You are on page 1of 93

BOUNDARY-LAYER METEOROLOGY

Han van Dop, September 2008

D 08-01
Contents

1 INTRODUCTION 3
1.1 Atmospheric thermodynamics . . . . . . . . . . . . . . . . . . . . 3
1.2 Statistical aspects of fluid mechanics . . . . . . . . . . . . . . . . 6

2 CONSERVATION LAWS 10
2.1 Governing equations . . . . . . . . . . . . . . . . . . . . . . . . . 10

3 THE REYNOLDS EQUATIONS 14


3.1 The average flow . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3.2 The mean-flow energy equation . . . . . . . . . . . . . . . . . . . 19
3.3 The turbulent kinetic energy equation . . . . . . . . . . . . . . . 20
3.4 Spectra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
3.5 Closure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

4 THE ATMOSPHERIC BOUNDARY LAYER 31


4.1 Phenomenology . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
4.2 The surface layer; Monin-Obukhov theory . . . . . . . . . . . . . 35
4.3 The neutral boundary layer . . . . . . . . . . . . . . . . . . . . . 38
4.4 The convective boundary layer . . . . . . . . . . . . . . . . . . . 44
4.5 The stable boundary layer . . . . . . . . . . . . . . . . . . . . . . 48
4.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

5 EXCHANGE OF HEAT AND WATER VAPOUR 54


5.1 The surface-energy budget . . . . . . . . . . . . . . . . . . . . . . 54
5.2 The profile method . . . . . . . . . . . . . . . . . . . . . . . . . . 56
5.3 The energy-balance method . . . . . . . . . . . . . . . . . . . . . 58
5.4 Estimating the evaporation . . . . . . . . . . . . . . . . . . . . . 59
5.5 Air-Sea interaction . . . . . . . . . . . . . . . . . . . . . . . . . . 61

6 HETEROGENEOUS BOUNDARY LAYERS 70


6.1 Thermal transitions . . . . . . . . . . . . . . . . . . . . . . . . . 70
6.2 Roughness transitions . . . . . . . . . . . . . . . . . . . . . . . . 72

7 EXERCISES BOUNDARY LAYERS AND TURBULENCE 77

2
Chapter 1

INTRODUCTION

This reader acts as a follow-up of the lecture notes of the Bachelor Hydro-
dynamics and Turbulence course. Some of that material directly relating to
boundary-layer meteorology, will be shortly discussed here.

1.1 Atmospheric thermodynamics

First Law of Thermodynamics

The change of energy dQ per unit mass of a closed system equals the sum of
the change of the internal energy dU , and the amount of work, pdV :

dQ = dU + p dV. (1.1)

The specific heat at constant volume, or constant pressure, cv and cp are:

(dQ)v = (dU )v ≡ cv dT
(dQ)p ≡ cp dT

where (. . . )v denotes that volume is held constant, and (. . . )p indicates a con-


stant pressure (cv,p in J kg−1 K −1 ).

Equation of state
Again for a unit mass we have

pV = RT (1.2)

or, using ρ = 1/V :

p = ρRT. (1.3)

3
An alternative for the First Law can now be written as:

1
dQ = cp dT − dp. (1.4)
ρ
where cv + R = cp , and R is the specific gas constant for dry air.

R = R? /ma , with R? , the universal gas constant (8.314 J K−1 mole−1 )


and ma the molecular weight of (dry) air (0.0288 kg). Thus R equals 288
J kg−1 K−1 .

Adiabatic lapse rate

The heat exchange between the atmosphere and its surroundings, for example
by molecular conduction or by radiation, is often negligible, implying dQ = 0.
From the first law of thermodynamics and the hydrostatic equation, it follows
that

cp dT + g dz = 0
dT g
− = ≡ Γd ≈ 10−2 Km−1 (1.5)
dz cp

This vertical temperature gradient is commonly referred to as the (dry) adia-


batic lapse rate.

Entropy

The entropy S is defined as

dQ dT dp
dS ≡ = cp −R
T T p

= cp d ln T − R d ln p.

Upon integration, this yields:

S = cp ln T − R ln p. (1.6)

Potential temperature

In adiabatic processes, dS = 0. In a process (1 → 2), this means that :

4
cp ln T1 − R ln p1 = cp ln T2 − R ln p2 . (1.7)

If we assign a constant pressure ps (e.g. 1000 mb) and temperature θ (potential


temperature) to the reference level 2, then (1.7) can be written as:

cp ln T − R ln p = cp ln θ − R ln ps , (1.8)

where the subscript 1 is omitted. This equation can be rewritten to obtain

ps κ
θ = T( ) , (1.9)
p

where κ = R/cp . This expression allows us to determine the potential tempera-


ture of any air mass as long as its temperature and pressure are known. In this
way, a vertical profile of potential temperature can be constructed. Using (1.6),
the entropy now becomes

S = cp ln θ − R ln ps , (1.10)

leading to the conclusion that, in case of an adiabatic process, there is a direct


connection between the potential temperature, θ, and the entropy S. Further-
more, Eq. (1.8) implies

dθ dT
≈ + Γd , (1.11)
dz dz
where the dry adiabatic lapse rate is given by

θ g g
Γd = ≈ . (1.12)
T cp cp

The above-mentioned equation poses a simple relation between the vertical tem-
perature gradient dT /dz and the vertical gradient of the potential temperature
dθ/dz.

Standard atmosphere

Using (1.5), an atmospheric temperature profile can be described:

T z
=1− , (1.13)
T0 H

5
T0 being the surface temperature, and H = cp T0 /g. We can now invoke the
equation of state (p = ρRT ) and the hydrostatic equation to derive equations
for vertical atmospheric profiles of pressure and air density:

p z cp
= (1 − )R (1.14)
p0 H
ρ z cp
= (1 − ) R −1 . (1.15)
ρ0 H
The atmospheric scale height is denoted by H. This approach implicates that
the atmosphere has a finite thickness (H ≈ 30 km) which is obviously not correct
but nevertheless yields a reasonable estimate of the thickness of the atmosphere.

1.2 Statistical aspects of fluid mechanics


The length scales relevant in a turbulent flow cover a wide range. The macroscale
and the smallest turbulent scale, known as the Kolmogorov microscale, are
related through:

3
l/lK = Re 4 (1.16)

In the atmospheric boundary layer, typical values of l = 1000 m and lK = 0.001


m, yield a Reynolds number of O(108 ). Solving all length scales up to 0.001 m
on a domain of 10 km x 10 km x 1 km using a numerical program would require
1020 grid points, whereas computation on a grid with 1010 points is currently
feasible. Thus, turbulence equations cannot be solved numerically in the entire
domain of length scales. As an alternative, turbulence equations could be solved
for averaged quantities only. In this way, the range of length scales, and thus
the required amount of grid points, can be reduced dramatically.

Stochastic variables, moments, probability density functions

In a turbulent flow, kinematic processes occur over a wide range of scales. As


a first step in a statistical treatment of turbulence, the average of a variable
(the first moment) can be determined. More detailed information of the flow
can be obtained by calculating higher moments of variables, or combinations of
variables (correlations). When analysing a turbulent flow in a statistical way,
correlations between flow variables at different locations or at different times
play a very important role. Examples of such correlations are the variance

u2 ≡ u(x
x, t)u(x
x, t) (1.17)

and the covariance

6
uw ≡ u(x
x, t)w(x
x, t), (1.18)

being statistical properties of flow velocity at location x and time t. The overbar
is used for the so-called ensemble-average of a quantity, i.e. the average obtained
by taking the average of a large amount of realisations under the same boundary
conditions.
If we treat velocity, pressure and other quantities in a turbulent flow as stochastic
variables, it is possible to define the flow by its moments. Let u be any variable,
and p(u) its probability density function. The moments are then given by

Z +∞
um ≡ p(u)um du (1.19)
−∞

A stationary stochastic process is defined as a process where the density function


p is invariant under a time translation. If a stochastic process is stationary,
p(u1 , u2 ; t1 , t2 ) = p(u1 , u2 ; t1 − t2 ) applies. The autocorrelation reads

x, t1 ) u(x
u(x x, t1 − t2 ).
x, t2 ) = ρ(x

Analogously, the spatial autocorrelation is defined as

x1 , t)u(x
u(x x2 , t).

In a homogeneous flow this function depends only on the separation:

x1 − x 2 , t) = u(x
ρs (x x1 , t)u(x
x2 , t).

where ρs is the (spatial) velocity autocorrelation function.

Reynolds decomposition and averaging processes

Reynolds decomposition

We can split the motion of a turbulent flow into a large scale and a small scale
component (here denoted with a prime). In this way, the average of a variable
and its fluctuation are discerned. Let u be the velocity. It can be split as follows

u = u + u0 (1.20)

’Averaging’, denoted by the overbar, means ensemble-averaging in this context.


We assume that the averaging process is a ’Reynolds operator’ which means
that

7
u = u (and thus u0 = 0), and that cu = cu (c is a constant)

Averaging

Features of ensemble-averaging are:

• Ensemble-averaging removes all turbulence from the description of the


flow: the variables are no longer chaotic. Any information about the
structure of the turbulence vanishes.
• Transport by eddies has to be parameterised, i.e. only the statistical (av-
erage) properties of the turbulence are featured in the resulting equations.
Such a parameterisation requires a good knowledge of the flow.
• Averages vary much less in time and space than the turbulence itself. Grid
point distances can therefore be chosen relatively large.

Time-averaging

The average value of a velocity variable over a period T is defined as

Z t+ 12 T
1
U T (x
x, t) = x, t0 )dt0 ,
u(x (1.21)
T t− 12 T

If T incorporates all turbulence time scales, the time-average would be a good


approximation of the ensemble-average (U will then be independent of T ).

Volume-averaging

Analogous to time-averaging, a variable can also be averaged over a certain


volume. The local values are split into a spatial average and a deviation from
this average

a = {a} + a0 (1.22)

where the volume-average is defined as

Z
1
{a} ≡ a dV. (1.23)
∆V ∆V

In more general terms, the volume-average can also be expressed as

Z +∞
{a}(x) ≡ G(x0 − x)a(x0 )dx0 (1.24)
−∞

8
where the filter function G is given by, for example,

G(x − x0 ) = 1 if |x0 − x| < L


= 0 otherwise.

Important properties of (1.23), that differ from the properties of ensemble-


averages, are

{{a}{b}} =
6 {a}{b}
{{a}} =
6 {a}
(1.25)

In practical applications these inequalities are often ignored, since errors made
by assuming that these quantities are equal appear to be small. Characteristics
of volume-averaging are:
1
• Volume-averaging only removes the scales smaller than V 3 ; it conserves
the larger scales.
1
• A so-called closure is needed for the scales smaller than V 3 . The advantage
of volume-averaging, compared to ensemble-averaging, is that the closure
hypothesis is less critical, since the small scales do not carry much energy
and have relatively well-known properties.

9
Chapter 2

CONSERVATION LAWS

We summarize the basic equations as follows:

The continuity equation

dρ ∂ui
+ρ = 0. (2.1)
dt ∂xi

The momentum equation

∂ui ∂ui ∂p ∂ 2 ui
ρ + ρuj =− − ρgδi3 + µ . (2.2)
∂t ∂xj ∂xi ∂xj ∂xj

The temperature equation

∂θ ∂ dθ ∂2θ
+ uj θ≡ =κ . (2.3)
∂t ∂xj dt ∂xj ∂xj

2.1 Governing equations


The continuity equation

As a good approximation, the continuity equation can be expressed as

∂ui
= 0,
∂xi
also implying that

10
1 dρ
≈ 0.
ρ dt
The density along the fluid-particle trajectories is approximately constant. The
flow can therefore be considered incompressible.

In an adiabatic process,
dρ 1 dp
= 2
dt c dt
c p 1
applies, where c (the velocity of sound) stands for ( cp ρ
)2 . When combined
v
with the Bernoulli equations, this yields

( )
∂ui 1 d 21 ui 2
= 2 + gu3 .
∂xi c dt

Choosing U and L as scale sizes of ui and xi respectively, then

L ∂ui u2 gL
= O( 2 ) + O( 2 ).
U ∂xi c c
Using the Boussinesq approximations, the right-hand side can be neglected.

Boussinesq approximations

The above approximations, u/c << 1 and L << c2 /g, which imply that at-
mospheric flow can be considered incompressible, form part of the so-called
Boussinesq approximations. This however does not mean that density differ-
ences are dynamically unimportant. They certainly are in combination with
gravity. The resulting equations are referred to as the Boussinesq equations.
They originate from a number of subtle arguments which will be given below.

We will suppose a reference state of the atmosphere (denoted by the subscript


0). The atmosphere is at rest (ui = 0), and horizontally homogeneous regarding
temperature, density and pressure. We further assume that thermodynamical
deviations from the reference state in a realistic atmosphere are small. We
further suppose that the actual atmospheric state does not deviate much from
the reference state, viz.

p −→ p0 + p̃
ρ −→ ρ0 + ρ̃
θ −→ θ0 + θ̃
ui −→ 0 + ũi ,

where p̃, ρ̃, θ̃ and ũi denote small deviations from the reference state. When we
substitute this in the equation of state, p = ρRθ, (we assume that in the ABL

11
θ ≈ T , see Eq. (1.9)). We obtain in zero order:

p0 = ρ0 Rθ0 , (2.4)

and to first order

p̃ = R(ρ0 θ̃ + ρ̃θ0 ).

