf (not shown).

25

ACCEPTED MANUSCRIPT

Figure 6 shows z0 calculated from the tower-data following Eq. 10 for each month of the year. As the leaves emerge in March and April of each year, z0 475

increases to higher summer values, and then with senescence in October, z0

CR IP T

drops. Overall, the conditionally sampling leaf-off and leaf-on periods on the tower we can identify a statistically significant (T-means test, p < 0.01) in-

crease by 19% from a mean normalized z0 /h = 0.119 (0.80 m) during leaves-off

to z0 /h = 0.142 (0.95 m) in the leaves-on season. LES determined z0 for the 480

full subset are 32% and 27% lower than the measured values on the tower, with

winter and summer vegetation, respectively. For subsets S1 and S3 which are

AN US

most representative for the tower footprint, the LES are 36% and 11% lower than tower measurements in summer, and 38% and 23% lower in winter, re-

spectively. We relate such trends to the lack of information in the LiDAR data 485

(e.g., fences, chimneys, power lines) when compared to reality – resulting in an aerodynamically smoother urban canopy – and to the gentle orography that was removed in simulations to enable the use of periodic boundary conditions in the

M

horizontal directions (x, y). Nevertheless, LES predictions fall in the interquartile range in all but two months (see Figure 6), validating both approaches and highlighting once more how vegetation effects on aerodynamic parameters can be significant.

ED

490

PT

4.3. Vegetation effects on integral turbulence statistics in the ISL Table 4 shows a comparison of dimensionless integral turbulence statistics evaluated at z = 24.4 m height between tower measurements and intrinsically DA LES over the S4 canopy. Note that DA LES statistics include dispersive

CE 495

contributions due to the chosen order of averaging operations (spatial averaging

AC

is performed first). Normalized streamwise and vertical velocity fluctuations (σu /u∗ , σw /u∗ ) and the turbulent momentum exchange efficiency (ruw , correlation coefficient between vertical wind and horizontal wind) indicate that tower

500

measurements are in the ISL irrespective of leaf situation, as they follow closely the neutral limits of the surface layer as predicted by MOS. No statistically significantly seasonal changes (T-means test, p < 0.01) is observed for such 26

ACCEPTED MANUSCRIPT

Table 4: Comparison of dimensionless integral turbulence statistics at z = 24.4 m height between tower measurements and DA LES. LES results are representative of flow over the S4 canopy, averaged over the west and south approaching wind directions and separated into “no vegetation”, “leaf-off” (winter) and “leaf-on” (summer) conditions. Tower-measured

refers to the surface layer neutral limits (Panofsky, 1984).

Tower

SL

Winter

Summer

No trees

Winter

Summer

σu /u∗

2.310

2.323

2.35

2.30

2.32

2.2

σw /u∗

1.206

1.209

1.16

1.17

1.175

1.25

ruw

-0.366

-0.365

-0.364

-0.372

-0.366

Skw

0.142

0.204

0.177

0.150

0.118

TKE / MKE

0.155

0.174

0.071

0.089

0.104

AN US

Parameter

LES

CR IP T

quantities are separated into “leaf-off” and “leaf-on” conditions as described in Sect. 2.3. SL

-0.36

M

quantities in measurements, and LES further support such a finding. Measurements predict a positive Skw , as is typical of surface layers (SLs), 505

which is observed in 90% and 87% of all cases in summer and winter, respec-

ED

tively, indicating the distribution is skewed to more positive values (updrafts less frequent but more intense). The skewness increases statistically significant

PT

(T-means test, p < 0.01) by 44% from a mean of 0.142 during leaf-off to a more skewed value of 0.204 in the leaf-on season (Tab.4). In contrast with mea510

surements, LES of flow over the S4 surface, averaged over the west and south

CE

approaching wind directions, predict a 15% and 33% decrease in Skw at z = 24.4 m from the no-vegetation to leaf-on and from the leaf-off to leaf-on cases, respectively. It should be noted that simulations over the S4 surface are characterized

AC

by a large max (hv , hb )/Lz ratio, which ensures that turbulence statistics (in-

515

cluding Skw ) up to z ≈ 5h are poorly dependent on the vertical extent of the computational domain (the sensitivity of flow statistics on Lz was tested using

half the current spatial resolution and is not shown). The most plausible reason for such a mismatch is hence the inability of LES to account for thermally

27

ACCEPTED MANUSCRIPT

95%

0.30

50% 25% 5%

0.25 0.20

Leaves-on

0.15 1035

615

0.10

CR IP T

TKE / MKE

75% Mean

J

967

537

504

318

452

226

514 844

660

F

549

Transition

Leaves-off

M

A

M

J

J

A

Month of the year

S

O

N

D

AN US

Figure 7: TKE to MKE ratio measured on the tower at a height of zt = 24.4 m and sorted

by month of the year (see Sect. 2.3 for additional details). The numbers under each category refer to the number of half hourly measurements represented in the statistics.

induced large-scale outer-layer structures, which are likely to influence Skw via 520

their interaction with the underlying canopy. The vertical structure of Skw from

quantity is provided.

M

LES is described and commented in Sect. 4.4, where further insight on such a

ED

Figure 7 shows the ratio of turbulent to mean kinetic energy measured at 24.4 m. The ratio TKE / MKE increases statistically significant (T-means test, 525

p < 0.01) by about 12% from a mean of 0.155 during leaf-off to a value of 0.174

PT

in the leaf-on season (Tab.4). Again, clearly as leaves emerge in March and April, the ratio TKE/MKE increases and in fall drops again. This is explained

CE

by the increased roughness of the leaved canopy which converts more MKE and large-scale TKE to small-scale “wake” TKE. TKE/MKE increases in LES

530

from no-vegetation to leaf-on (summer) cases, and displays a 17% gain between

AC

the leaf-off and leaf-on season, in qualitative agreement with results from tower measurements. However, not surprisingly, LES predicts lower TKE/MKE ratios when compared to measurements. This can be again justified by the presence of outer layer flow structures that the LES does not include, which are an

535

important contribution to the overall TKE and are not affected by the local

28

ACCEPTED MANUSCRIPT

surface as shown in Roth and Oke (1995) for this site. A significantly larger computational domain would be required to account for such structures, which

CR IP T

is beyond current computational capabilities. 4.4. Vegetation effects on momentum and energy transfer rates in the RSL The DA momentum equation reads

540

(12)

AN US

1 d −Fx = [−ϕ(h˜ u0 w ˜ 0 i + hτxz i)] |{z} ϕ dz | {z } | {z } Forcing SGS Turbulent Z b 1 ˜ ˜ + (f + τ xj nj ) dΓ + hfxv i , |{z} Vf (z) Γb (z) x | {z } Vegetation Buildings

where (·) denotes temporal averaging, ni is the normal-to-interface (Γb ) vector pointing into the fluid domain, ϕ(z) ≡ (1−λbp ) and Vf ≡ ϕ(Lx ×Ly ×∆z ) are the area fraction of fluid and the volume of fluid at a given height (z), h˜ u0 w ˜0 i

M

545

and hτxz i are the turbulent (including dispersive) and SGS kinematic stresses, b R 1 ˜ Vf (z) Γb (z) (f x +τ xj nj ) dΓ is the kinematic drag force that buildings are exerting

on the flow (sum of a normal and a deviatoric component), and hf˜xv i ≤ 0 is the

ED

kinematic drag force from trees.

