An Inter-comparison of Three Urban Wind Models Using Oklahoma City Joint Urban 2003 Wind Field Measurements

Embed Size (px)

Citation preview

  • 8/2/2019 An Inter-comparison of Three Urban Wind Models Using Oklahoma City Joint Urban 2003 Wind Field Measurements

    1/12

    An inter-comparison of three urban wind models using Oklahoma City Joint

    Urban 2003 wind field measurements

    Marina Neophytou a,n, Akshay Gowardhan b, Michael Brown b

    a Environmental Fluid Mechanics Laboratory, Department of Civil and Environmental Engineering, University of Cyprus, Cyprusb Los Alamos National Laboratory, New Mexico, USA

    a r t i c l e i n f o

    Available online 22 March 2011

    Keywords:

    Fast-response models

    Model evaluation

    Large eddy simulation

    Computational fluid dynamics

    Empirical-diagnostic model

    Field measurements

    a b s t r a c t

    Three wind models are compared to near-surface time-averaged wind measurements obtained in

    downtown Oklahoma City during the Joint Urban 2003 Field Campaign. The models cover severallevels of computational approximation and include in increasing order of computational demand: a

    mass-consistent empirical-diagnostic code, a Reynolds-averaged NavierStokes (RANS) computational

    fluid dynamics (CFD) model, and a Large Eddy Simulation (LES) CFD code. The models were run with

    identical inlet and boundary conditions using the same grid resolution; the choice of the specific

    computational set-up reflects demands for fast-response models, although it may be a sub-optimal

    choice for the more complex models. A qualitative comparison of the model-computed flow fields with

    the Joint Urban 2003 wind measurements shows that all three models compare favorably to the near-

    surface wind measurements in many locations, although there are often instances of winds being

    calculated poorly in specific locations. The CFD models, however, had clearly superior looking flow

    fields, whereas the empirical-diagnostic code produced fields that were less smoothly varying. The

    inter-comparison exercise was supported by point-by-point quantitative comparisons of the wind

    speed and wind direction and with statistical measures. The RANS-CFD code, for example, was within

    50% of the measured wind speed 62% of the time as compared to 53% for the LES model and 49% for the

    empirical-diagnostic code. For wind direction, the RANS-CFD code was within 301 of the measured

    wind direction 58% of the time as compared to 50% for the LES code and 43% for the empiricaldiagnostic code. It is noticeable that throughout the various IOP cases examined, and under the specific

    computational set-up used in the simulations for fast-response needs, there is no clear superiority of

    one model over another. In addition, for the LES model, which in theory should provide the most

    realistic representation of the flow field, it appears that further to the sub-optimal computational set-

    up, as well as the uncertainty and natural variability persistent in the real world, has resulted in

    diminished performance.

    & 2011 Elsevier Ltd. All rights reserved.

    1. Introduction

    Simulating urban wind flow for fast-response applications

    poses significant challenges for computational modeling; on one

    hand both the inhomogeneity of the urban geometry and thecomplexity of the resulting flows urge for more sophisticated

    models, on the other, however, more advanced or detailed models

    are associated with substantial increase in computational time

    that nearly removes the model application from the context of

    fast response. Expected or accepted run-up times range from a

    few seconds to tens of minutes.

    The complexity of the flows that develop in cities, around

    and in-between buildings lies in the three-dimensionality of the

    resulting flows (Boris et al., 2001). For example, vertically rotating

    vortices can develop between buildings in so-called street can-

    yons (e.g. DePaul and Sheih, 1986; Oke, 1988), horizontally

    rotating vortices may exist near the canyon-intersection interface

    (e.g., Hoydysh and Dabberdt, 1988; Pol and Brown, 2008),updrafts often occur on the backsides of tall buildings (e.g., Heist

    et al., 2004), and strong downdrafts on the front faces of tall

    buildings that stick up above others can lead to divergence at the

    surface and street-level winds in all directions (e.g. Hanna et al.,

    2006; Nelson et al., 2007; Princevac et al., 2010). In more open

    areas, the prevailing wind can penetrate down into the city and

    result in strong channel flow, bifurcating at 4-way intersections,

    T-intersections, and along side streets (e.g., Belcher, 2005). The

    channel winds interact, compete, and merge with the building-

    induced vortices, updrafts and downdrafts and may produce highly

    spatially variable winds at street level. As the prevailing wind

    directions shifts, the vertically rotating vortices can turn into spiral

    Contents lists available at ScienceDirect

    journal homepage: www.elsevier.com/locate/jweia

    Journal of Wind Engineeringand Industrial Aerodynamics

    0167-6105/$ - see front matter & 2011 Elsevier Ltd. All rights reserved.

    doi:10.1016/j.jweia.2011.01.010

    n Corresponding author.

    E-mail address: [email protected] (M. Neophytou).

    J. Wind Eng. Ind. Aerodyn. 99 (2011) 357368

    http://-/?-http://www.elsevier.com/locate/jweiahttp://dx.doi.org/10.1016/j.jweia.2011.01.010mailto:[email protected]://dx.doi.org/10.1016/j.jweia.2011.01.010http://dx.doi.org/10.1016/j.jweia.2011.01.010mailto:[email protected]://dx.doi.org/10.1016/j.jweia.2011.01.010http://www.elsevier.com/locate/jweiahttp://-/?-
  • 8/2/2019 An Inter-comparison of Three Urban Wind Models Using Oklahoma City Joint Urban 2003 Wind Field Measurements

    2/12

    vortices (e.g., Nakamura and Oke, 1988; Klein and Clark, 2007),

    downdrafts can dissipate in intensity as the prevailing wind hits tall

    buildings at oblique angles, and channel flow reduces in strength in

    some streets and increases in others, resulting in complicated non-

    linear responses to the changing wind direction. These complex

    building-induced 3D flow patterns can significantly alter the trans-

    port and dispersion of plumes, resulting in displaced plume cen-

    trelines (e.g., Theurer et al., 1996; MacDonald and Ejim, 2002),

    significant upwind transport (e.g., Hanna et al., 2006) and reducedstreet-level concentrations in some cases due to enhanced venting

    (e.g., Heist et al., 2004, 2009) and increased street-level concentra-

    tions in other cases due to trapping and sheltering effects

    (e.g., Chang and Meroney, 2003).

    Accurate transport and dispersion prediction at the city scale

    requires models that account for these effects on the wind field.

    Over the past several decades considerable progress has been

    seen in the application of computational fluid dynamics (CFD)

    models. Some of the first CFD simulations were performed around

    a few idealized buildings (e.g., Patankar, 1975; Murakami et al.,

    1992). With the enhancement in computer power, researchers

    are currently performing simulations in real cities that include

    hundreds to thousands of building obstacles (e.g., see Neophytou

    and Britter, 2005; Pullen et al., 2005; Hanna et al., 2006).

