You are on page 1of 10

Turbulent Diffusion

A Venkatram, University of California – Riverside, Riverside, CA, USA


S Du, California Air Resources Board, Sacramento, CA, USA
Ó 2015 Elsevier Ltd. All rights reserved.
This article is reproduced from the previous edition, volume 6, pp 2455–2466, Ó 2003, Elsevier Ltd.

Glossary
1 AU Astronomical Unit ¼ the mean Earth–Sun separation. SZA Solar zenith angle: the angle between the local vertical
Dobson Unit Unit for total column ozone and the Sun (SZA ¼ 90  solar elevation).
(1 DU ¼ 2.69  1016 molecule cm2). A column amount SPF Sun protection factor (for sunscreens).
of 300 Du, which is a typical global average that TOMS Total Ozone Mapping Spectrometer (a satellite-
corresponds to a 3-mm layer of pure ozone at standard borne ozone sensor).
temperature and pressure.
RAF Radiative amplification factor.

Nomenclature
~ Instantaneous concentration.
C t Travel time from source to receptor.
C Mean concentration. ~ Instantaneous wind speed.
U
c Concentration fluctuation defined as C ~  C. U Average wind speed.
Cpeak Peak concentration. u Wind speed fluctuation defined as U ~  U.
y
C Crosswind-integrated concentration. u* Surface friction velocity.
D Molecular diffusivity of species in air. ws Gravitational settling velocity of a heavy particle.
hs Source height. d Deviation of time-averaged mean about the ensemble
H0 Surface heat flux. mean.
K Eddy diffusivity. 3 Rate of dissipation of turbulent kinetic energy.

L Monin–Obukhov length. r Density of air.


l Length scale characterizing turbulent fluctuations. sc Standard deviation of concentration fluctuations.
N Brunt–Vaisala frequency. sv Standard deviation of crosswind velocity component.
Q Source strength for continuous emissions. sw Standard deviation of vertical velocity component.
Qm Total mass of a puff. sy Lateral dimension (or spread) for Gaussian plume
RLv Lagrangian autocorrelation for crosswind velocity. model.
s Puff dimension (or spread). sz Vertical dimension (or spread) for Gaussian plume
Tav Averaging interval. model.
Tc Timescale of concentration fluctuations. s Surface shear stress.
TLv Lagrangian timescale for crosswind velocity. sLeff Effective Lagrangian timescale of a heavy particle.
TLw Lagrangian timescale for vertical velocity. sparticle Relaxation timescale of a heavy particle.

Introduction Most of our understanding of turbulent dispersion is couched


in terms of semiempirical models that have been developed by
Turbulent diffusion or, more correctly, turbulent dispersion fitting tentative theories to observations. For example, models for
addresses the problem of estimating the concentration field in dispersion in the surface boundary layer are based on tracer
a turbulent flow. Our ignorance of the details of the turbulent experiments conducted in Prairie Grass, Nebraska, USA, in the
velocity field translates into uncertainty in estimating the 1950s. The development of models for dispersion in the
details of the concentration field. Thus, as in other problems convective boundary layer (CBL) has been guided by laboratory
related to turbulence, the goal of turbulent dispersion is limited experiments conducted by Willis and Deardorff in the 1970s.
to understanding the ensemble-averaged statistics of the Thus, to a large extent, our understanding of turbulent dispersion
concentration field and the associated fluxes that govern the consists of a patchwork of semiempirical models, each of which
field. describes a limited set of observations. The primary goal of this

Encyclopedia of Atmospheric Sciences 2nd Edition, Volume 6 http://dx.doi.org/10.1016/B978-0-12-382225-3.00441-2 277


278 Turbulence and Mixing j Turbulent Diffusion

article is to provide the reader an understanding of these semi- ground is taken to be the reference height, with the x-axis of
empirical models. It is only over the past 10 years that attempts the coordinate system aligned along the wind direction at the
have been made to develop methods that can be applied to a large source, empirical evidence indicates that the time-averaged
class of problems. These methods will be discussed in the last part (typically 1 h) concentration field can be described in terms
of this article. of the Gaussian distribution by eqn [5], where y is the cross-
This article assumes that the reader is familiar with the wind coordinate, Q is the source strength (mass/time), hs is the
fundamentals of turbulence in the atmospheric boundary layer. height of the source above ground, U is the time-averaged wind
To keep this article to a manageable size, we will not deal with speed at source height, and sy and sz are the plume spreads
several important topics including the effects of turbulence on corresponding to the Gaussian distribution.
chemical reactions. We begin with the statement of the problem " #
that the techniques of turbulent dispersion attempt to solve. Q ðz  hs Þ2 y2
Cðx; y; zÞ ¼ exp   2 [5]
2psy sz U 2s2z 2sy