Fig. 8 features selected components of the kinematic stress balance (Eq. 12). Profiles are qualitatively representative of the other LES runs. Below max hb , the largest contribution to the total canopy drag force is from

PT

550

buildings, and results in an approximately linear vertical momentum flux con-

CE

tribution, from the reference (negative) surface value to zero at z = max (hb ). Conversely, a significant portion of the vegetation drag force is exerted above

AC

555

max (hb ), resulting in a monotonic decrease of the corresponding momentum R hmax flux ( z v hf˜xv i dζ) at such heights, with direct impact on the turbulent flux

term h˜ u0 w ˜ 0 i, which experiences an equal decrease in magnitude (note that the to-

tal normalized momentum flux is the same among simulations). Throughout the runs, the downward flux of momentum from turbulent motions is decreased from the “no-vegetation” to the “leaves-on” cases within the UCL (z < max (hb , hv )).

29

CR IP T

ACCEPTED MANUSCRIPT

AN US

Figure 8: Vertical structure of the turbulent momentum flux (h˜ u0 w ˜ 0 i, black lines), total drag b R hmax 1 R ˜ b from buildings ( z (f + τ xj nj ) dΓdζ, gray lines), and from vegetation eleVf (ζ) Γb (ζ) x R max ments ( zhv hf˜xv i dζ, red lines). Profiles are representative of simulations over the S4 surface, west wind approaching angle. Different line styles denote different runs: S4NV W , solid lines; SV S4WV W , dashed lines; S4W , dot-dashed lines. SGS contributions are negligible throughout the

CE

PT

ED

M

boundary-layer and hence not shown. A smaller portion of the total domain height is shown.

uih˜ u0 w ˜ 0 i) 1 d(ϕh˜ , ϕ dz

AC

Figure 9: Turbulent transport of mean kinetic energy (Tt ≡

black lines),

production of MKE by the imposed pressure gradient (PΠ ≡ Fx h˜ ui, grey lines), and sink of

MKE from the work of vegetation drag (Pv ≡ hf˜xv ih˜ ui , red lines) for simulations over the S4 surface, west wind approaching angle. Different line styles denote different simulations: S4NV W ,

SV ∗ solid lines; S4WV W , dashed lines; S4W , dot-dashed lines. (·) is used to denote normalization

of each term by (u3∗ /h). A smaller portion of the total domain height is shown.

30

ACCEPTED MANUSCRIPT

To better quantify the impact of vegetation on the mean flow, Fig. 9 features

560

the vertical structure of selected terms of the mean kinetic energy (MKE) budget equation. It is apparent how vegetation has a twofold effect on MKE in the

CR IP T

considered urban canopies: on one hand, it is a direct sink of MKE via its work against the mean flow (Pv ), proportional to the LAI characterizing a given 565

canopy. On the other hand, its impact on h˜ u0 w ˜ 0 i and h˜ ui (via the work that

f˜iv performs against the mean flow) results in a significant decrease of MKE downward turbulent transport (Tt ) below max (hb ). At such heights where

people live and buildings consume energy, Tt is the major source of MKE (Tt  LAI results in a substantial decrease of wind speed magnitudes.

PT

ED

M

570

AN US

Fx h˜ ui > 0), and is thus no surprise that reductions in such a term with increasing

Figure 10: Vertical structure of the horizontal (left) and vertical (right) velocity skewness

CE

for LES cases over the S4 surface west (black lines) and south (red lines) approaching wind angles. Symbols: solid lines, no vegetation cases; dashed lines, leaves-off cases; dot-dashed

AC

lines, leaves-on cases.

Based on Fig. 10 it is clear how the upper RSL is dominated by ejections,

while the lower RSL and urban canopy layer show a dominance of sweeps in the turbulent momentum exchange, which is in agreement with several field studies within and/or above urban canopies (Rotach, 1993; Oikawa and Meng,

575

1995; Christen et al., 2007). As vegetation is added Sku and Skw increase in 31

CR IP T

ACCEPTED MANUSCRIPT

Figure 11: Difference between the two-dimensional histogram of h˜ u0 w ˜ 0 i at z = hb for simu-

AN US

WV NV SV lations S4NV ˜0 /u∗ and w ˜ 0 /u∗ sampled W , S4W (left) and S4W , S4W (right). Point clouds of u

in the plane z = hb have been uniformly binned in a two dimensional histogram for each simulation and the difference between the two dimensional histograms for the cases without and with vegetation are displayed. Red, white and black dots denote a decrease, no changes, and increase in the number of observations that fall within a given bin.

M

magnitude within the UCL, highlighting a more intermittent turbulence field, which favors u ˜0 > 0 and w ˜ 0 < 0 fluctuations (i.e., sweep motions). The height of

ED

the crossover for Sku and Skw also increases with the addition of vegetation and with increasing LAI, highlighting how the presence of irregular sparse roughness 580

with irregular heights results in a thicker RSL. Note that even in the absence of

PT

vegetation, the height of the crossover at z/h = 1.38 is higher than the crossover of z/h = 1.25 observed in simulations over cubical obstacles of uniform height

CE

(Coceal et al., 2007a) but comparable to the crossover in neutral conditions at z/h = 1.5 over a denser urban canopy without significant trees (Christen et al.,

585

2007).

AC

Figure 11 provides a qualitative description how turbulent fluctuations con-

tributing to h˜ u0 w ˜ 0 i at z = hb are modified by vegetation. The specific cases of flow over the S4 surfaces with wind approaching from west are considered WV SV (S4NV W , S4W and S4W ). As is apparent, including vegetation results in a pro-

590

gressive damping of streamwise fluctuations u ˜0 (red areas) contributing to both quadrant 2 and quadrant 4 events, reducing the variance of the u ˜0 probability 32

ACCEPTED MANUSCRIPT

density function. This behavior is consistent among the considered cases. Note also how the sparse vegetation suppresses ejections more effectively than sweeps, in line with the observed increase in the Sku and Skw magnitude at hb (see Fig. 10). In an absolute frame, sparse vegetation thus tends to make the turbulent

CR IP T

595

field more isotropic in the lower RSL. Accordingly, in addition to a direct fluid dynamical implication, including vegetation has potential benefits from an urban comfort perspective, as gusts affecting pedestrians, and dynamical pressure

600

5. Summary and Conclusions

AN US

variations that cause building energy losses are minimized (“sheltering effects”).