We assume, confirmed by observations and given the approximations already


made, that

p̃ << θ̃, ρ̃,

so that

ρ̃ θ̃
∼− . (2.5)
ρ0 θ0

Now we make the same substitutions in (2.2):


dũi ∂p0 ∂ p̃ ∂ 2 ũi
(ρ0 + ρ̃) =− − − (ρ0 + ρ̃)gδi3 + µ 2 .
dt ∂xi ∂xi ∂xj

In zero order this yields


∂p0
= −ρ0 g δi3 , (2.6)
∂xi
the hydrostatic equation, and to first order:
dũi ∂ p̃ ∂ 2 ũi
ρ0 =− − ρ̃g δi3 + µ 2 ,
dt ∂xi ∂xj

which can be rewritten, using (2.5), as

dũi 1 ∂ p̃ θ̃ ∂ 2 ũi
=− + g δi3 + ν ,
dt ρ0 ∂xi θ0 ∂x2j

where ν = µ/ρ0 , the kinematic viscosity. Back substitution of the original


variable θ yields
dũi 1 ∂ p̃ θ − θ0 ∂ 2 ũi
=− +( )gδi3 + ν . (2.7)
dt ρ0 ∂xi θ0 ∂x2j

This is a familiar expression of the Boussineq equations for a shallow boundary


layer. Essential is that density (temperature) deviations are only important in
combination with gravity. Otherwise density can be considered constant. So
the Boussinesq-approximations for the (shallow) ABL include:

12
u
<< 1
c
c2
L <<
g
θ≈T

<< 1
p
and we summarize the equations in Boussinesq form:

The continuity equation


∂ ũi
= 0. (2.8)
∂xi
The momentum equation

∂ ũi ∂ ũi 1 ∂ p̃ (θ − θ0 ) ∂ 2 ũi


+ uj =− + gδi3 − 2ijk Ωj uk + ν . (2.9)
∂t ∂xj ρ0 ∂xi θ0 ∂xj ∂xj
For the sake of completeness, the Coriolis term has been added, making the
equations apply for a rotating coordinate system, with an angular velocity Ω
(in rad/s).

The temperature equation

The temperature equation, neglecting molecular conduction, is changed to

dθ̃
= 0. (2.10)
dt
Together, eqs.(2.8, 2.9 and 2.10) constitute a set of 5 equations (with unknown
variables ũi , θ and p̃) that serves as a starting point for the study of the atmo-
spheric boundary layer.

13
Chapter 3

THE REYNOLDS
EQUATIONS

We will suppose that a flow consists of a laminar, average flow, and, superim-
posed, turbulent fluctuations, the so-called Reynolds decomposition (see sec-
tion 1.2).

3.1 The average flow


We will adopt the Navier-Stokes equation, Eq. 2.9,

∂ ũi ∂ ũi 1 ∂ p̃ (θ − θ0 ) ∂ 2 ũi


+ u˜j =− + gδi3 − 2ijk Ωj uk + ν . (3.1)
∂t ∂xj ρ0 ∂xi θ0 ∂xj ∂xj
and the continuity equation

∂ ũi
= 0. (3.2)
∂xi
We decompose the velocity ũi , the pressure p̃ and the temperature θ in a mean
and fluctuating (turbulent) component. After substitution of ũi = Ui + ui ,
p̃ = Π + π and θ = Θ + θ (with this substitution we redefine from hereon θ as a
temperature fluctuation) and averaging, we obtain

∂Ui ∂Ui 1 ∂Π (Θ − θ0 ) ∂ 2 Ui ∂ui uj


+Uj =− + gδi3 −2ijk Ωj Uk +ν − , (3.3)
∂t ∂xj ρ0 ∂xi θ0 ∂xj ∂xj ∂xj
and

∂Ui
= 0. (3.4)
∂xi

14
Figure 3.1: Frictional flow along a flat surface

The extra terms, ui uj , which represent turbulent momentum fluxes, are called
Reynolds stresses and are expressed in known variables. This procedure is called
closure (or ’K-theory’) and will be extensively treated in section 3.5:

∂Ui ∂Uj 
ui uj = −Km + (3.5)
∂xj ∂xi

The proportionality constant Km takes the dimension of viscosity (m2 s−1 ), but
is entirely determined by the turbulence itself.
Let l and u be the length and velocity scales of the turbulence, then it seems
reasonable to suppose that Km ∝ lu. This can be nicely illustrated by examining
a flow along a flat surface.

Neutral boundary-layer flow

The average flow is stationary and homogeneous in the x-direction ( ∂U ∂U


∂t , ∂x = 0).
The equations of motion (meglecting Coriolis forces) are (see 3.3)
∂ ∂U  1 ∂Π
ν − uw = (3.6)
∂z ∂z ρ0 ∂x
1 ∂Π ∂w2
− +g = 0. (3.7)
ρ0 ∂z ∂z
1 ∂Π
By differentiating (3.7) to x, we find that ρ0 ∂x cannot be a function of z. Now,
integrating the first equation yields:
∂U 
ν − uw ∝ z. (3.8)
∂z

15
Suppose that h is the ”thickness” of the boundary layer (the height at which
∂U/∂z ≈ 0)). Consider two cases:

a. Laminar flow close to the wall (uw = 0).

Eq. (3.8) gives


∂U
ν = az + b, (3.9)
∂z
where the constants a and b follow from the boundary conditions at z = 0 and
z = h. At z = h we assume ∂U/∂z = 0. b represents the friction force at the
surface (with dimension velocity squared). This defines the ’friction velocity’ u?
as
∂U
lim ν ≡ u2? , (3.10)
z→0 ∂z
so that we finally get
∂U z
ν ∝ u2? (1 − ), (3.11)
∂z h
implying that the velocity profile is linear for z << h. In this (thin) layer
molecular friction dominates and thus Re = O(1). If the thickness of that layer
is δ we have thus
u? δ
≈ 1,
ν
which yields δ ≈ ν/u? for the thickness of the ’laminar sub-layer’.

b. Turbulent flow: the logarithmic wind profile

Above the laminar sub-layer (ν/u? . z . h) turbulent friction dominates:

|uw| >> ν ∂U
∂z .

Eq. 3.8 now reads

−uw ∝ z

which, applying the boundary conditions −uw = 0 at z = h and −uw = u2? at


z = ν/u? , yields
z
−uw ∼ u2? (1 − ). (3.12)
h
(note that ν/u? << h). If z/h << 1, the vertical transport of momentum
(−uw ≈ u2? ) is approximately constant. This defines another sub-layer, the
’constant flux layer’, ν/u? . z << z/h.

16
We can now use the closure relation (3.5) to express the Reynolds stress in terms
of the average gradient in this layer:

∂U
−uw = Km = u2? . (3.13)
∂z
According to (3.13), the velocity profile would be linear (using a constant value
for Km ). This has not been confirmed experimentally: instead, logarithmic pro-
files are found. This only follows from (3.13) if the ’eddy’ diffusion coefficient
K is proportional to z. If we define

Km = κzu? ,

where κ is the Von Karman constant), the solution of (3.13) will be

U 1 z
= ln , (3.14)
u? κ z0
This is a logarithmic profile1 , where z0 is an integration constant. The pro-
portionality constant κ has a value of ≈ 0.4. The quantity z0 depends on
the roughness of the wall, and is called the roughness length. For an aerody-
namically smooth wall, z0 is defined as the height at which the corresponding
Reynolds number has the value of 1:

u? z0
(Re)z0 ≡ , (3.15)
ν
leading to z0 = ν/u? . If an aerodynamically rough wall is characterised by
irregularities of height h, the Reynolds number is given by

u? h
(Re)h ≡ . (3.16)
ν
A surface is called smooth when Reh < 1 and rough when Reh > 1. A couple
of typical values for z0 are listed in table 3.1. The roughness of a rough water
surface is treated later in these notes.
Under neutral conditions and over a large area of varying surface types, the wind
profile is approximately logarithmic from a height z0 to a couple of hundreds of
metres.
If the area is covered with higher objects (trees), it is possible to define a new
surface where the wind speed is 0. The wind profile is then given by

U 1 z−d
= ln , (3.17)
u? κ z0
1 There is as yet no exact derivation for the logarithmic wind profile. In section 4.3, it is

made plausible that logarithmic profiles occur in boundary layers

17
Table 3.1: Typical values for the roughness length z0

Surface type Roughness length (m)


Smooth water/ice 10−4
Short grass 10−2
Low vegetation 0.05
Countryside 0.20
Low built-up area 0.6
Forests/cities 1−5

where d is the so-called displacement height. The displacement height equals


roughly 80% of the object height (see figure 3.2). Equation (3.17) is very suitable
to calculate the wind speed at a given height, if the wind speed is known on any
other height:

U2 ln(z2 − d)/z0
= .
U1 ln(z1 − d)/z0

Drag coefficient
If we square (3.14) and subsequently multiply it by the density, we will find the
surface shear stress:

κ2
τ ≡ ρ u2? = ρ U 2.
ln2 (z/z0 )

The drag coefficient is thus defined as

κ2
Cd = 2 , (3.18)
ln (z/z0 )

implying that

τ = ρ Cd U 2 , (3.19)

This is a simple relationship to estimate the surface shear stress from the average
wind speed under neutral circumstances.

Power law

In practice, an algebraic formula for the wind profile is often applied:

18
Figure 3.2: Sketch of the wind speed profile over a homogeneous forest.

 p
U2 z2
= . (3.20)
U1 z1

In neutral conditions, the following rule approximately holds:

√ 
z1 z2
p ≈ ln−1 . (3.21)
z0

This relation can bep derived by the requirement that at the geometric mean
value of z1 and z2 , (z1 z2 ), the wind velocity and the first derivative are equal
in both formulations.
A frequently used, but not necessarily correct value of p is 1/7 (based on z1 ∼
z2 ∼ 10m and a roughness length of 8 cm).

3.2 The mean-flow energy equation


The starting point is (3.3). We shall neglect the Coriolis force since it does not
play a role in energy budget considerations:

∂Ui ∂Ui 1 ∂Π (Θ − θ0 ) ∂ 2 Ui ∂ui uj


+ Uj =− + gδi3 + ν − ,
∂t ∂xj ρ0 ∂xi θ0 ∂xj ∂xj ∂xj

Multiplying by Ui yields:

19
∂E ∂E
+ Uj
∂t ∂xj
1 ∂Π Ui (Θ − θ0 ) ∂ ∂Ui
= − Ui + gδi3 + {νUi − Ui ui uj }
ρ0 ∂xi θ0 ∂xj ∂xj
∂Ui ∂Ui 2
+ ui uj − ν( ) , (3.22)
∂xj ∂xj
where E ≡ 21 Ui2 . If the Reynolds number is high, the viscous terms can be
omitted.
When integrated over the volume of the flow, the result is

Ui (Θ − θ0 )
Z Z
dEtot ∂Ui
= ui uj dV + gδi3 dV. (3.23)
dt V ∂xj V θ0
The integrand of the first integral is negative in a shear flow, and is called the
deformation work. This term provides the translation of the average flow energy
to its fluctuations (i.e. the coupling between average flow and turbulence). The
second integral represents the conversion of potential into kinetic energy by the
mean motion in a stratified flow.

3.3 The turbulent kinetic energy equation


The energy transfer from the average flow to turbulence is provided by the
deformation work. We will now derive an expression for the average kinetic
energy of the fluctuations, q ≡ 21 ui 2 . We shall start from (2.9):

∂ ũi ∂ ũi 1 ∂ p̃ (θ − θ0 ) ∂ 2 ũi


+ u˜j =− + gδi3 + ν ,
∂t ∂xj ρ0 ∂xi θ0 ∂xj ∂xj
and for the temperature

∂θ ∂θ ∂2θ
+ u˜j =κ ,
∂t ∂xj ∂xj ∂xj
As in the derivation of the equation for the mean energy (3.3), we make again
the substitutions ũi = Ui + ui , etc. In order to arrive at an equation for the
fluctuations, we substract from this equation the equation for the mean flow
(3.3). We multiply the resulting equation with ui . Taking the average of this
equation, we finally obtain the equation for the energy of the fluctuation q ≡
1 2
2 ui :

∂q ∂q ∂Ui g ∂ 1 1 ∂ 1 u2
+Uj = −ui uj + wθ− { 2 ui ui uj + πuj −ν 2 i }−. (3.24)
∂t ∂xj ∂xj θ0 ∂xj ρ0 ∂xj

20
The first term on the right-hand side of (3.24) is the mechanical production term,
supplied by the average flow (normally positive). The second term represents
the thermal production of turbulent fluctuations due to differences in density
of the flow. In a gravity field, this term can transform potential energy into
kinetic and vice versa. The ratio between mechanical and thermal production
of turbulence is defined as the flux-Richardson number:

g
θ0 wθ
Rif = (3.25)
ui uj ∂U
∂xj
i

We discern three cases:

wθ > 0 unstable flow Rif < 0

wθ = 0 neutral flow Rif = 0

wθ < 0 stable flow Rif > 0

The use of the flux-Richardson number can be illustrated by the atmospheric


boundary layer that is heated by the Earth surface during the day, so that
wθ > 0. At daytime, the boundary layer is thus generally unstable. During
the night, the Earth surface will cool down by radiating out its energy. Thus,
wθ < 0, meaning that the boundary layer will be stable with little turbulence.
We also distinguish the gradient Richardson number, Ri, by applying K-theory
to the fluxes in Eq. 3.25. Similar to Eq. 3.5 we may write for wθ:
∂Θ
wθ = −Kh , (3.26)
∂z
so that Eq. 3.25 becomes
Kh
Rif = Ri, (3.27)
Km
where Ri is given by
g
θ0 ∂Θ/∂z
Ri = . (3.28)
(∂Ui /∂xj + ∂Uj /∂xi ) ∂Ui
∂xj

Apart from the flux-Richardson number, we also use the bulk-Richardson num-
ber, which follows from (3.28), by approximating gradients by finite differences
and assuming a uniform windfield, U (z),

g ∆Θ
Rib = ∆z. (3.29)
θ0 (∆U )2

21
Returning to Eq. 3.24, we observe that the redistribution term between braces
contains 3 energy fluxes: by turbulent fluctuations, by pressure fluctuations, and
by molecular fluctuations (the latter has already been omitted in the equation).
The last term represents the energy dissipation. It is the only negative term,
and therefore necessarily of the same order of magnitude as the other terms. If
not, then q could grow unlimitedly.
In a neutral, semi-stationary situation, (3.24) implies

∂Ui ∂ui 2
P ≡ −ui uj ≈ ν( ) ≡ ,
∂xj ∂xj
so P ≈ . The production and dissipation of turbulent energy are approximately
3
balanced. Also note that P = O( ul ), which means that

u3
 = O( ). (3.30)
l

3.4 Spectra
Turbulence can be viewed upon as a set of eddies of different sizes between lK
and L. In a turbulent flow, energy of a particular scale is transferred towards
larger and smaller scales. The average flow provides the energy for the turbu-
lence at the macroscale. The larger eddies are unstable and disintegrate into
smaller eddies of various sizes. This process is repeated several times. The
smallest eddies will ultimately lose its energy by molecular viscosity. The net
effect is that eddies tranfer energy from larger to smaller scales, known as the
energy cascade.
A turbulent velocity field consists of a superposition of fluctuations with a wide
range of temporal and spatial scales. To analyse the contribution of each scale,
a Fourier analysis can be exploited. Let the autocorrelation of a continuous
velocity function u(t) be defined as
Z + T2
1
ρ(τ ) = u(t) u(t + τ ) = lim u(t) u(t + τ ) dt. (3.31)
T →∞ T − T2

We have assumed stationary conditions so that ρ depends on the time difference


τ only. The energy spectrum S(ω) is, by definition, the Fourier transform of
the correlation function:
Z +∞
1
S(ω) ≡ e−iωτ ρ(τ ) dτ, (3.32)
2π −∞

with the associated inverse operation


Z +∞
ρ(τ ) ≡ eiωτ S(ω) dω, (3.33)
−∞

22
From (3.33), it immediately follows that for τ = 0
R +∞
u2 ≡ −∞
S(ω) dω.

The lefthand side represents the turbulent kinetic energy and this equation
shows why S is actually called the energy spectrum, since S(ω) equals the
contribution to u2 of S between the frequencies ω and ω + dω.

