Você está na página 1de 31

22 Nov 2002 11:42 AR AR159-FM35-19.tex AR159-FM35-19.

sgm LaTeX2e(2002/01/18) P1: IBC


10.1146/annurev.fluid.35.101101.161147

Annu. Rev. Fluid Mech. 2003. 35:469–96


doi: 10.1146/annurev.fluid.35.101101.161147
Copyright °
c 2003 by Annual Reviews. All rights reserved

FLOW AND DISPERSION IN URBAN AREAS


R. E. Britter
Department of Engineering, University of Cambridge, Cambridge CB2 1PZ,
United Kingdom: email: rb11@eng.cam.ac.uk
Annu. Rev. Fluid Mech. 2003.35:469-496. Downloaded from arjournals.annualreviews.org

S. R. Hanna
Hanna Consultants, Kennebunkport, Maine 04046; email: hannaconsult@adelphia.net
by University of Bologna on 08/24/10. For personal use only.

Key Words urban boundary layer, turbulence, pollutant dispersion


■ Abstract Increasing urbanization and concern about sustainability and quality
of life issues have produced considerable interest in flow and dispersion in urban
areas. We address this subject at four scales: regional, city, neighborhood, and street.
The flow is one over and through a complex array of structures. Most of the local
fluid mechanical processes are understood; how these combine and what is the most
appropriate framework to study and quantify the result is less clear. Extensive and
structured experimental databases have been compiled recently in several laboratories.
A number of major field experiments in urban areas have been completed very recently
and more are planned. These have aided understanding as well as model development
and evaluation.

1. INTRODUCTION

Much of the global population currently live and work in urban areas, and this
urbanization is expected to increase. This trend has recently inspired many urban-
centered studies. Many are of a fluid mechanical nature, either in isolation or in
combination with other disciplines such as chemistry, epidemiology, and pedestrian
and vehicular mobility.
Large-scale weather prediction and mesoscale meteorological models require
the parameterization of urban areas to provide boundary conditions. Regional
air-pollution models are used to estimate the transport of pollutants to and from
cities. Urban climatology (Oke 1987) addresses the mass, momentum, and en-
ergy transfers through an urban area and the resulting temperatures, humidities,
radiant fluxes, etc. These changes in surface energy balance and temperatures and
humidities in urban areas influence, for example, general urban planning, green-
space provision, and energy usage in cities.
Studies of urban air quality (Fenger et al. 1999) focus strongly on the wind flow
over and through the city and the sources of pollutants within and beyond the city.
A major pollutant source within the city is vehicle emissions, which lead to an

0066-4189/03/0115-0469$14.00 469
22 Nov 2002 11:42 AR AR159-FM35-19.tex AR159-FM35-19.sgm LaTeX2e(2002/01/18) P1: IBC

470 BRITTER ¥ HANNA

interaction between mobility, air quality, and the possible regulation of vehicles
in cities. The wind flow within cities, in particular, the local turbulence levels,
directly affect pedestrian mobility and comfort. This same wind flow, but on a
larger scale, represents the wind environment within which new buildings are to
be placed and is of concern both for wind-loading problems (Cook 1990) and for
the provision of clean air to the buildings and the removal of exhaust air from
the urban canopy. This wind environment and the building construction affect
some of the exchange processes between the building interior and exterior and,
consequently, building air quality and energy use. Hazardous materials in large
quantities are normally prohibited from heavily populated areas, but where this is
Annu. Rev. Fluid Mech. 2003.35:469-496. Downloaded from arjournals.annualreviews.org

not the case there is a need for emergency authorities and civil defense personnel
to have operational tools available to determine what action to take in case of an
accident. Very recently there has been increased concern about the nonaccidental
release of hazardous materials in urban areas.
by University of Bologna on 08/24/10. For personal use only.

Understanding the flow of the wind through and above the urban area and/or
the dispersion of material in that flow (Hanna & Britter 2002) is necessary. We
address these issues by considering the problem at different scales. At each of the
scales there are observations from the field and the laboratory that are interpreted in
terms of various physical (and possibly chemical) processes. These processes, once
recognized, are often combined and reformed into mathematical models that can
form a hierarchy of complexity or sophistication: Each model has its own regime
of applicability and accuracy. A detailed interpretation at one scale is commonly
parameterized to assist interpretation at the next larger scale. We use spatial scales
to describe the major urban flow features, although the spatial scales can be related
to time scales through the x = ut relation. Roughly speaking the timescales are the
spatial scales divided by an appropriate advection (wind) velocity. The discussion
is broken down into four ranges of length scales: regional (up to 100 or 200 km),
city scale (up to 10 or 20 km), neighborhood scale (up to 1 or 2 km), and street
scale (less than ∼100 to 200 m).
The regional scale is affected by the urban area. For example, the urban heat
island circulations, any enhanced precipitation, and the urban pollutant plume can
extend to these distances. At this scale the mean synoptic meteorological patterns
are given and the urban area represents a perturbation, causing deceleration and
deflection of the flow, as well as changes to the surface-energy budget and the
thermal structure.
The city scale represents the diameter of the average urban area. At these scales
the variations in flow and dispersion around individual buildings or groups of
similar buildings have been mostly averaged out. Wind flow models developed for
this range pay little attention to the details of the flow within the urban canopy
layer. Most of the mass of any pollutant cloud traveling over this distance will be
above the height of the buildings.
On the neighborhood scale buildings may still be treated in a statistical way;
however, the approach may be different to that on the city scale. At the
neighborhood scale we want to know more about the flow within the urban canopy.
30 Nov 2002 9:53 AR AR159-FM35-19.tex AR159-FM35-19.sgm LaTeX2e(2002/01/18) P1: IBC

FLOW IN URBAN AREAS 471

The wind flow, particularly within the canopy, may also be changing as it moves
from one neighborhood to the next. Much of the mass of a pollutant cloud traveling
over this distance may remain within the urban canopy.
The street (canyon) scale addresses the flow and dispersion within and near one
or two individual streets, buildings, or intersections. This would be of interest when
considering turbulence affecting pedestrian comfort and the direct exposure of
pedestrians and near-road residences to vehicular emissions. It can be of particular
interest when regulatory pollutant monitoring stations are placed within street
canyons.
Figure 1 is an illustration of the results from a model prediction of annual
Annu. Rev. Fluid Mech. 2003.35:469-496. Downloaded from arjournals.annualreviews.org

mean NO2 concentrations for greater London for the year 2005. The street scale
is obvious; the neighborhood scale is evident in Heathrow airport, which is to the
west of the center, whereas the city scale is an area of approximately 10 km in
radius. The relevant regulatory limit in the United Kingdom is 21 ppb (parts per
by University of Bologna on 08/24/10. For personal use only.

billion).

2. THE CITY SCALE AND THE REGIONAL SCALE

The city scale is, essentially, the urban area: an area distinguished from its sur-
roundings by its relatively large obstacles (buildings and other structures) and,
hence, by a large drag force; by the infusion of heat and, perhaps, moisture from
man’s activities; and by the large heat-storage capacity of concrete, other building
materials, and parking lots. The city scale can include variations in urban building
types and spacings and primarily concerns the boundary layer above the aver-
age building height, Hr. The regional scale is the larger surrounding area that is
influenced by or influences phenomena at the city scale.
Bornstein (1987) has observed that the flow is deflected over and around the
urban area. The vertical flow displacement sometimes is visible when it is marked
by formation of a cap cloud over the urban area, just as seen over a mountainous
island. The horizontal flow displacement can be seen in the turning of sea-breeze
or cold-front flow directions and in the bending of frontal surfaces. The deflections
are in part kinematic owing to the volume of the buildings and in part dynamic
owing to the drag forces on the buildings.
The surface-energy balance for the urban area and its surroundings is signifi-
cantly affected by differences in heat and moisture inputs due to human activities.
The additional heat fluxes and the heat storage of the urban surface lead to the
urban heat island phenomenon that has been widely observed and discussed (Oke
1987). This phenomenon causes a convergence of horizontal wind into the city
and vertical motions over the city. The mixing depth over a city is increased as a
result of the vertical motions as well as the enhanced heating.
The transport and dispersion of pollutants over the urban area is altered as a
result of increased mechanical turbulence caused by the relatively large obstacles
over which the pollutants must travel. Furthermore, the urban heat island causes the
22 Nov 2002 11:42 AR AR159-FM35-19.tex AR159-FM35-19.sgm LaTeX2e(2002/01/18) P1: IBC

472 BRITTER ¥ HANNA

boundary layer over an urban area to become more unstable as thermal turbulence
increases. Both of these effects enhance dispersion. The wind velocities are altered
in magnitude and direction. Wind speeds over the urban area are slower owing to
the increased roughness of these areas, and wind directions could change as a result
of the heat island circulations or the bending of the flow around and over the urban
area. Most urban areas have large “urban plumes” that can be easily observed 100
to 200 km downwind of the urban area.
Mathematical modeling of flows on the regional scale requires the parame-
terization of the effects of the urban surface on the flow. For example, weather
forecast models are routinely run with a horizontal grid scale of ∼20–40 km.
Annu. Rev. Fluid Mech. 2003.35:469-496. Downloaded from arjournals.annualreviews.org

Mesoscale meteorological models with horizontal grid scales of ∼2–8 km are


run for special purposes around urban areas and can better resolve the types of
land use (Brown 2000). The mesoscale models use a typical vertical grid spacing
of approximately 20 m near the surface and approximately 200 m in the middle
by University of Bologna on 08/24/10. For personal use only.

and top of the boundary layer. Such models often parameterize the urban area
by using a simple prescription of land use. Thus an “urban” land-use grid may
be assigned a single surface roughness length, soil moisture, albedo, and Bowen
ratio. At the city scale a similar need for the parameterization of the urban surface
exists.
Section 2.1 below addresses the effects of the momentum, the energy, and the
pollutant changes provided by the urban areas.