This research explored the role of trees over a realistic urban canopy using a combination of detailed LES, eight years of continuous measurements, and airborne LiDAR data to map and describe the detailed urban form and leaf area density. Unlike previous computational fluid dynamics simulations that investigated the effect of urban trees on pollutant dispersion in idealized streets,

M

605

the current study focused on how trees and changing leaf area density affect the neighborhood-averaged flow and turbulence fields. A number of surfaces were

ED

considered, representative of a typical suburban area of Vancouver, Canada, where a small fraction of trees are taller than buildings. LES and tower measurements showed that aerodynamic roughness lengths

610

PT

(z0 ) are sensitive to variations in the 0 ≤ LAI/λbf < 3 parameter for each of the considered canopies. Specifically, z0 increments were characterized by a

CE

sub-liner dependence on LAI/λbf for each of the considered canopies, indicative of additional sheltering occurring between vegetation elements as well as be-

615

tween vegetation and buildings as LAI increases. z0 from LES fell within the

AC

interquartile range of corresponding values from tower measurements, but mean z0 values were underpredicted by approximately 30% in the LES runs over the S4 surface (the most representative of what the tower sees). Such a behavior is mostly due to lack of information in the LiDAR measurements when compared

620

to real settings (power lines, chimneys, fences) and to the gentle orography that

33

ACCEPTED MANUSCRIPT

was not considered in simulations. Specifically, the aerodynamic roughness determined from tower measurements was seasonally changing with a statistically significant 19% increase from leaf-off season (z0 /h = 0.119) to the leaf-on season

625

CR IP T

(z0 /h = 0.142). Similarly, LES results over the S4 subset averaged over south and west wind approaching directions found a seasonal change in z0 with an

increase by 27% from the winter (LAI/λbf = 0.74, z0 /h = 0.081) to the summer

season (LAI/λbf = 2.24, z0 /h = 0.104). LES also showed that in the absence

of vegetation z0 would experience a significant decrease, down to over 50% its reference summer value in most cases.

The zero-plane displacement d, computed as center of action of the total

AN US

630

surface drag from LES, was found to also increase as a function of the LAI/λbf parameter, mostly due to the disproportionate amount of drag that the (few) relatively taller trees exert on the flow.

Integral statistics σu /u∗ , σw /u∗ and ruw at the measurement height of the 635

tower (z/h = 3.75, where h ≡ mean(hb ) + σhb ) were not subject to any seasonal

M

change (evidence provided by both measurements and LES). They followed predicted surface layer values in summer and winter, hence at such a height and

ED

above, we can expect a well-developed ISL in which the neutral limits of MOS are fulfilled. Statistically significant increases between leaf-off and leaf-on cases 640

were found at z/h = 3.75 in the ratio of TKE to MKE (+12% in measurements,

PT

+17% in LES over the S4 canopy averaged between west and south wind approaching direction), but the magnitude from LES were over 50% smaller than

CE

the corresponding measured quantity. Furthermore, LES and tower measurements predicted a positive vertical velocity skewness (Skw > 0) at the tower

645

sampling location, as is typical of atmospheric surface layers where ejections of

AC

low momentum fluid are the main contribution to the total momentum flux. However, tower data were characterized by an increase in Skw from winter to

summer, whereas LES data predicted a decrease in such a quantity in that same period. The most plausible reason for such a mismatch in both TKE/MKE and

650

in Skw is the inability of LES to account for thermally induced large-scale outer-layer structures, which are likely to influence such flow statistics via their 34

ACCEPTED MANUSCRIPT

interaction with the underlying canopy. The analysis of the vertical transfer of momentum in the LES showed how trees exert a significant portion of the total drag in both the winter and the summer season, thus reducing pressure loads on buildings. In most cases, the

CR IP T

655

total drag from vegetation elements during the summer season was even larger than that of buildings. In addition, as vegetation is added Sku and Skw increase in magnitude within the UCL, highlighting a more intermittent turbulence field

favoring sweep motions, and resulting in higher locations in sign crossover for 660

both statistics. From the quadrant analysis that was performed at z = hb , it is

AN US

also apparent that including vegetation results in a progressive damping in the magnitude of streamwise fluctuations u0 , thus reducing extreme wind events in the UCL, with potential benefits from an urban comfort perspective.

Lastly, the analysis of selected mean kinetic energy budget terms in the LES 665

showed that urban trees influence the flow in the RSL twofold: on one hand, trees are a direct sink of MKE via the work of their drag force against the

M

mean flow; on the other hand, they reduce downward turbulent transport of high momentum fluid from higher atmospheric layers. Given that most of the

670

ED

MKE in the lower RSL is generated by downward transport from large organized turbulent motions, it is apparent how the specific shape, clumping and density of urban vegetation controls flow at the height where people live in a city, hence

PT

suggesting the need for careful landscape architecture design strategies. With more than 50% of the global population now living in cities (United

CE

Nations, 2014), the atmosphere which the majority of us experiences is not a 675

classical surface layer, but the roughness sublayer embedded within and above urban canopies. In many cities of the world, trees are important elements of the

AC

urban form, and also a design tool to mitigate adverse effects on pedestrians, provide shelter and shade, mitigate heat stress, and conserve energy in buildings. However, trees also affect the dispersion of air pollutants emitted at ground level

680

negatively (e.g. Salmond et al., 2013). Our current ability to incorporate the effect of trees in ULSMs is limited, but the current study shows that the role of trees and their seasonally changing leaf area density (a) in cities has a crucial 35

ACCEPTED MANUSCRIPT

effect on both, turbulence and flow in the ISL above vegetated cities and flow an turbulence in the urban roughness sublayer, including at street level and 685

these effects should be considered in next-generation ULSMs. This may help in

CR IP T

future to make improved predictions that benefit the management of air quality, heat stress and generally assist the planning of energy efficient and intelligently designed cities.

Acknowledgements

This research was funded by the Swiss National Science Foundation (SNSF-

690

AN US

200021-134892), the Competence Center for Environmental Sustainability (CCESSwissEx) of the ETH domain, the NSERC Discovery Grant program (Parlange, Christen). Selected instrumentation was supported by the Canada Foundation for Innovation (Grant 33600). Numerical simulations were supported by a grant 695

from the Swiss National Supercomputing Centre (CSCS) under project ID s633.

M

We are grateful to the peer reviewers whose comments helped to improve the overall quality of the manuscript.

ED

Appendix A. Grid resolution requirements To assess the resolution requirements for the LES runs the subset S1 is selected with west wind approaching direction and summer vegetation conditions

PT

700

(similar to case S1SV W ). Four different resolutions are considered, as indicated

CE

in Fig. 12, and quantities such as the mean turbulent kinetic energy (TKE ≡ (1/2)h˜ u0i u ˜0i i) and the mean rate of TKE shear production (Ps ≡ h˜ u0 w ˜ 0 i(dh˜ ui/dz)) are inter-compared. The second quantity is often used as indicator of the quality of a numerical solution, as it encodes information on both mean flow and tur-

AC

705

bulence. The grid stencil in the vertical direction (∆z = 0.5 m) is not modified across the considered runs and only the horizontal grid stencils are varied, from a maximum of ∆x ≈ ∆y = 3.14 (m) to a minimum of ∆x ≈ ∆y = 1.04 (m).

Varying only the horizontal grid stencil is justified considering that the vertical 710

resolution is much higher than the horizontal across the various cases. Note 36

Figure 12: (1/2)h˜ u0 u ˜0 i)

AN US

CR IP T

ACCEPTED MANUSCRIPT

Comparison of time- and space-averaged turbulent kinetic energy (TKE ≡

(left plot) and TKE production rate (Ps ≡ h˜ u0 w ˜ 0 i(dh˜ ui/dz)) (right plot) for

case S1SV W (see Table 1 for details) at different spatial resolutions.

that the coarser among the considered horizontal grid stencils is approximately

M

60% larger than that used in the runs over the S4 surface (see Table 1), the second coarser resolution case (Nx × Ny × Nz = 192 × 96 × 168) corresponds to approximately 10% stencil increased when compared to that used in the S4 runs, and the second finer resolution (Nx × Ny × Nz = 320 × 160 × 168) corresponds

ED

715

to that used in all the other runs (flow over the S1, S2, and S3 cases).