    A number of teams have successfully applied Reynolds-averaged

    NavierStokes (RANS) codes and large-eddy simulation (LES)

    codes to neighborhood- and city-scale transport and dispersion

    problems (e.g., Coirier et al., 2006a, b). In addition, progress has

    been made in highly parameterized (non-NavierStokes) numer-

    ical wind field models for use in emergency response systems, or

    even for assessments and planning where many cases must be

    run over a short time period (Boris et al., 2001). These codes are

    based on potential flow or diagnosticempirical approaches

    (e.g. Rockle, 1990; Kaplan and Dinar, 1996). These types of models

    attempt to approximate the impacts of large numbers of buildings

    on the wind flow, providing a semi-realistic 3D flow field around

    buildings, but at a fraction of the run time as compared to CFD

    codes. The substantial decrease of run time makes such models

    particularly useful; however, it is extremely important to deter-

    mine: (i) what kind of wind flow features are missed using such

    simplifications and assumptions and consequently (ii) how much

    these predictions from simpler models differ from predictions of

    higher-complexity models.

    This paper reports results from an inter-comparison between

    three urban wind models, each of different level\of complexity,

    using data from time-averaged wind field measurements obtai-

    ned in downtown Oklahoma City during the Joint Urban 2003

    Field Campaign; the exercise is performed for 3 specific test cases

    (corresponding to three different Intensive Observation Periods)

    using a specific computational set-up reflecting fast-response

    standards. The paper is structured as follows: a brief summary

    of the Joint Urban 2003 field campaign including the specific

    Intensive Observation Periods (IOP) that were used in our ana-lyses is given in Section 2. The models in increasing order of

    computational cost and complexity include: (i) a mass-consistent

    empirical-diagnostic code, (ii) a RANS-CFD code, and (iii) a LES-CFD

    code with a more detailed description of the models presented

    in Section 3. It is important to note that the specific computa-

    tional set-up used in this exercise was selected to reflect realistic

    running conditions for fast-response applications but may be sub-

    optimal for the more complex RANS and LES models. Finally, a

    systematic comparison of the models with the wind field data is

    presented and discussed in Section 4 while Section 5 concludes

    with the overall findings of this work considering carefully the

    difficulty in arguing for clear superiority of one model over

    another based on a certain number of cases, as well as suggestions

    for further work.

    2. Oklahoma City Joint Urban 2003 (JU2003)

    Wind measurements obtained from the Joint Urban 2003

    (JU03) field experiment were used to evaluate the wind models.

    The JU03 experiment was held in Oklahoma City in July 2003 with

    the goal of providing information useful for testing and evaluation

    of the next generation of urban transport and dispersion models.

    The experiment consisted of a large number of tracer releases, a

    network of concentration samplers and meteorological sensorsplaced in and around the city. An overview of the field campaign

    is provided in Allwine et al. (2004).

    Although the Central Business District (CBD) in Oklahoma

    City is only a little over a kilometer on a side (Fig. 1), a number

    of relatively tall buildings exist there, including the Bank One

    Building (152 m), the FNC Building (123 m), the Oklahoma Tower

    (117 m), the Kerr-McGee Building (115 m), and City Place

    (107 m). The plan area fraction in the CBD is 0.4 and the average

    building height in the southern half of the CBD is 27 m, while in

    the northern half it is 65 m ( Burian et al., 2005). There are some

    trees in the CBD along streets and in the Botanical Gardens Park in

    the southwest sector of the CBD. During the July experimental

    period, temperatures were generally warm, winds were strong,

    and the prevailing wind direction consistently had a southerly

    component.

    About one hundred and fifty 2D and 3D sonic anemometers

    were placed throughout the city on towers, tripods, light poles,

    and building rooftops and nine sodars were placed in and around

    the city to provide upper-air information. Although a majority of

    the wind instrumentation was operating during the entire period

    of the experiment, there were intensive operating periods (IOPs)

    where more equipment was placed in the field. Within the central

    business district (CBD), the instrument distribution below roof-

    level included twenty four anemometers in the eastwest running

    street canyons (e.g., Park Avenue) and twenty four anemometers

    in northsouth running street channels (most in intersections,

    Depicted domain in comparative

    results (Figures 2,3, and4)

    350m

    Fig. 1. A Google Earth aerial view of the entire Central Business District in

    Oklahoma City. The extent of the image corresponds to the size of the modeling

    domain used for this study.

    M. Neophytou et al. / J. Wind Eng. Ind. Aerodyn. 99 (2011) 357368358

    http://dx.doi.org/10.1016/j.jweia.2011.01.010
  • 8/2/2019 An Inter-comparison of Three Urban Wind Models Using Oklahoma City Joint Urban 2003 Wind Field Measurements

    3/12

    however) during the IOPs. For this model evaluation study, we have

    used thirty-minute average sonic measurements made near street

    level on tripods (22.5 m agl), towers (2.53.5 m) and on light

    poles (7.58 m agl) from IOPs 2, 8, and 9. Note that in several

    locations two anemometers were placed at the same x, y location,

    but a meter or so apart vertically. In the vector plots to follow, this

    explains what appears to be two vectors at the same location in

    several cases.

    Towers and remote sensing sodars located upwind (i.e., south)of the CBD were used to create inflow wind profiles. The closest

    upwind location was about 200 m due south of our modeling

    domain. A Dugway Proving Ground (DPG) propeller anemometer

    was located at that location on a tower 35 m above the ground

    and 25 m above the roof of a post office building. The anem-

    ometer was not operating, however, during IOP 2. Eight sonic

    anemometers from Indiana University (IU) were located between

    2 and 80 m above ground level on several towers about 5 km

    south of our domain. For IOP2, only two sonics provided informa-

    tion, however. A sodar from the Pacific Northwest National

    Laboratory (PNNL) located about 1 km south-southwest of the

    CBD was used for the winds above 100 m. The Argonne National

    Laboratory (ANL) sodar located in the Botanical Gardens in the

    southwest corner of our domain also provided reasonable inflow

    information if the winds were from the southwest, but not if they

    were from the southeast (probably due to larger buildings being

    directly southeast of the sodar). In addition, we used two other

    ANL sodars that were downwind of the city to cross correlate with

    the upwind measurements.

    3. Description of computational models and model set-up

    In this study, we have used three different urban wind models,

    each representing a different level of complexity: an empirical-

    diagnostic code, a simple Reynolds-averaged NavierStokes (RANS)

    code, and a large eddy simulation (LES) codeall implemented as an

    option in the Quick Urban & Industrial Complex (QUIC) dispersion

    modeling system (Nelson and Brown, 2006).

    3.1. Empirical-diagnostic wind solver

    The QUIC-URB (Updated Rockle-style Building-aware) wind

    solver produces a mass-consistent flow field and uses empirical

    parameterizations to produce a velocity field that maintains

    important features of the time-averaged flow around buildings.

    The wind solver is based on Rockle (1990) in which various

    empirical relationships based on the building height, width, and

    length and the spacing between buildings are used to initialize

    the velocity fields in the regions around buildings (e.g., upwind

    rotor, downwind cavity and wake, street canyon vortex, and

    rooftop vortex). This initial flow field is then forced to satisfy

    mass conservation. More information about the QUIC-URB windsolver and modifications made to the code for high density urban

    areas can be found in Pardyjak and Brown (2003), Singh et al.

    (2008), Brown et al. (2009), and Gowardhan et al. (2010). Version

    5.4 of the QUIC-URB wind solver has been evaluated in this study.