2.1. Momentum: Surface Shear-Stress and Wind Profiles


Urban obstacles exert a relatively large drag force on the atmosphere. This can
be treated by standard atmospheric boundary-layer formulas (Stull 1997), as long
as the mean building height, Hr, is small compared to the surface boundary-layer
depth, which is usually approximately 100 or 200 m, and the surface has some
statistical homogeneity.
Figure 2, adapted from Grimmond & Oke (2002), depicts a cross section of a
typical urban area, showing the three major sublayers: the urban canopy sublayer,
the roughness sublayer, and the inertial sublayer. The inertial sublayer is the area
where the boundary layer has adapted to the integrated effect of the underlying
urban surface, and it is treated by standard atmospheric boundary-layer formulas.
In the urban canopy sublayer the flow at a specific point is directly affected by
the local obstacles, and in the roughness sublayer the flow is still adjusting to the
effects of many obstacles.
The surface shear stress (averaged over the urban surface) defines a friction
velocity, u∗ , that can be used to derive wind and turbulence profiles. It is assumed
that, regardless of the underlying surface, the wind speed at the top of the boundary
layer (i.e., at a height of ∼500 to 1000 m) is approximately equal to the equilibrium
wind speed defined by the geostrophic wind speed, which is based on the synoptic
pressure gradient. For larger surface roughness the drag force (or u∗ ) is larger, and
the wind speed at any given level in the boundary layer is smaller.
22 Nov 2002 11:42 AR AR159-FM35-19.tex AR159-FM35-19.sgm LaTeX2e(2002/01/18) P1: IBC

FLOW IN URBAN AREAS 473


Annu. Rev. Fluid Mech. 2003.35:469-496. Downloaded from arjournals.annualreviews.org
by University of Bologna on 08/24/10. For personal use only.

Figure 2 Schematic of the flow through and over an urban area. Adapted
from Grimmond & Oke (2002).

The friction velocity, u∗ , is the key scaling velocity in Monin-Obukhov simi-


larity theory (Stull 1997). The other key scaling parameter is the Monin-Obukhov
length, L, which accounts for the effects of atmospheric stability and is propor-
tional to u3∗ divided by the surface turbulent (or sensible) heat flux from the ground
surface: Hs: L = − (u3∗ /κ)/(gHs/CpρT ), where g is the acceleration due to gravity,
Cp is the specific heat of air at constant pressure, ρ and T are the air density and
temperature, respectively, and κ is von Karman’s constant taken to be 0.40. Hs is
positive in the day and is negative at night (discussed in more detail in Section 2.3).
The wind-speed profile conforms to Monin-Obukhov similarity theory, with
the addition of the two scaling lengths, z0 (the surface roughness length) and d (the
surface displacement length):
u = (u∗ /κ)(ln [(z − d)/z0 ] + ψ[z/L]), (1)
where ψ(z/L) is a universal dimensionless function (Stull 1997) that equals zero
in neutral or adiabatic conditions (i.e., when L = ∞ or z/L = 0).
For the purposes of the following discussion of the urban boundary layer, neutral
conditions are assumed for which Equation 1 reduces to
u = (u∗ /κ) ln [(z − d)/z0 ]. (2)
Thus, given a wind-speed observation at a height greater than approximately 2Hr
(see Figure 3) and given an estimate of z0 and d, then u∗ can be estimated from
Equation 2. Note that u∗ and u at some reference height will define a drag coefficient
22 Nov 2002 11:42 AR AR159-FM35-19.tex AR159-FM35-19.sgm LaTeX2e(2002/01/18) P1: IBC

474 BRITTER ¥ HANNA


Annu. Rev. Fluid Mech. 2003.35:469-496. Downloaded from arjournals.annualreviews.org
by University of Bologna on 08/24/10. For personal use only.

Figure 3 The spatially averaged mean velocity profile near an urban


area. Adapted from Bottema (1997).

(for meteorologists) or skin-friction coefficient (for engineers). Consequently,


these coefficients relate directly to the parameters z0 and d.
Parameterizations for z0 and d are possible using several approaches. Land-use
methods are used in most applied dispersion models. These methods are based on
the tables of z0 versus descriptive land-use types (e.g., Stull 1997). The problem for
those interested in urban sites is that the land-use categories are usually very broadly
defined, as they have to cover all land uses, ranging from ice flats and deserts to
water surfaces and crops and forests. For example, the categories suggested by
Stull (1997) do not account for commercial or industrial sites, even though there
are four separate categories for towns and cities, with z0 ranging from 0.3 m for
“outskirts of towns” to 2 m for “centers of cities with very tall buildings.”
The methods by Auer (1978), based on land-use categories within a radius of
3 km from the source and on approximate descriptions of building geometries, are
used in the U.S. Environmental Protection Agency’s Guidelines and form the basis
for the decision as to whether the rural or urban dispersion correlations should be
used but not to assign a z0.
Davenport et al. (2000) have proposed a more detailed description of the relation
of z0 to urban/industrial land use. A portion of their table is given in Table 1,
where surface roughness lengths, z0, are listed for five categories of buildings and
industrial obstacles. The z0 values range between approximately 0.1 and 2 m, in
agreement with the z0 suggestions for “towns and cities” by Stull (1997).
A set of 12 urban land-use types has been proposed by Grimmond & Oke
(1999), based on an extensive review of wind-profile observations in many urban
areas. Their data show that z0/Hr ranges from approximately 0.06 to 0.20, and d/Hr
ranges from approximately 0.35 to 0.85. A further complementary set is provided
by Theurer (1999).
22 Nov 2002 11:42 AR AR159-FM35-19.tex AR159-FM35-19.sgm LaTeX2e(2002/01/18) P1: IBC

FLOW IN URBAN AREAS 475

TABLE 1 Updated surface roughness lengths, z0, for five urban and industrial categories∗
Surface roughness
Category length, z0 Urban/industrial site description
Roughly open 0.1 m Moderately open country with occasional obstacles
(e.g., isolated low buildings) at relative separations of
at least 20 obstacle heights.
Rough 0.25 m Scattered buildings and/or industrial obstacles at
relative separations of 8 to 12 obstacle heights.
Analysis may need displacement length, d.
Annu. Rev. Fluid Mech. 2003.35:469-496. Downloaded from arjournals.annualreviews.org

Very rough 0.5 m Area moderately covered by low buildings and/or


industrial tanks at relative separations of 3 to 7
obstacle heights. Analysis requires displacement
length, d.
by University of Bologna on 08/24/10. For personal use only.

Skimming 1.0 m Densely built-up area without much obstacle height


variation. Analysis requires displacement length, d.
“Chaotic” 2.0 m City centers with mixture of low-rise and high-rise
buildings. Analysis by wind tunnel advised.

This table is an abridged version of a table in Davenport et al. (2000).

More precise estimates of z0 and d can be made using information about building
sizes and spacing. The total building plan area, Ap, and the total building frontal
area, Af , in a building lot area, AT, can be used to define the “lambda parameters”
that are used in many empirical urban boundary-layer formulas: λp = Ap/AT;
λf = Af/AT.
The dimensionless frontal area, λf, is more important to drag because it repre-
sents the surface facing the wind flow. Typical values of λf are ∼0.1 for areas with
a moderate density of buildings and 0.3 for downtown areas. λf would be larger in
urban downtown areas were it not for the presence of large parking lots and parks.
Note that if the spacing between square buildings of dimension Hr equals 1/2 and
1.0 times Hr, then λf = 0.44 and 0.25, respectively. Grimmond & Oke (1999)
review and evaluate many competing techniques for determining z0 and d from λp
and λf . They also acknowledge three types of urban flow. At small λf the buildings
act in isolation, at larger λf the building wakes interfere with each other, and at
even larger λf a skimming flow over the buildings has limited direct penetration
into the spaces between the buildings.
Hanna & Britter (2002) have considered several field and laboratory data sets
as well as theoretical and empirical formulas in the literature and recommend the
following formulas:
z0 /Hr = λf for λf < 0.15, (3a)

z0 /Hr = 0.15 for λf > 0.15, (3b)

d/Hr = 3λf for λf < 0.05, (4a)


22 Nov 2002 11:42 AR AR159-FM35-19.tex AR159-FM35-19.sgm LaTeX2e(2002/01/18) P1: IBC

476 BRITTER ¥ HANNA

d/Hr = 0.15 + 5.5(λf − 0.05) for 0.05 < λf < 0.15, (4b)

and

d/Hr = 0.7 + 0.35(λf − 0.15) for 0.15 < λ f < 1.0. (4c)

Of course, λp and λf are not necessarily adequate to fully describe urban areas.
One aspect omitted from most techniques used for describing urban areas is the
great variability in building heights. Ratti et al. (2002) show that the ratio of the
standard deviation of building heights to the mean building height, Hr, ranges up
Annu. Rev. Fluid Mech. 2003.35:469-496. Downloaded from arjournals.annualreviews.org

to 1.0 for some urban areas. As a consequence the skimming-flow regime may not
be well represented by laboratory studies that use obstacles of constant height.
Because of the difficulties involved in processing geometrical data from multiple
individual buildings that are needed to apply some of the morphological methods,
by University of Bologna on 08/24/10. For personal use only.

these methods have not been widely used in the past for estimating the roughness
length for air-quality modeling applications. However, very recently there has been
demand for more detailed modeling, and this has led to the increased use of digital
elevation models of cities (Ratti et al. 2002) for morphological studies.
A more extensive discussion on urban parameterization schemes for mesoscale
models may be found in the report by Brown (2000). These schemes are directly
applicable to city-scale modeling. The parameterization is often similar to the use
of “wall functions” in engineering-based computational fluid dynamics.
It is important to recognize that, for all techniques described, an equilibrium
boundary layer develops only after the air has flowed over many individual obsta-
cles or rows of obstacles. This roughness change problem has been well studied
(see Smits & Wood 1985 for a review), and correlations for the growth of the inter-
nal boundary layer are readily available; the internal boundary layer is the region of
the flow that adjusts to the change of roughness. Within the internal boundary layer
there is an equilibrium layer that can be broken into the new roughness sublayer
and a new inertial sublayer. In a laboratory study Cheng & Castro (2001) found
that it took ∼160z0 for evidence of approximate similarity and ∼300z0 for the equi-
librium layer to reach the upper limits of the roughness sublayer. These distances
correspond to approximately four to seven rows of obstacles. This finding led them
to question whether urban areas ever allowed sufficient fetch for an inertial sublayer
to develop and whether z0 was an appropriate scaling parameter for urban areas.
However, urban areas might be better represented as regions (or neighborhoods)
of gradually varying roughness. The flow then is continuously readjusting to these
changing surface conditions. The effective roughness length over terrain that con-
sists of well-defined repeating patches of two different roughness surfaces was
studied by Goode & Belcher (1999), who used a linearized perturbation model.
They define the “blending height” as the top of the highest extent of the internal
boundary layers from individual obstacles, above which the flow is “fully adjusted”
to the combined roughnesses. In one of their numerical modeling tests, where the
roughness patches alternated every 50 m with values of z0 = 0.004 m (typical
of mowed grass) and z0 = 0.4 m (typical of an area of industrial or residential
22 Nov 2002 11:42 AR AR159-FM35-19.tex AR159-FM35-19.sgm LaTeX2e(2002/01/18) P1: IBC

FLOW IN URBAN AREAS 477

buildings measuring 4–5 m in height), wind-speed perturbations of as much as


20% occur at heights of 3 m. Slower wind speeds were found over the rough sur-
face, as expected. At heights of 10 m or above the wind-speed perturbation was
calculated to be less than 1%, implying that the blending height or the top of the
internal boundary layer was at a height of ∼10 m, or 2Hr, for this combination of
roughness lengths and other conditions. In this study the roughness was imposed
as a boundary condition rather than by physical obstacles. However, the study
does emphasize that the internal boundary layer can be limited to lie within the
roughness sublayer and that a z0 for the combined surfaces may be appropriate.
Thus there is some evidence for the pragmatic approach of determining a z0 and a
Annu. Rev. Fluid Mech. 2003.35:469-496. Downloaded from arjournals.annualreviews.org

d for various areas of a city or for the entire city. Very few operational transport and
dispersion models applied to the city scale allow inputs of space-varying z0. Instead,
it may be more robust to simply estimate an average or representative z0 over an
area. For example, an average z0 can be assigned to 30◦ wind sectors and to distance
by University of Bologna on 08/24/10. For personal use only.

increments representing at least 20% of the along-wind distance of interest.