The Problem of Turbulent Dispersion The effect of the ground on concentrations is accounted for
by making sure that there is no flux of material through the
The evolution of the concentration field of a species is governed plane at z ¼ 0. The mathematical trick to achieve this is to
by the mass conservation equation, eqn [1], where the place an ‘image’ source at a distance z ¼ hs; the upward flux
‘squiggle’ overbar refers to the instantaneous field, and D is the from this image source essentially cancels out the downward
molecular diffusivity of the chemical species in the fluid. flux from the real source. Then, the concentration can be
described by eqn [6].
~
vC v  ~ ~ ~
v2 C
þ Ui C ¼ D [1] " #
vt vxi vxi vxi Q y2
Cðx; y; zÞ ¼ exp  2
Because the instantaneous velocity field is unknown, we 2psy sz U 2sy
8 " # " #9 [6]
have to limit ourselves to estimating the ensemble-averaged < ðz  hs Þ2 ðz þ hs Þ2 =
mean and the associated statistics of the concentration field.  exp  þ exp 
: 2s2z 2s2z ;
To define an ensemble, we express the instantaneous
velocity field as the sum of a known velocity field, Ui, and the
unknown deviation, ui, as in eqn [2]. In the real atmosphere, dispersion in the upward direction
is limited by the height of the atmospheric boundary layer. This
~ i ¼ Ui þ u i
U [2] limitation of vertical mixing is incorporated into the Gaussian
Then, an ensemble is defined as the infinite set of possible formulation by reflecting off the top of the mixed layer. When
concentrations corresponding to a given Ui. Note that the the material is ‘reflected’ from both the ground as well as the
definition of an ensemble is arbitrary because it depends on top of the mixed layer, it is necessary to account for the infinite
what we know about the velocity field. Each member of this set of ‘reflections’ from the two surfaces. This can be readily
ensemble corresponds to the component of the unknown accounted for in the Gaussian formulation.
velocity field, ui; we assume that we know something about the The point source solution is the kernel for the integral that is
statistics of this unknown field. The concentration associated used to estimate dispersion from a variety of sources, including
with each member of the ensemble can be written as eqn [3], line and volume sources. Most dispersion models that apply to
where C is the average obtained over all possible concentration spatial scales of tens of kilometers are based on eqns [5] and
fields in the ensemble. [6]. The plume spreads or s values are empirically derived from
observations. It turns out that we continue to use plume-spread
~ ¼ Cþc
C [3] formulations first recommended by Pasquill in the 1960s.
Substituting eqn [3] into eqn [1] and averaging over the These formulations, largely based on the Prairie Grass experi-
ensemble yield eqn [4], where the overbar refers to the ment, relate plume spread to surface meteorological measure-
ensemble average, the subscripted index refers to coordinate ments such as wind speed and cloud cover. The advances in
direction, and repeated indices imply summation. This equa- micrometeorology during the 1970s provided the incentive to
tion can be solved only by modeling the turbulent flux term, develop dispersion models that relied on theoretical under-
ui c, using known properties of the velocity field. Before we take standing of dispersion.
up models for this term, let us consider the problem in The preceding expression for the concentration field is
turbulent dispersion that has to a large extent motivated the essentially an empirical description of observations. Equation
development of the field. [5] is a formal solution to the point source problem only when
the turbulence field is homogeneous, and the velocity distri-
vC v   v2 C bution is Gaussian. Under these circumstances, the concentra-
þ Ui C þ ui c ¼ D [4]
vt vxi vxi vxi tion field can be analyzed using a statistical approach, first
proposed by Taylor.
The Point Source in the Atmospheric Boundary Layer
Statistical Analysis of Dispersion
The classical problem of turbulent dispersion in the atmo-
spheric boundary is that of a continuous source emitting Consider a source located at ‘s’ emitting particles continuously
material at some height above the ground (see Figure 1). If the into a turbulent flow. If the mean flow and turbulence are
Turbulence and Mixing j Turbulent Diffusion 279

Plume

z y

Concentration
distribution
is assumed to
hs be Gaussian in
z x and y directions

Figure 1 Turbulent dispersion from a point source.