23
In practice the energy or power spectrum of the function u(t) is determined
in a more direct way: we rewrite Eq. 3.31 as
Z +∞
1
ρ(τ ) = lim v(t) v(t + τ ) dt, (3.34)
T →∞ T −∞

where v(t) is defined as

v(t) = u(t) for|t| < T /2


v(t) = 0 otherwise.

The Fourier tranform of v(t) is


Z +∞
1
ṽ(ω) = e−iωt v(t) dt, (3.35)
2π −∞
with the inverse transformation
Z +∞
v(t) = eiωt ṽ(ω) dω.
−∞

We use the last expression to rewrite Eq. 3.34 as


Z +∞ Z +∞
1
ρ(τ ) = lim v(t) ṽ(ω) eiω(t+τ ) dω dt.
T →∞ T −∞ −∞

Rearranging terms we get


Z +∞ Z +∞
1
ρ(τ ) = lim ṽ(ω) eiωτ v(t)eiωt dt dω,
T →∞ T −∞ −∞
or

Z +∞
ρ(τ ) = lim ṽ(ω) ṽ(−ω) eiωτ dω.
T →∞ T −∞

Since ṽ(−ω) = ṽ ? (ω) we have


2π +∞
Z
ρ(τ ) = lim |ṽ(ω)|2 eiωτ dω.
T →∞ T −∞

Comparing this result with Eq. 3.33 yields



S(ω) = lim |ṽ(ω)|2 ,
T →∞ T
which can be rewritten as
˛Z T ˛2
2π ˛˛ + 2 iωt
˛
S(ω) = lim u(t) e dt˛ . (3.36)
˛
˛
T →∞ T ˛ − T ˛
2

This equation, known as the Wiener-Khintchin theorem, provides a direct


relation between the velocity signal u(t) and its power spectrum. The Fast
Fourier Transform (FFT) is an efficient numerical way to determine the
spectrum of an arbitrary stationary process based on a discrete represen-
tation of Eq. 3.36.

Some of the properties of S are

• S(ω) = S(−ω)

24
• S(ω) is real and ≥ 0.
R∞
• It also follows from the definition that S(0) = π1 0 ρ(τ ) dτ = 1
π (u
2 Tu ),
where Tu denotes the time scale of the signal u(t).

Spectral gap

During the 1950s, the ideas about boundary-layer meteorology were dominated
by the assumption that (small-scale) turbulent kinetic energy mainly occurred
at scales comparable to the boundary layer height (1-2 km). At larger scales,
hardly any energy would be available, whereas the energy would increase at
mesoscale and sub-synoptic scales. This implied that the turbulence energy
spectrum would contain a minimum, separating boundary-layer turbulence from
turbulence at larger scales. This would also justify the process of Reynolds
decomposition. Now that much more experimental data are available, this view
has become challenged. An example of a spectral analysis of the energy, recorded
during an airplane flight, is given in figure 3.4. In this figure, the spectral energy
of the u− and v velocity components continue to increase at larger scales. Other
turbulent variables, like temperature, water vapour content, and liquid water
concentration show comparable behaviour. Only the spectrum of the vertical
velocity w seems to level off towards larger scales.
That the spectral energy at larger scales does not vanish, has some inconvenient
implications for the integral properties of the spectrum. For example, the total
variance of u, being defined as

Z ∞
u2 ≡ Eu (k)dk, (3.37)
0

cannot be determined if Eu (k) continues to grow for k → 0. In these cases, the


ogive is defined as

Z ∞
Ogu (k0 ) ≡ Eu (k)dk, (3.38)
k0

so that integral properties of the spectrum can be determined anyway. The


choice of k0 is an arbitrary one.

3.5 Closure
Expressing the stress in terms of the average flow is called closure. More gener-
ally, closure means that higher moments are expressed in terms of lower ones.
It is called closure because the set of equations of motion can be solved after
closure. In the previous sections, some simple examples of 1st -order closure have
already been demonstrated. More elaborate ways of closure exist, which will be
shortly discussed below.

25
Figure 3.3: Kaimal’s schematic representation of the energy spectrum. (A)
production subrange; (B) the inertial subrange, where production and dissi-
pation are not important; (C) dissipation subrange. The second Kolmogorov
hypothesis conjectures that in the inertial subrange, the dissipation  and the
wavenumber k are the dominating parameters. The spectrum E(k) can be
determined using dimensional analysis based on  and k only. The result is
2 5
E(k) = CK  3 k − 3 .

First-order closure (K-theory)

First-order closure is often applied since it is a fast method. It implies that the
0
K-theory’ is used to determine the Reynolds-stress term (see, e.g., Eq. 3.5 ).
The diffusion coefficient K is typical for the flow and may thus change in time
and space. At the surface layer, we acceptably showed that Km = κzu? (see
chapter 3.3, Eq. 3.13). Above the surface layer, there are many possible ways to
follow. At the top of the boundary layer, the vertical exchange is small, leaving
Km approximatly zero. Somewhere within the boundary layer, Km should have
a maximum. A possible formulation for Km is:

  12
2 ∂U 2 ∂V 2
Km = l + . (3.39)
∂z ∂z
In this equation, l is the mixing length, that will be proportional to the height
in the surface layer, l = κz, and converge to a constant value, say λ, for greater
heights. This choice for Km will give the desired behaviour in the surface layer,
where V v 0 and −uw are constant (u2? ). It now follows from

∂U
uw = −Km (3.40)
∂z

26
Figure 3.4: 1-D spectra of the horizontal (u and v) and vertical (w) wind-speed
components as a function of the wave number k. Measurements from an airplane
at approx. 150 m above the sea.

27
after substitution in (3.39), that

∂U u?
= ,
∂z κz
(see Eq. 3.13) which provides the desired logarithmic wind profile. The height-
dependent growth of the length scale can be delimited by choosing the following
parameterisation:

κz
l= (3.41)
1 + κz
λ

where λ is an empirical heigth scale (typically ∼100 m). The eddie diffusion
parameter Km (3.39) can easily be corrected for the stability by adding an
empirical function F :

  12
2 ∂U 2 ∂V 2
K=l + F (Ri), (3.42)
∂z ∂z

F being a function of the stability through the Richardson number. This pa-
rameterisation from 1979 is still used in present-day global circulation models
like the ECMWF-model, that calculate weather forecasts.

1.5-order closure

This type of closure makes use of the energy equation (3.24), which is a second-
moment equation. Suppose that Km obeys

1
Km ≈ lq 2 , (3.43)

where q is given by the turbulent kinetic-energy equation (3.24). The terms


present in that equation can be approximated in the following way, if the flow
is assumed to be horizontally homogeneous:

∂Ui ∂U 2
−ui uj = −Km
∂xj ∂z
g g ∂Θ 
wθ = − Kh
θ0 θ0 ∂z
∂ πuj  ∂ ∂q
2 ui ui uj +
1
= − Km
∂xj ρ0 ∂z ∂z
3
 = q 2 l−1 .

Substituting the above equations into the stationary-energy equation yields:

28
∂U 2 g ∂Θ ∂ ∂q 3
0 = Km + Kh + Km − q 2 l−1 . (3.44)
∂z θ0 ∂z ∂z ∂z
An equation for the turbulent kinetic energy q emerges. The set of equations
(3.3, 3.5, 3.41, 3.43 and 3.44) can now be solved. The name ’1.5-order closure’
is used, since only one second-order equation is used for the closure algorithm.

Second-order closure

In this case, all second moments, like ui uj , ui θ, are incorporated in the closure
algorithm. These equations contain terms of the third order, which are either
parameterised or neglected. As a result, the number of equations that are to
be solved increases rapidly. Sometimes, this method yields better results. On
the other hand, the mathematical complexity, the large number of empirical
constants and the lack of a solid theory make second-order closure techniques
less favourable.

Large Eddy Simulation (LES)

LES makes use of a fundamentally different approach. The variables in the


Navier-Stokes equation are averaged over a small volume. Contrary to the
variables in the Reynolds-averaged equations, the LES-variables describe the
turbulent behaviour of the flow at scales larger than the scales of the averaging
process. The starting point is the momentum equation (see 3.1), where we
take for simplicity a neutral flow and neglect the Coriolis-term and molecular
viscosity. The volume-average is:

∂{ũi } ∂{ũi } 1 ∂{p̃} ∂Rij


+ {u˜j } =− + , (3.45)
∂t ∂xj ρ0 ∂xi ∂xj
where Rij is a tensor that describes the impact of the flow behaviour within
the volume on that outside the volume. This is called the subgrid scale stress,
defined as Rij = {ũi }{u˜j } − {ũi u˜j }. Comparable to Reynolds-averaging, this
term is approximated by relating it to volume-averaged variables:

 ∂{ũ } ∂{u˜ } 
i j
Rij = −νt + (3.46)
∂xj ∂xi
In a mathematical sense, LES is almost identical to the method of Reynolds-
averaging. The difference between the two processes can be found in the way
the eddy viscosity is formulated. Reynolds-averaging appoints an order of mag-
nitude u.l to K, so that the Reynolds number, based on K, is of the order of 1.
The choice for the eddy viscosity in LES, νt , depends on the volume over which
the averaging process takes place. This volume is usually of the same order of

29
magnitude as the grid point distance of the numerical scheme, ∆. The Reynolds
number, ul/νt ≈ l/∆ is now much larger than 1 (approximatly 102 to 103 using
current computer capacity).
A well-known choice for the eddy viscosity is formulated by Smagorinsky (1963),
which is based on a mixing hypothesis (like the K-theory):

νt = (Cs ∆)2 |S| (3.47)


1
where Cs is the Smagorinsky constant (Cs = 0.2), |S| = (2Sij Sij ) 2 and
 
∂{u˜ }
Sij = 12 ∂{∂xu˜ji } + ∂xji ,

the deformation tensor.


The most important applications of LES can be found in the simulation of
convective turbulence, where transports are dominated by large-scale motions
in the atmosphere.

Direct Numerical Simulation (DNS)

This technique aims to find solutions for the Navier-Stokes equation itself, with-
out introducing closure or averaging. At large Reynolds numbers, this approach
faces difficulties, since the range of turbulence scales increases rapidly. In order
to find numerical solutions, the grid points distances have to be approximately
equal to the smallest scale. For calculating practical situations, a large number
of grid points are required. Flows with Reynolds numbers up to 100-1000 can
be simulated with currently available computer power.

30
Chapter 4

THE ATMOSPHERIC
BOUNDARY LAYER

4.1 Phenomenology
Boundary-layer meteorology comprises the dynamics and physics of the atmo-
spheric layer that is closest to the Earth surface. Exchange processes between
the Earth surface and the ’free atmosphere’ occur in the boundary layer. The
state of the boundary layer is influenced by flow in the free atmosphere on the
one hand, and by boundary conditions imposed by the Earth surface on the
other hand.
Dominating boundary layer processes are the vertical exchange of momentum
τ = −ρ0 uw, heat H = ρ0 cp wθ and water vapour E = ρ0 wq.
The free atmosphere (see figure 4.3) is the layer above the uppermost boundary
of the (turbulent) boundary layer.
The behaviour of the boundary layer has an enormous impact on values of the
maximum and minimum temperature, wind speed gradient (wind shear), fog
and cloudiness. In aviation and architecture (wind engineering), both wind
shear, fog and the occurrence of wind gusts are important factors to keep in
mind. Turbulence, radiation, surface properties and thermodynamics are the
key ingredients of boundary-layer meteorology. We will first give a qualitative
overview.

Slightly above the Earth surface, a thin laminar layer exists: the ’viscous sub-
layer’. This layer adjusts the balance between the Earth surface and the surface
layer. The surface layer (or inner layer) is a couple of tens of metres thick. The
surface layer is topped by the (planetary) boundary layer, sometimes also re-
ferred to as the Ekman layer or outer layer. The transition between the surface
layer and the Ekman layer is smooth. The scales in the surface layer are very
different from those in the Ekman layer.
The boundary-layer dynamics are not only influenced by the average horizontal

31
Figure 4.1: Schematic overview of the troposphere.

32
flow, but also by turbulent processes. The most important of these processes
are:
a. Mechanical turbulence
b. Buoyant or convective turbulence.
The stability of the atmosphere influences the boundary layer, which has a
typical height of 100-2000 m. The flux-Richardson number is a measure for
the atmospheric stability. It is the ratio between buoyancy production and
mechanical production

g
θ0 wθ
Rif =
ui uj ∂U
∂xj
i

(see section 3.3).