Because the wind flow and the cloud dispersion rates are relatively insensitive
to minor (say, a factor of 2) variations in z0, it is sufficient to use a single averaged
value of z0 for situations with minor variations of z0 that are associated with land
use (e.g., forest to industrial site to suburb). However, if z0 varies by a few orders
of magnitude (e.g., dense urban area to bay to industrial park), then a weighting
scheme can be used to estimate a representative z0. Following Goode & Belcher
(1999), Hanna & Britter (2002) argue that a representative ln(z0) and ln(d ) can be
estimated by summing the distance-weighted values of ln(z0) and ln(d ) over the
area of interest.
The flow at a particular position depends on the urban surface upwind and,
consequently, on the direction of the wind and on the relative influence of different
upwind portions of that urban surface. Therefore different flow conditions may
need to be specified for the different wind sectors surrounding a point of interest.
In most cases there are no observations of wind in the urban area, but there
may be some at a nearby airport or other measurement site. A simple method is
available for using the wind speed at a height of, say, 10 m, at a nearby site to a
height at the urban area. If we assume that the wind speed over both sites is the same
at a height of 30 m, for example, then Equation 1 can be applied at both sites to
link them. If the atmospheric stabilities are neutral, Equation 2 suffices. Because
the roughness length at the airport is a few orders of magnitude less than that
at the urban area, the wind speed at a height of 10 m is greater at the airport than
at the urban area. However, because of the increased drag over the urban area, the
friction velocity and the turbulence components are larger.

2.2. Momentum: Urban Turbulence


Because the rate of dispersion of a pollutant cloud is proportional to the turbu-
lent velocity components (σ u, σ v, σ w), investigators have expressed much inter-
est in observations of these turbulent velocity components over urban, suburban,
22 Nov 2002 11:42 AR AR159-FM35-19.tex AR159-FM35-19.sgm LaTeX2e(2002/01/18) P1: IBC

478 BRITTER ¥ HANNA

commercial, and industrial surfaces. The results of an extensive measurement pro-


gram in the St. Louis, Missouri, area, as reported by Clarke et al. (1978) show
that during the night the turbulence over the residential and commercial surfaces
is approximately twice that over the rural surface. During the day the difference is
less, approximately 20% or 30%. These diurnal differences are expected because
at night the roughness obstacles not only generate more turbulence, but also force
the atmosphere to be less stable. In addition at night human activities add heat to
the atmosphere.
Rotach (1995) and Roth (2000) have measured and reviewed, respectively, the
turbulence above urban areas. The turbulent velocity components, scaled on a u∗
Annu. Rev. Fluid Mech. 2003.35:469-496. Downloaded from arjournals.annualreviews.org

calculated from the local height-dependent Reynolds stress, were approximately


constant. Hanna & Britter (2002) confirm that this is consistent with σ u = 2.4u∗ ,
σ v = 1.9u∗ , and σ w = 1.3u∗ , where u∗ is based on the surface shear stress, usu-
ally assumed for the atmospheric boundary layer at heights above approximately
by University of Bologna on 08/24/10. For personal use only.

2Hr. At lower heights measured in the roughness sublayer the turbulent velocity
components decrease somewhat as the ground surface is approached.

2.3. Urban Surface-Energy Balance


Figure 2 can be used to understand the surface-energy balance in an urban area. This
surface (surface layer, actually) is warmed by the net radiation flux, Q∗ , which is the
sum of incoming short-wave solar radiation and the difference between incoming
and outgoing long-wave radiation. The surface layer is cooled by the loss of heat
due to the sensible heat flux, Hs, that is positive upwards during the day, and a latent
heat flux, He, that is positive upwards when water is evaporating from the surface.
These three fluxes may be defined at a height of approximately 2Hr (i.e., at the top
of the roughness sublayer). Oke (1987) and Grimmond & Oke (2002) define a
heat-storage term, 1Qs, that describes the net change in heat within the surface
layer. This change in heat applies to the layer consisting of the buildings and the
air around them as well as to the ground surface down to a depth of approximately
1 or 2 m where there is insignificant diurnal change in temperature.
A hysteresis in the heat-storage term, 1Qs, has been found in urban areas. Large
positive values are noted in the morning. Negative values begin in early afternoon,
and then 1Qs is equal to the net radiation flux, Q∗ , from late evening through the
rest of the night. The latter equality leads to the conclusion that the sensible heat
flux, Hs, equals zero; therefore, neutral (adiabatic) conditions prevail for most of
the night in urban areas.
A major result of the urban surface-energy balance is that urban boundary layers
do not stabilize and cool down at night, which contributes to formation of the urban
heat island. Sometimes the urban area is 10◦ C warmer than the surrounding area,
although a more common magnitude of the urban heat island is 2◦ or 3◦ C. This
warm urban core may cause a thermal circulation with winds tending to flow
inward at the surface, forming a convergence zone. The heat island perturbations
can be seen as far as 10 to 20 km outside of the city.
22 Nov 2002 11:42 AR AR159-FM35-19.tex AR159-FM35-19.sgm LaTeX2e(2002/01/18) P1: IBC

FLOW IN URBAN AREAS 479

As the air flows out of the urban area and over the surrounding rural area, the
boundary layer at night will restabilize (cool) at the surface. Consequently, there
may be a warm urban plume aloft (at heights of 100 to 200 m) over downwind
rural surfaces. This urban plume may contain pollutants and is measurable up to
100 km downwind.

2.4. Urban Moisture Effects


In the absence of irrigation, the urban area tends to be drier than its surroundings
because of the pavement and buildings. This could lead to smaller latent heat
Annu. Rev. Fluid Mech. 2003.35:469-496. Downloaded from arjournals.annualreviews.org

flux, He, and a larger sensible heat flux, Hs. Oke (1987) discussed how cities
in some arid environments (especially in the United States) can be moister than
their surroundings because of the use of irrigation in urban parks, lawns, and
other vegetation. Consequently, the latent heat flux can be larger than the sensible
by University of Bologna on 08/24/10. For personal use only.

heat flux in irrigated cities. Changnon et al. (1977) summarized the results of the
St. Louis Metromex field study, where the excess heat flux from urban areas was
shown to combine with the convergence zone associated with an urban heat island
to form updrafts. These upward vertical motions can lead to condensation and
cloud formation, sometimes with precipitation. Cloud physicists and climatologists
claim to have found significant increases in clouds and precipitation over and
just downwind of urban areas. In some cases, the urban-enhanced clouds and
precipitation do not occur until 10 or 20 km downwind of the urban area.

2.5. Urban Stability


Several physical effects have been described above that all conspire to force the
stability over urban areas toward neutral (adiabatic) conditions. For example, the
fundamental stability parameter in Monin-Obukhov similarity theory is z/L, which
is proportional to the sensible heat flux, Hs, and inversely proportional to the cube
of the friction velocity, u∗ . Because Hs is not overly enhanced over urban areas, and
u ∗ can be much larger (because of the increased drag force due to the roughness
obstacles) over urban areas, z/L is forced closer to zero by the dominant u3∗ effect
in the denominator of z/L. At night, the urban storage term causes the heat flux
to be close to zero; hence, nearly neutral stability is assured. This assumption is
made in several urban dispersion models.

2.6. Urban Dispersion Characteristics


At the regional scale, the urban plume is observed to extend downwind of urban
areas. A complex mix of pollutants is present, and there are likely to be chemical
reactions and gas-to-particle conversions. The sides of the urban plume generally
grow at a rate of approximately 0.5 m/s, and the maximum vertical extent of the
plume is the daytime mixing depth, usually at a height of 500–1000 m. In the
urban plume at night, the near-surface layer may be relatively pollutant free and
stable over downwind rural surfaces, whereas the air aloft (heights of 100 m or
22 Nov 2002 11:42 AR AR159-FM35-19.tex AR159-FM35-19.sgm LaTeX2e(2002/01/18) P1: IBC

480 BRITTER ¥ HANNA

more) may be well mixed and polluted. Observations by satellites, aircraft, and
surface monitors have shown that the urban plume can sometimes be detected
several hundred kilometers downwind of the urban area and may have a width of
100 or 200 km (White et al. 1983).
At the city scale, where it is assumed that a pollutant plume extends vertically
over a layer of depth at least 2Hr, there is no need to account for specific effects
around individual buildings. Consequently, dispersion can be calculated with stan-
dard approaches that apply for general boundary layers. Larger surface roughness
produces larger z0, u∗ and turbulence levels, and these lead to greater dilution of
a plume and reduced concentrations downwind (Hanna et al. 1982, Roberts et al.
Annu. Rev. Fluid Mech. 2003.35:469-496. Downloaded from arjournals.annualreviews.org

1994).
There is a hierarchy in complexity for mathematical models for treating disper-
sion in urban areas on the city scale. The simplest (and most commonly used in
operational modeling) are of the Gaussian plume/puff type with empirical (urban
by University of Bologna on 08/24/10. For personal use only.

category) correlations for the growth with distance of the plume dimensions. Less
empiricism is apparent when the plume growth rates are determined from univeral
functions that are based on the turbulence levels (and these can be related to u ∗ dis-
cussed earlier in this section) and estimates of the Lagrangian integral timescales
(Hanna et al. 1982).
Lagrangian stochastic dispersion models have been developed for urban situa-
tions. Rotach & de Haan (1997) reported on improved model performance when
a more detailed description of the flow in the roughness sublayer is incorporated.
Computational fluid dynamics models (i.e., those including turbulence closure
models) abound. Many mesoscale flow models include a dispersion extension that
may be attained through an Eulerian modeling approach or a Lagrangian stochastic
approach, and these models or the modeling approach have been extended to the
city scale (Schatzmann et al. 1997).
However, there has been a lack of comprehensive experimental data for dis-
persion in urban areas, particularly from low-level sources (apart from routine air
quality monitoring), and this has restricted the evaluation and development of
dispersion modeling. Very recently several urban dispersion experiments have
been undertaken. An experiment in Birmingham, United Kingdom, is discussed in
Section 3; results from more comprehensive experiments in San Diego, California;
Los Angeles, California; and Salt Lake City, Utah are awaited. A follow-up to the
Salt Lake City experiment is planned for Oklahoma City in 2003. The results from
these experiments will likely lead to a reassessment of dispersion modeling in
urban areas.