steady, the ensemble-averaged concentration at ‘r,’ C(r) can be describe the distribution of particle positions within a ‘puff, ’
shown to be which usually describes an entity at an instant of time.
ZN To make progress, we need expressions for sx , sy , and sz.
While the mean particle positions are determined by the mean
CðrÞ ¼ Q pðrjs; tÞdt [7]
flow, the standard deviations depend on the characteristics of
0
the turbulent flow. Taylor derived expressions for the variance
where Q is the mass emission rate of particles, and p(rjs,t)dV is of particle positions as a function of travel time from a fixed
the probability that a particle released at ‘s’ will be found in release point in a steady flow in which turbulent statistics do
a volume dV surrounding ‘r’ after a travel time t from release. not depend on location. His expressions for the asymptotic
Then, the problem of calculating the concentration reduces to behavior of plume spread are
estimating the probability density function, p(rjs,t), of particle
sy ¼ sv t for t  TLv
positions as a function of travel time from the source. A good [9]
sy ¼ sv ð2tTLv Þ1=2 for t[TLv
approximation for this function is the Gaussian distribution.
Placing the origin of our coordinate system at the source, we where TLv is the Lagrangian timescale, which can formally defined
can express the distribution as in terms of the statistics of the turbulent flow. For our purposes, it
1 is sufficient to interpret the timescale as roughly the time over
pðrjs; tÞ ¼ 3=2 which a particle retains its initial velocity. For small travel times,
ð2pÞ sx sy sz
!! a particle’s velocity remains essentially unchanged from its value
1 ðxr  xÞ2 ðyr  yÞ2 ðzr  zÞ2 at the release point, and the particle trajectory is a straight line.
 exp  þ þ
2 s2x s2y s2z This explains the result that, for small travel times, the spread of
particles is proportional to the travel time from the source (eqn
[8]
[9]). On the other hand, when the travel time is large compared
where sx, sy, and sz are the standard deviations of particle to the Lagrangian timescale, the plume spread is proportional to
positions about their mean positions x, y, and z after a travel the product of the ‘average’ step size, sv TLv , and the square root of
time, t, from release, and xr, yr, and zr are the receptor coordi- the number of steps, t/TLv, taken by the particle.
nates. These statistics are derived by averaging over an infinite In order to obtain an expression for the concentration, we
number of particles for a flow with fixed mean and turbulent still have to integrate eqn [7] after inserting eqn [8] with
characteristics; the statistics are functions of travel time, t. appropriate expressions for plume spread. Let us first consider
The probability distribution function, eqn [8], represents an an idealized flow that is used to model dispersion in the
ensemble average over all possible particle positions for a fixed atmospheric boundary layer. In this flow, the mean wind U is
travel time from the source; the travel time is the difference along the x-axis, and the turbulence is homogeneous and
between the arbitrary time at which the particle is released and stationary. These assumptions lead to
time at which the particle is observed at a location. In principle,
x ¼ Ut; y ¼ 0; and z ¼ 0 [10]
it can be constructed experimentally by releasing particles
serially from a source, and recording the coordinates of these If we make the assumption that along-wind dispersion, sx,
particles at specified travel times. Thus, eqn [8] does not is small compared to transport by the mean wind, the
280 Turbulence and Mixing j Turbulent Diffusion

exponential term in eqn [8], associated with downwind conservation equation. Let us examine this approach in some
dispersion, becomes a Dirac delta function in the limit of sx detail because it yields useful results for dispersion in the
going to zero. This allows us to integrate eqn [7] for arbitrary sy surface boundary layer.
and sz to obtain
!!
Q 1 y 2 z2
CðrÞ ¼ exp  þ [11] Solving the Species Conservation Equation
2psy sz U 2 s2y s2z
The species conservation equation (eqn [4]) can be rewritten as
where the plume spreads are evaluated at t ¼ xr/U. This equa-
eqn [14].
tion is identical to the empirical expression presented earlier.
We can obtain an expression for the concentration even when vC v v   v2 C
þ ðUi CÞ ¼ ui c þ D [14]
downwind dispersion is not small if we can express plume vt vxi vxi vxi vxi
spreads in terms of the asymptotic limits of eqn [9].
One way of modeling the turbulent flux term is to postulate
Observations of plume spread from elevated releases are
the concept of eddy diffusivity. It is based on an analogy with
often summarized in the form
molecular diffusion, in which the flux of material in any
sv t direction is proportional to the gradient of the concentration.
sy ¼ [12]
ð1 þ t=2TLv Þ1=2 For example, the turbulent flux of species is according to eqn
[15], where Ki is the so-called eddy diffusivity.
to ensure consistency with theory of eqn [9]. For ground-level
releases, there is no simple way of relating travel time to vC
turbulent flux ¼ ui c ¼ Ki [15]
distance because the velocity varies rapidly with height near the vxi
ground. Under these circumstances, we need to use methods
where the bar over i negates the summation convention. This
that are discussed next.
relationship cannot be justified rigorously for turbulent trans-
port. However, it has heuristic value, and is useful for devel-
Dispersion in an Inhomogeneous Boundary Layer oping semiempirical models of turbulent transport. The use of
the eddy diffusivity in the mass conservation equation is often
The theory presented thus far applies to a boundary layer in referred to as K-theory.
which the mean and turbulent properties are constant in space With eqn [15], eqn [14] can be rewritten in the form of eqn
and time. To apply it to a real boundary layer in which the [16],
properties are highly inhomogeneous, we can use one of two  
vC v v vC
approaches. The first is to average the turbulence and mean þ ðUi CÞ ¼ Ki [16]
vt vxi vxi vxi
properties over the region of interest, and use the average
properties in the (homogeneous) formulations discussed where we have neglected the molecular diffusion term in
earlier. This is the most straightforward approach, except that comparison to turbulent diffusion. While molecular diffusion
the averaging procedure is necessarily arbitrary. The validity of can often be ignored in calculating the ensemble mean, it plays
the method needs to be established by comparing the results a major role in determining the statistics of concentration
obtained from the formulations with observations or theory fluctuations, as we will see later.
that accounts for inhomogeneity more explicitly. In general, One way of checking whether the use of the eddy diffusivity
empirical knowledge derived from observations plays a major is plausible is to see whether eqn [16] yields solutions that are
role in the development of practical models of dispersion. As in compatible with observations. We will apply eqn [16] to the
most turbulence research, theory can suggest plausible forms point source problem. If we assume that transport along the
for a dispersion model, but the model almost always contains mean wind dominates over the corresponding diffusion term,
parameters that have to be estimated from observations. and turbulent properties are homogeneous, eqn [16] can be
Even if we could treat the boundary layer as vertically reduced to the form given by eqn [17].
homogeneous, the presence of boundaries, such as the ground vC v2 C v2 C
and the top of the mixed layer, makes it difficult to estimate the U ¼ Kz 2 þ Ky 2 [17]
vx vz vy
Lagrangian timescale, TLv, from a priori considerations. Thus,
the timescale is often treated as an empirical parameter that is It turns out that eqn [17] yields the empirical Gaussian
derived by fitting eqn [12] to observations. Alternatively, we solution of eqn [5] if we ensure that the eddy diffusivity is
can postulate an expression for TLv in terms of a known length related to the plume spread according to eqn [18].
scale l as shown in eqn [13].
1 ds2i
Ki ¼ U [18]
TLv ¼ al=sv [13] 2 dxi
The parameter a has to be obtained by fitting estimates of If we take plume spread to follow the behavior described in
plume spread to observations. In unstable conditions, l, usually eqn [9], we see from eqn [18] that the eddy diffusivity is
scales with the depth of the boundary layer, while in stable proportional to the travel time, x/U, from the source, for travel
conditions, the relevant length scale is taken to be sw =N, where times less than the governing Lagrangian timescale. What this
N is the Brunt–Vaisala frequency. means is that the eddy diffusivities corresponding to two
The second approach to accounting for inhomogeneity in different sources displaced along the wind will have different
the boundary layer is based on the solution of the species values at the same location. It is only at large travel times that
Turbulence and Mixing j Turbulent Diffusion 281