The surface layer is heated when the surface heat flux wθ0 is positive. This
happens during daytime as a consequence of solar radiation. Convection may
also take place above a warm sea surface, over which cooler air is flowing.
If the heat flux is negative, the boundary layer will cool down. This happens
during the night. These two situations differ substantially, having quite some
implications for the dynamics (e.g. the growth of the boundary layer height:
if wθ > 0, then Rif < 0, so the atmosphere is unstable. This implies that
gravity is generating turbulence, and the boundary layer height increases as a
consequence. On the other hand, if wθ < 0, so Rif > 0, the boundary layer is
stable. Gravity will suppress turbulence. The boundary layer height will not
change notably. We conclude that there is a daily cycle of boundary layer height
over land surfaces (see figure 4.2).
• Unstable boundary layer
Incoming shortwave radiationwill heat the Earth surface which will heat
and lift the surface-layer air: convection (thermals) will start to emerge.
In this manner, the boundary layer becomes unstable. Normally, there is
a stable layer on top of the boundary layer, which limits the growth of
convection (see figure 4.3). When convective cells penetrate into a stable
layer, they will lose their kinetic energy and won’t ascend any further.
These convective cells mix the air of the boundary layer with the air
above. In this way, the thickness of the boundary layer increases. This
process is called entrainment. The strong turbulence makes the mixing
effective, resulting in a rather uniform distribution of momentum, heat
and water vapour.
A (moist) convective cell that reaches its LCL (Lifting Condensation
Level) will condensate and form a cumulus cloud. As soon as the cell
reaches its LFC (Level of Free Convection) the clouds can grow upward
until the next stable atmospheric layer. The limit of this entrainment pro-
cess is at the level of the tropopause, an extremely stable layer at a height
of about 10-15 kms.

33
Figure 4.2: Typical diurnal progress of the boundary layer over a land surface.
(From: Stull, An Introduction to Boundary-Layer Meteorology, 1988).

Figure 4.3: Characteristic profiles of average potential temperature, wind speed,


vapour and a random trace gas in an unstable boundary layer (From: Stull, An
Introduction to Boundary-Layer Meteorology, 1988).

34
Figure 4.4: Typical profiles of potential temperature and wind speed in a stable
boundary layer (From: Stull, An Introduction to Boundary-Layer Meteorology,
1988).

• Stable boundary layer


A stable boundary layer emerges when the net longwave radiation of the
Earth surface is negative, which cools down the air above it. Vertical
motions are often hampered in a stable boundary layer, since the (nega-
tive) buoyancy slows them down. The turbulent kinetic energy and the
length scales are much smaller than in an unstable boundary layer. The
turbulence is not well-structured. The height of a stable boundary layer is
determined by the balance between turbulence production and dissipation.
Since turbulence is suppressed, little mixing takes place in the boundary
layer. Gradients of heat, momentum, and water vapour are therefore much
larger (see figure 4.4).

4.2 The surface layer; Monin-Obukhov theory


The Monin-Obukhov length

In a non-neutral boundary layer, the surface heat flux (wθ)0 is not zero. To-
gether with the variables z and u? , the surface heat flux, or rather the buoyancy
g g
θ0 (wθ)0 plays a significant role. Since u? and θ0 (wθ)0 can be more or less con-
stant over an hour or longer, they are considered to have a big influence on
the dynamical structure of the surface layer. Based on these variables, we can
derive a stability parameter, the Monin-Obukhov (MO) length scale L:

u? 3
L=− .
κ(g/θ0 )(wθ)0

The turbulent kinetic energy equation (3.24) can be simplified to

35
∂U g
0 = −uw + wθ − ,
∂z θ0

and it can be made dimensionless through multiplication by κ z/u3? :

uw κ z ∂U z
0=− − − φ .
u2? u? ∂z L

We used the definition of L to arrive at the latter equation, and denoted the
dimensionless dissipation of energy by φ . Since in the surface layer −uw ≈ u2? ,
we find:

z
0 = φm − − φ ,
L
in which we wrote the dimensionless wind speed profile as

κ z ∂U
φm = . (4.1)
u? ∂z
In this simplified form, the energy equation can be interpreted as a balance
between mechanical and buoyancy production on the one hand, and the dissi-
pation of energy on the other. The equation is expressed in only a few scales
(z, L, and u? ). From dimensional analysis, we can also define a dimensionless
temperature gradient φh :

κz dΘ
φh (z/L) = . (4.2)
θ? dz

The temperature scale θ? is defined as θ? ≡ −wθ0 /u? .


The functions φm and φh have been determined experimentally. In case of
instability (z/L < 0), the following relations have been found:

φm = (1 − 16z/L)−1/4
φh = (1 − 16z/L)−1/2

and for the stable case (z/L > 0):

z
φm = φh = 1 + 5 .
L
An analogous function for the water vapour profile is often set equal to the
expression for φh .

36
Figure 4.5: A summary of dimensionless gradients of wind speed (a) and tem-
perature (b) as they have been experimentally determined for the surface layer
(From: Yaglom, blm 1977).

Upon integration of (4.1) and (4.2), we obtain

u?
U (z) = {ln(z/z0 ) − Ψm (z/L)}
κ
θ?
Θ(z) − Θ0 = {ln(z/z0 ) − Ψh (z/L)}
κ
If z/L < 0, the expressions for Ψm and Ψh become

1 + x2
   
1+x
Ψm = 2 ln + ln − 2 tan−1 x + π/2
2 2
 
1+y
Ψh = 2 ln ,
2
and otherwise (if z/L > 0)

Ψm = Ψh = −5z/L,
where x = (1 − 15z/L)1/4 and y = (1 − 9z/L)1/2 (see figure 4.5).
If, for whatever practical reason, a power law for the wind speed profile is
preferred, the exponent p is derived from the previous equations using (3.21):

φm (z12 /L)
p= ,
ln z12 /z0 − Ψ(z12 /L)

where z12 = z1 z2 .

Turbulence in the surface layer

37
The following (empirical) relations hold in a stable, flat surface layer:

σu /u? = 2.39 ± 0.03


σv /u? = 1.92 ± 0.03
σw /u? = 1.25 ± 0.03

In unstable conditions, this relation becomes

 1/3
σw 3z
= 1.25 1 − .
u? L

In the unstable situation, the horizontal-velocity fluctuations are too strongly


influenced by variations on a larger scale. The Monin-Obukhov theory starts to
lose its validity. For the temperature variance we have
z −1
σθ /|θ? | = 0.95( ) 3.
−L
Expressions for the unstable situation, that also apply to the surface layer, will
be given in section 4.4. (see also Garratt section 3.5)

Spectra

There exists a plethora of spectra and spectral data in the surface layer, for
example in ?. For our purposes, it suffices to know that

f Su (f ) 102 n
=
u2? (1 + 33n)5/3
f Sv (f ) 17 n
=
u2? (1 + 9.5n)5/3
f Sw (f ) 2.1 n
= ,
u2? (1 + 5.3n)5/3
(4.3)

in which n is the dimensionless frequency (n = f z/U ).

4.3 The neutral boundary layer


The Ekman layer

In this section, we will consider a neutral, homogeneous and stationary bound-


ary layer. Unlike previously, the Coriolis force will also be taken into account.
We will assume that the flow above the boundary layer is geostrophic, i.e. the

38
Figure 4.6: Force balance in the Ekman layer. P is the pressure gradient, Co
the Coriolis force, and F r is friction.

pressure gradient and the Coriolis force balance each other. From the momen-
tum equation, Eq. 3.3, neglecting the buoyancy term, it follows that

∂uw
0 = f (V − Vg ) −
∂z
∂vw
0 = −f (U − Ug ) − (4.4)
∂z
(see figure 4.6). Ug and Vg are the components of the geostrophic wind, by
definition being equal to the wind speed while neglecting the turbulent stress in
(4.4). These components are expressed as

1 ∂Π
Ug = −
ρ0 f ∂y
1 ∂Π
Vg = .
ρ0 f ∂x
The Coriolis parameter, f , equals 2Ω sin φ, where φ represents the latitude, Ω
the angular velocity (rad/s) of the Earth axis. In order to solve eqs. (4.4),
K-theory is applied. Suppose that

∂U
uw = −Km
∂z
∂V
vw = −Km .
∂z
After substitution into (4.4), we obtain

∂2U
0 = f (V − Vg ) + Km
∂z 2
∂2V
0 = −f (U − Ug ) + Km 2 ,
∂z

39
Figure 4.7: The Ekman spiral, or the Leipzig wind profile (Mildner, 1932).

with the solution


U = Ug (1 − e−γz cos γz)
V = Ug e−γz sin γz. (4.5)
p
(γ = f /2Km ).


These equations are solved by multiplying one of them by −i (= −1) and
add it to the other. Using the substitution w = U − Ug + i(V − Vg ), the
result is
∂2w f
= −i w,
∂z 2 Km
In a coordinate system where G is chosen along the x-axis (so that Vg = 0;
see figure 4.7), the boundary conditions for w are given by w = −Ug at
z = 0, and w = 0 at z = ∞. The solution which satisfies the boundary
conditions is
w = −Ug exp (i − 1)γz,
p
with γ equal to f /2Km . Equating the real and imaginary parts yields
the solution (4.5).

The Ekman spiral is only sporadically observed, since the conditions of the
theoretical derivation are seldomly met in practice. The assumption that Km
is constant in the entire layer is incorrect, as are the assumptions of first-order
closure. (For example, Km would rather be proportional to z close to the
surface).
Using the parameter γ defined in the box above, we can define the boundary
layer height, for example as the height at which the wind speed is within 5 % of
the geostrophic wind. In this way, we can choose a boundary layer height like

h ∝ (Km /f )1/2 .
If, on the other hand, the eddy diffusivity parameter is supposed to be the
product of a length scale (h) and a velocity scale (u? ), Km = hu? , then the
boundary layer height can be expressed as:

40
h = c1 u? /f. (4.6)

If we substitute the values c1 = 0.3, f ∼ 10−4 s−1 , and u? = 0.3m/s, then the
neutral boundary layer height amounts to approximately 1000 m.
The equations for the Ekman profile can now be rewritten by using (4.6):

√ p
U − Ug = −Ug e− c1 /2 z/h cos( c1 /2 z/h)
√ p
V = Ug e− c1 /2 z/h sin( c1 /2 z/h).

As a result, the wind profile can be written as the product of the geostrophic
wind speed and dimensionless functions of z/h. This solution doe not hold for
the surface layer, where a logarithmic wind speed profile is found, both in theory
and in practice. Apparently, other parameters apply close to the surface. We
will come back to this is the next section.

Resistance laws

In this paragraph, we will discuss the relation betwen the surface layer param-
eter u? , and the external pressure gradient G. A stationary, homogeneous and
neutral boundary layer is considered. Resistance laws for non-neutral boundary
layers will be shortly touched later on.
The starting point is again a stationary, homogeneous and neutral boundary
layer. The negative x-axis is directed towards the shear stress. The equations
of motion are

∂uw
0 = f (V − Vg ) −
∂z
∂vw
0 = −f (U − Ug ) − , (4.7)
∂z
The boundary conditions that apply are given in table 4.1. Note that the
lower boundary of the layer is denoted by z = z0 , the roughness length. We
are looking for a relation between
q the friction at the Earth surface (u? ), and
the geostrophic wind speed G ≡ Ug 2 + Vg 2 . Dimensional analysis shows that,
apart from G/u? , only one other dimensionless parameter can be found, namely
Ro = u? /f z0 , the so-called friction Rossby number, implying that

Ug
= kx (Ro) (4.8)
u?
Vg
= ky (Ro)
u?
where kx,y, are unknown function of ρ0 .

41
Table 4.1: Boundary conditions

z = z0 z=h
U =0 U = Ug
V =0 V = Vg
−uw = u? 2 uw = 0
vw = 0 vw = 0

These relations are called resistance laws. We can construct these by adapting
the solution for z ≈ h to the solution for z = z0 . This procedure is called
’matching’. Let us define two dimensionless height coordinates as η = z/h, and
ξ = z/z0 where h = u? /f , the boundary layer height. We implicitly assume
that z0 << h = u? /f , being equivalent to taking Ro → ∞. When z ≈ h, the
dynamical equations (see 4.4) in this coordinate system are:

U − Ug dvw/u? 2
=− ≡ gx (η) (4.9)
u? dzf /u?
V − Vg duw/u? 2
= ≡ gy (η). (4.10)
u? dzf /u?

In case z0 . z << h and τ (−uw0 , 0) (note that the shear stress at the surface
is chosen along the negative x-axis):

U
= fx (ξ) (4.11)
u?
V
≈ 0. (4.12)
u?

If we subtract (4.9) from (4.11), we are left with

Ug
= fx (ξ) − gx (η), (4.13)
u?
which can be proven to be equal to
Ug 1
= (ln Ro − A0 ) . (4.14)
u? κ

42
Proof: using Eq. 4.8, Eq. 4.13 can be written as
kx (Ro) = fx (Ro η) − gx (η). (4.15)
Note that ξ = Ro η. Now, f, g and k must be continuous and logarithmic
functions. This can be demonstrated for f by first differentiating (4.15) to
Ro and subsequently to η. We then obtain the differential equation
0 00
0 = fx + Ro η fx ,
that has a logarithmic solution. A similar procedure shows that also g is
a logarithmic function. Substituting the logarithmic solutions into (4.9)
yields:
1 Ug 1
ln ξ − = ln η + a constant.
κ u? κ
which can be rewritten to
Ug 1
= (ln Ro − A0 ) , (4.16)
u? κ
where A0 is a constant.

For Vg we have

Vg
= ky (Ro)
u?
which for Ro → ∞, must converge to a constant. If we assign to this constant
the value −B0 /κ, the equation can be written as

Vg B0
=− (4.17)
u? κ
q
Combining the above-mentioned equations and substituting G = Ug 2 + V g 2
gives:

u? κ
=p . (4.18)
G (ln Ro − A0 )2 + B02

This is the resistance law under neutral circumstances. A0 ≈ 1.7 and B0 ≈ 4.5.
The relation between frictional velocity and the geostrophic wind forcing is
only roughly approximated by resistance laws. If the layer is not neutral, the
’constants’ A0 and B0 in (4.18) are simply assumed to relate to h/L, implying
the following empirical relation (see figure 4.8):

u? κ
=p ,
G (ln Ro − A(h/L))2 + B 2 (h/L)
p
in which h stands for u? /f , Ro = u? /(z0 f ) and τ = uw2 + vw2 .

43
The angle between the geostrophic wind and the shear stress at the surface is
given by α = arctan(Vg /Ug ). Substitution leads to

B
α = arctan .
ln Ro − A
Thus, the angle α is a measure for the wind turning in the boundary layer. The
constants A and B have been determined experimentally (see figure 4.8).