PT

TKE and Ps profiles convergenge as the grid stencil is refined. Assuming the finer resolution as reference, largest departures in the TKE magnitude occur in the near surface region for the coarsest run, and at z/h ≈ 2 for the other cases. Large errors in max (Ps ) are also localized in the near surface region

CE

720

for the coarsest case, whereas the remaining runs are well converged as such a location, and greatest departures occur near max (Ps ). The latter height is

AC

where thin shear layers separate from buildings. Such flow features are regions of strong TKE production where small and highly energetic motions set tough

725

requirements in terms of grid resolution. The grid convergence study indicates that the chosen resolution for the runs over the S1, S2, and S3 surfaces is fine enough to correctly capture the TKE and max (Ps ) profiles within the RSL. The

37

ACCEPTED MANUSCRIPT

resolution chosen for the S4 runs is slightly larger, and the current convergence tests shows that local variations in the TKE and in Ps with respect to the exact 730

value are bounded to a maximum of 16% and 8% respectively. Such modest

and validate the quality of the proposed LES results.

References

CR IP T

percentage errors are well within the level of accuracy required for the study,

Albertson, J.D., Parlange, M.B., 1999a. Natural integration of scalar fluxes Adv Water Resour 23, 239–252.

S0309-1708(99)00011-1.

doi:10.1016/

AN US

from complex terrain.

735

Albertson, J.D., Parlange, M.B., 1999b. Surface length scales and shear stress: implications for land-atmosphere interaction over complex terrain. Water Resour Res 35, 2121–2132. doi:10.1029/1999WR900094.

Anderson, W., 2016. Amplitude modulation of streamwise velocity fluctuations

M

740

in the roughness sublayer: evidence from large-eddy simulations. J Fluid

ED

Mech 789, 567–588. doi:10.1017/jfm.2015.744. Anderson, W., Li, Q., Bou-Zeid, E., 2015. Numerical simulation of flow over urban-like topographies and evaluation of turbulence temporal attributes. J Turbul 16, 809–831. doi:10.1080/14685248.2015.1031241.

PT

745

Anderson, W., Meneveau, C., 2010.

A large-eddy simulation model for

CE

boundary-layer flow over surfaces with horizontally resolved but vertically unresolved roughness elements.

Boundary-Layer Meteorol 137, 397–415.

AC

doi:10.1007/s10546-010-9537-5.

750

Anderson, W., Meneveau, C., 2011. Dynamic roughness model for large-eddy simulation of turbulent flow over multiscale, fractal-like rough surfaces. J Fluid Mech 679, 288–314. doi:10.1017/jfm.2011.137.

Anderson, W., Passalacqua, P., Port´e-Agel, F., Meneveau, C., 2012. Largeeddy simulation of atmospheric boundary-layer flow over fluvial-like land38

ACCEPTED MANUSCRIPT

scapes using a dynamic roughness model. Boundary-Layer Meteorol 144,

755

263–286. URL: http://link.springer.com/10.1007/s10546-012-9722-9, doi:10.1007/s10546-012-9722-9. Turbulence in sparse, organized vegetative

CR IP T

Bailey, B.N., Stoll, R., 2013.

canopies: A large-eddy simulation study. Boundary-Layer Meteorol 147, 369– 400. doi:10.1007/s10546-012-9796-4.

760

Balcz´ o, M., Gromke, C., Ruck, B., 2009. Numerical modeling of flow and pollutant dispersion in street canyons with tree planting. Meteorologische

AN US

Zeitschrift 18, 197–206.

Barlow, J., 2014. Progress in observing and modelling the urban boundary layer. Urban Clim (In Press) 44. URL: http://linkinghub.elsevier.com/

765

retrieve/pii/S2212095514000558, doi:10.1016/j.uclim.2014.03.011. Belcher, S.E., 2005. Mixing and transport in urban areas. Philosophical Trans-

M

actions of the Royal Society A 363, 2947–2968.

Bou-Zeid, E., Meneveau, C., Parlange., M.B., 2004. Large-eddy simulation of neutral atmospheric boundary layer flow over heterogeneous surfaces: blend-

ED

770

ing height and effective surface roughness. Water Resour Res 40, W02505. doi:10.1029/2003WR002475.

PT

Bou-Zeid, E., Meneveau, C., Parlange, M.B., 2005. A scale-dependent Lagrangian dynamic model for large eddy simulation of complex turbulent flows. Phys Fluids 17, 025105. doi:10.1063/1.1839152.

CE

775

Bou-Zeid, E., Overney, J., Rogers, B.D., Parlange, M.B., 2009. The effects

AC

of building representation and clustering in large-eddy simulations of flows

780

in urban canopies. Boundary-Layer Meteorol 132, 415–436. doi:10.1007/ s10546-009-9410-6.

Bowler, D.E., Buyung-Ali, L., Knight, T.M., Pullin, A.S., 2010. Urban greening to cool towns and cities: A systematic review of the empirical evidence. Landscape and Urban Planning 97, 147–155. 39

ACCEPTED MANUSCRIPT

Britter, R.E., Hanna, S.R., 2003. Flow and dispersion in urban areas. Annu Rev Fluid Mech 35, 469–496. doi:10.1146/annurev.fluid.35.101101.161147. 785

Brunet, Y., Finnigan, J.J., Raupach, M.R., 1994. A wind tunnel study of air flow

CR IP T

in waving wheat: Single-point velocity statistics. Boundary-Layer Meteorol 70, 95–132. doi:10.1007/BF00712525.

Brutsaert, W., 1982. Evaporation into the atmosphere. Springer Netherlands, Dordrecht. doi:10.1007/978-94-017-1497-6. 790

Brutsaert, W., Parlange., M.B., 1992. The unstable surface layer above forest:

1029/92WR01860.

AN US

regional evaporation and heat flux. Water Resour Res 28, 3129–3134. doi:10.

Brutsaert, W., Parlange, M.B., Gash, J.H.C., 1989. Neutral humidity profiles in the boundary layer and regional evaporation from sparse pine forest, in: Ann Geophys, pp. 623–630.

795

M

Calaf, M., Meneveau, C., Meyers, J., 2010.

Large eddy simulation study

of fully developed wind-turbine array boundary layers.

URL: http://scitation.aip.org/content/aip/journal/pof2/

ED

015110.

Phys Fluids 22,

22/1/10.1063/1.3291077, doi:10.1063/1.3291077. 800

Calaf, M., Parlange, M.B., Meneveau, C., 2011. Large eddy simulation study

PT

of scalar transport in fully developed wind-turbine array boundary layers. Phys Fluids 23, 126603. URL: http://scitation.aip.org/content/aip/

CE

journal/pof2/23/12/10.1063/1.3663376, doi:10.1063/1.3663376. Cava, D., Katul, G.G., 2008. Spectral short-circuiting and wake production within the canopy trunk space of an alpine hardwood forest. Boundary-Layer

AC

805

Meteorol 126, 415–431. doi:10.1007/s10546-007-9246-x.