eddy diffusivities become independent of travel time (eqn The eddy diffusivity formulation is almost exclusively used
[19]), where TLv and TLw are the Lagrangian timescales for the in comprehensive Eulerian air quality models that include
horizontal and vertical velocity fluctuations, respectively. details of atmospheric process such as gas- and aqueous-phase
chemistry. The main reason is that the species conservation
Ky ¼ s2v TLv and Kz ¼ s2w TLw [19] equation, formulated in terms of the eddy diffusivity, is
The eddy diffusivity, Kz, can be related to turbulent flow a convenient framework to incorporate a number of processes,
properties by appealing to ‘mixing length’ theory, which including nonlinear chemistry. The resulting mass conservation
suggests the relationship [20], where sw is the standard devia- can be solved using numerical methods. Results from compre-
tion of the vertical velocity fluctuations, and lz is the ‘length hensive air quality models indicate that modeling dispersion
scale’ of turbulence for vertical transport, defined by eqn [21], with the eddy diffusivity model has some practical value, even
which is consistent with eqn [13]. though the underlying theoretical justification is weak.
We saw earlier that the eddy diffusivity concept is likely to be
Kz ¼ sw lz [20] most applicable when the turbulent length scales are smaller than
or comparable to the concentration space scales. Thus, we might
lz ¼ sw TLw [21] expect it to apply to dispersion in the surface boundary layer,
We are now in a position to make some additional state- where plume spread in the vertical spread is comparable to the
ments on the applicability of the eddy diffusivity concept. We length scale of the eddies responsible for vertical transport. It
saw earlier that eqn [19] is valid when the travel time is much turns out that K-theory provides useful results for dispersion in
larger than the Lagrangian timescale, expressed by the rela- the surface boundary layer even though it is characterized by steep
tionship [22]. gradients in temperature and velocity. However, the gradients of
fluxes and turbulence levels are negligible. In the surface
x
t ¼ [TLw [22] boundary layer, semiempirical theories, referred to as Monin–
U
Obukhov similarity, provide useful relationships between
If we combine this condition with the expression for plume velocity and temperature gradients and the corresponding heat
spread, eqn [9], and use eqn [21], we obtain eqn [23]. and momentum fluxes. These relationships are cast in terms of
length and velocity scales, which are the surface friction velocity
sz [lz [23]
u* and the Monin–Obukhov length, L, defined by eqns [24],
Thus, the eddy diffusivity concept is most applicable when rffiffiffi
the scale of concentration variation, sz, is much larger than the s
u ¼
scale of the eddies responsible for plume spreading. r
[24]
Equation [20] is useful because we can guess at the T0 u3 rCp
appropriate form of lz, and then see whether the consequences L ¼ 
g kH0
of our assumption agree with observations. Over the years, we
have developed enough experience with different types of where s is the surface shear stress, r is the air density, Cp is the
flows to prescribe useful forms for the mixing length (or eddy specific heat of air, T0 is the surface temperature, k is the von
diffusivity) for these flows. Our initial guesses for K are usually Karman constant, g is the acceleration due to gravity, and H0 is
based on measurement of fluxes and the associated gradients the surface heat flux. These relationships can be used to derive
for a limited set of situations. This K is then extrapolated to eddy diffusivities for heat and momentum.
situations different from those used to derive it. For example, Using the eddy diffusivity for heat in the mass conservation
we can derive a K for heat flux, and find out whether it works equation has provided concentration estimates that compare
for pollutant transport. It is this type of semiempirical argu- well with observations made in field experiments conducted in
ments that form the basis of practical calculations for turbu- Prairie Grass, Nebraska, in the 1950s. We note that data from
lent flows. this experiment, conducted with relatively primitive equip-
Equation [15] represents only one possible approach to ment, are still the most complete for the analysis of surface
expressing the turbulent flux. In principle, we can write layer dispersion.
conservation equations for the turbulent fluxes, but these The solutions of the mass conservation equations, using the
equations contain ‘third-order’ terms that are essentially the eddy diffusivity, have a number of useful asymptotic forms for
y
fluxes of the second-order terms. These third-order terms have the crosswind-integrated concentration, C , as shown by eqn
y
to be parameterized using some sort of flux-gradient approxi- [25], where x ¼ x=jLj C ¼ C u jLj=Q, with Q representing the
mation. At this point, there is no compelling evidence to source strength. These asymptotic forms are useful because they
suggest that these approaches yield much better results than the can be patched together to obtain analytic expressions that
closure of eqn [15]. One way of improving upon a simple span the entire range of stability. Unknown parameters in these
prescription of the eddy diffusivity is to formulate semi- expressions have been obtained by fitting them to observations
empirical conservation equations for the components of eddy from Prairie Grass.
diffusivities: the turbulent velocity and the length scale in eqn
[20]. In practice, the turbulent velocity is related to the turbu- C  x1 for neutral conditions
 x2=3 for stable conditions [25]
lent kinetic energy, k, and the length scale is related to the
 x2 for unstable conditions
turbulent dissipation rate, 3 . While the k  3 approach is
popular in modeling turbulent flows, it has found limited These expressions for crosswind-integrated concentrations
application in modeling dispersion. can be converted to yield centerline concentrations through
282 Turbulence and Mixing j Turbulent Diffusion