    3.2. RANS computational fluid dynamics model

    The Q-CFD(RANS) solver is based on the 3D Reynolds-Aver-

    aged NavierStokes (RANS) equations for incompressible flow

    using a zero equation (algebraic) turbulence model based on

    Prandtls mixing length theory (Gowardan et al., in press). The

    selection of zero-equation turbulence model was made so as to

    reduce the run time of the CFD simulation, and therefore making

    it more closely adapted for a fast-response application (Chen and

    Xu, 1998). Computational time using a zero-equation model can

    be reduced by 28 times from that using a more complex

    turbulence modelkeeping all other computational settings such

    as discretization schemes and mesh size the same (REF). It is

    accepted however, that more complex turbulence models could

    also be considered, however, given that there has been no

    evidence of clear superiority or appropriateness of one model

    over another for all wind flow applications in real urban geome-

    tries (REF), using a zero-equation model was considered accep-table in the context of this exercise.

    The governing RANS equations are solved explicitly in time

    until steady state is reached using a projection method. At each

    time step of the projection method, the divergence-free condition

    is not strictly satisfied to machine precision levels, but rather

    when steady state is reached incompressibility is recovered. This

    makes the method comparable to the artificial compressibility

    method (Chorin, 1967). The RANS equations are solved on a

    staggered mesh using a finite volume discretization scheme that

    is second-order accurate in space (central difference) and time

    (AdamsBashforth). The law-of-the-wall was imposed at all of the

    solid surfaces. The pressure Poisson equation was solved using

    the successive over-relaxation method (SOR). A free slip condition

    was imposed at the top boundary and the side boundaries, while

    an outflow boundary condition is used at the outlet. More

    information on the numerical scheme and parameterizations

    can be found in Gowardhan et al. (in press).

    3.3. LES computational fluid dynamics model

    In large eddy simulation (LES) application of a spatial filter to

    the NavierStokes equations is generally used to partition the

    solution space into resolved and subgrid parts. The large and

    energetic scales of turbulence are calculated explicitly, while the

    small scales are represented by a model. The QUIC-LES code uses the

    Smagorinsky subgrid-scale eddy viscosity model. The NavierStokes

    equations are discretized using the second-order QUICK scheme

    in space and a second-order AdamsBashforth scheme in time ona staggered grid using a finite volume technique. The algorithm is

    based on the fractional step method (Kim and Moin, 1985). The

    Poisson equation for pressure p is solved by a multigrid method.

    Inflow velocity profile is specified at the inlet and outflow boun-

    dary condition is specified at the outlet, while the boundary

    condition at the top of the domain is free slip. At all solid surfaces,

    the local profile of the tangential velocity component is taken to

    be logarithmic and the normal velocity component is zero. The

    LES code used in this exercise is based on the development

    by Gowardhan et al. (2007) and has been validated for various

    turbulent flow problems.

    3.4. Model set-up

    The modeling domain covered most all of the Oklahoma City

    central business district (CBD) and was 1.2 km 1.2 km in size for

    all model calculations (see Fig. 1); the horizontal grid size was

    set to 5 m and the vertical grid size was 3 m resulting in a

    236 242 64 grid cell domain (3.66 million cells in total) for all

    models, with a blockage ratio of 15%. Such a computational set-up

    (e.g. of this mesh and cell size) is considered typicalreflect-

    ing realistic scenarios of fast-response applications (Boris et al.,

    2001; Patnaik and Boris, 2007). It should be noted that this may

    not necessarily be optimal for a standard application of CFD, e.g.

    LES or RANS, as the focus of this study is rather the application of

    these models in a realistic scenario of a fast-response situation for

    a corresponding typical computational set-up for such situation

    (Boris et al., 2001; Patnaik and Boris, 2007). Sensitivity studies

    M. Neophytou et al. / J. Wind Eng. Ind. Aerodyn. 99 (2011) 357368 359

  • 8/2/2019 An Inter-comparison of Three Urban Wind Models Using Oklahoma City Joint Urban 2003 Wind Field Measurements

    4/12

    were not performed due to the extremely high computational

    cost. The 3D building data were obtained from the Defense Threat

    Reduction Agency and the University of Oklahoma. Parking

    garages were set as solid buildings and although there were some

    trees in the downtown area, they were ignored in the simulations.

    The inflow wind speed profile was created by fitting a log-law

    to thirty-minute average wind measurements from the IU sonics,

    the DPG prop-vane and the PNNL sodar measurements (see

    Table 1). The specification of the inflow wind direction profile

    was somewhat more difficult as the different instrumentation

    often showed up to 301 differences in wind direction, with little

    consistent bias between instrument location and time periods. To

    determine the inflow wind direction profile, we plotted thirty-

    minute average measurements from the IU sonics, the DPG prop-

    vane, the PNNL sodar, and the three ANL sodars and specified the

    wind direction based on where the majority of measurements fell,

    ignoring outliers (similar to a modal average). Note that the

    inflow profile for the LES calculation did not vary with time, so

    that the inflow turbulence will be underestimated (e.g. Xie and

    Castro, 2006, 2009; DeCroix and Brown, 2002). The QUIC-LES code

    was run for 1800 s (physical time) for the flow to develop and the

    results were then averaged over the next 1800 s (actual duration

    of an IOP) so that it can be compared with the field data.

    In order to test the models under different inflow conditions,

    simulations were performed for three distinct time periods when

    the ambient winds were from the south-southwest, the south, and

    the south-southeast. These time periods corresponded to IOP-2,

    Release 1 (10:0010:30 CST, July 2), IOP-8, Release 2 (00:0000:30

    CST, July 25), and IOP-9, Release 2 (01:0001:30 CST, July 27). All

    simulations were performed assuming neutral stability. Within the

    urban canopy this assumption is likely valid, but above the canopy

    there may be a stable layer during IOP8 and an unstable layer during

    IOP2. Winds were fairly brisk during all three IOPs; hence, it is

    likely that the stability was not too far from neutral. Additionally,

    numerous studies have shown that a several hundred meter well-

    mixed neutrally stratified layer often exists above larger-sized city

    centers due to enhanced thermal and mechanical mixing effects. Thedowntown district of Oklahoma City is rather small, however, and

    thus it is not clear if the well-mixed layer is deep or shallow.

    4. Results and discussion

    4.1. Overview

    The results in this section are focused on the central part of the

    domain, which corresponds to the core of the Central Business

    District with dimensions roughly 400 m 500 m. This area con-

    tains a good number of the near-surface wind measurements in

    the most built-up region of the city. Additionally, from the

    computational modeling point of view, it avoids regions nearer

    the edges of the computational domain, which may be influenced

    by the application of specific boundary conditions. Fig. 1 shows an

    aerial view of the computational domain. An important feature of

    the flow field determining pollutant transport is the resulting

    wind flow directions, which subsequently determine flow struc-

    tures (e.g. channeling, recirculation, etc). Therefore, we have focused

    on several different types of flow regimes within the CBD: (i) flow

    channeling regions, (ii) street intersections, (iii) sheltered regions,

    and (iv) open areas. In the subsequent discussion and the modelinter-comparison, emphasis is placed on the ability of the model to

    capture such flow features discerned from the wind measurements.