Chen, F., Bornstein, R., Grimmond, S., Li, J., Liang, X., Martilli, A., Miao, S., Voogt, J., Wang, Y.H., 2012. Research priorities in observing and modeling urban weather and climate. Bull Am Meteorol Soc 93, 1725–1728. doi:10.

810

1175/BAMS-D-11-00217.1. 40

ACCEPTED MANUSCRIPT

Cheng, H., Castro, I.P., 2002.

Near wall flow over urban-like roughness.

Boundary-Layer Meteorol 104, 229–259. doi:10.1023/A:1016060103448. Cheng, W.C., Port´e-Agel, F., 2013. Evaluation of subgrid-scale models in large-

CR IP T

eddy simulation of flow past a two-dimensional block. Int J Heat Fluid Flow 44, 301–311. doi:10.1016/j.ijheatfluidflow.2013.06.007.

815

Cheng, W.C., Port´e-Agel, F., 2015. Adjustment of turbulent boundary-layer

flow to idealized urban surfaces: A large-eddy simulation study. BoundaryLayer Meteorol 155, 249–270. doi:10.1007/s10546-015-0004-1.

AN US

Chester, S., Meneveau, C., 2007. Renormalized numerical simulation of flow

over planar and non-planar fractal trees. Environ Fluid Mech 7, 289–301.

820

doi:10.1007/s10652-007-9026-7.

Chester, S., Meneveau, C., Parlange, M.B., 2007. Modeling turbulent flow over fractal trees with renormalized numerical simulation. J Comput Phys 225,

825

M

427–448. doi:10.1016/j.jcp.2006.12.009.

Chorin, A.J., 1968. Numerical solution of the Navier-Stokes equations. Math

ED

Comput 22, 745–762. doi:10.2307/2004575. Christen, A., van Gorsel, E., Vogt, R., 2007. Coherent structures in urban roughness sublayer turbulence. International Journal of Climatology 27, 1955–

830

PT

1968.

Christen, A., Rotach, M.W., Vogt, R., 2009. The Budget of Turbulent Kinetic

CE

Energy in the Urban Roughness Sublayer. Boundary-Layer Meteorology 131, 193–222.

AC

City of Vancouver,

835

egy,

2014.

City of Vancouver.

City of Vancouver Urban Forest StratURL: http://vancouver.ca/files/cov/

Urban-Forest-Strategy-Draft.pdf.

Coceal, O., Dobre, A., Thomas, T.G., Belcher, S.E., 2007a. Structure of turbulent flow over regular arrays of cubical roughness. Journal of Fluid Mechanics 589, 375–409. 41

ACCEPTED MANUSCRIPT

Coceal, O., Thomas, T.G., Belcher, S.E., 2007b. Spatial Variability of Flow Statistics within Regular Building Arrays. Boundary-Layer Meteorology 125,

840

537–552.

CR IP T

Crawford, B., Christen, A., 2015a. Spatial source attribution of measured urban

eddy covariance CO2 fluxes. Theor Appl Climatol , 733–755doi:10.1007/ s00704-014-1124-0. 845

Crawford, B.R., Christen, A., 2015b. Spatial source attribution of measured urban eddy covariance CO2 fluxes. Theoretical and Applied Climatology .

AN US

Crawford, B.R., Christen, A., Ketler, R., 2012. Processing and quality control

procedures of turbulent flux measurements during the Vancouver EPiCC experiment – EPiCC Technical Report No. 1. Technical Report. The University of British Columbia. URL: https://dx.doi.org/10.14288/1.0103593.

850

Di Sabatino, S., Solazzo, E., Paradisi, P., Britter, R., 2008. A simple model

M

for spatially-averaged wind profiles within and above an urban canopy. Boundary-Layer Meteorol 127, 131–151. doi:10.1007/s10546-007-9250-1.

ED

Dwyer, M., Patton, N., Shaw, R.H., 1997. Turbulent kinetic energy budgets from a large-eddy simulation of airflow above and within a forest canopy.

855

PT

Boundary-Layer Meteorology 84, 23–43. Fang, J., Diebold, M., Higgins, C., Parlange, M.B., 2011. Towards oscillationfree implementation of the immersed boundary method with spectral-like

CE

methods. J Comput Phys 230, 8179–8191. doi:10.1016/j.jcp.2011.07.017.

Feigenwinter, C., Vogt, R., Parlow, E., 1999. Vertical structure of selected

AC

860

turbulence characteristics above an urban canopy. Theoretical and Applied Climatology 62, 51–63.

Finnigan, J.J., Shaw, R.H., Patton, E.G., 2009a. Turbulence structure above a vegetation canopy. Journal of Fluid Mechanics 637, 387.

42

ACCEPTED MANUSCRIPT

865

Finnigan, J.J., Shaw, R.H., Patton, E.G., 2009b.

Turbulence structure

above a vegetation canopy. J Fluid Mech 637, 387. URL: http://www. journals.cambridge.org/abstract{_}S0022112009990589, doi:10.1017/

CR IP T

S0022112009990589. Germano, M., Piomelli, U., Moin, P., Cabot, W., 1991. A dynamic subgrid-scale eddy viscosity model. Phys Fluids 3, 1760–1765. doi:10.1063/1.857955.

870

Giometto, M.G., Christen, A., Meneveau, C., Fang, J., Krafczyk, M., Par-

lange, M.B., 2016. Spatial characteristics of roughness sublayer mean flow

AN US

and turbulence over a realistic urban surface. Boundary-Layer Meteorol 160, 425–452. 875

Grimmond, C.S.B., Blackett, M., Best, M., Baik, J., Belcher, S.E., Beringer, J., Bohnenstengel, S.I., Calmet, I., Chen, F., Coutts, A., Dandou, A., Fortuniak, K., Gouvea, M.L., Hamdi, R., Hendry, M., Kanda, M., Kawai, T.,

M

Kawamoto, Y., Kondo, H., Krayenhoff, E.S., Lee, S.H., Loridan, T., Martilli, A., Masson, V., Miao, S., Oleson, K., Ooka, R., Pigeon, G., Porson, A., Ryu, Y.H., Salamanca, F., Steeneveld, G.J., Tombrou, M., Voogt, J.A., Young,

880

ED

D.T., Zhang, N., 2010. Initial results from Phase 2 of the international urban energy balance model comparison. International Journal of Climatology 31,

PT

244–272.

Grimmond, C.S.B., King, T., Roth, M., Oke, T.R., 1998. Aerodynamic roughness of urban areas derived from wind observations. Boundary-Layer Meteo-

885

CE

rology 89, 1–24.

AC

Grimmond, C.S.B., Oke, T.R., 2002. Turbulent heat fluxes in urban areas:

890

observations and a local-scale urban meteorological parametrization scheme (LUMPS). Journal of Applied Meteorology 41, 792–810.

Gromke, C., Blocken, B., 2015. Influence of avenue-trees on air quality at the urban neighborhood scale. Part II: Traffic pollutant concentrations at pedestrian level. Environmental Pollution 196, 176–184.

43

ACCEPTED MANUSCRIPT

Gromke, C., Ruck, B., 2008. On the Impact of Trees on Dispersion Processes of Traffic Emissions in Street Canyons. Boundary-Layer Meteorology 131, 19–34.

895

CR IP T

Hertwig, D., Patnaik, G., Leitl, B., 2016a. LES validation of urban flow, part

I: flow statistics and frequency distributions. Environ Fluid Mech 17, 1–30. doi:10.1007/s10652-016-9507-7.