eqn [26], where sy is the crosswind spread, and the crosswind The simplest puff model is the Gaussian puff model that
distribution is taken to be Gaussian. relates concentration at receptor (x, y, z) at time t due to a puff
y
! released from origin at time 0 by eqn [30], where Qm is the total
C y2 mass of the puff and s is the puff spread corresponding to the
Cðx; y; 0Þ ¼ pffiffiffiffiffiffiffiffiffiffiffi exp  2 [26]
2psy 2sy Gaussian distribution.
" #
Equation [25] can be used to derive expressions for the Qm ðx  UtÞ2 þ y2 þ z2
vertical plume spread. These expressions depend on distance Cðx; y; z; tÞ ¼ exp  [30]
ð2pÞ3=2 s3 2s2
from the release location because travel time has little meaning
near the ground. In fact, most dispersion models used in In practical applications, a continuous release in a wind
practical applications express the concentration in terms of field that varies in space and time can be modeled through
a Gaussian distribution, where the plume spreads are empir- a series of puffs, each of which is allowed to follow a different
ically derived functions of source–receptor distance and trajectory. The concentration at a receptor at any given time is
micrometeorology. calculated by summing the contributions from these puffs. One
advantage of this approach is that it can deal with situations
when the mean wind is calm.
Puff Dispersion

In the preceding sections, we have discussed dispersion of Dispersion of Heavy Particles


plumes, which refers to a continuous release from a source.
Often we are interested in concentrations associated with puffs The previous discussions implicitly assumed that the material
of material that are released almost instantaneously, as in an being dispersed by turbulence has the same density as air. This
explosion. The concentration of material in the plume is assumption is clearly not valid for aerosol particles, whose
determined by the spread of the material about the center of densities are typically over 1 g cm3. Two effects influence
mass of the moving puff. The analysis of such puffs is more dispersion of such particles. One is related to the finite time
complicated than that of plumes because the spread depends required by the particle to respond to turbulent velocity fluc-
on both space and time correlations between particles in the tuations. The other is the so-called trajectory crossing effect
puff; as Taylor’s analysis indicates, plume dispersion can related to particle trajectories being different from fluid parcel
consider the motion of particles to be independent of each trajectories because of gravitational settling. Let us consider
other. These space–time correlations depend on the properties each of these effects.
of the turbulent eddies that contribute to puff spread at any The particle inertia effect is related to the difference between
instant of time. The length scale of the relevant eddies is the fluid velocity, uf, and the particle velocity, up, whose
roughly proportional to the size of the puff. Eddies smaller difference is proportional to the reaction time to turbulent
than this length scale disperse the material within the puff, velocity fluctuations. This can be expressed symbolically by eqn
while eddies larger than the puff transport the puff as a whole. [31], where the relaxation timescale of the particle is given by
When the puff size is of the order of the Kolmogorov eqn [32].
microscale, the puff spreads by molecular diffusion, which  
implies that the puff spread, s, is proportional to the square uf  up sparticle
 [31]
root of time. When the puff size is comparable to eddies in uf sturbulence
the inertial subrange, dimensional considerations suggest that
ws
the rate of puff spread is represented by eqn [27], where 3 is the sparticle  [32]
g
dissipation rate of turbulent kinetic energy.
Here, ws is the gravitational settling velocity, which is
ds
 ðεsÞ1=3 [27] a function of the size and density of the particle and the
dt
viscosity of the fluid. The turbulence timescale can be expressed
Integration yields expression [28]. as [33], where l is a measure of eddy size and sw is the associ-
s  ε1=2 t 3=2 [28] ated velocity fluctuation.
l
This rapid growth phase ends when the puff size is sturbulence  [33]
sw
comparable to the largest eddy of dimension L. Then, the puff
spread can be written as [29], where sv refers to the standard Thus, inertia effects can be neglected if the ratio of these two
deviation of the turbulent velocity fluctuations in the direction scales is small (see eqn [34]).
of the spread. ws sw
1 [34]
s  ðsv LÞ1=2 t 1=2 [29] gl
These three regimes of puff growth can, in principle, be If we take small to mean 0.1, and consider 100 mm particles
patched together to provide a continuous description of puff with settling velocities of the order of 1 m s1, and take the
spread. But this is rarely done in practice, and one usually turbulent velocity to be 1 m s1, the inertia effect is not likely to
resorts to empirical descriptions of puff spread. The actual be important for length scales greater than 1 m. However, it
concentration in a puff is usually estimated with a Gaussian could play a role in the dispersion of large particles under very
distribution about the puff center of mass. stable conditions, close to the ground.
Turbulence and Mixing j Turbulent Diffusion 283