    4.2. Comparative qualitative results

    4.2.1. IOP2southwesterly inflow

    For the IOP2 southwesterly inflow case (2151), the three

    models agree qualitatively with measurements in several loca-

    tions, and disagree in several other places (Fig. 2). Qualitatively

    the flows created by the Q-CFD(RANS) and Q-LES codes look

    smoother and are less noisy than the Q-URB solution. All three

    models capture extremely well the west to east channeling along

    Sheridan just north of the Convention Center, as well as the

    change in direction of the flow from west-northwesterly to

    southeasterly as one goes from the western side of the Broad-

    way-Main St. intersection to the northern side. As one travels up

    Broadway just past the Bank One Building, the Q-LES and

    Q-CFD(RANS) models show much stronger south-to-north chan-

    neling flow as compared to Q-URB. It appears that the CFD models

    overestimate the magnitude in this region, while the empirical

    Q-URB model underestimates it (this is discussed further in

    subsequent sections). All three models show a downwind cavity

    on the north side of the Bank One building (just to the west of

    the Broadway-Park Ave. intersection), however the CFD codes

    better represent the strength and direction of the single wind

    measurement there.

    There are some distinct differences between the models

    near the RobinsonSheridan intersection in the SW sector of

    the domain. First, in the southwestern corner of Fig. 2 plots, in

    the open region immediately west of the Convention Center, the

    Q-LES model has produced near zero winds, the Q-URB model

    shows the flow in this region being virtually identical to the

    ambient SW inflow, and Q-CFD(RANS) shows flow in the same SW

    direction, but it has been significantly retarded. Although there is

    only one measurement in this region it appears that the Q-URB

    and Q-CFD(RANS) codes have best matched the flow direction,

    and Q-CFD(RANS) has best matched the wind speed. The stagna-

    tion region produced by the Q-LES model appears to have resulted

    from strong southerly backflow on the front faces of tall buildings

    a few blocks north that are just outside Fig. 2 western plot

    boundary.

    Just to the north near the RobinsonMain St. intersection,the Q-CFD(RANS) model appears to be doing the best, capturing

    the westerly flow along Main St. and the southerly flow along

    Robinson. The Q-LES model also has the westerly flow penetrating

    through the intersection and matching the measurement along

    Main St., however, it also shows weak northerly winds on

    Robinson at the measurement location, not the relatively strong

    southerly winds measured there. Q-URB, on the other hand, is not

    able to propagate the westerly flow through the Robinson-Main

    St. intersection, but does correctly produce the southerly winds to

    the immediate north of the intersection on Robinson.

    Further north along Robinson at the Park Ave. cross street, the

    Q-CFD(RANS) model is seen to compute the merger of the south-

    erly flow along Robinson and the westerly flow along Park Avenue

    exceptionally well. The Q-LES model also does well, but appears

    Table 1

    The inflow wind profiles used to describe the different IOP cases in the numerical

    models.

    Test case: IOP2

    Wind inlet profile (logarithmic) Uref5 m/s, zref50 m, wind angle 215;

    Surface roughness z00.6 m

    Test case: IOP8

    Wind inlet profile (logarithmic) Uref8.5 m/s, zref80 m, wind angle 165;

    Surface roughness z01 m

    Test Case: IOP9

    Wind inlet profile (logarithmic) Uref7 m/s, zref50 m, wind angle 180;

    Surface roughness z01 m

    M. Neophytou et al. / J. Wind Eng. Ind. Aerodyn. 99 (2011) 357368360

  • 8/2/2019 An Inter-comparison of Three Urban Wind Models Using Oklahoma City Joint Urban 2003 Wind Field Measurements

    5/12

    to overestimate the strength of the westerly flow to the west of

    the intersection and underestimate the strength of the southerly

    flow to the south of the intersection. Q-URB actually matchesthe wind directions fairly well point by point in the Robinson

    Park Ave. intersection, but the flow field in this region is

    unsatisfactorily noisy.

    Within the western half of the Park Ave. street canyon

    between Robinson and Broadway, the Q-URB model has the flow

    going down the street the right way (westerly) at about the right

    magnitude, but individual flow vectors can be off by ninety

    degrees. Both the Q-CFD(RANS) and Q-LES models capture the

    flow direction well on the western half of Park Ave., but the

    strength of the flow is overestimated. The Q-CFD(RANS) code is

    able to resolve the easterly reverse flow at the eastern end of the

    street canyon, whereas the Q-LES model incorrectly shows the

    westerly flow extending through the entire length of the street

    canyon.

    The Q-LES model appears to perform better on the northern

    side of the domain. The backflow produced by Q-LES along McGee

    Ave. matches the measurements at McGee & Broadway and atMcGee & Robinson (note that there are tall buildings immediately

    north beyond the plot boundary). Both Q-URB and Q-CFD(RANS)

    are off by 901 at these two locations. It appears that the models

    slightly underestimated the spatial extent of the backflow as

    only a few grid cells away the model-computed flow is in the

    right direction. Another big difference between Q-URB and the

    two CFD codes is apparent in the northeast quadrant downwind

    of the wide building running along Kerr Avenue between Broad-

    way and Gaylord. Here, both CFD models show a rather short

    downwind cavity region, whereas Q-URB shows relatively strong

    backflow. Similarly, the long rectangular building directly south

    parallel to Gaylord shows southerly flow on the downwind side

    near the back wall, whereas Q-URB shows a thin region of

    northerly flow.

    Fig. 2. Measured (in black) and predicted (in gray) velocity field in the Central Business Area using QUIC-URB (a), Q-CFD(RANS) (b) and Q-LES (c) for IOP2 (wind direction:

    2151).

    M. Neophytou et al. / J. Wind Eng. Ind. Aerodyn. 99 (2011) 357368 361

  • 8/2/2019 An Inter-comparison of Three Urban Wind Models Using Oklahoma City Joint Urban 2003 Wind Field Measurements

    6/12

    4.2.2. IOP8south-southeasterly inflow

    For the IOP8 south-southwesterly inflow case (1651) depicted

    in Fig. 3, the three models do fairly well reproducing the flow

    along Sheridan near the Convention Center. The flow is straigh-

    tened to a pure southerly flow east of the Convention Center on

    Gaylord Blvd., is close to background flow direction to the west of

    the Convention Center on Robinson, and shows the easterly

    backflow in the Park Ave.Sheridan intersection immediately

    north of the Convention Center. The strength of the backflow offof the building immediately north of the Convention Center

    (x750, y 575) is weakest for Q-URB and strongest for Q-LES.

    All three models show southerly channeling along sections of

    Broadway, but there are differences in the strength of the flow,

    where the southerly flow is interrupted, and whether the south-

    erly flow covers the entire width of the street. The Q-LES model,

    for example, shows westerly flow penetrating through the Main

    St.Broadway intersection (in contradiction to the measurements

    along Main St. near the intersection), while the Q-CFD(RANS)

    shows the entire intersection consisting of southerly flow and

    Q-URB has the western half of the street with southerly flow and

    the eastern half being northerly. All of the models have some level

    of disagreement with the measurements along Main St., in some

    cases being 901801 out of phase.