Hertwig, D., Patnaik, G., Leitl, B., 2016b. LES validation of urban flow, part II: eddy statistics and flow structures. Environ Fluid Mech 17, 1–28. doi:10.

900

AN US

1007/s10652-016-9504-x.

Huang, J., Cassiani, M., Albertson, J.D., 2009. The effects of vegetation density on coherent turbulent structures within the canopy sublayer: A largeeddy simulation study. Boundary-Layer Meteorol 133, 253–275. doi:10.1007/ s10546-009-9423-1.

905

M

Hultmark, M., Calaf, M., Parlange, M.B., 2013. A new wall shear stress model for atmospheric boundary layer simulations. J Atmos Sci 70, 3460–3470.

ED

URL: http://journals.ametsoc.org/doi/abs/10.1175/JAS-D-12-0257. 1, doi:10.1175/JAS-D-12-0257.1. 910

Inagaki, A., Castillo, M.C.L., Yamashita, Y., Kanda, M., Takimoto, H., 2012.

PT

Large-eddy simulation of coherent flow structures within a cubical canopy. Boundary-Layer Meteorol 142, 207–222. URL: http://link.springer.com/

CE

10.1007/s10546-011-9671-8, doi:10.1007/s10546-011-9671-8. Jackson, P.S., 1981. On the displacement height in the logarithmic velocity profile. J Fluid Mech 111, 15–25. URL: http://journals.cambridge.org/

AC

915

abstract{_}S0022112081002279, doi:10.1017/S0022112081002279.

J¨arvi, L., Grimmond, C.S.B., Christen, A., 2011. The Surface Urban Energy and Water Balance Scheme (SUEWS): Evaluation in Los Angeles and Vancouver. Journal of Hydrology 411, 219–237.

44

ACCEPTED MANUSCRIPT

920

Johansson, E., Emmanuel, R., 2006. The influence of urban design on outdoor thermal comfort in the hot, humid city of Colombo, Sri Lanka. International Journal Of Biometeorology 51, 119–133. Large-eddy simulations on the effects of surface ge-

CR IP T

Kanda, M., 2006.

ometry of building arrays on turbulent organized structures.

Boundary-

Layer Meteorol 118, 151–168. URL: http://link.springer.com/10.1007/

925

s10546-005-5294-2, doi:10.1007/s10546-005-5294-2.

Kanda, M., Hino, M., 1994. Organized structures in developing turbulent flow

AN US

within and above a plant canopy, using a Large Eddy Simulation. BoundaryLayer Meteorol 68, 237–257. doi:10.1007/BF00705599. 930

Kanda, M., Inagaki, A., Miyamoto, T., Gryschka, M., Raasch, S., 2013. A New Aerodynamic Parametrization for Real Urban Surfaces. Boundary-Layer Meteorology 148, 357–377.

M

Kanda, M., Moriwaki, R., Kasamatsu, F., 2004. Large-eddy simulation of turbulent organized structures within and above explicity resolved cube arrays.

Boundary-Layer Meteorol 112, 343–368.

ED

935

URL: http://link.

springer.com/article/10.1023/B:BOUN.0000027909.40439.7c,

doi:10.

1023/B:BOUN.0000027909.40439.7c.

PT

Katul, G.G., Albertson, J.D., 1998. An investigation of higher-order closure models for a forested canopy. Boundary-Layer Meteorology 89, 47–74. Katul, G.G., Parlange, M.B., 1992. Estimation of bare soil evaporation us-

CE

940

ing skin temperature measurements. J Hydrol 132, 91–106. doi:10.1016/

AC

0022-1694(92)90174-T.

Kellett, R., Christen, A., Coops, N.C., van der Laan, M., Crawford, B.R., Tooke,

945

T.R., Olchovski, I., 2013. A systems approach to carbon cycling and emissions modeling at an urban neighborhood scale. Landscape and Urban Planning 110, 48–58.

45

ACCEPTED MANUSCRIPT

Krayenhoff, E.S., Christen, A., Martilli, A., Oke, T.R., 2014. A multi-layer radiation model for urban neighbourhoods with trees. Boundary-Layer Meteorology 151, 139–178. Krayenhoff, E.S., Santiago, J.L., Martilli, A., Christen, A., Oke, T.R., 2015.

CR IP T

950

Parametrization of Drag and Turbulence for Urban Neighbourhoods with Trees. Boundary-Layer Meteorology 156, 157–189.

Lee, S.H., Lee, H., Park, S.B., Woo, J.W., Lee, D.I., Baik, J.J., 2016. Impacts of in-canyon vegetation and canyon aspect ratio on the thermal environment of

AN US

street canyons: Numerical investigation using a coupled WRF-VUCM model.

955

Q J R Meteorol Soc , 2562–2578doi:10.1002/qj.2847.

Lee, S.H., Park, S.U., 2008. A vegetated urban canopy model for meteorological and environmental modelling. Boundary-Layer Meteorol 126, 73–102. doi:10. 1007/s10546-007-9221-6.

Li, D., Bou-Zeid, E., 2013. Synergistic Interactions between Urban Heat Islands

M

960

and Heat Waves: The Impact in Cities Is Larger than the Sum of Its Parts*.

ED

Journal of Applied Meteorology and Climatology 52, 2051–2064. Li, Q., Bou-Zeid, E., Anderson, W., 2016a. The impact and treatment of the Gibbs phenomenon in immersed boundary method simulations of momentum and scalar transport. J Comput Phys 310, 237–251. doi:10.1016/j.jcp.

PT

965

2016.01.013.

CE

Li, Q., Bou-Zeid, E., Anderson, W., Grimmond, S., Hultmark, M., 2016b. Quality and reliability of LES of convective scalar transfer at high Reynolds numbers.

AC

970

Int J Heat Mass Transf 102, 959–970.

doi:10.1016/j.

ijheatmasstransfer.2016.06.093.

Lilly, D.K., 1992. A proposed modification of the Germano subgridscale closure method. Phys Fluids 4, 633–635. doi:10.1175/JAS3989.1.

Liss, K., Tooke, T.R., Heyman, E., Coops, N.C., Christen, A., 2013. Vegetation characteristics at the Vancouver EPiCC experimental sites – EPiCC Technical 46

ACCEPTED MANUSCRIPT

Report No. 3. Technical Report. The University of British Columbia. URL:

975

https://dx.doi.org/10.14288/1.0103589. Lu, H., Port´e-Agel, F., 2010. A modulated gradient model for large-eddy simu-

015109. doi:10.1063/1.3291073. 980

CR IP T

lation: Application to a neutral atmospheric boundary layer. Phys Fluids 22,

Macdonald, R.W., Griffiths, R.F., Hall, D.J., 1998. An improved method for the estimation of surface roughness of obstacle arrays. Atmos Environ 32, 1857–1864. doi:10.1016/S1352-2310(97)00403-2.

AN US

Masson, V., 2005. Urban surface modeling and the meso-scale impact of cities. Theoretical and Applied Climatology 84, 35–45. 985

Meneveau, C., Katz, J., 2000. Scale-invariance and turbulence models for largeeddy simulation. Annu Rev Fluid Mech 32, 1–32. doi:10.1146/annurev. fluid.32.1.1.