reach the vicinity of the ground. Numerical experiments


l , ws indicate that to a useful degree of approximation, the
TLw ~ particle ~
w
g velocity of these particles can be taken to be constant at the
value at the release point.
This assumption allows one to express the crosswind-
If settling y
ws
l integrated ground-level concentration, C , in terms of the pdf
velocity ws >> w, ~
l
Leff
ws of the vertical velocity, p(wjhs), at the height of release, hs, by
eqn [37], where the vertical velocity corresponds to that
w required to bring material from the release point to the ground-
level receptor at x, given by [38].

y 2Q
C ¼ Pðwjhs Þ [37]
x

w ¼ Uhs =x [38]
Figure 2 Timescales that affect dispersion of heavy particles. The simple formulation, which can be readily modified to
account for the presence of the mixed layer, provides an
The trajectory crossing effect can be examined by excellent description of laboratory observations of dispersion
considering the extreme case when the time taken for in the CBL. Note that eqn [37] reproduces the empirical
a falling particle to traverse an eddy is much smaller than Gaussian distribution if we make the reasonable assumption
the Lagrangian timescale for vertical dispersion. Then, the that the pdf of vertical velocity fluctuations is normal, giving
effective Lagrangian timescale becomes this traversal time eqn [39], where sz is expressed by eqn [40].
because it corresponds roughly to the decorrelation time. rffiffiffiffi

The effective Lagrangian timescale and the associated eddy y 2 Q h2


C ¼ exp  s2 [39]
diffusivity for the heavy particle can now be represented by p Usz 2sz
eqn [35].
sz ¼ sw x=U [40]
l s2 l
sLeff ¼ and Keff ¼ w [35] The Gaussian formulation is not reliable in the CBL
ws ws
because the pdf is positively skewed. The associated negative
Figure 2 shows the timescales that govern dispersion of mode of the pdf leads to the descent of the plume centerline
heavy particles. In practice, the effects of particle settling are when the release is elevated, and leads to concentrations that
not important in determining plume spread because settling are about 30% higher than that predicted with the Gaussian
velocities for most particles are generally much smaller than formulation.
turbulent velocities. However, even particle velocities of the In principle, if we could simulate all the scales of turbulent
order of a few centimeters per second lead to mean downward motion, there would be no need for models of turbulent
motion of the plume, and hence increase concentrations at dispersion. We could use the Navier–Stokes equations to
ground level. This effect can be described approximately by generate an ensemble of flows, obtain the corresponding
‘tilting’ the plume toward the ground, represented by eqn concentration fields from the species conservation equation,
[36]. and average over them to obtain the ensemble average as well
ws x as the statistics of concentration fluctuations. With the rapid
hs ðwith particlesÞ ¼ hs  [36]
U increases in computing power, direct numerical simulation
Alternatively, the entire concentration profile can be moved (DNS) is becoming a reality, and we have been able to obtain
‘into’ the ground by a distance wsx/U after computing the useful information for low Reynolds number flows. However,
concentration corresponding to passive dispersion. It is simple it will be some time in the future before we will be able to use
to account for mean motion of particles in the eddy diffusivity the technique for routine applications.
formulation through the advection term, ws(vC/vz). The large eddy simulation (LES) technique avoids the
computational demands of DNS by only simulating the
energy-containing eddies. The effects of the unresolved scales of
Other Approaches to Modeling Dispersion motion are modeled using a variety of parameterizations. It is
believed that the most important features of flow are insensi-
The eddy diffusivity approach does not generally apply to tive to these parameterizations, because the subgrid scales
sources far removed from the ground. For example, it is contain a small fraction of the total energy. This assumption
difficult to justify its application to the convective atmo- has been vindicated by LES of CBLs that was pioneered by
spheric layer, where the turbulent length scales are large Deardorff in the 1970s. Velocity fields generated by LESs
compared to the spatial scales of the concentration field. The compare well with observations, and continue to provide
pdf approach can provide useful results under these information that is difficult to obtain in the field. Lamb used
circumstances. Studies show that dispersion in the CBL is the velocity fields generated by Deardorff to simulate disper-
strongly influenced by the relative longevity of convective sion in the CBL. The simulation consisted of releasing a large
downdrafts and updrafts. The majority of particles released number of particles and tracing their motion using the LES
in downdrafts travel continuously downward until they velocity field. Then, the crosswind concentration averaged over
284 Turbulence and Mixing j Turbulent Diffusion

eqn [42] is equivalent to the Langevin equation in which


a particle is subject to a linear viscous force and a random
pressure force.
The particle position is traced through eqns [43].
Dz ¼ wðt þ DtÞDt
[43]
Dx ¼ U Dt
Δz
Q In homogeneous turbulence, we can show that a and b are
U given by eqns [44],
   