    The channeling of the flow on Broadway north of the

    BroadwayMain St. intersection, the reverse flow in the lee ofthe Bank One Building, and the penetration of the flow into Park

    Ave. seems to be best captured by the Q-CFD(RANS) model,

    although the Q-LES does a fair job as well. Q-URB completely

    misses the reverse flow on the downwind side of the Bank One

    Building, but other than that appears to get the wind directions

    more or less correct in this region. On the western side of

    Park Ave there is clearly divergent flow resulting from

    Fig. 3. Measured (in black) and predicted (in gray) velocity field in the Central Business Area using QUIC-URB (a), Q-CFD(RANS) (b) and Q-LES (c) for IOP8 (wind direction:

    1651).

    M. Neophytou et al. / J. Wind Eng. Ind. Aerodyn. 99 (2011) 357368362

  • 8/2/2019 An Inter-comparison of Three Urban Wind Models Using Oklahoma City Joint Urban 2003 Wind Field Measurements

    7/12

    south-southeasterly flow impinging on the front face of the

    relatively tall City Place building. The CFD models capture this

    feature best, although Q-URB also captures some aspects of

    the flow.

    The flow computed by all three models near the Robinson

    Main St. intersection is in good agreement with the two measure-

    ments there. Further north, at the RobinsonPark Ave. intersec-

    tion, the southerly flow is also well predicted by all three models

    and in the case of the two CFD models the easterly outflow fromPark Ave as well. All three models, however, incorrectly show a

    southerly flow through the RobinsonKerr intersection, whereas

    the wind measurement indicates an easterly flow there.

    The southerly flow measured near the BroadwayKerr intersec-

    tion is captured by the Q-URB and Q-LES models, but the Q-

    CFD(RANS) code shows easterly flow there. Further north, at the

    BroadwayMcGee intersection, the Q-LES model is in best

    agreement to the weak westerly flow measured there, while

    Q-URB and Q-CFD(RANS) are both off by 901 with a strong

    southerly flow.

    4.2.3. IOP9southerly inflow

    This case with winds coming from 1801 proved to be fairly

    difficult for Q-URB, while the Q-CFD(RANS) code appears to have

    done fairly well and the Q-LES somewhere in between ( Fig. 4). Atthe BroadwaySheridan intersection just north of the Convention

    Center, the Q-URB model has likely overestimated the length of

    the downwind cavity reverse flow region (however, the measure-

    ments are inconclusive as one of the measurements at this

    location shows weak reverse flow, while the other does not).

    Q-LES shows a combination of westerly channel flow, backflow off

    the tall building to the north (x 750, y 575), and weak reverse

    Fig. 4. Measured (in black) and predicted (in gray) velocity field in the Central Business Area using QUIC-URB (a), Q-CFD(RANS) (b) and Q-LES (c) for IOP9

    (wind direction:1801).

    M. Neophytou et al. / J. Wind Eng. Ind. Aerodyn. 99 (2011) 357368 363

  • 8/2/2019 An Inter-comparison of Three Urban Wind Models Using Oklahoma City Joint Urban 2003 Wind Field Measurements

    8/12

    flow in the lee of the Convention Center in reasonable agreement

    with the measurements. Q-CFD(RANS) shows similar features as

    the Q-LES calculation, except that the westerly channeling is

    stronger and no evidence of reverse flow is apparent immediately

    downwind (north of) the Convention Center.

    Both Q-URB and Q-LES incorrectly predict southerly flow

    through the BroadwayMain St. intersection, while Q-CFD(RANS)

    correctly computes the easterly flow seen there (resulting from

    strong street-level backflow off the front face of the tall Bank Onebuilding). All three models are in agreement with the strong

    southerly channeling north of the BroadwayMain intersection.

    Both Q-CFD(RANS) and Q-LES predict backflow on the lee (north)

    side of the Bank One building in agreement with measurements,

    but the flow field circulation is very different in each case. Q-URB

    computes some backflow in this region, but the locations of the

    backflow do not agree with the measurements.

    Both Q-LES and Q-CFD(RANS) capture the penetration of the

    winds from Broadway into Park Ave as depicted in the measure-

    ments, while Q-URB does not. Q-CFD(RANS) is the only model

    that shows the end vortex in this region in agreement with the

    easterly flow on the north side of the street and westerly flow on

    the southern side. Both Q-LES and Q-CFD(RANS) agree with the

    measurements showing westerly flow in the center of the street

    canyon and both have produced a counter rotating vorticity in the

    street canyon near the RobinsonPark Ave intersection (although

    it is not clear from the measurements if such a vortex exists

    there). Overall, Q-URB performed poorly in Park Ave.

    The southerly winds along Robinson, however, were well

    predicted by Q-URB, as well as Q-LES. Q-CFD(RANS) did well also,

    but incorrectly computed a westerly wind at the RobinsonMain

    St. intersection, whereas the measurement revealed a southerly

    wind. All three models incorrectly predicted a southerly wind at

    the RobinsonMcGee intersection, whereas the actual wind was

    easterly. The direction of the wind one block to the east, also on

    McGee, was correctly computed by Q-CFD(RANS), but not by

    Q-URB and Q-LES. It is obvious that Q-URB has severely over-

    estimated the southerly channel flow along Broadway resulting in

    the incorrect wind flow at the McGeeBroadway intersection.

    Overall, it could be noted that the models have reasonable

    agreement with measurements in some locations and significant

    differences in others. Although an attempt has been made to

    group the zones of weak agreement and provide possible reasons

    for the discrepancies, due to the relatively limited number of

    points, areas and cases we avoided to deduce a more generalized

    conclusion on such issue of stronger/weaker agreement and their

    possible related explanation. For instance, the inherent transient

    effects of urban aerodynamics are predominant in certain loca-

    tions than others and that could explain perhaps some discre-

    pancies especially for RANS.

    4.3. Comparative quantitative results

    A quantitative comparison has been made by examining

    various wind scatter plots produced by the three models. An

    important feature that is examined here is the wind direction, in

    addition to wind speed, as this determines substantially the flow

    pattern as well as features produced within the urban area, which

    thereby determine the transport and dispersion. Figs. 5 and 6

    show the scatter plots for the wind speed and wind direction for

    each of the 3 IOP cases IOP2, IOP8, IOP9, respectively and

    depict the bounds for the 710%, 25%, 50% and 100% percentage

    error for wind speed as well the bounds within 7151, 301, 451and 901 for wind direction. From these plots, the percentage of the

    model predictions within the specified error bounds was calcu-

    lated. The results from each IOP case and each model are listed in

    Table 2, while Table 3 lists the average results for all three IOP

    cases for each model and Table 4 lists the correlation coefficients

    (between observed and predicted) derived from each scatter plot

    (each case and model).

    As indicated in Tables 2 and 3, the RANS code was within 50%

    of the measured wind speed 62% of the time (on average

    considering all three IOP cases, ranging from 57% in IOP9 to 71%

    IOP2) as compared to 53% for the LES model (on average

    considering all three IOP cases, ranging from 41% in IOP2 to 60%in IOP8) and 49% for the empirical-diagnostic code (on average

    considering all three IOP cases, ranging from 47% in IOP2 to 51%

    IOP8). Considering all the cases individually, there does not seem

    any one proving consistently more difficult than others for the

    models to capture. In turn, considering the level of quantitative

    differences in the computed errors, no model seems to be

    performing clearly better than other in predicting the wind speed.