4.4 The convective boundary layer


The dynamics in the convective boundary layer are determined by wind shear
and convection (wθ > 0). As soon as the convection becomes so strong that
the contribution of the wind shear to the turbulence becomes negligible, the
convection is usually called ’free convection’. The velocity scale u? is then of
minor importance. Relevant parameters in this layer are the boundary layer
height (h) and the surface heat flux. Using these, we can define a convective
velocity and temperature scale as follows:

 1/3
g
w? = (wθ)0 h
θ0
T? = (wθ)0 /w? .

Besides these, z and h are the length scales. The time scale h/w? amounts
to approximately 1000 s, which is one order of magnitude smaller than 1/f .
The influence of the Coriolis force on the dynamics in the convective boundary
layer is thus only important in processes that play at a scale of hours or more.
Concerning the wind profile in the convective layer, it can only be remarked
that the wind profile gradient will be small due to the strong mixing that will
occur. Usually, a uniform (constant) wind is assumed.

Turbulence characteristics

In the convective boundary layer, the following expressions are used. They are
derived using dimensional analysis, the constants follow from observations:

σw /u? = (−z/L)1/3 ,

and in the upper part of the boundary layer:

σw /u? = (−h/L)1/3 .

44
Figure 4.8: The functions A and B are dependent on the stability parameter µ ≡
h/L, given that L ≥ 0. The curves numbered 1-5 are based √ on empirical data.
3 µ √
The curves labeled 6 are the relations A(µ) = 1.7 + ln(1 + 3.4 ) − 2.55 µ and

3 µ
B(µ) = 4.5 + 1.7 . The hatched areas represent the spread in the observations.
(from: ?).

45
Figure 4.9: Schematic representation of the flow, temperature and eddy struc-
ture in the convective boundary layer.

The horizontal variances are expressed empirically as follows:

 1/3
σu σv h
≈ = 12 − 0.5
u? u? L
These equations also apply to the surface layer.

For the temperature variance we have


|σθ |  z −1/3
= 1.34 ,
θ? h
which applies in the lower half of the CBL.

Dynamics

In a convective boundary layer, the air is lifted locally (thermals). When these
pockets of air reach the uppermost boundary layer, they penetrate into a more
stable part of the atmosphere. As a consequence, they will start to descend
again. The net vertical transport is approximately zero. This process also causes
warmer air to be pulled into the turbulent layer. This is called entrainment. In
this way, the boundary between strong and weak turbulence, the inversion layer,
can move upward. The mechanical turbulence may also contribute to the decay
of the stable layer above, and thus giving way to the growth of the mixing layer.
Because the convection is controlled by the heat flux at the surface, which has
a strong diurnal cycle, this process is highly unstationary. In the course of the
morning, a convective layer will build up if sufficient solar radiation is available.

46
Figure 4.10: Structure of the unstable boundary layer (From: Holtslag and
Nieuwstadt, 1986).

Its vertical size will reach a maximum at the end of the afternoon. For this
growth and decay, a simple mathematical models exists, that is based on three
simplified dynamical equations: one for the boundary layer temperature (θ), the
inversion strength (∆), and the boundary layer height (h). In figure 4.11a two
subsequent temperature profiles, separated by a time difference ∆t, in the CBL
are drawn. The surface ∆ δh represents the amount of heat that is captured by
the boundary layer by entrainment, so that the total amount of heat absorbed
by the boundary layer in a period ∆t equals

hδθ = (wθ)0 δt + ∆ δh, (4.19)

where the surface heat flux, (wθ)0 , is supposed to be constant. In a differential


form:

∂θ ∂h
h = (wθ)0 + ∆ . (4.20)
∂t ∂t
From geometric considerations, the two subsequent profiles imply:

∂∆ ∂h ∂θ
=γ − , (4.21)
∂t ∂t ∂t

47
∂θ
where γ is the temperature gradient above the inversion ( ∂z ). Concludingly, we
will assume that the ratio between entrainment and surface heat flux, ε, remains
constant, leaving us with

∂h
∆ ≡ −(wθ)i = −ε(wθ)0 . (4.22)
∂t
These equations, containing the unknown variables h, ∆ and θ provide an ex-
pression for h:

 2  
h h 2(1 − 2ε)(wθ)0
− 1 − 2ε ( )1/ε − 1 = t. (4.23)
h0 h0 γh20
In this equation, h(t) represents the inversion height, h0 its initial value, γ the
initial temperature gradient (= dθ/dz), and ε is the ratio between the heat flux
at the top and the bottom of the boundary layer (see figure 4.11), (wθ)i /(wθ)0 .
In the derivation of (4.23), it is assumed that ∆(t = 0) = 0.
If the mechanical turbulence cannot be neglected as a fraction of the thermally
produced turbulence, the entrainment process is parameterised using:

u3?
(wθ)i = ε (wθ)0 + εm .
hg/θ0
The constants ε and εm have the values 0.2 and 2.5, respectively.

4.5 The stable boundary layer


In a stable boundary layer, the turbulence is suppressed by Archimede or buoy-
ancy forces. Especially during nighttime above a land surface, the cooling of
the Earth surface induces a negative heat flux (directed to the surface). The
turbulent energy will decrease dramatically, and vertical motion in the bound-
ary layer will be suppressed. This process will ultimately result in a cooled
boundary layer and a stable potential-temperature profile (dθ/dz > 0).

Nocturnal boundary-layer height

The vertical exchange being small, the growth of the stable layer is much weaker
than the growth of an unstable layer. Observations indicate that a stable layer
rapidly builds up in the beginning of the evening, which remains approximately
the same during the rest of the night. The thickness of the layer can be estimated
using dimensional analysis:

p
h = c3 u∗0 L/f .

48
Figure 4.11: Geometry of an expanding boundary layer. In the uppermost panel,
the idealized potential temperature profile is drawn at two subsequent moments.
In the rightmost picture, the heat flux as a function of height is depicted. The
entrainment process is symbolically depicted by the hatched area. The middle
and lower panes give a more realistic picture of the entrainment process (From:
?).

49
Figure 4.12: Structure of the stable boundary layer (From: Holtslag and Nieuw-
stadt, 1986).

The constant c3 approximately equals 0.4.


A more general estimate would be the solution of

 2
f h h N h
+ + = 1,
Cn u? Cs L Ci u?
q
g ∂Θ
where Cn,s,i are 0.5, 10 and 20, respectively, and N ≡ θ0 ∂z , the Brunt-Vaisala
frequency.

The local length scale (Λ)

In a stable boundary layer, the vertical exchange is small. As a consequence,


the dynamics are governed by local parameters. We will therefore introduce the
local Monin-Obukhov length:

τ 3/2
Λ≡− ,
κ(g/θ0 )wθ

where τ and wθ are now dependent on z.


As a result of the high stability, the vertical exchange over large distances is
very limited. Small eddies having a more isotropic structure dominate the tur-
bulent processes. The y-component of the shear stress will be significant as well.

50
p
For that reason, the total shear stress τ (= uw2 + vw2 ) is preferred to the
frictional velocity u? .
Vertical exchange within the stable boundary layer can be so small that the
distance to the surface becomes a meaningless parameter. In theoretical treat-
ments, the Richardson number is assigned its maximal constant value of 0.25.
Another occurence is called intermittant turbulence. In that case, flow can
become laminar at times. Wavelike processes are sometimes observed as well.

Wind maximum in the stable boundary layer

It regularly happens that, especially during nighttime above a land surface,


the wind speed relatively close to the surface can be 10-15 m/s, while at the
surface, it is very calm. This is when the so-called low-level jet occurs. It is a
consequence of the transition from a convective turbulent boundary layer to a
stable boundary layer at nightfall. The boundary layer rapidly becomes stable,
reducing the vertical transport of momentum strongly. As starting equations,
we will take the time-dependent boundary layer equations in a horizontally
homogeneous situation:

∂U
= fV
∂t
∂V
= −f (U − G).
∂t
We chose a coordinate system with the positive x-axis directed towards the
geostrophic wind, implying G = (G, 0)).
Multiply the latter equation by i, add both, and introduce the new variable
w ≡ U + iV . This will result in:

U − G = V0 sin f t + (U0 − G) cos f t (4.24)


V = V0 cos f t − (U0 − G) sin f t.

As starting condition, it is assumed that at t = 0 : (U, V ) = (U0 , V0 ). Squaring


and adding these equations will lead to (U − G)2 + V 2 = (U0 − G)2 + V0 2 .
This equation can be represented by a circle in a (U,V)-coordinate system, as
depicted in figure 4.13.
The wind vector will move in a clockwise direction over the circle in a pe-
riod 2π/f (approx. 15 hours at middle latitudes). It can thus become super-
U | > G). q
geostrophic (|U
The radius of the circle ( V0 2 + (U0 − G)2 ) increases when the difference be-
tween U 0 and the geostrophic wind is larger (close to the surface). The boundary
condition U = 0 at z = 0, will cause a maximum for the wind speed at a certain
height (see figure 4.14).

51
V

(U0,V0)

G U

Figure 4.13: Graphical representation of the behaviour of the wind vector during
a low-level jet. The initial situation is depicted (subscript 0). The wind vector
will start to move clockwise over the circle, and will in the course of the process
become larger than the geostrophic wind G .

Figure 4.14: Simple representation of the low-level jet.

52
4.6 Summary
An overview of the most important parameters in the atmospheric boundary
layer is given in the following table.

The wind profile

It is not easy to formulate general expressions describing the wind profile in the
boundary layer. It does turn out, however, that the wind in the surface layer
can be extrapolated into the layers above satisfactorily. An empirical relation
that performs quite well is that from ?:

 
ln z2 /z0 − Ψm (z2 /L)
U2 = U1 .
ln(z1 /z0 ) − Ψm (z1 /L)

This expression can be used up to 200 m, both in neutral and in diabatic situ-
ations. The deviations from this equation are usually less than 1 m/s.

Table 4.2: Anatomy of the atmospheric boundary layer.

STABILITY PARAMETERS HEIGHT INTERVAL


viscous sublayer all u? , ν 0 − ν/u?
turbulent sublayer all u? , z0 ν/u? − z0
surface layer all z, u? , θ? z0 << z < L
quasi-neutral boundary layer neutral h, z, u? , f 0.1h - h
convective boundary layer unstable h, w? 0.1h - h
free convection layer unstable z, w? 0.01h - 0.1h
local scaling layer stable z, τ, wθ 0.1h-0.5h
z-less scaling layer very stable τ, wθ 0.1h-h
intermittency layer very stable N 0-h
entrainment layer stable ∆U, ∆θ, u? , w? 0.8h - 1.2h

53
Chapter 5

EXCHANGE OF HEAT
AND WATER VAPOUR

5.1 The surface-energy budget


Radiation is the driving force of heating and evaporation at the Earth sur-
face (see figure 5.1). It is composed of longwave radiation originating from the
Earth and the atmosphere itself, and of shortwave solar radiation. Furthermore,
a distinction is made between incoming (downward) and outgoing (upward) ra-
diation. The incoming shortwave radiation (S ↓) equals the sum of the directly
incoming solar radiation and the diffuse radiation due to scattering processes
(by clouds and atmospheric molecules). The outgoing shortwave radiation (S ↑)
is the upward reflected shortwave solar radiation. The longwave incoming radi-
ation at the Earth surface (L ↓) has its origin in the atmospheric water vapour.
It is mainly determined by the amount of water vapour and the temperature
of the atmosphere. The outgoing longwave radiation at the Earth surface is
directly linked to the surface temperature of the Earth.
The net radiation (RN ) at the surface is defined positive when the radiation is
directed towards the surface, and given by the sum of all radiation components:

RN = S ↓ −S ↑ +L ↓ −L ↑ .

The ratio of outgoing and incoming shortwave radiation is called the albedo, a,
in formula:
S↑
a= .
S↓
The albedo (reflectivity) of a grass surface is approximately 0.2, it is 0.1-0.5 for a
water surface and 0.7-0.95 for fresh snow. The incoming and outgoing radiation
exhibit a daily cycle. For a clear day, the typical behaviour of RN is shown in
figure 5.3.

54
Figure 5.1: Longwave and shortwave radiation in the atmosphere.

Rn

λE

Figure 5.2: The energy balance.

55
At climatological time scales, we assume that the energy budget at the Earth
surface is approximately in a state of thermal equilibrium. The components of
the energy balance are the net radiation (RN ) on the one hand, and on the other
hand the energy that is removed by flow or diffusion: the sensible heat flux (H),
evaporation (E), and the surface heat flux (G). The total energy balance at the
Earth surface can then be expressed as:

RN = H + λE + G, (5.1)

(see figure 5.3), where λ represents the latent heat of evaporation, and E is the
water vapour flux. The terms on the right-hand side of 5.1 are given by

H = ρ0 cp (wθ)0
E ρ0 wq 0
=
 
∂Ts
G = k ,
∂z 0

where k is the molecular diffusion coefficient of the soil. RN has a daily cycle
above a land surface, see fig. 5.3. As a result of this, λE, H and G also exhibit
a daily cycle. During the day, RN is positive. A typical value at a clear summer
day would be 400 Wm−2 . H is also positive during daytime, as well as λE. G is
positive, but very small. It amounts to around 10 % of the net radiation. During
nighttime, RN is negative, typically ∼ 50 Wm−2 in a clear night. Because there
is only little evaporation during night (rather would there be dew formation),
λE is practically zero. At night, H is negative, while G can be relatively large
(50 % of the net radiation).
Several possibilities exist to determine λE and H from measurements of other
quantities. For example, temperature and vapour profile measurements can be
used, or measuring all other components of the energy balance. Other estimates
have been made by Penman and Monteith. A summary will be given below.

5.2 The profile method


When using the profile method, temperature and humidity are measured at two
levels above the surface. Wind speed is measured at one level, and the surface
roughness is supposed to be known.
Using Monin-Obukhov similarity theory, we find that:

u?
U (z2 ) = {ln(z2 /z0 ) − Ψm (z2 /L)}
κ
θ?
Θ(z2 ) − Θ(z1 ) = {ln(z2 /z1 ) − Ψh (z2 /L) + Ψh (z1 /L)} (5.2)
κ
q?
q(z2 ) − q(z1 ) = {ln(z2 /z1 ) − Ψq (z2 /L) + Ψq (z1 /L)}.
κ

56
Figure 5.3: Daily cycle of the components of the energy balance on a clear day
over a maize crop. Source: ?.

T2
u2
q2

T1
q1

z0

Figure 5.4: The profile method experimental setup. At level 1, temperature and
humidity are measured. At level 2, also the wind speed is recorded.