Dt 2 Dt
a ¼ 1 and b ¼ [44]
TLw TLw
where the Lagrangian timescale, TLw, can be related to the
turbulent dissipation rate according to eqn [45], where C0 is
a constant.
2 s2w
y Fraction of particles TLw ¼ [45]
U Δz C = that pass through Δz
Q C0 ε
While this simple model has produced useful results, it is
Figure 3 Calculation of concentrations using Lagrangian stochastic not applicable to boundary layers with gradients in turbulence
simulation of particle trajectories. properties. New developments in LSS are best described by
recasting eqn [42] as eqn [46], where a is given by eqn [47], and
the dz is a normal random variable with zero mean and vari-
a vertical distance Dz is given by eqn [41], where f is the fraction
ance dt.
of the particles released that pass through Dz.
dw ¼ aw dt þ ðC0 εÞ1=2 dz [46]
y Q
C ¼ f [41]
U Dz C0 ε
a ¼ [47]
Figure 3 justifies this equation. 2s2w
Lamb’s simulations provided valuable insight into disper-
During the 1980s, several ad hoc formulations for the term
sion in the CBL, including the observation that the locus of the
‘a’ in eqn [46] were proposed to account for flow complexities
maximum concentration descended toward the ground. This
such as inhomogeneity and non-Gaussianity of the turbulent
behavior is related to the negative mode of the pdf of the
velocities. This unsatisfactory situation was remedied by
vertical velocity fluctuations.
Thomson in 1987 when he proposed a systematic method for
A technique that is not as computationally demanding as
constructing formulations for ‘a’ by invoking the constraint
DNS or LES is called Lagrangian stochastic simulation (LSS). It
that the model for this term should preserve a well-mixed
is attractive because it only uses the statistics of the velocity field
concentration field. This is equivalent to insisting that the
such as velocity variance, and dissipation rate of turbulent
Lagrangian pdf for fluid particles marked at the source should
kinetic energy. Because this technique is being used to examine
become identical to that of unmarked fluid particles at large
routine dispersion problems, we discuss it in some detail in the
travel times. The evolution of the pdf is governed by the
next section.
Fokker–Planck equation corresponding to eqn [46]. Using the
solution of this equation, Thomson was able to derive
formulations for the term ‘a’ that accounted for inhomoge-
Lagrangian Stochastic Models
neity and non-Gaussianity of the turbulent velocities. For
example, the modified form of eqn [46] can be represented by
In Lagrangian stochastic modeling, turbulent dispersion is
eqn [48].
modeled by tracing the motion of a large number of fluid
 
particles that are tagged at the source; these particles are treated C0 ε 1 w2 vs2w
mathematically as points. The evolution of the velocity of each dw ¼  2 w dt þ 1þ 2 dt þ ðC0 εÞ1=2 dz [48]
2sw 2 sw vz
particle depends on turbulence properties at the current loca-
tion of the particle. To illustrate the basic ideas, let us trace It turns out that ‘a’ can be expressed uniquely only for one-
a particle that is only affected by vertical velocity fluctuations. dimensional turbulence or isotropic turbulence; for two- and
Then, the vertical velocity of a parcel at time t þ Dt is related to three-dimensional turbulent flows, ‘a’ is not a unique function
the velocity at time t through eqn [42], where w0 is a random of the velocity field. This nonuniqueness problem can be se-
velocity drawn from the distribution of vertical velocity rious because for a given turbulent flow, substantially different
fluctuations. concentration fields can be obtained from two models, both of
which satisfy the well-mixed condition.
wðt þ DtÞ ¼ awðtÞ þ bw0 [42]
One way of alleviating this problem is to drive Lagrangian
Thus, the future velocity of the particle consists of stochastic models with velocity fields obtained from large eddy
a deterministic component that depends on its current simulation because large eddies that contribute to inhomoge-
velocity, and a random component that depends on the neity and anisotropy of the velocity field are treated explicitly in
turbulence at the location of the particle. We can show that both LES and LSS. The subgrid-scale eddies can be considered
Turbulence and Mixing j Turbulent Diffusion 285