3. THE NEIGHBORHOOD SCALE


The neighborhood scale is a spatial scale of 1–2 km; a spatial scale at which a gross
parameterization of the flow can be attempted and also a scale at which detailed
computational study is feasible although, currently, at some expense. It is also a
scale over which some statistical homogeneity may be anticipated; the city then is
22 Nov 2002 11:42 AR AR159-FM35-19.tex AR159-FM35-19.sgm LaTeX2e(2002/01/18) P1: IBC

FLOW IN URBAN AREAS 481

seen as being composed of these neighborhoods. Dispersion studies on this scale


will likely require a more refined knowledge of the flow within and immediately
above the urban canopy. This is of particular relevance when considering the
consequences and risks associated with the accidental or deliberate release of
hazardous materials within cities.
There are two often-studied aspects of the flow within and near the urban canopy
when viewed on the neighborhood scale. First, the flow is assumed to have a long
fetch over a statistically homogeneous surface, and some quasi-equilibrium flow
has been established. Second, the flow is assumed to have developed as a result of
a change from one to another region (Smits & Wood 1985). Of course, real urban
Annu. Rev. Fluid Mech. 2003.35:469-496. Downloaded from arjournals.annualreviews.org

morphology will not provide such convenient distinctions, and some pragmatism
is required when addressing real problems.
by University of Bologna on 08/24/10. For personal use only.

3.1. Flow Characteristics


Observations from field and laboratory experiments have been used to develop an
understanding of flow characteristics. In wind tunnels and water flumes, laboratory
experiments allow for detailed flow measurement, the use of idealized urban areas
consisting of simple geometric shapes in ordered arrays, as well as the modeling of
real urban areas. Field measurements are far less common, difficult and expensive
to perform, and provide limited data. They are, however, essential in providing
data to ensure that the laboratory modeling has a sound basis.
The flow over and through urban areas is characterized in Section 2; isolated,
wake interference, and skimming are used as descriptive terms. Real urban areas
often have large variations in building heights. Hall et al. (1996) noted the impor-
tance of height variability in inhibiting the skimming-flow regime. This suggests
a fundamental difference between flows in an urban canopy and those in a typical
vegetative canopy (Finnigan 2000).
Raupach et al. (1980) and Rotach (1995) have highlighted the existence of
a roughness sublayer within the atmospheric boundary layer below the inertial
sublayer. The roughness sublayer is thought of as a region in which the underlying
buildings lead to a spatial horizontal inhomogeneity of the flow. Raupach et al.
(1980) provide a criterion based on obstacle height and spacing, whereas Rotach
(1995), concerned particularly with cities, uses building height alone. For typical
urban areas these approaches are similar. These, and other similar comments, have
commonly been interpreted as a restriction that the roughness sublayer extends
to approximately twice the average building height. The mean velocity profile in
Figure 3 is, by implication, a horizontally spatially averaged profile. The profile
well above the urban canopy is of a conventional logarithmic form. However,
closer to the buildings, individual velocity profiles deviate as they respond directly
to the real surface that produces the spatial inhomogeneity (see the wind tunnel
simulation of a section of Nantes, France, by Kastner-Klein et al. 2000).
The “twice the average building height” criterion may be too demanding when
considering the variation of the horizontally spatially averaged profile from the
22 Nov 2002 11:42 AR AR159-FM35-19.tex AR159-FM35-19.sgm LaTeX2e(2002/01/18) P1: IBC

482 BRITTER ¥ HANNA

logarithmic velocity profile, noted above, extrapolated downward. Hanna & Britter
(2002) suggest that this profile will approximate the logarithmic form down to 1.5
times the average building height or even lower. Here “approximate” means that
the error in estimating an advection wind speed for a dispersion model would be of
no practical consequence. At the same time velocity measurements at these heights
should not be used for profile fitting to obtain gross surface parameters such as u∗ ,
z0 and d (Snyder 1981).

3.2. Shear-Stress Profiles


Annu. Rev. Fluid Mech. 2003.35:469-496. Downloaded from arjournals.annualreviews.org

At positions away from a roughness change and for constant height obstacles
the maximum shear stress occurs at approximately the height of the obstacles
(Macdonald et al. 2000, Cheng & Castro 2002). Below this height the shear
stress carried by the fluid decreases to zero as the buildings take up part of the
by University of Bologna on 08/24/10. For personal use only.

stress through the drag forces on them. This is consistent with studies of vegeta-
tive canopies (Kaimal & Finnigan 1994). The shear stress approaches zero at the
underlying surface for small λp or λf , although it approaches zero at elevation for
large λp or λf. Above the obstacle height the shear stress decreases with height with
some limited evidence of a constant shear-stress region. Finite fetch experiments
make conclusions concerning a constant shear-stress region difficult.
A change of roughness (from small to large roughness, say) viewed at the neigh-
borhood scale appears as a high velocity flow impinging on an array of obstacles.
This produces a large drag force on the most upstream obstacles (producing a
large surface shear stress) and a divergence of the flow as it is turned vertically
up and laterally out of the array. Further downstream the velocity within the array
decreases to a level in some quasi-equilibrium with that above the urban canopy.
This vertical deflection near the upper edge of the roughness change produces
an elevated maximum shear stress quite distinct from that discussed above. A
change of roughness from large to small produces a decline in the elevation of the
maximum shear stress.
Field measurements by Louka (1998) with an array of four canyon-type build-
ings placed the maximum shear stress close to the height of the buildings. In a
wind tunnel study of an array of street canyons Brown et al. (2000) found similar
results but noted that the maximum shear stress was larger and more confined in
the first few rows; this shear stress then became smaller and more diffuse fur-
ther downstream, a behavior consistent with such a roughness change. Rafailidis
(1997) had observed much the same behavior in a similar study but with various
roof geometries.
Field (Rotach 1993a, 1993b, 1995; Oikawa & Meng 1995; Feigenwinter et al.
1999) and laboratory measurements with a model of a 400 m diameter section
of central Nantes, France, (Kastner-Klein & Rotach 2001) in which the build-
ing heights vary produce somewhat different results. A maximum is evident in
the shear stress although this occurs well above the average building height. For
varied-height obstacles the maximum shear stress should occur at approximately
the height of the highest obstacle, decreasing to zero shear stress in much the
22 Nov 2002 11:42 AR AR159-FM35-19.tex AR159-FM35-19.sgm LaTeX2e(2002/01/18) P1: IBC

FLOW IN URBAN AREAS 483

same manner as for constant-height obstacles. This statement relies on the disper-
sive stresses, those due to the inhomogeneity of the mean flow, being negligible,
as demonstrated by Cheng & Castro (2002). The maximum shear stress should
equate to the surface shear stress and determine the surface friction velocity. As
a consequence the maximum shear stress occurs (well) above the mean height of
the buildings.

3.3. Mean-Velocity Profiles


Some disagreement exists in the literature over the form of the (spatially averaged)
Annu. Rev. Fluid Mech. 2003.35:469-496. Downloaded from arjournals.annualreviews.org

mean-velocity profile over an urban canopy or very rough surface. There is agree-
ment that above the surface roughness layer the velocity profile has a conventional
logarithmic form based on u∗ , z0 and d (for the neutrally stratified boundary layer).
Below this level Raupach et al. (1980) showed clearly that the velocity is increased
by University of Bologna on 08/24/10. For personal use only.

above that expected by extrapolation toward the surface of the logarithmic pro-
file. However, other velocity profiles (Macdonald 2000) show the mean velocity
to be decreased (below that given by the extrapolated logarithmic form) in the
surface roughness layer at positions above the mean building height and increased
at positions below the mean building height.
This apparent inconsistency has arisen because of the difficulty of specifying
u∗ , z0, and d from experimentally determined velocity profiles. All three cannot be
determined with any accuracy from curve-fitting. The determination of u∗ from di-
rect measurement of the Reynolds stress is often made. In the work by Macdonald
(2000) a further constraint was added. The fitted logarithmic velocity profile had
to contain the same volume flux as the measured profile, an approach that requires
there be regions of velocity excess and deficit (compared to the extrapolated log-
arithmic form). This problem observed with laboratory data will be even more
apparent with field measurements.
There is general agreement that the wind speeds are more uniform with height
below rather than above the average building height, except for positions very
close to the underlying surface where the velocity must decrease to zero. Cionco
(1965) developed a model for the velocity profile within a vegetative canopy of
constant height; this model has been extended to the urban canopy. Using a simple
mixing-length turbulence model he analyzed the flow through a regular array of
obstacles of constant cross-section and constant height. The wind velocity u(z),
scaled on the velocity at the (constant) building height, had an exponential profile:
u/u Hr = exp [−a(1 − z/Hr )]. (5)

The profile requires specification of an empirical constant a. Macdonald (2000)


observed exponential profiles in laboratory experiments with λf between 0.05 and
0.20. He was also able to link the constant a to λf directly with a = 9.6λf . This
approach is simple and describes the laboratory experiments well; however, it does
rely on knowledge of the wind speed at the building height (the height at which
the velocity profile is changing most rapidly), which may not be easily defined.
The approach might easily be extended to a real urban canopy with a statistical
22 Nov 2002 11:42 AR AR159-FM35-19.tex AR159-FM35-19.sgm LaTeX2e(2002/01/18) P1: IBC

484 BRITTER ¥ HANNA

description of the buildings, thereby reducing the velocity gradients in the vicinity
of the average building.
An alternative, even simpler, approach is to define a spatially and temporally
averaged characteristic velocity within the urban canopy. Bentham & Britter (2002)
showed, with some reasonable assumptions, that this can be related to u∗ , λf , and
the average drag coefficient for the buildings CDB with
−1/2
Uc /u ∗ = (CDB /2)−1/2 λ f . (6)

Comparison of this result with laboratory data produced good agreement with
Annu. Rev. Fluid Mech. 2003.35:469-496. Downloaded from arjournals.annualreviews.org

an assumed drag coefficient of unity. This approach is simple and direct, as is the
exponential profile approach above, but for both it is unlikely that λf is the only
parameter required to describe the urban canopy. Laboratory data show that flat
plates produce a larger surface stress than cubes with the same λf , that cubes in
by University of Bologna on 08/24/10. For personal use only.

a staggered array produce a larger surface stress than cubes in an aligned array
with the same λf and λp, and that variable-height elements produce a larger sur-
face stress than cubes with the same λf and λp (Macdonald et al. 2000, Cheng &
Castro 2002).
Macdonald (2000), using the exponential profile, and Hanna & Britter (2002),
using the characteristic velocity approach, connected the in-canopy profile with
the velocity profile above the surface roughness layer by direct extrapolation of
the velocity profile above the surface roughness layer down through that layer.
Both regard the deviations of the velocity profile within the surface roughness
layer from the downward extrapolated profile as not being essential for simple
modeling purposes.
For the flow above the roughness elements the change of surface roughness
is interpreted as the growth of an internal boundary layer. At lower levels near
the roughness elements there are large alterations to the flow field at the upwind
edge of the roughness change that then relax back to a nearly unchanging, fully
developed velocity field downwind. Jerram et al. (1994), using laboratory and
field data with cubes, showed that a distance corresponding to approximately five
obstacle heights was necessary to approach a fully developed flow. Macdonald
et al. (2000), also using cubes and with λf = 0.16, found that the velocity field
near to the roughness elements showed little further development beyond the first
or second row of obstacles, a distance into the array of only two to four obstacle
heights. The flow within the obstacle arrays adjusted more rapidly than the flow
well above the array. On the basis of these limited experiments it is concluded
that an approximately fully developed velocity field within the array will have
been attained within a distance of approximately five obstacle heights into the
array.