    Similarly, for wind direction, the RANS-CFD code was within 301of the measured wind direction 58% of the time (on average

    considering all IOP cases, ranging from 54% in IOP8 to 63% in

    IOP2) as compared to 50% for the LES code (on average consider-

    ing all IOP cases, ranging from 50% in IOP9 to 57% in IOP2 and

    IOP8) and 43% for the empirical diagnostic code (see Tables 2

    and 3). If stricter bounds are considered, for example wind speed

    errors within 25%, the RANS-CFD was within this error threshold

    33% of the time, the LES model 29% and the empirical-diagnostic

    code 28%, considering all three IOP cases on average. For the average

    wind direction error, the model predictions were within 151 of the

    measured wind direction 39% of the time for the RANS-CFD code, as

    compared to 34% for the LES code and 27% for the empirical

    diagnostic code. For cases where the plume is being channeled down

    a street one direction or 1801 in the other direction, wind errors at

    street-level of up to7601 can in many cases actually be considered a

    success for street-level plume transport and dispersion since the

    plume will be transported in the right direction within a street

    canyon if the winds are within this error bound. That is, for

    successful plume model transport one may consider whether the

    model computations have the same easterly or westerly component

    as the wind measurements in an eastwest running street or the

    same northerly or southerly component in a northsouth running

    street. It is interesting to note that although some of the vector plots

    may show a fairly good agreement (in the flow patterns), it appears

    that a relatively good agreement in vector plots does not necessarily

    mean a good scatter plot for wind speeds. For example, although

    the values of the correlation coefficients for wind direction may

    reach as high as 0.860 (that is for Q-URB in the IOP 2 case), for

    exactly the same case, the same model predictions produce a

    correlation coefficient for the wind speeds as low as 0.068

    (see Table 4). Given the range of correlation coefficient values across

    the various cases, it appears that wind directions are overall better

    captured than the wind speeds. All the correlation coefficients

    between the observed and simulated values for each presented

    scatter plot (of Figs. 5 and 6) are listed in Table 4.It is important to consider that given the natural variability

    and uncertainties in the field, these levels of quantitative differ-

    ences in the various computed statistical measures do not support

    clearly a better performance of one model over others. It is also

    worth noting the computational times required to run the

    corresponding code. For example, the computational time for

    the simulations using the Q-URB code for the three IOP cases were

    of the order of 1 min, for the Q-CFD(RANS) of the order of 30-min

    and for the Q-LES 30 h using a standard PC for Q-URB and Q-

    CFD(RANS) and a parallel cluster of 8 nodes for the Q-LES.

    Although the accuracy of Q-URB may not appear as good as

    that of the Q-CFD(RANS) and Q-LES codes, it is only by a rela-

    tively small percentage fraction when compared to the two

    CFD codes.

    M. Neophytou et al. / J. Wind Eng. Ind. Aerodyn. 99 (2011) 357368364

  • 8/2/2019 An Inter-comparison of Three Urban Wind Models Using Oklahoma City Joint Urban 2003 Wind Field Measurements

    9/12

    Moreover, accounting for the inherent uncertainty and natural

    variability in the wind field, it is difficult to argue in some cases

    (especially when differences in percentage fraction of successes

    are of the order of a few percentage points) for clear superiority of

    one model over another. Especially, with regard to evaluationof urban wind models, it is necessary to specify the purpose of

    model application and thereby judge its appropriateness through

    specific performance evaluation criteria and metrics (COST732,

    2007). For example, given that the empirical diagnostic code runs

    23 orders of magnitude faster and its performance in the

    quantitative comparisons, the empirical diagnostic code appears

    to be a viable option for applications where fast turnaround time

    is required or where many cases must be run.

    It is important to note that for the purpose of consistent

    comparisons, the LES code is most likely run under sub-optimal

    conditions, e.g., the grid cell size near the walls should be smaller

    and the inlet boundary conditions should be fully turbulent and

    match the atmospheric conditions during the experiment. It is the

    case, however, that conditions and parameters representative of

    the field and necessary as input to the LES code are often not

    available from the available field measurements and in some

    cases the needed grid resolution is not operationally plausible.

    5. Concluding remarks

    Three computational wind models of different level of com-

    plexity were systematically tested and compared with wind

    measurements from the Oklahoma City Joint Urban 2003 field

    experiment. The comparative exercise included only near-street

    level data collected within the Central Business District (between

    2 and 7.5 m) using 30 min averages. The size of the input domain,

    grid resolution, building dimensions, wind inflow profiles and

    other relevant parameters within the different codes were

    selected to reflect fast-response modeling needs and were matched

    to allow for consistent comparisons.

    Overall, the results show that qualitatively all three models

    compare favorably to the near-surface wind measurements in

    IOP2 QUIC-URB

    IOP2 QUIC-RANS

    IOP2 QUIC-LES

    IOP9 QUIC-URBIOP8 QUIC-URB

    IOP9 QUIC-RANSIOP8 QUIC-RANS

    IOP9 QUIC-LESIOP8 QUIC-LES

    Fig. 5. Scatter plots for wind speed for measurements and corresponding predictions in the Central Business Area using QUIC-URB (top), Q-CFD(RANS) (middle) and Q-LES

    (bottom) for IOP2, IOP8, IOP9 (left, center, right, respectively). The plotted lines show the bound lines for percentage error within 710%, 725%, 750%, 7100%.

    M. Neophytou et al. / J. Wind Eng. Ind. Aerodyn. 99 (2011) 357368 365

  • 8/2/2019 An Inter-comparison of Three Urban Wind Models Using Oklahoma City Joint Urban 2003 Wind Field Measurements

    10/12

    IOP2 QUIC-URB

    IOP2 QUIC-RANS

    IOP2 QUIC-LES

    IOP9 QUIC-URBIOP8 QUIC-URB

    IOP8 QUIC-RANS IOP9 QUIC-RANS

    IOP9 QUIC-LESIOP8 QUIC-LES

    Fig. 6. Scatter plots for wind velocity direction for measurements and corresponding predictions in the Central Business Area using QUIC-URB (top), Q-CFD(RANS) (middle)

    and Q-LES (bottom) for IOP2, IOP8, IOP9 (left, center, right, respectively). The plotted lines show the bound lines for wind direction error within 7151, 7301, 7451 and

    7901.

    Table 2

    Calculated percentage error for each model in each IOP case, for wind speed and wind direction; the symbols OBS and SIM represent the observed and simulated/predicted

    values, respectively.