57
Figure 5.5: The setup for the Bowen-ratio method. Net radiation, surface heat
flux, temperature, and humidity (at two levels) are measured.

By measuring wind, temperature and humidity at two levels, the three unknown
parameters (u? , θ? and q? ) can be determined. From these equation, the heat
and water-vapour fluxes immediately follow:

H = −ρ0 cp u? θ?
E = −ρ0 u? q?

5.3 The energy-balance method


This method will also exploit the energy balance equation(Eq. 5.1). The ex-
perimental setup is as follows: at two levels, temperature and humidity are
measured. At one level, net radiation is measured. At the surface, the surface
heat flux G is recorded.
We introduce the Bowen ratio, β ≡ H/(λE). From the energy balance, it follows
that:

1
λE = (RN − G)
1+β
β
H = (RN − G).
1+β
The Bowen ratio is calculated from the temperature and humidity-difference
measurements. From the definition of the Bowen ration, it can be concluded
that:

∆θ
β=γ ,
∆q

58
cp
with γ = λ and where (5.2) and (5.3), as well as the assumption that Ψm ≈ Ψq .

5.4 Estimating the evaporation


In this section, we will present some methods to calculate the evaporation. This
can applied usefully in, for example, agriculture. In all different formulations,
the type and state of the surface play an important role: the Penman formulae
can be applied for a wet surface without vegetation, and the Penman-Monteith
formulae can be used for vegetation. Other practical evaporation formulae have
been developed by Priestly and Taylor. All of these formulations will be dis-
cussed below.

The Penman method

It is assumed that RN − G is known, and that the evaporation pressure of a wet


surface equals the saturation value, q0 = qs (T0 ). The temperature and humidity
of the atmosphere and the surface, as well as the wind speed, are supposed to
be known. From (5.2), the following applies in a stratified surface layer:
θ? ∆θ
=
u? U
q? ∆q
= .
u? U
Note that ∆q = q1 − q0 , which combines with the definitions of H and λE to
give:

u2?
H = −ρ0 cp ∆θ (5.3)
U
u2?
λE = −ρ0 λ ∆q . (5.4)
U
If we define a drag coefficient as cd ≡ (u? /U )2 , we get:

H = −ρ0 cp ∆θ cd U (5.5)
λE = −ρ0 λ ∆q cd U. (5.6)
Using the following equations, we can obtain a simple estimate of the evapo-
ration rate. We will use Taylor expansion (subscripts 0 and 1 refer to the two
measurements levels):
∂qs
qs (T1 ) = qs (T0 ) + (T1 − T0 )( )0 + . . . , (5.7)
∂T
∂qs
which we’ll rewrite using the notation s ≡ ∂T 0 , and the substitution T1 − T0 ≈
θ1 − θ0 ≡ ∆θ = −H/(ρ0 cp cd U ):
∂qs
qs (T1 ) − qs (T0 ) = ∆θ ,
∂T 0

59
yielding
sH
qs (T0 ) = qs (T1 ) + . (5.8)
ρ 0 cp cd U

The difference in relative humidity, ∆q, can now be written as:

sH
∆q = q(T1 ) − qs (T0 ) = q(T1 ) − qs (T1 ) − . (5.9)
ρ0 cp cd U

Substituting this equation into(5.4), replacing H by RN −G−λE and reordering


yields:

γ s
λE = ρ0 λcd U [qs (T1 ) − q(T1 )] + (RN − G). (5.10)
s+γ s+γ

The term cp /λ is denoted as γ. The first term on the right-hand side in (5.10) is
called the aerodynamical evaporation, whereas the second term represents the
so-called thermodynamical evaporation.
In order to calculate the evaporation, in addition to RN − G, both temperature,
humidity and wind speed have to be known at the reference level (1). These are
standard meteorological data that are collected every 1 to 3 hours worldwide.
The Penman-Monteith method

This approach is best suited for a surface with complete coverage of vegetation.
Heat exchange and evaporation within the vegetation layer is not taken into
account. On the other hand, evaporation is assumed to depend on the vegetation
type. The crops have their own resistance against evaporation through stomata
(pores) in the leaves. The vapour level is saturated inside the stomata of the
crop leaves. Depending on the amount of available water, the stomata will be
more or less closed. In this way, there will be more or less resistance against
evaporation. This does not apply to the exchange of heat. If, analogous to
Ohm’s law, the heat and vapour difference is written as the product of the heat
(or vapour) flux and a resistance, we can rewrite (5.6) as

ρ0 cp ∆θ
H = − (5.11)
rh
ρ0 λ ∆q
λE = − ,
rh
where rh , the ’resistance’, equals 1/(cd U ). As a next step, we have to realize
that transport of moisture within the crops is an extra resistance. This can be
easily accounted for by replacing rh with rh + rs in the second equation, where
rs is the so-called crop resistance. Using the procedure of the previous section,
we get an expression for the evaporation:

60
γδ sδ
λE = − ρ0 λcd U [q(T1 ) − qs (T1 )] + (RN − G), (5.12)
sδ + γ sδ + γ
where δ = rh /(rh + rs ).
If sufficient water is available in the upper soil layer (where the plant roots are),
the plant will optimize its evaporation by minimizing the crop resistance. Each
crop is characterized by its own crop resistance. The maximum evaporation is
the potential evaporation rate of the crop. At this rate, the crops grow fastest.
This is an important quantity in agriculture: if there is drought, the crops have
to be irrigated until the potential evaporation rate is reached. Any irrigation
surplus is useless since the plant has already reached the potential evaporation
rate.
The Priestley-Taylor method

The Priestley-Taylor evaporation method makes use of the Penman-Monteith


equation (5.12). The crop resistance is supposed to be zero, and the relative
humidity is chosen at 100 %: q(T1 ) = qs (T1 ) . From (5.12), it can be deduced
that the equilibrium evaporation is given by:

s
λE = (RN − G). (5.13)
s+γ

5.5 Air-Sea interaction


The surface layer theory also applies over a water surface. There are some
aspects that have to be kept in mind, however. When estimating the stability
of the surface layer, evaporation plays an important role, since it influences the
atmospheric density (at a constant temperature, humid air has a lower density
than dry air). To simulate this density effect, a somewhat higher temperature
is assigned to humid air, related to the water vapour content in the humid air.
This temperature is called virtual temperature, Tv . Without deriving it here,
the following expression is generally valid for the virtual temperature:

Tv = T (1 + 0.61 q) ≈ T + 0.61θ0 q, (5.14)

where q is the water vapour mixing ration (in kg/kg), and θ0 is a reference
temperature (e.g. 287 K). In the surface layer, the same formulation holds for
the virtual potential temperature. θv :

θv = θ(1 + 0.61 q) ≈ θ + 0.61θ0 q, (5.15)

If we replace θ by θv in all Navier-Stokes equations, we also take dynamical


effects of humid air into account. This excludes the mechanism of condensa-
tion! Should one want to formulate a stability criterion, it is again possible

61
to use (wθ)v > 0 for an unstable atmosphere, etcetera (cf. section 4.1). The
Richardson number is now given by:

g
θ0 wθv
Rif = . (5.16)
ui uj ∂U
∂xj
i

(see Eq. (3.25)). The virtual-temperature flux follows from (5.15):

wθv = wθ + 0.61θ0 w q.

In order to determine the potential temperature, the water-vapour flux wq has


to be known. It can be estimated from the water-surface temperature and the
humidity (see section 5.4). The atmosphere will usually be neutral above a sea
surface, since the air flow above it will gradually adapt to the water temperature,
which only varies slightly in time and place.
The exchange above a sea surface can also be crudely estimated by using drag
coefficient relations, where transport is proportional to temperature and to hu-
midity differences. If we take the wind shear stress ρ0 u? 2 and the drag coeffi-
cient cd ≡ (u? /U10 )2 (using wind speed at 10 m), the momentum exchange can
be described by:

τ ≡ ρ0 u? 2 = ρ0 cd U10 2 .

Measurements show that the drag coefficient is ≈ 1.3 10−3 for neutral conditions.
In a stable surface layer, we find lower values, and higher values can be found
under unstable circumstances. A simple stability correction is cd = cdn (1 − Ri),
where the Richardson number can be estimated from:

g
(T10 − Ts )
Ri = 10 θ0
U10 2
again assuming the reference height at 10 m.
Similar to a land surface, the heat and water-vapour transport can be estimated
using:

H = −ρ0 cp ch U10 (T10 − Ts ) (5.17)

and

λE = −ρ0 λ cq U10 (q10 − qs ) (5.18)

The coefficients ch and cq both adapt a constant value of 1.4 10−3 .

62
Unlike at open sea, stability effects do play a role near the coast. We will come
back to that later. A complicating difference between land and sea surfaces is
the waves at the sea surface. The observed wave pattern usually is a result of
both wind-generated waves (wind waves) and waves that have been generated
elsewhere (swell). On top of the wave pattern lies a turbulent layer, again
characterized by a vertical momentum flux ρ0 uw, analogous to the situation
above a land surface. Also at sea, a logarithmic wind profile in the surface layer
is a suitable assumption:

U (z) 1 z
= ln + C. (5.19)
u? κ r
(see 3.14). We chose on purpose another parameters , since the length scale r
and the constant C will depend on the state of the sea surface and atmospheric
turbulence. Because the exact theory is very complicated, we will use scale
analysis. A spectral analysis of wind waves can give us a wave number kp that
corresponds to the characteristic wave length of the wind waves, 2π/kp (e.g. the
value where the spectrum
p shows a maximum). In this way, the phase velocity Cp
is determined: Cp = g/kp . Its dimensionless counterpart Cp /u? is sometimes
called wave age.
The average wave height of a stationary windy sea, Hav , can be derived using
scale analysis: Hav ∝ u2? /g.
Another important wind wave parameters is the so-called fetch, the length of
the trajectory that a wave has travelled from its ’birth’ to its full-grown size.
There exists a dimensionless relation between the fetch and the phase velocity.
The presence of waves enhances the momentum exchange from the atmosphere
to the ocean. Short, steep waves play a relatively important role, since pressure
forces can easily clamp to the waves (unlike the land surface, where only the
shear stress accounts for the momentum exchange).
Momentum exchange above sea

The functions C and r in (5.19) depend on u? , g and Cp , according to the


previous section.
An alternative formulation of 5.19 would be:

U (z) 1 gz
= ln 2 + C, (5.20)
u? κ u?

where r ∝ u2? /g from dimensional considerations. The proportionality constant


has been included in the new constant C, and may still depend on the wave age
(Cp /u? ). A constant value of 11.3 turns out to apply reasonably well, as will be
shown later.
If C is assumed constant, it is also possible to include C into the logarithm, and
rewrite (3.17) analogous to the land surface case:

63
Figure 5.6: An ensemble of measurements of the wind-shear stress relation.
(a) lower wind speeds, and (b) moderate to strong winds. The dashed line
represents Eq. (5.20). (From ?)

U (z) 1 z
= ln ,
u? κ z0

where z0 = 0.011 u2? /g represents the roughness of a windy sea. This formulation
(Charnock’s Law) combines the parameters r and C without taking into account
their functional dependence on Cp /u? . Under non-neutral circumstances, these
formulations can be adjusted in very much the same way as over a land surface
(see section 4.2).
Charnock’s formulation is pretty succesful, despite its simplicity (see figure 5.6).
We can conclude that Charnock’s Law (including stability correction!) applies
for a relatively large range of wind speeds (6-25 m/s), giving a satisfying relation
between U and u? .
The influence of wave age has been studied in more detail. We then assume
that z0 = Ac u2? /g, where Ac depends on Cp /u? . A large number of experiments
support the applicability of the following empirical relation (see figure 5.7):

 
Cp Cp −2.59 Cp −4.59
Ac = 1.89( )−1.59 1.0 + 47.2 ( ) + 11.8 ( ) . (5.21)
u? u? u?

A simple model for ocean-atmosphere exchange

We will assume a homogeneous, thin, and well-mixed layer (for example, the

64
Figure 5.7: Several measurements of Cp and z0 . The solid line represents the
relation in 5.21. (From ?)

65
Ta qa U
H LE

d
Shallow Lake Ts ρs

Figure 5.8: Schematic representation of ocean-atmosphere exchange.

oceanic mixing layer, or a shallow lake). Heat and water vapour transport are
the most important processes. The water temperature Ts can be described by:

∂Ts
ρs cps d = −(H + λE), (5.22)
∂t
where the subscript s indicates sea-related variables (see fig. 5.8).
Using the definition from the previous section (5.17, 5.18), we know that:

 
λ
H + λE = ρ0 cp ch U10 Ts − Ta + (qs (Ts ) − qa ) . (5.23)
cp
λ
We will introduce the wet-bulb temperature, Tw ≡ Ta − cp (qs (Tw ) − qa ).

This expression is directly obtained by integrating the First Law of ther-


modynamics for an adiabatic process, cp dT + λdq = 0, from the initial
condition T, qa until the state Tw , qs (Tw ). This describes the physical pro-
cess in an air parcel in which liquid water evaporates until it saturates.

Developing qs into a Taylor series

∂qs
qs (Ts ) = qs (Tw ) + (Ts − Tw ) ( )s + . . . , (5.24)
∂T
and substituting into (5.23) yields

∂Ts
ρs cps d = −ρ0 ch U10 (Ts − Tw ) (cp + sλ). (5.25)
∂t
This is a differential equation exhibiting an exponential behaviour if all coeffi-
cients are assumed constant. From (5.25), we will find the time constant τ :

ρs cps d
τ= , (5.26)
ρ0 ch U10 (cp + sλ)

66
which is proportional to d/U10 . The cooling rate is proportional to the layer
thickness, and inversely proportional to the wind speed.
We can also include the heat loss through radiation. The energy budget then
becomes:

∂Ts Ts − Tw
=Q− (5.27)
∂t τ
where Q represents the net longwave radiation term due to radiation loss of the
water layer (∝ Ts4 ) and the atmospheric back-radiation (∝ Ta4 ), implying

∂Ts Ts − Tw
= αTa4 − βTs4 − . (5.28)
∂t τ
Ts4 can be approximated as

Ts4 = (Tw + Ts − Tw )4 ≈ Tw4 + 4 Tw3 (Ts − Tw ),

yielding

∂Ts Ts − Tw
= αTa4 − βTw4 − , (5.29)
∂t τ0
with a time constant τ 0 = τ /(1 + 4βTw3 τ ). The response time turns out to
decrease due to the longwave radiation. The value of τ 0 on mid-latitudes turns
out to be 1 day for a depth of 1 m, and 10 days for a depth of 10 m. At these
depths, water temperatures can apparently fluctuate quite rapidly.