3.4. Turbulence Intensities


Rotach (1995), Roth (2000), and Macdonald et al. (2000) showed that above the
average building height the local turbulence intensities scale with the square root
22 Nov 2002 11:42 AR AR159-FM35-19.tex AR159-FM35-19.sgm LaTeX2e(2002/01/18) P1: IBC

FLOW IN URBAN AREAS 485

of the local kinematic Reynolds stress. The local Reynolds stress is not the same
as the surface stress from which u∗ is determined; however, they are comparable
in the region near the top of the roughness elements or building heights. Below
the average building height this scaling is less appropriate; the Reynolds stress
can go to zero above the underlying surface, whereas the turbulence intensities do
not. Scaling on the surface shear stress and surface friction velocity may be more
appropriate.
Only limited data is available: W.H. Snyder (unpublished), Macdonald et al.
(2000), and Kastner-Klein et al. (2000) all provide results from laboratory exper-
iments. The turbulence intensities vary (decrease) slowly with height inside the
Annu. Rev. Fluid Mech. 2003.35:469-496. Downloaded from arjournals.annualreviews.org

canopy, and spatially averaged (up to Hr) turbulence intensities may be sensibly
defined. Snyder’s results at a λf of 0.027 and Macdonald’s results at a λf of 0.0625
and 0.16 are similar and average out to approximately σ u/u∗ = 1.6, σ v/u∗ = 1.4,
and σ w/u∗ = 1.1, respectively. Kastner-Klein et al. (2000) used a model of a cen-
by University of Bologna on 08/24/10. For personal use only.

tral part of Nantes, France, with a large λp of 0.6–0.7. Her results in Figure 4
show roughly uniform turbulence intensities within the canopy, and they can be
approximated as σ u/u∗ = 1.0, σ v/u∗ = 1.0, σ w/u∗ = 0.8.
Broadly speaking, these results suggest that the turbulence levels may be as-
sumed to be approximately uniform throughout the canopy, to scale on u∗ , and
to be less than that above the canopy. The turbulence levels, when scaled on u∗ ,
appear to decrease with increasing λf. If this slight decrease was ignored, then σ u,
σ v, σ w/(characteristic canopy velocity Uc) must vary as (λf/2)+1/2. This ratio of
turbulence levels to advective velocity is an important variable in the dispersion
of pollutants. The turbulence levels within the canopy may be estimated using a
production-dissipation balance argument. The conclusions depend on the choice
of length scale and averaging procedures; however, it can be argued that the above
ratios should depend on (λf/2)+1/2 in the isolated obstacle regime, (λf/2)+1/3 in

Figure 4 Measured vertical profiles of the turbulent velocity components (σ u, σ v,


σ w) at various positions within a wind-tunnel study in Nantes, France. Printed with
permission from P. Kastner-Klein.
22 Nov 2002 11:42 AR AR159-FM35-19.tex AR159-FM35-19.sgm LaTeX2e(2002/01/18) P1: IBC

486 BRITTER ¥ HANNA

the wake-interference regime, and (λf/2)0, i.e., a constant for the skimming-flow
regime. This simple analysis is consistent with the observed decrease in σ u, σ v,
σ w/u∗ with increasing λf. There will be an additional contribution to turbulent
dispersion arising from the spatial variation of the mean velocity in the horizontal
plane as a result of the buildings. Thus “effective” levels of turbulence within the
canopy are comparable to those above the canopy.
The discussion above has assumed that it is a flow through the canopy that
generates the turbulence. This view is less appropriate for cities in which λf or λp
are large. The turbulence may be more attributable to the interaction of the high-
speed flow near the top of the urban canopy with the building tops and subsequent
Annu. Rev. Fluid Mech. 2003.35:469-496. Downloaded from arjournals.annualreviews.org

advection of this turbulence into the canopy. Similarly, turbulence may be generated
by the shear layer atop a street canyon for a skimming flow and advected into the
canyon.
Another approach to studying the flow in urban areas is to consider exchange
by University of Bologna on 08/24/10. For personal use only.

processes. The surface shear stress is a measure of the momentum exchange from
the flow to the surface. The surface stress nondimensionalized with the fluid density
and a reference mean wind speed (such as the geostrophic wind or the wind speed
at a convenient reference height) becomes a surface-drag coefficient for meteorol-
ogists or a skin-friction coefficient for engineers. The two disciplines have slightly
different definitions; engineers conventionally include a factor of one half in the
denominator that is often omitted by the meteorologists. Note that the surface-
drag coefficient (and the reference height for which the nondimensionalizing wind
speed is measured) can be directly related to z0 and d. The aerodynamic conduc-
tance reported by Grimmond & Oke (1999) is an exchange velocity that transfers
momentum at the reference level to the surface where the velocity is zero. Using
the meteorologists definition of CD researchers note that the surface conductance
is equal to CDu (z = zref) and the ratio of the surface conductance to the wind
speed at the reference level is just CD. Grimmond & Oke (1999) estimate CD as
0.008 for residential and warehouse sites, 0.016 for a central site in Mexico city,
and between 0.03 and 0.05 for a downtown site in Vancouver.
This approach can be usefully extended to consider an exchange velocity be-
tween the in-canopy and the above-canopy flow, that is, an exchange velocity
relevant for momentum transfer into and out of the canopy. Here we view the
momentum exchange as being in two stages: an exchange between the reference
level and the in-canopy flow and an exchange between the in-canopy flow and
the surface (including the building surfaces). On this basis and choosing a refe-
rence level of 2.5Hr the exchange velocity can be calculated exactly, and it is
typically between 0.2 and 0.3u∗ for a wide range of scenarios.
This exchange velocity, based on momentum transfer, is also the exchange ve-
locity for ventilating the canopy and for the convective exchange of heat, moisture,
and pollutant between the canopy and the flow above. The drag coefficient CD is
not the same as the equivalent coefficients for heat or moisture because the transfer
processes from the flow to the surfaces are quite different for momentum and heat
or moisture. For the same reason z0 calculated for temperature or moisture profiles
22 Nov 2002 11:42 AR AR159-FM35-19.tex AR159-FM35-19.sgm LaTeX2e(2002/01/18) P1: IBC

FLOW IN URBAN AREAS 487

can be several orders of magnitude different from the z0 for the velocity profile
(Voogt & Grimmond 2000).

3.5. Dispersion Characteristics


It was shown in Section 2 that an increase in surface roughness produces a sig-
nificant increase in turbulence levels. These cause a reduction in ground-level
concentrations that arise from ground-level releases for situations where the pol-
lutant cloud depth is much larger than the height of the surface obstacles. It is less
clear whether the same conclusion can be drawn when the pollutant cloud is of a
Annu. Rev. Fluid Mech. 2003.35:469-496. Downloaded from arjournals.annualreviews.org

depth comparable to or smaller than the height of the surface obstacles.


Laboratory experiments (Davidson et al. 1995; Macdonald et al. 1997, 1998)
showed that a conventional Gaussian plume model provides an appropriate struc-
ture for the problem. The increased turbulence levels within the urban canopy (over
by University of Bologna on 08/24/10. For personal use only.

those occurring in the absence of obstacles) produce larger dispersion coefficients


that tend to reduce concentrations. However, the accompanying reduction in the
advection velocity within the canopy tends to increase concentrations. The rela-
tive magnitudes of these opposing effects determine whether the obstacles lead to
increased or decreased concentrations as the roughness is increased.
The Kit Fox field experiments (Hanna & Chang 2001) showed clear evidence of
substantial reductions in ground-level concentrations from a ground-level source
for conditions in which the plume centroid was comparable to or smaller than the
obstacle heights. The reductions were a factor of approximately 3 as z0 changed
from 0.002 m to 0.02 m and were a further factor of approximately 3 as z0 changed
from 0.02 m to 0.2 m. These experiments used a dense gas, carbon dioxide, as the
pollutant; however, the preceding statement is based only on experiments at the
highest wind velocities when the dense gas effects would be minimal.
Other small-scale field experiments (Davidson et al. 1995, Macdonald et al.
1997) and related laboratory experiments (Davidson et al. 1996, Hall et al. 1996)
found that the obstacle arrays produced little effect on the maximum ground-level
concentrations.
The results noted above for the advection velocity and the turbulence levels
allow for an estimate of the effect of increasing surface roughness on the ground-
level concentrations that arise from a ground-level source within the urban canopy.
If we use the turbulence levels as surrogates for the plume growth rates and the
characteristic in-canopy velocity as the advection velocity, it is straightforward
to show that the ground-level concentration should be proportional to (λf/2)−1/2
in the isolated-building regime, to (λf/2)−1/6 in the wake-interfering regime, and
to (λf/2)+1/2 in the skimming-flow regime. Thus, increasing surface roughness
initially tends to decrease the ground-level concentration that will go through a
minimum and then increase, possibly returning to become equal or larger than that
for an unobstructed case.
Dispersion-field experiments from low-level sources in real cities are rare. The
St. Louis experiments (McElroy & Pooler 1968) were conducted with low-level
22 Nov 2002 11:42 AR AR159-FM35-19.tex AR159-FM35-19.sgm LaTeX2e(2002/01/18) P1: IBC

488 BRITTER ¥ HANNA


Annu. Rev. Fluid Mech. 2003.35:469-496. Downloaded from arjournals.annualreviews.org
by University of Bologna on 08/24/10. For personal use only.