    Wind speed error, Error9OBSSIM9

    0:5 OBS SIM

    Wind direction error, Error 9OBSSIM9

    o100% o50% o25% o 10% o901 o451 o301 o151

    IOP 2

    Q-URB 71% 47% 27% 14% 77% 65% 48% 32%

    Q-CFD 88% 71% 37% 14% 84% 75% 63% 38%

    Q-LES 82% 41% 27% 10% 77% 68% 57% 38%

    IOP 8

    Q-URB 85% 51% 35% 13% 77% 52% 43% 30%

    Q-CFD 85% 58% 33% 16% 86% 59% 54% 38%

    Q-LES 87% 60% 18% 5% 77% 68% 57% 38%

    IOP 9

    Q-URB 79% 48% 23% 4% 77% 57% 39% 20%

    Q-CFD 90% 57% 30% 13% 91% 68% 57% 41%

    Q-LES 84% 57% 43% 23% 91% 70% 50% 34%

    M. Neophytou et al. / J. Wind Eng. Ind. Aerodyn. 99 (2011) 357368366

  • 8/2/2019 An Inter-comparison of Three Urban Wind Models Using Oklahoma City Joint Urban 2003 Wind Field Measurements

    11/12

    many locations, although there are several instances of winds

    being calculated poorly in specific locations. This qualitative

    assessment was bolstered by the point-by-point quantitative

    comparisons of the wind speed and wind direction. The RANS-

    CFD code, for example, was within 50% of the measured wind

    speed 62% of the time as compared to 53% for the LES model and

    49% for the empirical-diagnostic code. For wind direction, the

    RANS-CFD code was within 301 of the measured wind direction

    58% of the time as compared to 50% for the LES code and 43% for

    the empirical diagnostic code. It should be pointed out that there

    were noticable differences between the wind fields produced by

    the LES and RANS-CFD codes in specific regions of the domain,

    especially in intersections, in the strength of backflow resulting

    from flow impingement on the front faces of buildings, and the

    strength of channel flow. It is important to consider that given the

    variability and uncertainties in the field, the level of quantitative

    differences in the computed errors do not support a clearlysuperior performance of one model over others. However, it is

    interesting to note that the simpler model seems to perform

    almost nearly as well as the more computationally demanding

    models. It should also be noted that the computational set-up

    selected for all the models reflects typical realistic settings and

    requirements for fast-response modeling and may not necessarily

    reflect optimal settings and performance, particularly for RANS

    and LES in more general applications. Moreover, accounting for

    the inherent uncertainty and natural variability in the wind field,

    it is difficult to argue in some cases for clear superiority of one

    model over another. Given that the empirical diagnostic code is

    23 orders of magnitude faster than the RANS and LES-CFD codes,

    it appears that in this context it can be considered a practical

    option for studies where time is of concern. Future work on the

    comparison of model-computed and measured concentrations

    from the Joint Urban 2003 tracer field study using the three

    different types of wind models to drive a plume transport and

    dispersion model will help to answer whether or not the differ-

    ences in the wind fields translates into significant differences in

    the concentration fields.

    Acknowledgements

    Dr. M. Neophytou wishes to acknowledge financial supportgranted by the United Nations Educational, Science and Cultural

    Organization (UNESCO) for a 6-month exchange visit to US during

    which this research has been conducted. Dr. Neophytou wishes

    also to acknowledge the Systems Engineering and Integration

    Group of Los Alamos National Laboratory (LANL) for hosting and

    facilitating her 3-month visit at LANL.

    References

    Allwine, K.J., Leach, M.J., Stockham, L.W., Shinn, J.S., Hosker, R.P., Bowers, J.F.,Pace, J.C., 2004. Overview of Joint Urban 2003an atmospheric dispersionstudy in Oklahoma City. In: Proceedings of the AMS Symposium on Planning inNowcasting and Forecasting in Urban Zones, Seattle, WA.

    Belcher, S., 2005. Mixing and transport in urban areas. Philos. Trans. R. Soc. A 363,29472968.

    Boris, J., et al., 2001. Simulation of fluid dynamics around complex urbangeometries. In: Proceedings of the 39th Aerospace Sciences Meeting, AmericanInstitute Aeronautics and Astronautics, Reston, VA.

    Brown, M., Gowardhan, A., Nelson, M., Williams, M., Pardyjak, E., 2009. Evaluationof the QUIC wind and dispersion models using the Joint Urban 2003 fieldexperiment. In: Proceedings of the AMS eighth Symposium Urban Environ-ment, Phoenix, AZ.

    Burian, S., McKinnon, A., Hartman, J., Han, W., 2005. National Building StatisticsDatabase: Oklahoma City, Los Alamos National Laboratory Report LA-UR-05-8153,17 pp.

    Chang, C., Meroney, R., 2003. Concentrations and flow distributions in urban streetcanyons: wind tunnel and computational data. J. Wind Eng. Ind. Aerodyn. 91,11411154.

    Chen, Q., Xu, W., 1998. A zero-equation turbulence model for indoor airflowsimulation. Energy Buildings 28 (2), 137144.

    Chorin, A.J., 1967. A numerical method for solving incompressible viscous flow

    problems. J. Comput. Phys. 2, 1226.Coirier, W.J., Kim, S., Chen, F., Tuwari, M., 2006b. Evaluation of urban scalecontaminant transport and dispersion modeling using loosely coupled CFDand mesoscale models. In: Proceedings of the sixth Symposium on the UrbanEnvironment, American Meteorological Society, Atlanta, GA.

    COST Action 732, 2007. Quality Assurance of Microscale Meteorological Models.Background and Justification Document to support the model evaluationguidance and protocol. 1 May 2007, Rex Britter, Michael Schatzmann (Eds.).

    DePaul, F., Sheih, C., 1986. Measurements of wind velocities in a street canyon.Atmos. Environ. 20, 455459.

    DeCroix, D., Brown, M., 2002. A large-eddy simulation of urban-scale dispersionand comparison to URBAN 2000 Field Program measurements, LA-UR-02-4755,120 pp.

    Gowardhan, A., Brown, M., Pardyjak, E.R., 2010. Evaluation of a fast responsepressure solver for flow around an isolated cube. J. Environ. Fluid Mech. 10,311328.

    Gowardhan, A., Pardyjak, E., Senocak, I., Brown, M.J., 2007. Investigation ofReynolds stresses in a 3D idealized urban area using large eddy simulation.In: Proceedings of the AMS Seventh Symposium Urban Environment, San

    Diego, CA, paper J12.2, 8 pp.

    Table 3

    Calculated overall average (based on the 3 IOP cases) percentage error for each model, for wind speed and wind direction; the symbols OBS and SIM represent the observed

    and simulated/predicted values, respectively.

    Average wind spee d error, ErrorOBSSIMj j

    0:5 OBSSIM

    Average wind direction Error, Error OBSSIMj j

    o 100% o50% o25% o10% o901 o451 o301 o151

    Q-URB 78% 49% 28% 10% 77% 58% 43% 27%

    Q-CFD 88% 62% 33% 14% 87% 67% 58% 39%Q-LES 84% 53% 29% 13% 82% 64% 50% 34%

    Table 4

    Calculated correlation coefficient between the predicted and observed values of

    wind speed and wind direction for each model in the 3 IOP cases; the symbols OBS

    and SIM represent the observed and simulated/predicted values, respectively,

    while s represents the standard deviation of the corresponding values.