locally homogeneous and locally isotropic. Because only the Tc roughly corresponds to the time taken for the instanta-
particle motion associated with subgrid energy is modeled with neous plume to pass a fixed observer. If we take the instanta-
LSS, errors associated with the nonuniqueness of the LSS model neous plume size to be 100 m and the wind speed to be
can be minimized. 5 m s1, the timescale, Tc, is 20 s. Assuming that we are inter-
ested in an averaging time of 1 h, d works to be about 25% of
the mean for a peak-to-mean ratio of 10. This exercise identifies
Concentration Fluctuations and Model Uncertainty the variables that might govern the deviation between model
estimates and observations. Actual comparisons between
Any given observed concentration will deviate from the model estimates and observations indicate that the error is
ensemble average. This deviation is caused by the intrinsic much larger because of model formulation and input errors. A
variability of the concentration field called concentration fluc- model is generally considered adequate if its estimates are
tuations. In principle, the statistics of these concentration consistently within a factor of two of the observations.
fluctuations can provide insight into the expected deviation of We can use a simple model to discuss the role of molecular
the model-predicted ensemble mean from a corresponding diffusion in determining concentration fluctuations. Assume
observed concentration. that we release Qm mass units in an initial volume of V0, so that
To understand the effect of concentration fluctuations, the initial concentration is Qm/V0. As the released material is
consider the following model of the concentration time series stretched over a larger volume, V, in space, the volume marked
in which the concentration is a peak value, Cpeak, or zero. Then, by the material remains unchanged if the fluid is incompress-
we can show that the variance of the instantaneous concen- ible and molecular diffusion is negligible (see Figure 4). The
tration about the ensemble-averaged mean is given by eqn [49]. mean concentration, corresponding to material spread over V,
is proportional to 1/V, while the peak concentration remains
s2c Cpeak
¼ 1 [49] unchanged from the initial value. Then, the peak-to-mean
C2 C ratio, which determines the concentration variance, is simply
Because the peak-to-mean concentration ratio can be as V/V0. In the presence of molecular diffusion, the material is no
large as 10 or even 100, especially close to an elevated source, longer confined to its initial volume, and the peak concentra-
the standard deviation of the concentration fluctuations can be tion has to decrease with time. Thus, molecular diffusion
several times the mean. If we are interested in predicting a time- decreases concentration fluctuations relative to the mean value.
averaged concentration, we have to estimate the timescale, Tc The meandering plume model, used to estimate the vari-
that governs the concentration fluctuations. This then allows us ance of concentration fluctuations, is an extension of the
to estimate the number of independent concentration events previous concept. Here, the time-averaged plume is assumed to
that we are likely to observe during the averaging interval Tav, as be composed of instantaneous plumes whose dimensions are
Tav/Tc. Then, the deviation, d, of the time-averaged mean about determined by relative dispersion. Then, the peak concentra-
the ensemble mean is given by eqn [50]. tion is determined by the concentration within the instanta-

1=2 neous plume, which is inversely proportional to s2, where s
d Tc sc refers to the spread by relative dispersion. Then, if s is the time-
 [50]
C Tav C averaged spread of the plume, the peak-to-mean ratio, and the

Volume of marked material


is still V0 but material is
spread over V in space

Qm

V0
V
Qm
C0 ~ Qm
V0 C~ Peak concentration
V is decreased
by molecular
diffusion through
strands of marked
material

Figure 4 Factors that affect concentration fluctuations. Cpeak  C0 in the absence of molecular diffusion, and s2c =C2  Cpeak =C  V=V0 .
286 Turbulence and Mixing j Turbulent Diffusion

normalized concentration variance is of the order of s2 =s2 as Further Reading


long as we can neglect molecular diffusion within the instan-
taneous plume. Molecular diffusion will eventually smear the Arya, S.P., 1999. Air Pollution Meteorology and Dispersion. Oxford University Press,
concentration over the instantaneous plume and thus decrease New York.
the normalized concentration variance. Csanady, G.T., 1973. Turbulent Diffusion in the Environment. Reidel, Dordrecht,
Holland.
The statistics of concentration fluctuations can be used in Nieuwstadt, F.T.M., van Dop, H. (Eds.), 1982. Atmospheric Turbulence and Air
estimating the air quality impacts of species for which small Pollution Modeling. Reidel, Dordrecht, Holland.
time exposures are important. These statistics can be used to Pasquill, F., Smith, F.B., 1983. Atmospheric Diffusion, third ed. Ellis Horwood Limited,
estimate the probability that a certain threshold is exceeded. John Wiley & Sons, New York.
Rodean, H.C., 1996. Stochastic Lagrangian Models of Turbulent Diffusion. American
Thus, there is great interest in formulating models for
Meteorological Society, Boston, MA.
concentration fluctuations. Some of the more recent models Seinfeld, J.H., Pandis, S., 1999. Atmospheric Chemistry and Physics of the Atmo-
have combined a version of LSS with LES velocity fields to sphere. Wiley-Interscience, New York.
model relative dispersion and thus concentration fluctuations. Taylor, G.I., 1921. Diffusion by continuous movements. Proceedings of the London
In the future, we are likely to see more use of DNS in under- Mathematical Society 20, 196–211.
Tennekes, H., Lumley, J., 1972. A First Course in Turbulence. MIT Press,
standing turbulent dispersion. This does not mean that our Cambridge, MA.
ability to predict concentrations will improve substantially, Venkatram, A., Wyngaard, J. (Eds.), 1988. Lectures on Air Pollution Modeling.
because the nature of turbulence places practical limits on American Meteorological Society, Boston, MA.
predicting individual realizations of concentrations.

See also: Aerosols: Observations and Measurements; Aerosol


Physics and Chemistry. Boundary Layer (Atmospheric) and Air
Pollution: Modeling and Parameterization; Overview. Aviation
Meteorology: Clear Air Turbulence. Numerical Models:
Parameterization of Physical Processes: Turbulence and
Mixing. Turbulence and Mixing: Overview; Turbulence, Two
Dimensional.

You might also like