Figure 5 The concentration-time history at a receptor ∼1 km downwind of a short-


duration tracer release in central Birmingham, United Kingdom. Two tracers, PMCH
and PMCP, were used.

sources. A study in Birmingham, United Kingdom, (Britter et al. 2002) was only
able to run three experiments owing to resource constraints. The recent and planned
dispersion-field studies referred to in Section 2 are intended to provide data on the
neighborhood scale in particular.
Figure 5 is a concentration-time history for a receptor in central Birmingham
∼1 km away from the source. The release was a 20-min finite-duration release.
The concentration history shows a time delay consistent with an estimated advec-
tion velocity and a rapid rise in concentration. There is also some evidence for
a plateau as the release behaves like a continuous plume and then a slow decay.
A time constant of ∼4 min can be determined from the early part of the decay;
this is an obvious result of pollutant trapped in recirculating regions among the
buildings. It should be possible to relate the time constant to the morphological
characteristics of the surface even though this has not been done. In addition there
appears to be another, much longer, time constant as the concentration approaches
the background level. This may be due to pollutant that is taken into buildings and
released on a building ventilation timescale.
Model development at the neighborhood scale has proceeded along several
diverse lines. Jerram et al. (1994) developed a linearized perturbation model with
a force distribution representing the drag on the buildings. Theurer (1995) based
a Gaussian plume model on extensive wind-tunnel data; this work allowed for
the plume direction to be influenced by the urban layout. Kaplan & Dinar (1996)
22 Nov 2002 11:42 AR AR159-FM35-19.tex AR159-FM35-19.sgm LaTeX2e(2002/01/18) P1: IBC

FLOW IN URBAN AREAS 489

derived a mass-consistent wind model accounting for flow recirculation behind


buildings and combined this with a Lagrangian particle model for dispersion.
The extensive laboratory experiments by Hall et al. (1996) have been used to
develop a simple urban dispersion model (Hall et al. 1997) based principally on
data correlations. Similarly the arguments in Hanna & Britter (2002) have been
recast into another simple urban dispersion model.
Computational techniques of much greater complexity including large eddy
simulation [called LES only recently (Boris 2002)] have been used. It is only
now that data that allow formal scientific evaluation of such models are becoming
available. At present there has been only limited effort (e.g., Soulhac et al. 2002)
Annu. Rev. Fluid Mech. 2003.35:469-496. Downloaded from arjournals.annualreviews.org

to use advanced computational techniques to develop a general understanding of


the flow and to give guidance as to how best to view and parameterize the flow as
distinct from a direct application to specific scenarios.
by University of Bologna on 08/24/10. For personal use only.

4. THE STREET SCALE


The street (canyon) scale is particularly studied in the context of urban air quality
where the dominant source of pollution, vehicle emissions, are in close proximity to
the pollutant receptors of concern, people (and the regulatory pollutant-monitoring
stations). In cities both the source and the receptor are within a very confined geo-
metry, the street canyon, and this confined geometry exhibits a sheltering effect
from the diluting influence of the wind. Flow and dispersion near street intersec-
tions are also of interest as it is the acceleration of vehicles away from traffic lights
and/or pedestrian crossings that gives rise to large pollutant emission rates and
consequently poor urban air quality. The location of building ventilation intakes
is sensitive to the anticipated distribution of pollutants on the spatial scale of the
buildings or the street.

4.1. Flow and Dispersion


The archetypal street-canyon flow is basically that of a turbulent shear flow above
a rectangular cavity with the mean flow direction perpendicular to the axis of the
street canyon. For a canyon of near unity aspect ratio, and provided that the roof-
level wind speed exceeds 1.5–2 m/s, a recirculating flow is set up in the canyon
driven by momentum transport from the shear layer above. The flow at the bottom
of the canyon is then upwind toward the lee wall of the canyon. Observations in
the laboratory (see Hosker 1985) and the field (Nakamura & Oke 1988) support
this structure. Typically the recirculating velocities relative to the roof-level wind
speed are of the order of 1/3–1/2, and the turbulence levels are of the order of
1/10. The recirculating flow is neither steady nor symmetric. The asymmetric flow
appears as a stronger, more concentrated downflow on the downwind wall and as
a weaker, more extensive upflow on the lee wall.
There are also variations on this archetypal flow as the geometry and flow
directions vary. For example, regular rectangular street canyons with a small height
22 Nov 2002 11:42 AR AR159-FM35-19.tex AR159-FM35-19.sgm LaTeX2e(2002/01/18) P1: IBC

490 BRITTER ¥ HANNA

to width ratio allow for possible reattachment of the separating shear layer off the
lee wall to the floor of the street canyon. For large aspect ratios there can be a
counter-rotating vortex below the main recirculation flow. Similarly and even for
aspect ratios near unity there can be a counter-rotating vortex in the corner between
the lee wall and the floor. However, these latter possibilities are more common in
idealized laboratory experiments rather than in field experiments. Irregular street
canyons with the lee wall and the downwind wall having different elevations
introduce further complications to the flow, but the resulting flows are generally
qualitatively consistent with expectations (Pavageau et al. 2001). When the wind
flow direction is not perpendicular to the street axis, a recirculating flow is still
Annu. Rev. Fluid Mech. 2003.35:469-496. Downloaded from arjournals.annualreviews.org

present as is an alongstreet flow. The flow in the street canyon can also depend on
the characteristics of the flow above, e.g., whether it is a fully developed rough wall
boundary layer, what turbulence scales are present, or whether the street canyon
is one of several canyons in parallel. These considerations might reflect an overly
by University of Bologna on 08/24/10. For personal use only.

detailed view of a complex, essentially inhomogeneous problem and might be


based on a view of the problem that is too idealized to ever be relevant to real
situations.
Recent work has been done on how thermal effects will influence the flow. Can
street canyon walls heated by the sun influence the mechanically driven recircu-
lating motion? Although laboratory flows and computational studies do show an
effect, this seems unlikely to be operationally important in most scenarios. Any
effect is far less evident in field measurements (Louka et al. 2002) probably be-
cause, in the field, the physical width of the free convective boundary layer on the
heated wall is small compared with the scale of the mechanically driven motion,
which is on the street scale. Of course when the external wind is very weak these
thermal effects are of greater consequence.
Thus the basic flow in a street canyon is a recirculating flow filling the canyon
unless the canyon has a large aspect ratio, in which case the recirculating flow may
not reach to the floor. Another basic aspect of the flow is an exchange of air between
the street canyon and the flow above. This might be estimated assuming that a free
shear layer separates the street canyon from the flow above (Soulhac et al. 2002;
F. Caton, personal communication). Surprisingly few experiments, in the labora-
tory or in the field, direct attention to the exchange mechanism, to exchange fluxes,
or to determining an exchange velocity (the velocity that is ventilating the street
canyon) (but see Louka et al. 2000, Robins et al. 2001).
These two aspects are the principal phenomena that require inclusion in op-
erational models of street-scale pollution from vehicles (Yamartino & Wiegand
1986, Berkowicz 2000). Laboratory experiments can be used to estimate the mag-
nitude of adjustable parameters; however, tuning and testing of the models is more
commonly done with field experiments by using roadside monitoring stations and
extensive traffic data (Lohmeyer et al. 2002).
Tracer dispersion measurements in wind tunnels (e.g., Hoydysh & Dabberdt
1988, Dabberdt & Hoydysh 1991) are qualitatively consistent with expectations.
A low-level pollutant source produces a vertical concentration profile that is
22 Nov 2002 11:42 AR AR159-FM35-19.tex AR159-FM35-19.sgm LaTeX2e(2002/01/18) P1: IBC

FLOW IN URBAN AREAS 491

approximately exponential. For wind flows perpendicular to the street axis the
lee-side concentration is larger than the windward side by a factor of ∼2. This is
true except for a step-down configuration when the windward side may have the
larger concentration. Maximum concentrations occur for wind directions perpen-
dicular and parallel to the street axis with a shallow minimum between. Studies
similar to those for street canyons have also been undertaken for more complex
intersections where pollutant emission rates are likely to be relatively high (Robins
et al. 2001).
Street-scale problems are amenable to computational fluid mechanics, and many
such studies may be found in the literature. They typically produce reasonable
Annu. Rev. Fluid Mech. 2003.35:469-496. Downloaded from arjournals.annualreviews.org

qualitative results, but the performance, when compared with laboratory or field
experiments, is little better than the simple operational models described above. Of
course the solution using computational fluid dynamics allows for the treatment
of more complicated geometrical arrangements.
by University of Bologna on 08/24/10. For personal use only.

4.2. Traffic-Produced Turbulence


Poor urban air quality is associated with low wind-speed conditions; however,
operational urban air quality models commonly perform poorly under these con-
ditions. Typically the models are structured so as to have an inverse relationship
between the concentration field and an externally imposed wind speed, and this re-
lationship leads to substantial, and critical, concentration overestimation of models
at low wind speeds. Other sources of turbulence are thermal production from the
environment or from vehicles, as well as mechanical production from the motion
of vehicles, also known as traffic produced turbulence (TPT).
Laboratory measurements of the turbulence produced by modeled traffic in a
street canyon (Kastner-Klein 1999) showed that TPT can be a significant source
of turbulence when compared with that arising from the mean wind. There is an
asymmetry of the distributions that reflect the vortex-like structure within the street
canyon owing to the external wind. This asymmetry pushes the added turbulence
toward the lee wall of the canyon.
Vachon et al. (2002) described generally the same influences of traffic motions
on the turbulence field in an urban street canyon. The measurements were done on
a street in Nantes, France. In the lower part of the street canyon increased levels of
turbulent kinetic energy, which could be attributed to turbulence created by vehicle
motions, have been found. Close to the traffic region turbulence enhancement has
been observed on the leeward and windward side of the street canyon. However, on
the leeward side the influence has been more pronounced, and the vertical extent
of the region with increased turbulence levels has been much larger than on the
windward side. This indicates once again that the advection of turbulence created
in the traffic layer toward the leeward wall is due to the wind-induced street-canyon
vortex.
The magnitude of, or at least the appropriate scaling for, the TPT may be esti-
mated from a production-dissipation balance for turbulent kinetic energy.
22 Nov 2002 11:42 AR AR159-FM35-19.tex AR159-FM35-19.sgm LaTeX2e(2002/01/18) P1: IBC

492 BRITTER ¥ HANNA

DiSabatino et al. (2002) considered three regimes: isolated vehicles, wake-inter-


fering vehicles, and congested traffic (analogous to the isolated, wake-interference,
and skimming-flow regimes for buildings) to produce estimates of TPT. The tur-
bulence levels depend on the traffic density to the 1/2, 1/3, and zero power as the
traffic density increases.
When the laboratory results by Kastner-Klein (1999) were scaled with the (ap-
propriate) wake-interfering correlation, the leading coefficient was found to be
close to unity. In Kastner-Klein et al. (2002) these results were tested in full-scale
observations in Gottingerstrasse, Hanover, Germany, where street-side measure-
ments of NO2 were available over several years. The measurements were short-
Annu. Rev. Fluid Mech. 2003.35:469-496. Downloaded from arjournals.annualreviews.org

term, that is, half-hour, averaged. In contrast to the predictions in Figure 1 of


annual mean concentrations other regulations are noted in terms of the statistics of
short-term-averaged measurements. Extreme events are likely to occur when the
ambient wind speed is low, and models that ignore TPT will significantly overpre-
by University of Bologna on 08/24/10. For personal use only.

dict the magnitude of these events. The incorporation of TPT into an operational
model led to a marked improvement in the model performance, in particular for
the extreme events.