Finally, we will give an example of the effect of a periodic forcing of the water
temperature by the atmosphere, q cos ωt, where q is the temperature forcing
(in Km−1 ):

∂Ts Ts − Teq
= q cos ωt − , (5.30)
∂t τ
where ω may represent the annual cycle of temperature, and Teq the annual
average temperature. The general solution of this equation is:

qτ qτ
Ts − Teq = Ce−t/τ − e−t/τ + p cos(ωt + φ), (5.31)
1 + (ωτ )2 1 + (ωτ )2

where C is an integration constant, and φ = arctan ωτ . If t >> τ , this relation


can be rewritten as

67
Figure 5.9: Phase difference between the annual atmosperic temperature cycle
and the water temperature at two depths (5 and 15 m).

68

Tseq − Teq = p cos(ωt + φ). (5.32)
1 + (ωτ )2

We can conclude that the water temperature is out of phase with respect to
the radiative atmospheric forcing. There are two extreme cases, ωτ → 0 and
ωτ → ∞. In the first case, the phase difference is 0, and in the latter, it is 90o .
In reality, values in between these extremes will be found (see figure 5.9).

69
Chapter 6

HETEROGENEOUS
BOUNDARY LAYERS

In the previous sections, the Earth surface has always been assumed to be flat
and uniform. However, the largest part of the land surface on Earth is not
flat. The geometry of the terrain may be very irregular, and the characteristics
of the surface may change. We will discuss two types of change in terrain
characteristics: water-vapour and heat-exchange characteristics, and roughness
changes.

6.1 Thermal transitions


A thermal transition occurs when the air mass flows from a warm surface to
a cold surface or v.v. As the thermal characteristics of the underlying surface
change, the atmospheric stability can change dramatically. Turbulence and
momentum exchange are influenced by the underlying surface as well. Thermal
transitions typically happen at a land-sea boundary, if warm and convective air
mass flows out over a relatively cold sea (offshore flow or land breeze). A new,
more stable boundary layer will emerge within the existing, unstable boundary
layer. On the other hand, relatively cold air may also flow over a warm sea
surface. The warmer sea water will start a convection process and create a fast-
growing, turbulent internal boundary layer. These processes will of course also
occur over a land surface, when the wind is onshore (sea breeze) (see fig. 6.1).
In
√ both cases, the growth of the internal boundary layer will be proportional to
x, x being the distance to the coast.
Both the diurnal cycle of land temperature and the seasonal cycle will influence
these thermal transition processes. Moreover, the presence of clouds can have a
moderating effect on these processes. In general, the meteorological situation in
a coastal area (over sea and land) will be highly complex and variable. Although
changes in the turbulent structure may be significant, changes in the average
flow field are usually neglected.

70
Figure 6.1: The development of an internal boundary layer (a) onshore wind. A
stable layer will develop into a convective boundary layer over land, hb (x). (b)
The inverse situation, where a stable boundary layer emerges from a convective
layer over sea (From: ?).

71
Figure 6.2: Numerical simulation of land-sea breeze circulation. (from: A.B.C.
Tijm, PhD dissertation, (IMAU) Universiteit Utrecht, 1999)

A typical feature involving thermal transitions is the land-sea breeze circulation,


which can be witnessed when wind speeds are low and when the difference
between land and sea temperatures is large. The different rates of warming
and cooling of both surfaces lead to a sea breeze during daytime, and to a land
breeze at night. This local circulation may extend tens of kilometres from the
coast (see figure 6.2).

6.2 Roughness transitions


Transitions in the terrain roughness are relevant in wind turbine technology,
since large changes in the wind field may arise due to changes in terrain rough-
ness. In this case, we will suppose a sudden transition from one uniform surface
roughness to another. Furthermore, the wind is considered to blow perpen-
dicular to the boundary between both surfaces. We will have to discriminate
between the transition smooth → rough and rough → smooth. In the first case,
the air flow will slow down, in the latter it will accelerate. Initially, this effect
is only felt at the surface, but it will gradually expand to higher altitudes, as
a new internal boundary layer is developed (see figure 6.3). At the boundary
between the surfaces, pressure effects will arise causing vertical accelerations,
being upward if the transition is from smooth to rough, and downward when
rough → smooth. At a certain distance from the boundary, pressure effects are
neglected.

Dynamics of the internal boundary layer

We will make an approximation that is once again based on dimension analysis.

72
Let us assume a neutral boundary layer (subscript 1) having a logarithmic wind
profile:

U (z) 1 z
= ln .
u∗1 κ z01
After the terrain transition (indicated by subscript 2), a new boundary layer will
develop. Its growth in the vertical direction is determined by the new velocity
scale u∗2 . The growth can simply be described by

dh u∗2
=B .
dx U (h)

where x denotes the horizontal coordinate relative to the surface boundary, and
B is a constant (≈ 1.3). If we substitute the logarithmic wind profile of terrain
2 into U (h), a simple differential equation emerges, having an implicit solution
for h(x):

 
h h x
1+ ln − 1 = Bκ
z02 z02 z02

If z < h(x), the wind profile has adapted to the new terrain roughness, whereas
the wind profile remains unchanged for z > h(x). This is schematically depicted
in figure 6.4.
Empirical approximations are usually exploited to smoothen the sudden tran-
sition in the wind profiles (figure 6.4). A fairly satisfying agreement turns out
to exist between measured and modelled profiles, as is shown in figure 6.5. It
is nonetheless difficult to determine the new shear stress u∗2 from the new,
smoothened profiles. The distance from the surface boundary should be larger
in order to do so.

Change of frictional velocity

The change of the frictional velocity is determined by wind tunnel measure-


ments, complemented by numerical simulations. A compilation of measure-
ments and calculations is shown in figure 6.7. The fast adaptation to the new
situation is striking, as is the overshoot that occurs. A simple estimation that
turns out to perform fairly well, is obtained by demanding u1 (h) = u2 (h) at
z = h, yielding:

u∗2 ln(z01 /z02 )


=1− .
u∗1 ln h/z02 )

The ratio will only depend on h(x).

73
Figure 6.3: Simple representation of a transition from a smooth to a rough
surface.

z z z

h(x)
100 100 100

10 10 10

1 1 1

z02 z02
0.1 0.1 0.1

0.01 0.01 0.01

z01 U(z) z01 U(z) z01 U(z)


x

Figure 6.4: Schematic representation of a smooth → rough transition. The ver-


tical scale is logarithmic, so the logarithmic profiles are represented by straight
lines.

74
Figure 6.5: Wind profiles as measured in a wind tunnel at several distances from
the surface boundary. (a) smooth → rough, z01 = 0.02 mm and z02 = 2.5 mm;
(b) rough → smooth, z01 = 2.5 mm and z02 = 0.02 mm (From: ?)

Turbulence changes
We will have a look at momentum transport and the variance as a function of
height after a transition from a smooth to a rough surface has occurred. We can
rely on wind tunnel measurements. In figure 6.6, the variance σu ≡ u2 , σw ≡ w2
and the momentum transport uw are shown. We can see that the value of uw
increases strongly at the surface, followed by a decrease to an equilibrium value
(compare with figure 6.7). A maximum can be found at a certain height, that
seems to move away from the surface. This behaviour cannot be seen when
looking at the variances. The development of h is very similar in both graphs
and resembles the development of h(x) that followed from the wind profile.
We can formulate the equation for the average flow using (3.3):

∂U ∂u2 ∂uw
U =− − .
∂x ∂x ∂z
We have assumed that the flow is stationary, and that the pressure gradient
occurring at the roughness transition can be neglected. Figure 6.6 shows that
the right-hand side of the equation is smaller than zero, causing the average
flow to slow down.

75
Figure 6.6: Development of the shear stress (a) and the variance (b) in a smooth
to rough transition, as a function of the distance from the surface boundary. The
variables have been normalized with respect to the wind tunnel velocity. (From:
?)

Figure 6.7: Some calculations (lines) and wind tunnel data (points) of the change
in surface shear stress (u∗2 /u∗1 )2 (From: ?)

76
Chapter 7

EXERCISES BOUNDARY
LAYERS AND
TURBULENCE

CHAPTER 1

Exercise 1
The first law of Thermodynamics states:

dQ = (dU )v + pdV

a) Why is in this form of the first law of Thermodynamics dU = (dU )v used?


b) From this derive:

1
dQ = (cv + R)dT − dp
ρ

c) Derive cv + R = cp and equation 1.4.

Exercise 2
Derive the pressure and density profile for an isothermal atmosphere (T (z) =
T0 ).

77
Exercise 3
a) Show from equation 1.4 that for an adiabatic atmosphere we obtain equation
1.5.
b) Use this to obtain the adiabatic temperature profile (equation 1.13).
c) Use again the adiabatic form of equation 1.4 and the equation of state to
obtain (eqs. 1.14, 1.15).

Exercise 4
a) Derive from equation 1.8:

dθ θ dT g
= ( + )
dz T dz cp

b) Show now that eqs. 1.11, 1.12 are valid for the boundary layer.

Exercise 5
Consider a stochastic 1-D velocity field u(x). We define the velocity correlation
between two arbitrary points located at x1 and x2 as

ρ0 (x1 , x2 ) ≡ u(x1 )u(x2 )

The field is called homogeneous if its statistical properties are the same in all
regions of the field. Thus

ρ0 (x1 + L, x2 + L) = ρ0 (x1 , x2 ),

where L is an arbitrary displacement. Hence the correlation is invariant under


the translation L.

a Show that in a homogeneous field ρ0 is only a function of the separation


r = x1 − x2 .
b Show for the homogeneous case that the velocity variance σ defined as
σ = ρ0 (x, x) = u2 (x) must be constant.
c Make a sketch of the velocity correlation in a turbulent flow as function
of r. Explain the shape.
d Give a similar consideration for stationarity of a stochastic process u(t).

78
CHAPTER 2

Exercise 6
Starting with dQ = cp dT − ρ1 dp derive for an adiabatic process:

dρ 1 dp
=
dt c2 dt
cp p
with c2 = cv ρ

CHAPTER 3

Exercise 7
The Boussinesq momentum equation (eq. 2.9) (without the Coriolis term) is
given by:
The momentum equation

∂ ũi ∂ ũi 1 ∂ p̃ (θ − θ0 ) ∂ 2 ũi


+ uj =− + gδi3 + ν . (7.1)
∂t ∂xj ρ0 ∂xi θ0 ∂xj ∂xj

a Use this equation to substitute the Reynolds decomposition (we call this
equation A), and take the average of this whole equation to obtain

∂Ui ∂Ui 1 ∂Π (Θ − θ0 ) ∂ 2 Ui ∂ui uj


+ Uj =− + gδi3 + ν − ,
∂t ∂xj ρ0 ∂xi θ0 ∂xj ∂xj ∂xj

which is the first equation in section 3.2

b Subtract this equation from equation A to obtain the fluctuation equation:

∂ui ∂Ui ∂ui ∂ui ∂ui uj 1 ∂π θ ∂ 2 ui


+ uj + Uj + uj − =− + gδi3 + ν
∂t ∂xj ∂xj ∂xj ∂xj ρ0 ∂xi θ0 ∂xj ∂xj

c Check that the average of this fluctuation equation is zero.

d Multiply this whole fluctuation equation by ui , and take the average of


this result to obtain equation (3.24). Explain all the terms.

79
Exercise 8
In general our chosen frame of reference forms a curvilinear co-ordinate system
which is fixed to the Earths surface. As a result two apparent forces are intro-
duced, the curvature terms and the Coriolis forces. If we take the co-ordinate
system such that x is positive to the East, y positive to the North and z positive
upwards, we can write the momentum equations:

du 1 ∂p uw uv tan ϕ ∂2u
= − − + + f v − lw + ν 2
dt ρ0 ∂x R R ∂xj
dv 1 ∂p vw u2 tan ϕ ∂2v
= − − − − fu + ν 2
dt ρ0 ∂y R R ∂xj
dw 1 ∂p u2 + v 2 ∂2w
= − −g+ + lu + ν 2
dt ρ0 ∂z R ∂xj

where R is the radius of the Earth, ρ0 the constant density, f = 2ω sin ϕ and
l = 2ω cos ϕ (see e.g. page 35 in J. R. Holton, An introduction to dynamic
meteorology).
a) Derive an equation for the kinetic energy (u2 + v 2 + w2 )/2 for the above set
of momentum equations.
b) Discuss why the curvature terms and Coriolis forces do not appear in this
equation.

Exercise 9
The flow in a vegetation canopy with height h0 is considered with the assumption
that the leaves are diffusely distributed, so that they act as a continuum-like
sink for momentum. The horizontal equation of motion in the absence of a
pressure gradient is then
∂τ
− − Dc = 0
∂z
where τ (z) is the horizontal shear stress in the air and Dc is the momentum sink
term, or the drag experienced by the foliage per unit volume of air. Further it
is assumed that the shear stress is proportional to the velocity gradient
∂U
τ = −ρ0 K
∂z
and that Dc is proportional to U 2 as
1
Dc = ρ0 Cd Ac U 2
2
with the the drag coefficient, Cd ≈ 0.25 and Ac is the surface area of leaves per
unit volume of air.

80
a) With the assumption of a constant mixing length l in the canopy so that
K = l2 |∂U/∂z|, and Cd , Ac is constant, solve the velocity profile in the canopy
(hint: substitute an exponential function). Use U = U (h0 ) at z = h0 .
b) Solve the corresponding expression for the exchange coefficient.
c) Assuming that the velocity profile and its derivative match at h0 with the
logarithmic wind profile kU/u? = ln((z − d)/z0 ). Derive an expression for z0 in
terms of the displacement height (d), h0 , etc. What is a typical value for z0 /h0 ?
Take d/h0 = 2/3 and h0 (Cd Ac /(4l2 ))1/3 ∼ 3.