    Wind speed correlation

    coefficient,

    R OBSOBSSIMSIM

    sOBSsSIM

    Wind direction correlation

    coefficient,

    R OBSOBSSIMSIM

    sOBSsSIM

    IOP 2

    Q-URB 0.068 0.860

    Q-CFD 0.463 0.776Q-LES 0.160 0.676

    IOP 8

    Q-URB 0.352 0.425

    Q-CFD 0.644 0.618

    Q-LES 0.653 0.755

    IOP 9

    Q-URB 0.356 0.402

    Q-CFD 0.561 0.802

    Q-LES 0.513 0.784

    M. Neophytou et al. / J. Wind Eng. Ind. Aerodyn. 99 (2011) 357368 367

  • 8/2/2019 An Inter-comparison of Three Urban Wind Models Using Oklahoma City Joint Urban 2003 Wind Field Measurements

    12/12

    Gowardhan, A., Pardyjak, E., Senocak, I., Brown, M (in press). A CFD-based wind

    solver for an urban fast response transport and dispersion model, Environ

    Fluid Mech, doi:10.1007/s10652-011-9211-6.Hanna, S.R., Brown, M.J., Camelli, F.E., Chan, S.T., Coirier, W.J., Hansen, O.R.,

    Huber, A.H., Kim, S., Reynolds, R.M., 2006. Detailed simulations of atmospheric

    flow and dispersion in downtown Manhattan. Bull. Am. Meteorol. Soc. 87 (12),17131726.

    Heist, D.K., Brixey, L.A., Richmond-Bryant, J., Bowker, G.E., Perry, S.G., Wiener, R.W.,

    2009. The effect of a tall tower on flowand dispersion through a modelurban neighborhood, Part 1. Flow characteristics. J. Environ. Monit. 11,

    21632170.Hoydysh, W., Dabberdt, W., 1988. Kinematics and dispersion characteristics offlows in asymmetric street canyons. Atmos. Environ. 22, 26772689.

    Heist, D.K., Perry, S.G., Bowker, G.E., 2004. Evidence of enhanced vertical disper-

    sion in the wakes of tall buildings in wind tunnel simulations of LowerManhattan. In: Proceedings of the AMS Fifth Symposium on the Urban

    Environment, Vancouver, BC Canada.Kaplan, H., Dinar, N., 1996. A Lagrangian dispersion model for calculating

    concentration distribution within a built-up domain. Atmos. Environ. 30

    (24), 41974207.Kim, J., Moin, P., 1985. Application of a fractional-step method to incompressible

    NavierStokes. J. Comput. Phys. 59, 308323.Klein, P., Clark, J., 2007. Flow variability in a North American downtown street

    canyon. J. Appl. Met. Clim. 46, 851877.Macdonald, R.W., Ejim, C.E., 2002. Hydraulic flume measurements of mean flow

    and turbulence in a scale model array of 4:1 aspect ratio obstacles. In:Proceedings of the Fifth UK Conference on Wind Engineering (WES 02),

    University of Nottingham, England, 46 September 2002.Murakami, S., Mochida, A., Hayashi, Y., Sakamoto, S., 1992. Numerical study of

    velocity-pressure field and wind forces for bluff bodies by ke, ASM and LES.

    J. Wind Eng. Ind. Aerodyn. 4144, 28412852.Nakamura, Oke, T., 1988. Wind, temperature, and stability conditions in an

    eastwest oriented urban canyon. Atmos. Environ. 22, 26912700.Nelson, M.A., Pardyjak, E.R., Klewicki, J.C., Pol, S.U., Brown, M.J., 2007. Properties of

    the Wind Field within the Oklahoma City Park Avenue Street Canyon. Part I:mean flow and turbulence statistics. J. Appl. Meteorol. Climatol. 46 (12),

    20382054.

    Neophytou, Britter, 2005. Modelling of atmospheric dispersion in complex urbantopographies: a computational-fluid-dynamics simulation of the central Lon-don area. In: Proceedings of the Fifth International Congress on ComputationalMechanics, Limassol, Cyprus.

    Nelson, M.A., Brown, M.J., 2006. The QUIC Start Guide (v. 5.6). LA-UR-10-01062(available from /http://www.lanl.govS).

    Oke, T.,1988. Boundary Layer Climates, Routledge.Pardyjak, E.R., Brown, M., 2003. QUIC-URB v. 1.1: Theory and Users Guide, LA-UR-

    07-3181, 22 pp (available from /http://www.lanl.govS).Patankar, S.V., 1975. Numerical prediction of 3D flow, in Studies in Convection. In:

    Launder (Ed.), Academic Press, London, pp. 175.

    Patnaik, G., Boris, J., 2007. Fast and Accurate CBR Defense for Homeland Security:Bringing HPC to the First Responder and Warfighter. HPCMP-UGC pp.120-126,DoD High Performance Computing Modernization Program Users GroupConference, 2007, IEEE Computer Society, /http://doi.ieeecomputersociety.org/10.1109/HPCMP-UGC.2007.33 S.

    Pol, S., Brown, M., 2008. Flow patterns at the ends of a street canyon: measure-ments from the Joint Urban 2003 field experiment. J. Appl. Met. Clim. 47,14131426.

    Princevac, M., Baik, J.J., Xiangyi, L., Pan, H., Park, S.B., 2010. Lateral channelingwithin rectangular arrays of cubical obstacles. J. Wind Eng. Ind. Aerodyn. 98,377385.

    Pullen, J., Boris, J.P., Young, T.R., Patnaik, G., Iselin, J.P., 2005. A comparison ofcontaminant plume statistics from a Gaussian puff and urban CFD model fortwo large cities. Atmos. Environ. 39, 10491068.

    Rockle, R., 1990. Bestimmung der stomungsver-haltnisse im Bereich KomplexerBebauugsstruk-turen. Ph.D. Thesis, Vom Fachbereich Mechanik, der Tech-nischen Hochschule Darmstadt, Germany.

    Singh, B., Hansen, B., Brown, M., Pardyjak, E., 2008. Evaluation of the QUIC-URB

    fast response urban wind model for a cubical building array and wide buildingstreet canyon. Environ. Fluid Mech. 8, 281312.Theurer, W., Plate, E., Hoeschele, K., 1996. Semi-empirical models as a combination

    of wind tunnel and numerical dispersion modeling. Atmos. Environ. 30,35833597.

    Xie, Z., Castro, I., 2009. Large-eddy simulation for flow and dispersion in urbanstreets. Atmos. Environ. 43, 21742185.

    Xie, Z., Castro, I.P., 2006. Large-eddy simulation for urban micro-meteorology.J. Hydrodyn. B, 18, 259264 doi:10.1016/S1001-6058(06)60062-0.

    M. Neophytou et al. / J. Wind Eng. Ind. Aerodyn. 99 (2011) 357368368

    http://www.lanl.gov/http://www.lanl.gov/http://doi.ieeecomputersociety.org/10.1109/HPCMP-UGC.2007.33http://doi.ieeecomputersociety.org/10.1109/HPCMP-UGC.2007.33http://doi.ieeecomputersociety.org/10.1109/HPCMP-UGC.2007.33http://doi.ieeecomputersociety.org/10.1109/HPCMP-UGC.2007.33http://www.lanl.gov/http://www.lanl.gov/