5. PROBLEMS AND PROGRESS

The literature addressing flow and dispersion in urban areas is spread over many
disciplines (meteorology, engineering, geography, and others), with each having a
fundamental and an operational aspect. In preparing this review it became apparent
that there was no clear coherent framework within which the study of the urban
area was taking place, which reflects the various disciplines participating, their
different goals, and the complexity and essential heterogeneity of the problem.
Attention to the development of a common, accepted framework would likely be
rewarded.
The sensitivity of operational modeling procedures to the uncertainty in the
input variables or, possibly, the conceptual underpinning of the procedures is often
unclear. This may be due to the lack of data, particularly in the field; however,
there is currently much progress in this area, and more formalized model evalua-
tion procedures are being developed. These will also use the extensive laboratory
databases that have been recently constructed by D.J. Hall, R.W. Macdonald,
M. Schatzmann, and others.
In a fluid-mechanical context the most pressing problems include the treatment
of atmospheric stability in urban areas, the specification of reference variables
(e.g., the wind speed well above the urban area, the wind speed just above or
at the average building height, or the wind speed within the urban canopy), and
the treatment of arbitrary spatial variations in surface roughness. For dispersion
studies it is still unclear how best to address the neighborhood scale and its con-
nections with the street and the city scale, particularly when addressing transient
problems.
22 Nov 2002 11:42 AR AR159-FM35-19.tex AR159-FM35-19.sgm LaTeX2e(2002/01/18) P1: IBC

FLOW IN URBAN AREAS 493

ACKNOWLEDGMENTS
The authors thank M.K. Neophytou and S. DiSabatino for assistance in the prepa-
ration of this manuscript. Figure 1 was kindly provided by CERC Ltd., Cambridge,
United Kingdom.

The Annual Review of Fluid Mechanics is online at http://fluid.annualreviews.org

LITERATURE CITED
Annu. Rev. Fluid Mech. 2003.35:469-496. Downloaded from arjournals.annualreviews.org

Auer AH. 1978. Correlation of land use and ISWS B-62, Illinois State Water Survey.
cover with meteorological anomalies. J. 260 pp.
Appl. Meteorol. 17:636–43 Cheng H, Castro IP. 2001. Near-wall flow devel-
Bentham T, Britter RE. 2002. Spatially aver- opment after a step change in surface rough-
aged flow velocity within large groups of ob- ness. Bound.-Layer Meteorol. 6:1–21
by University of Bologna on 08/24/10. For personal use only.

stacles. Atmos. Environ. Submitted Cheng H, Castro IP. 2002. Near-wall flow over
Berkowicz R. 2000. OSPM—a parameterised urban-like roughness. Bound.-Layer Meteo-
street pollution model. J. Environ. Monit. As- rol. In press
sess. 65:323–31 Cionco RM. 1965. Mathematical model for air
Boris J. 2002. The threat of chemical and biolo- flow in a vegetative canopy. J. Appl. Meteo-
gical terrorism. Comput. Sci. Eng. Mar/Apr: rol. 4:517–22
22–32 Clarke JF, Ching JKS, Binkowski FS, Godo-
Bornstein R. 1987. Mean diurnal circulation witch JM. 1978. Turbulent structure of the
and thermodynamic evolution of urban boun- urban surface boundary layer. Proc. Int.
dary layers. In Modeling the Urban Bound- Tech. Meet. Air Pollut. Model. Its Appl., 9th,
ary Layer. Boston, MA: Am. Meteorol. Soc. NATO/CCMS. New York/London: Plenum
Bottema M. 1997. Urban roughness modeling Cook N. 1990. The Designers Guide to Wind
in relation to pollutant dispersal. Atmos. En- Loading of Building Structures Part 2—
viron. 31:3059–75 Static Structures. London: Butterworth Publ.
Britter RE, Disabatino S, Caton F, Cooke K, Dabberdt WF, Hoydysh WG. 1991. Street
Simmonds P, Nickless G. 2002. Results from canyon dispersion: sensitivity to block shape
three field tracer experiments at the neighbor- and entrainment. Atmos. Environ. 25A:1143–
hood scale in the city of Birmingham, UK. 53
Water Air Soil Pollut.-Focus. In press Davenport AG, Grimmond CSB, Oke TR,
Brown MJ. 2000. Urban parameterizations for Weiranga J. 2000. Estimating the roughness
mesoscale meteorological models. In Mesos- of cities and scattered country. Conf. Appl.
cale Atmospheric Dispersion, ed. Z. Boy- Climatol., 12th, Asheville, NC, 2000, pp. 96–
beyi, Chapter 5, pp. 193–255. Southhampton, 99. Boston, MA: Am. Meteorol. Soc.
UK: WIT Press Davidson MJ, Mylne KR, Jones CD, Phillips
Brown MJ, Lawson RE, Descroix DS, Lee RL. JC, Perkins RJ, et al. 1995. Plume disper-
2000. Mean flow and turbulence measure- sion through large groups of obstacles—a
ments around a-D array of buildings in a field investigation. Atmos. Environ. 29:3245–
wind tunnel. Conf. Appl. Air Pollut. Meteo- 56
rol., 11th, Long Beach, Calif., 2000. Boston, Davidson MJ, Snyder WH, Lawson RE, Hunt
MA: Am. Meteorol. Soc. JCR. 1996. Plume dispersion from point
Changnon SA, Huff FA, Schickedanz PT, Vo- sources upwind of groups of obstacles—
gel JL. 1977. Summary of METROMEX wind tunnel simulations. Atmos. Environ. 30:
Vol 1: weather anomalies and impacts. 3715–25
22 Nov 2002 11:42 AR AR159-FM35-19.tex AR159-FM35-19.sgm LaTeX2e(2002/01/18) P1: IBC

494 BRITTER ¥ HANNA

DiSabatino S, Kastner-Klein P, Berkowicz R, field data to analyze dense gas dispersion


Britter R, Fedorovich E. 2002. The modeling modeling issues. Atmos. Environ. 35:2231–
of turbulence from traffic in urban dispersion 41
models—Part I: theoretical considerations. J. Hanna SR, Ramsdell J, Cramer H. 1987. Urban
Environ. Fluid Mech. In press Gaussian diffusion parameters. Modeling the
Feigenwinter C, Vogt R, Parlow E. 1999. Ver- Urban Boundary Layer. Boston, MA: Am.
tical structure of selected turbulence charac- Meteorol. Soc.
teristics above an urban canopy. Theor. Appl. Hosker RP. 1985. Flow around isolated struc-
Climatol. 62:51–63 tures and building clusters: a review.
Fenger J, Hertel O, Palmgren F. 1999. Urban ASHRAE Trans. 91:1671–92
Air Pollution—European Aspects. London: Hoydysh WG, Dabberdt WF. 1988. Kinemat-
Annu. Rev. Fluid Mech. 2003.35:469-496. Downloaded from arjournals.annualreviews.org

Elsevier ics and dispersion characteristics of flows in


Finnigan J. 2000. Turbulence in plant canopies. asymmetric street canyons. Atmos. Environ.
Annu. Rev. Fluid Mech. 32:519–71 22:2677–89
Goode K, Belcher SE. 1999. On the parame- Jerram N, Perkins RJ, Fung JCH, Davidson MJ,
by University of Bologna on 08/24/10. For personal use only.

terisation of the effective roughness length Belcher SE, Hunt JCR. 1994. Atmospheric
for momentum transfer over heterogeneous flow through groups of buildings and disper-
terrain. Bound.-Layer Meteorol. 93:133–54 sion from localized sources. Proc. NATO Adv.
Grimmond CSB, Oke TR. 1999. Aerodynamic Study Inst., Karlsruhe, ed. EJ Plate, JE Cer-
properties of urban areas derived from anal- mak. Dordrecht: Kluwer
ysis of surface form. J. Appl. Meteorol. Kaimal JC, Finnigan JJ. 1994. Atmospheric
38:1261–92 Boundary Layer Flows: Their Structure and
Grimmond CSB, Oke TR. 2002. Turbulent heat Measurement. Oxford, UK: Oxford Univ.
fluxes in urban areas: observations and a Press
local-scale urban meteorological parameter- Kaplan H, Dinar N. 1996. A Lagrangian dis-
isation scheme. J. Appl. Meteorol. In press persion model for calculating concentration
Hall DJ, Macdonald R, Walker S, Spanton AM. distribution within a built-up domain. Atmos.
1996. Measurement of dispersion within sim- Environ. 24:4197–207
ulated urban arrays—a small scale wind tun- Kastner-Klein P. 1999. Experimentelle Unter-
nel study. BRE Client Rep. 178/96, Build. suchung der strömungsmechanischen Trans-
Res. Establ., Garston, Watford, UK portvorgänge in Strassenschluchten. PhD
Hall DJ, Spanton AM, Macdonald R, Walker S. thesis. Inst. Hydromech. Univ., Karlsruhe,
1997. A simple model for estimating disper- Ger.
sion in urban areas. BRE Client Rep. 169/97, Kastner-Klein P, Ketzel M, Berkowicz R, Fe-
Build. Res. Establ., Garston, Watford, UK. dorovich E, Britter R. 2002. The modeling
91 pp. of turbulence from traffic in urban disper-
Hanna SR, Briggs GA, Hosker RP. 1982. Hand- sion models—Part II: evaluation based on
book on Atmospheric Diffusion. DOE/TIC- laboratory and full-scale concentration mea-
11223 (DE82–002045). Springfield, VA: surements in street canyons. J. Environ. Fluid
NTIS/USDOC Mech. In press
Hanna SR, Britter RE. 2002. Wind Flow and Kastner-Klein P, Rotach M. 2001. Parameteri-
Vapor Cloud Dispersion at Industrial Sites. zation of wind and turbulent shear stress pro-
New York: Am. Inst. Chem. Eng. files in the urban roughness sublayer. Proc.
Hanna SR, Chang JC. 1992. Boundary layer pa- Int. Conf. Urban Air Qual., 3rd, Loutraki,
rameterizations for applied dispersion mod- Greece, 2001. London: Inst. Phys.
eling over urban areas. Bound.-Layer Mete- Kastner-Klein P, Rotach M, Fedorovich E.
orol. 58:229–59 2000. Experimental study on mean flow
Hanna SR, Chang JC. 2001. Use of the Kit Fox and turbulence characteristics in an urban
22 Nov 2002 11:42 AR AR159-FM35-19.tex AR159-FM35-19.sgm LaTeX2e(2002/01/18) P1: IBC