M

Meneveau, C., Lund, T.S., Cabot, W.H., 1996. A Lagrangian dynamic subgridscale model of turbulence.

J Fluid Mech 319, 353–385.

doi:10.1017/

S0022112096007379.

ED

990

Mittal, R., Iaccarino, G., 2005. Immersed boundary methods. Annu Rev Fluid

PT

Mech 37, 239–261. doi:10.1146/annurev.fluid.37.061903.175743. Monin, A.S., 1977. Statistical fluid mechanics: The mechanics of turbulence.

CE

volume 45. doi:10.1119/1.10870. 995

Moonen, P., Gromke, C., Dorer, V., 2013. Performance assessment of Large

AC

Eddy Simulation (LES) for modeling dispersion in an urban street canyon with tree planting. Atmospheric Environment 75, 66–76.

Nelson, M.A., Pardyjak, E.R., Klewicki, J.C., Pol, S.U., Brown, M.J., 2007.

1000

Properties of the wind field within the Oklahoma City Park Avenue street canyon. Part I: Mean flow and turbulence statistics. Journal of Applied Meteorology and Climatology 46, 2038–2054. 47

ACCEPTED MANUSCRIPT

Nikora, V., McLean, S., Coleman, S., Pokrajac, D., McEwan, I., Campbell, L., Aberle, J., Clunie, D., Koll, K., 2007. Double-averaging concept for roughbed open-channel and overland flows: theoretical background. J Hydraul Eng

CR IP T

133, 884–895. doi:10.1061/(ASCE)0733-9429(2007)133:8(884).

1005

Nowak, D.J., Crane, D., 2002. Carbon storage and sequestration by urban trees in the USA. Environmental Pollution 116, 381–389.

Oikawa, S., Meng, Y., 1995. Turbulence characteristics and organized motion

in a suburban roughness sublayer. Boundary-Layer Meteorology 74, 289–312. Oke, T.R., 1988. Street design and urban canopy layer climate. Energy Build

AN US

1010

11, 103–113. doi:10.1016/0378-7788(88)90026-6.

Oke, T.R., 1989. The Micrometeorology of the Urban Forest. Philosophical Transactions Of The Royal Society Of London Series B-Biological Sciences 324, 335–349.

Oke, T.R., Mills, G., Christen, A., Voogt, J.A., 2017. Urban Climates. Cam-

M

1015

ED

bridge University Press, In Press.

Orszag, S.A., 1969. Numerical methods for the simulation of turbulence. Phys Fluids 12, II–250. doi:10.1063/1.1692445.

PT

Orszag, S.A., 1970. Transform method for the calculation of vector-coupled sums: application to the spectral form of the vorticity equation. J Atmos Sci

1020

CE

27, 890–895. doi:10.1175/1520-0469(1970)027<0890:TMFTCO>2.0.CO;2. Panofsky, H.A., 1984. Atmospheric Turbulence: Models and Methods for Engi-

AC

neering Applications. Wiley.

Parlange, M.B., Brutsaert, W., 1989. Regional roughness of the landes forest

1025

and surface shear stress under neutral conditions. Boundary-Layer Meteorol 48, 69–81. URL: http://link.springer.com/10.1007/BF00121783, doi:10. 1007/BF00121783.

48

ACCEPTED MANUSCRIPT

Parlange, M.B., Brutsaert, W., 1990. Are Radiosonde Time Scales Appropriate to Characterize Boundary Layer Wind Profiles? 29, 249–255.

1030

J Appl Meteorol

URL: http://journals.ametsoc.org/doi/abs/10.1175/

CR IP T

1520-0450{%}281990{%}29029{%}3C0249{%}3AARTSAT{%}3E2.0.CO{%}3B2, doi:10.1175/1520-0450(1990)029<0249:ARTSAT>2.0.CO;2.

Parlange, M.B., Eichinger, W.E., Albertson, J.D., 1995. Regional scale evaporation and the atmospheric boundary layer. doi:10.1029/94RG03112.

1035

Rev Geophys 33, 99–124.

AN US

Patton, E.G., Sullivan, P.P., Shaw, R.H., Finnigan, J.J., Weil, J.C., 2016. At-

mospheric Stability Influences on Coupled Boundary Layer and Canopy Turbulence. Journal of the Atmospheric Sciences 73, 1621–1647.

Poggi, D., Porporato, A., Ridolfi, L., Albertson, J.D., Katul, G.G., 2004. Interaction between large and small scales in the canopy sublayer. Geophys Res

1040

M

Lett 31, L05102. doi:10.1029/2003GL018611.

Port´e-Agel, F., 2004. A scale-dependent dynamic model for scalar transport in

ED

large-eddy simulations of the atmospheric boundary layer. Boundary-Layer Meteorol 112, 81–105. URL: http://link.springer.com/10.1023/B:BOUN. 0000020353.03398.20, doi:10.1023/B:BOUN.0000020353.03398.20.

1045

PT

Port´e-Agel, F., Meneveau, C., Parlange, M.B., 2000. A scale-dependent dynamic model for large-eddy simulation: application to a neutral atmospheric bound-

CE

ary layer. J Fluid Mech 415, 261–284. doi:10.1017/S0022112000008776. Ramamurthy, P., Pardyjak, E.R., 2011. Toward understanding the behavior of carbon dioxide and surface energy fluxes in the urbanized semi-arid Salt Lake

AC

1050

Valley, Utah, USA. Atmos Environ 45, 73–84.

Ramamurthy, P., Pardyjak, E.R., 2015. Turbulent transport of carbon dioxide over a highly vegetated suburban neighbourhood. Boundary-Layer Meteorol 157, 461–479. doi:10.1007/s10546-015-0074-0.

49

ACCEPTED MANUSCRIPT

1055

Raupach, M.R., 1992. Drag and drag partition on rough surfaces. BoundaryLayer Meteorol 60, 375–395. URL: http://link.springer.com/article/ 10.1007/BF00155203, doi:10.1007/BF00155203.

Stress. Boundary-Layer Meteorology 65, 1–28. 1060

CR IP T

Rotach, M.W., 1993. Turbulence Close to a Rough Urban Surface .1. Reynolds

Roth, M., 1993. Turbulent transfer relationships over an urban surface. II.

Integral statistics. Quarterly Journal of the Royal Meteorological Society 119, 1105–1120.

AN US

Roth, M., 2000. Review of atmospheric turbulence over cities. Quarterly Journal of the Royal Meteorological Society 126, 941–990. 1065

Roth, M., Oke, T.R., 1993. Turbulent transfer relationships over an urban surface. I. Spectral characteristics. Quarterly Journal of the Royal Meteorological Society 119, 1071–1104.

M

Roth, M., Oke, T.R., 1995. Relative efficiencies of turbulent transfer of heat, mass, and momentum over a patchy urban surface. Journal of the Atmospheric Sciences 52, 1863–1874.

ED

1070

Ryu, Y.H., Bou-Zeid, E., Wang, Z.H., Smith, J.A., 2015. Realistic Representa-

PT

tion of Trees in an Urban Canopy Model. Boundary-Layer Meteorology 159, 193–220.

CE

Sadique, J., Yang, X.I.A., Meneveau, C., Mittal, R., 2016. Aerodynamic Properties of Rough Surfaces with High Aspect-Ratio Roughness Elements: Effect

1075

AC

of Aspect Ratio and Arrangements. Boundary-Layer Meteorol doi:10.1007/

s10546-016-0222-1.

Salim, S.M., Buccolieri, R., Chan, A., Di Sabatino, S., Cheah, S.C., 2011. Large

1080

eddy simulation of the aerodynamic effects of trees on pollutant concentrations in street canyons. Procedia Environmental Sciences 4, 17–24.

50

ACCEPTED MANUSCRIPT

Salmond, J.A., Williams, D.E., Laing, G., Kingham, S., Dirks, K., Longley, I., Henshaw, G.S., 2013. The influence of vegetation on the horizontal and vertical distribution of pollutants in a street canyon. Science of the Total

1085

CR IP T

Environment 443, 287–298. Schmid, H., Cleugh, H.A., Grimmond, C.S.B., Oke, T.R., 1991. Spatial variability of energy fluxes in suburban terrain. Boundary-Layer Meteorology 54, 249–276. Scotti, A., Meneveau, C., Lilly, D.K., 1993.

2308.

1090

Phys Fluids A Fluid Dyn 5, 2306–

AN US

sky model for anisotropic grids.

Generalized Smagorin-

URL: http://scitation.aip.org/content/aip/journal/pofa/5/

9/10.1063/1.858537, doi:10.1063/1.858537.

Sharma, V., Calaf, M., Lehning, M., Parlange, M.B., 2016. adaptive wind turbine model for an LES framework.

Wind Energy

URL: http://onlinelibrary.wiley.com/doi/10.1002/

M

19, 939–952.

Time-

we.1608/fullhttp://doi.wiley.com/10.1002/we.1877, doi:10.1002/we.

1095

ED

1877, arXiv:arXiv:1006.4405v1.

Shashua-Bar, L., Hoffman, M.E., 2000. Vegetation as a climatic component in the design of an urban street. Energy And Buildings 31, 221–235.

PT

Shaw, R.H., Patton, E.G., 2003.

Canopy element influences on resolved-

and subgrid-scale energy within a large-eddy simulation. Agric For Mete-

1100

CE

orol 115, 5–17. URL: http://linkinghub.elsevier.com/retrieve/pii/ S016819230200165X, doi:10.1016/S0168-1923(02)00165-X.

AC

Shaw, R.H., Schumann, U., 1992a. Large-eddy simulation of turbulent flow

1105

above and within a forest. Boundary-Layer Meteorol 61, 47–64. URL: http: //link.springer.com/10.1007/BF02033994, doi:10.1007/BF02033994.

Shaw, R.H., Schumann, U., 1992b.

Large-eddy simulation of turbulent

flow above and within a forest.

Boundary-Layer Meteorol 61, 47–

51

ACCEPTED MANUSCRIPT

64. URL: http://link.springer.com/10.1007/BF02033994http://link. springer.com/article/10.1007/BF02033994, doi:10.1007/BF02033994. 1110

Smagorinsky, J., 1963.

General circulation experiments with the primitive

091<0099:GCEWTP>2.3.CO;2.

CR IP T

equations. Mon Weather Rev 91, 99–164. doi:10.1175/1520-0493(1963)

Stensrud, D.J., 2007. Parameterization schemes. Cambridge University Press, Cambridge. doi:10.1017/CBO9780511812590. 1115

Stewart, I.D., Oke, T.R., 2012. Local climate zones for urban temperature

AN US

studies. Bulletin of the American Meteorological Society 93, 1879–1900.

Thom, A.S., 1968. The exchange of momentum, mass, and heat between an artificial leaf and the airflow in a wind???tunnel. Q J R Meteorol Soc 94, 44–55. doi:10.1002/qj.49709439906. 1120

Tooke, T.R., Coops, N.C., Voogt, J.A., Meitner, M.J., 2011. Tree structure

M

influences on rooftop-received solar radiation. Landscape and Urban Planning

ED

102, 73–81.

Tseng, Y.H., Meneveau, C., Parlange, M.B., 2006. Modeling flow around bluff bodies and predicting urban dispersion using large eddy simulation. Environ Sci Technol 40, 2653–2662. doi:10.1021/es051708m.

PT

1125

United Nations, 2014. Concise Report on the World Population Situation 2014

CE

, 1–38.

Vranckx, S., Vos, P., Maiheu, B., Janssen, S., 2015. Impact of trees on pollutant

AC

dispersion in street canyons: A numerical study of the annual average effects

1130

in Antwerp, Belgium. Science of the Total Environment 532, 474–483.

Wilson, J.D., 1988. A second-order closure model for flow through vegetation. Boundary-Layer Meteorol 42, 371–392. doi:10.1007/BF00121591.

52

ACCEPTED MANUSCRIPT

Xie, Z., Castro, I.P., 2006. LES and RANS for turbulent flow over arrays of wall-mounted obstacles. Flow, Turbul Combust 76, 291–312. doi:10.1007/ s10494-006-9018-6.

1135

CR IP T

Xie, Z., Coceal, O., Castro, I.P., 2008. Large-eddy simulation of flows over random urban-like obstacles. Boundary-Layer Meteorol 129, 1–23. doi:10. 1007/s10546-008-9290-1. Xie, Z.T., Castro, I.P., 2009. persion in urban streets.

1140

Large-eddy simulation for flow and dis-

Atmos Environ 43, 2174–2185.

URL: http:

AN US

//linkinghub.elsevier.com/retrieve/pii/S135223100900034X, doi:10. 1016/j.atmosenv.2009.01.016.

Yang, X.I.A., 2016. On the mean flow behaviour in the presence of regional-scale surface roughness heterogeneity.

Boundary-Layer Meteorol 161, 127–143.

URL: http://link.springer.com/10.1007/s10546-016-0154-9, doi:10.

1145

M

1007/s10546-016-0154-9.

Yang, X.I.A., Sadique, J., Mittal, R., Meneveau, C., 2016.

Exponential

ED

roughness layer and analytical model for turbulent boundary layer flow over rectangular-prism roughness elements.

J Fluid Mech 789, 127–165.

doi:10.1017/jfm.2015.687.

1150

PT

Yue, W., Meneveau, C., Parlange, M.B., Zhu, W., Kang, H.S., Katz, J., 2008. Turbulent kinetic energy budgets in a model canopy: Comparisons between

CE

LES and wind-tunnel experiments. Environ Fluid Mech 8, 73–95. doi:10. 1007/s10652-007-9049-0.

Yue, W., Parlange, M.B., Meneveau, C., Zhu, W., Hout, R., Katz, J., 2007.

AC

1155

Large-eddy simulation of plant canopy flows using plant-scale representation. Boundary-Layer Meteorol 124, 183–203. URL: http://link.springer.com/ 10.1007/s10546-007-9173-x, doi:10.1007/s10546-007-9173-x.

Zhu, X., Iungo, G.V., Leonardi, S., Anderson, W., 2016. 1160

Parametric

study of urban-like topographic statistical moments relevant to a pri53

ACCEPTED MANUSCRIPT

ori modelling of bulk aerodynamic parameters. Boundary-Layer Meteorol URL: http://link.springer.com/10.1007/s10546-016-0198-x, doi:10.

AC

CE

PT

ED

M

AN US

CR IP T

1007/s10546-016-0198-x.

54