FLOW IN URBAN AREAS 495

roughness sublayer. 14th Symp. Bound. Oke TR. 1987. Boundary Layer Climates. Lon-
Layer Turbul., Aspen, CO, 2000. Boston, don, UK: Routledge
MA: Am. Meteorol. Soc. Pavageau M, Rafailidis S, Schatzmann M.
Lohmeyer A, Mueller WJ, Baechlin W. 2002. A 2001. A comprehensive experimental data-
comparison of street canyon predictions by bank for the verification of urban car emis-
different modelers: final results now avail- sion dispersion models. Int. J. Environ. Pol-
able from the Podbi-exercise. Atmos. Envi- lut. 15:417–25
ron. 36:157–58 Rafailidis S. 1997. Influence of building, areal
Louka P. 1998. Measurements of airflow in an density and roof shape on the wind character-
urban environment. PhD thesis. Univ. Read- istics above a town. Bound.-Layer Meteorol.
ing, UK 85:255–71
Annu. Rev. Fluid Mech. 2003.35:469-496. Downloaded from arjournals.annualreviews.org

Louka P, Belcher SE, Harrison RG. 2000. Cou- Ratti CS, Disabatino S, Britter R, Brown M,
pling between airflow in streets and the well- Caton F, Burian S. 2002. Analysis of urban
developed boundary layer aloft. Atmos. Env- databases with respect to pollutant disper-
iron. 34:2613–21 sion for a number of European and American
by University of Bologna on 08/24/10. For personal use only.

Louka P, Vachon G, Sini JF, Mestayer PG, cities. Water Air Soil Pollut.-Focus. In press
Rosant JM. 2002. Thermal effects on the air- Raupach MR, Thom AS, Edwards I. 1980. A
flow in a street canyon—Nantes ’99 experi- wind-tunnel study of turbulent flow close
mental results and model simulation. Water to regularly arrayed rough surfaces. Bound.-
Air Soil Pollut.-Focus. In press Layer Meteorol. 18:373–97
Macdonald RW. 2000. Modelling the mean Roberts PT, Fryer-Taylor RE, Hall DJ. 1994.
velocity profile in the urban canopy layer. Wind-tunnel studies of roughness effects on
Bound.-Layer Meteorol. 97:25–45 gas dispersion. Atmos. Environ. 28:1861–
Macdonald RW, Carter S, Slawson PR. 2000. 70
Measurements of mean velocity and tur- Robins AG, Scaperdas A, Grigoriadis D. 2001.
bulence statistics in simple obstacle arrays Spatial variability and source-receptor rela-
at 1:200 scale. Therm. Fluids Rep. 2000-1, tions at a street intersection. 2001. Water Air
Therm. Dep. Mech. Eng., Univ. Waterloo, Soil Pollut.-Focus. In press
Can. Rotach MW. 1993a. Turbulence close to a
Macdonald RW, Griffiths RF, Cheah SC. 1997. rough urban surface. Part I: Reynolds stress.
Field experiments of dispersion through reg- Bound.-Layer Meteorol. 65:1–28
ular arrays of cubic structures. Atmos. Envi- Rotach MW. 1993b. Turbulence close to a rough
ron. 31:783–95 urban surface. Part II: variances and gradi-
Macdonald R, Griffiths RF, Hall DJ. 1998. A ents. Bound.-Layer Meteorol. 66:75–92
comparison of results from scaled field and Rotach M. 1995. Profiles of turbulence statistics
wind tunnel modelling of dispersion in ar- in and above an urban street canyon. Atmos.
rays of obstacles. Atmos. Environ. 32:3845– Environ. 29:1473–86
62 Rotach M, de Haan P. 1997. On the urban aspect
McElroy JL, Pooler F. 1968. St. Louis Disper- of the Copenhagen dataset. Int. J. Environ.
sion Study Volume II—Analysis. Arlington, Pollut. 8, 3–6:279–86
VA: US Dep. HEW Roth M. 2000. Review of atmospheric turbu-
Nakamura Y, Oke TR. 1988. Wind, temperature lence over cities. Q. J. R. Meteorol. Soc.
and stability conditions in an E-W oriented 126:941–90
canyon. Atmos. Environ. 22:2691–700 Schatzmann M, Rafailidias S, Britter R, Arend
Oikawa S, Meng Y. 1995. Turbulence charac- M. 1997. Database, Monitoring and Mod-
teristics and organized motion in a subur- elling of Urban Air Pollution: Inventory
ban roughness layer. Bound.-Layer Meteorol. of Models and Datasets. Luxembourg: Eur.
74:289–312 Comm. 109 pp.
22 Nov 2002 11:42 AR AR159-FM35-19.tex AR159-FM35-19.sgm LaTeX2e(2002/01/18) P1: IBC

496 BRITTER ¥ HANNA

Smits AJ, Wood DH. 1985. The response of ments for urban air pollution modelling. At-
turbulent boundary layers to sudden per- mos. Environ. 33:4057–66
turbation. Annu. Rev. Fluid Mech. 17:321– Vachon G, Louka P, Rosant J-M, Mestayer
58 P, Sini J-F. 2002. Measurements of traffic-
Snyder WH. 1981. Guidelines for fluid mod- induced turbulence within a street canyon
eling of atmospheric diffusion. Rep. EPA- during the Nantes 1999 experiment. Water
600/8-81-0009, EPA, Res., Triangle Park, Air Soil Pollut.-Focus. In press
NC Voogt JA, Grimmond CSB. 2000. Modeling
Soulhac L, Mejean P, Perkins RJ. 2002. Mod- surface sensible heat flux using surface ra-
eling transport and dispersion in a street- diative temperatures in a simple urban area.
canyon. Int. J. Environ. Pollut. In press J. Appl. Meteorol. 39:1679–99
Annu. Rev. Fluid Mech. 2003.35:469-496. Downloaded from arjournals.annualreviews.org

Stull RB. 1997. An Introduction to Boundary White WH, Patterson DR, Wilson WE. 1983.
Layer Meteorology. Dordrecht: Kluwer At- Urban export to the nonurban troposphere:
mos. Publ. 670 pp. results from Project MISTT. J. Geophys. Res.
Theurer W. 1995. Point sources in urban areas: 88:10745–52
by University of Bologna on 08/24/10. For personal use only.

modeling of neutral gas clouds with semi- Yamartino RJ, Wiegand G. 1986. Development
empirical models. Wind Climate in Cities, of simple models for the flow, turbulence and
pp. 485–502. Dordrecht: Kluwer pollutant concentration fields within an urban
Theurer W. 1999. Typical building arrange- street canyon. Atmos. Environ. 20:2137–56
29 Nov 2002 13:42 AR AR159-19-COLOR.tex AR159-19-COLOR.SGM LaTeX2e(2002/01/18) P1: GDL

Annu. Rev. Fluid Mech. 2003.35:469-496. Downloaded from arjournals.annualreviews.org


by University of Bologna on 08/24/10. For personal use only.

Figure 1 Model predictions of annual mean NO2 concentrations


for 2005 in Greater London.
P1: FDS
November 22, 2002 11:16 Annual Reviews AR159-FM

Annual Review of Fluid Mechanics


Volume 35, 2003

CONTENTS
STANLEY CORRSIN: 1920–1986, John L. Lumley and Stephen H. Davis 1
Annu. Rev. Fluid Mech. 2003.35:469-496. Downloaded from arjournals.annualreviews.org

AIRCRAFT ICING, Tuncer Cebeci and Fassi Kafyeke 11


WATER-WAVE IMPACT ON WALLS, D. H. Peregrine 23
MECHANISMS ON TRANSVERSE MOTIONS IN TURBULENT WALL
FLOWS, G. E. Karniadakis and Kwing-So Choi 45
by University of Bologna on 08/24/10. For personal use only.

INSTABILITIES IN FLUIDIZED BEDS, Sankaran Sundaresan 63


AERODYNAMICS OF SMALL VEHICLES, Thomas J. Mueller
and James D. DeLaurier 89
MATERIAL INSTABILITY IN COMPLEX FLUIDS, J. D. Goddard 113
MIXING EFFICIENCY IN STRATIFIED SHEAR FLOWS, W. R. Peltier
and C. P. Caulfield 135
THE FLOW OF HUMAN CROWDS, Roger L. Hughes 169
PARTICLE-TURBULENCE INTERACTIONS IN ATMOSPHERIC CLOUDS,
Raymond A. Shaw 183
LOW-DIMENSIONAL MODELING AND NUMERICAL SIMULATION
OF TRANSITION IN SIMPLE SHEAR FLOWS, Dietmar Rempfer 229
RAPID GRANULAR FLOWS, Isaac Goldhirsch 267
BIFURCATING AND BLOOMING JETS, W. C. Reynolds, D. E. Parekh,
P. J. D. Juvet, and M. J. D. Lee 295
TEXTBOOK MULTIGRID EFFICIENCY FOR FLUID SIMULATIONS,
James L. Thomas, Boris Diskin, and Achi Brandt 317
LEVEL SET METHODS FOR FLUID INTERFACES, J. A. Sethian
and Peter Smereka 341
SMALL-SCALE HYDRODYNAMICS IN LAKES, Alfred Wüest
and Andreas Lorke 373
STABILITY AND TRANSITION OF THREE-DIMENSIONAL BOUNDARY
LAYERS, William S. Saric, Helen L. Reed, Edward B. White 413
SHELL MODELS OF ENERGY CASCADE IN TURBULENCE, Luca Biferale 441
FLOW AND DISPERSION IN URBAN AREAS, R. E. Britter and S. R. Hanna 469

vii
P1: FDS
November 22, 2002 11:16 Annual Reviews AR159-FM

viii CONTENTS

INDEXES
Subject Index 497
Cumulative Index of Contributing Authors, Volumes 1–35 521
Cumulative Index of Chapter Titles, Volumes 1–35 528

ERRATA
An online log of corrections to Annual Review of Fluid Mechanics chapters
may be found at http://fluid.annualreviews.org/errata.shtml
Annu. Rev. Fluid Mech. 2003.35:469-496. Downloaded from arjournals.annualreviews.org
by University of Bologna on 08/24/10. For personal use only.

Você também pode gostar