Exercise 10
In the RANS approximation the buoyancy production, B, in the turbulent ki-
netic energy equation is given by:
g
B= wθ
θ0
The buoyancy production is equal to the release of potential energy of the sys-
tem.
a) Give an expression for the vertically integrated potential energy (P) of the
atmosphere.
b) Of course we are only interested in the change of potential energy in time.
Therefore, we only consider the change in potential energy compared to some
reference state (P0 ). Show that P − P0 per unit area is approximately given by:
Z z
g
P − P0 = −ρ0 (θ − θ0 ) z 0 dz 0
0 Θ 0

c) Using the Reynolds averaged equation for the potential temperature (horizon-
tally homogeneous boundary layer, see Section 3.3), derive the expression for the
change of potential energy of the system in terms of the buoyancy production
(B):

Z tZ z
∂B
P − P0 = −ρ0 z0 dz 0 dt
0 0 ∂z 0

d) Apply this result to the convective boundary layer (without shear) with a
given surface heat flux. Discuss the vertically integrated kinetic energy equation
and its relation with the change in potential energy of the boundary layer.

81
Exercise 11
The goal is to ’derive’ the mixing length expression for the exchange coefficients
(Km and Kh ). Neglecting the material derivative (dq/dt) and neglecting the
transport term in the Turbulent Kinetic Energy (TKE) equation we obtain:
∂Ui g
0 = −ui uj + wθ − 
∂xj θ0
We parametrize the momentum flux, potential temperature flux and viscous
dissipation as:
 ∂U ∂Uj 
i
−ui uj = Km + ≡ Km Sij
∂xj ∂xi
∂Θ
−wθ = Kh
∂z
3
q2
 = c
l
where l is a length scale, q the turbulent kinetic energie and c a constant.
a) If we assume that the exchange coefficients Km,h can be written as the
1
product of a length and a velocity scale: Km,h = cm,h l q 2 , then solve the
kinetic energy q in terms of l, c , cm,h , Rig and the deformation tensor Sij .
Where Rig is given by:
g ∂Θ
θ0 ∂z
Rig = 1 2
2 Sij

b) Give the corresponding expression for Km,h .


c) Which value of the critical gradient Richardson number is obtained if ch /cm ≈
3?

Exercise 12
In a 3-dimensional Large Eddy Simulation (LES), the large scale turbulent mo-
tions (anisotropic) are explicitly resolved whereas the smaller scale motions
(isotropic) have to be parametrized. In an LES model with grid resolution
∆, the subgrid fluxes for momentum and heat are parametrized as:
 ∂U ∂Uj 
i
−ui uj = Km +
∂xj ∂xi
∂Θ
−wθ = Kh
∂z
where Km,h are the turbulence exchange coefficients. Uppercase variables de-
note averages over the grid volume. The resolution (∆) of the LES has to be
in the inertial subrange (see Fig. 3.3). The exchange coefficients are written
1
as the product of a length and a velocity scale: Km,h = cm,h ∆qm2
, where qm
is the sub-grid kinetic energy and cm,h are constants. For an LES model this

82
expression can exactly be derived from the inertial subrange theory, for which
the turbulence is isotropic.
The spectrum of the turbulence kinetic energy, E(k) and potential temperature
variance Eh (k) in the inertial subrange is given by:
2 5
Em (k) = αm
3
k− 3
−1 5
Eh (k) = βh m 3 k − 3
where α ≈ 1.5 and β ≈ 1.0 are constants, k is the wavenumber, m is the dissi-
pation rate of turbulence kinetic energy and h the dissipation rate of potential
temperature variance. The subgrid turbulence kinetic energy (qm ) and potential
temperature variance (qh ) are given by:
Z ∞
qm = Em (k) dk (2a)
k
Z c∞
qh = Eh (k) dk (2b)
kc

where kc (= 2π/λc ) and λc ≈ 2∆ is the cut-off wavenumber. For isotropic


turbulence dissipation rates are given by:
Z kc
m = 2Km k 2 Em (k) dk (3a)
0
Z kc
h = Kh k 2 Eh (k) dk (3b)
0

a) Derive from (2a,b) an expression for qm and qh in terms of kc , m and h .


b) Use (3a,b) to derive an expression for Km and Kh in terms of kc and m ,
and give the ratio Km /Kh .
c) Note that the expressions for the turbulence exchange coefficient can be
1
written as: Km,h = cm,h ∆qm 2
. Derive the values of cm,h in terms of α and β.
The results obtained above can be substituted in the following simplified subgrid
turbulence kinetic energy equation:
∂Ui g
0 = −ui uj + wθ − 
∂xj θ0
d) Which critical gradient Richardson number is then obtained?

83
Exercise 13
Consider the TKE (3.24).
a) Give the physical interpretation of all terms and derive the equation for a
stationary, homogeneous surface-layer. Neglect the transport term. Note that
in the surface layer we have

−uw = u2? = constant


−θw = θ? u? = constant.
b) The flux Richardson number is defined as
g
θ0 wθ
Rif = .
uw dU
dz

Interpret this number in view of the results obtained under a. Show that Rif ≤ 1
and express Rif as a function of z/L. Make a sketch of this relationship using
the empirical functions

κz dU z 1
φm ≡ = (1 − 16 )− 4
u? dz L
z
φm = 1+5 ,
L
for z/L < 0 and z/L > 0, respectively.
c) Motivate the following closure hypotheses

dU
u2? = Km
dz
3
q2
ε = ,
l
1
where Km = l q 2 and l a length scale.
d) Express Km in the parameters u? , l and Rif .
e) Assume l = κz. Show that in a neutral case U (z) ∝ ln z. Derive an asymp-
totic expression for Km in the limit u? → −∞.

CHAPTER 4

Exercise 14
Make a sketch and discuss in an unstable and stable boundary-layer the vertical
profiles of
a) mean wind and potential temperature
b) variances of velocity and temperature fluctuations
c) momentum and heat flux

84
Focus on differences between stable and unstable conditions. List a number of
scaling parameters in the various regions of the ABL.

Exercise 15
Consider the scaling of local turbulence parameters in the convective boundary
layer (h/ − L >> 1).
a) Which scaling parameters apply in the Monin-Obukhov theory? In which
part of the CBL does this theory apply? Apply this scaling to the vertical
1
velocity variance, σw ≡ w2 2 .
b) Which parameters are important in convective scaling and in which part of
the CBL does this apply. Get again an expression for σw in this region.
c) How would you model σw in the intermediate region?

Exercise 16
a) What is Monin-Obukhov scaling? Apply it to the following quantities:
b) the wind and temperature gradient, dU/dz and dΘ/dz, respectively,
Show that with the results obtained in a, the momentum and heat flux at the
surface can be obtained from measurements of mean temperature and wind
speed.

Exercise 17
The momentum equations for a homogeneous boundary layer in a rotating frame
are
∂U 1 ∂Π ∂uw
= fV − −
∂t ρ0 ∂x ∂z
∂V 1 ∂Π ∂vw
= −f U − −
∂t ρ0 ∂y ∂z
1 ∂Π
0 = − −g
ρ0 ∂z


a) Introduce the definition of the geostrophic wind, Vg . Show that in a barotropic
ABL the geostrophic wind does not vary with z.


b) Show that in a flow with Vg = 0 the mean kinetic energy vanishes. (Formulate
the energy equation, integrate it over the interval z = 0−h and use U (0), V (0) =
0, uw(h), vw(h) = 0 and the closure: uw = −KdU/dz, vw = −KdV /dz).

Exercise 18
For steady state conditions and neglecting the turbulent transport and advection

85
terms, the TKE equation reduces in horizontally homogeneous conditions to:
∂U ∂V g
0 = −uw − vw + wθ −  (4)
∂z ∂z θ0
which states that the turbulence kinetic energy budget is completely determined
by local parameters. We express the fluxes in terms of their local gradients using
the turbulent exchange coefficient Km :
∂U
uw = −Km
∂z
∂V
vw = −Km
∂z
Using the definition
1
τ = (uw2 + vw2 ) 2
3
and multiplying (4) with κz/τ 2 , we obtain (κ is von Karman’s constant ∼ 0.4):
z
0 = φm − − φ
Λ
where
κz ∂U 2 ∂V 2 1
φm = 1 [( ) +( ) ]2
τ2 ∂z ∂z
3
τ2
Λ = − g
κ θ0 wθ
κz
φ = 3 
τ2
This equation states that the local dimensionless wind shear (φm ) and dissi-
pation (φ ) only depend upon the stability parameter z/Λ. This is the most
general form of local scaling and is used in the neutral and stable boundary
layer and the surface layer.
Typically stability is discussed in terms of the stability parameter z/Λ or the
gradient Richardson number (Ri).

a) Derive that the gradient Richardson number is related to z/Λ through


Rig = Λz φφ2h .
m
b) In the very stable limit (z/Λ >> 1 ), it is typically assumed that both φm
and φh increase linearly with z/Λ, φm,h ∼ βm,h z/Λ (with βm,h ∼ 5). What
does this mean for the gradient and flux Richardson number in the very stable
limit?
c) Sometimes it is argued that the flux Richardson number (Rif ) has a critical
behaviour and the gradient Richardson number has not . What could be the
physical argument for this?
(d) We (empirically) assume a critical Rif number and that Rig is related to
the turbulent Prandtl number (P rt ) through P rt = 1 + 4Rig (Kim and Mahrt,
1992).

86
What does this imply for φm , φh and the critical gradient Richardson number?
Reference: Kim J. and L. Mahrt, 1992; Simple formulation of turbulent mixing
in the stable free atmosphere and nocturnal boundary layer. Tellus 44A, 381-
394.)

Exercise19
Consider a stationary and horizontally homogeneous boundary layer over a flat
surface with geostrophic wind components (Ug , 0). Given the momentum flux
profile with components

u2? (1 − b hz )

for z < h
−uw =
0 for z > h
−vw = 0

with b a constant > 1.


Sketch the mean wind profile (U, V ) between z = 0 and z = 2h.

Exercise 20
In case of a geostrophic wind balance, the wind blows parallel to the isobars,
this is typically observed in the free troposphere. Due to surface friction the
wind in the boundary layer (for the northern hemisphere) is backing relative
to the geostrophic wind. For a high and low pressure system the wind in the
boundary layer is schematically shown below. For the low pressure situation
we get a convergence of the horizontal wind in the boundary layer, and as a
result vertical advection. In order to study the relative importance of vertical
advection and turbulent mixing we consider a stationary and horizontally homo-
geneous boundary layer over a flat surface with geostrophic wind (Ug , Vg ). We
take the x-axis parallel to the wind at the surface such that the y-component of
the surface stress is zero. The angle between the surface and geostrophic wind
is denoted by α.
a) Show from the momentum equations

∂uw
0 = f (V − Vg ) −
∂z
∂vw
0 = −f (U − Ug ) −
∂z
that the net transport of air mass (integrated over the whole boundary layer
(height h of about 500 m) and relative to the geostrophic wind) has a component
in the y-direction only. Give the net transport also in the ’geostrophic wind’
co-ordinate system.
b) Assume a circle symmetric flow and a geostrophic wind in a solid-body ro-
tation (Vg = ωr with ω = 10−5 s−1 ). Derive an equation for the horizontal

87
Figure 7.1: Wind field of the boundary layer (northern hemisphere) with a low
and high pressure system.

mean-vertical velocity at the top of the boundary layer from the continuity
equation which reads in cylindrical coordinates:

∂w 1 ∂ru
+ =0
∂z r ∂r
c) Assume that the integral of U with height is proportional to ωrh. Estimate
the upwelling rate (w) at the top of the boundary layer due to ’Ekman pumping’.
d) What are the relative time scales of Ekman pumping and turbulence in these
boundary layers. Or, in other words: how long does it take to bring material
from the surface to the top of the boundary layer?

Exercise 21
In this exercise we will study the mixed-layer dynamics of the horizontally ho-
mogeneous dry boundary layer driven by a constant surface heat flux (wθ)0
and subsidence. Assuming a mean vertical velocity (W (z) ), the mixed layer
equations can be summarised as:
∂h
= w e + Wh
∂t
∂Θm (wθ)0 − (wθ)h
=
∂t h
∂∆ ∂θm
= γθ w e −
∂t ∂t
where the entrainment velocity is defined as we ≡ −(wθ)h /∆, θm is the constant
potential temperature in the mixed layer, ∆ is the potential temperature jump
at the inversion, h the boundary layer depth, γθ the potential temperature lapse
rate above the inversion, Wh is the mean vertical velocity at h and β a constant
(about 0.2). In order to close the set of equations we assume

(wθ)h = −β(wθ)0

88
a) Assume that the large scale subsidence is given by: W (z) = −D z, where D
(equal to the divergence of the horizontal wind, typically about 10−5 s−1 ) is a
constant larger than zero. If we assume h constant in time what is the solution
for h and ∆?
b) Give a physical interpretation of the obtained solution.
c) In order to investigate the effect of the surface friction on the wind jumps we
consider the corresponding mixed-layer equations for the wind (positive x-axis
in the mean mixed-layer wind direction):
∂Um (uw)0 − uwh
= f (Vm − Vg ) + (5a)
∂t h
∂Vm (vw)0 − vwh
= −f (Um − Ug ) + (5b)
∂t h
where we assume the geostrophic wind components (Ug , Vg ) constant with height
and time. Find the stationary solution for the wind jumps (∆U, ∆V ) at the
boundary layer height from (5a,b) in terms of D, h, f and (uw)0 . Give a physical
interpretation of the obtained solution.

Exercise 22
Here we address the mixed-layer dynamics of the horizontally homogeneous dry
boundary layer driven by a given surface sensible heat flux wθ0 . The surface
sensible heat fluxes is positive. The mixed-layer equations for potential temper-
ature θm can be summarised as:
∂h
= we (6a)
∂t
∂θm
h = (wθ)0 − (wθ)h (6b)
∂t
∂∆ ∂θm
= γθ we − (6c)
∂t ∂t
where the entrainment velocity is defined as we ≡ −(wθ)h /∆, θm is the potential
temperature in the mixed layer, ∆(t) is the potential temperature jump at the
inversion, h(t) the boundary layer depth, γθ the potential temperature lapse
rate above the inversion and β a constant (about 0.2). In order to close the set
of equations we assume
(wθ)h = −β(wθ)0
a) We assume that the initially that θm (0) = 302 K, ∆(0) = 7K, h(0) = 800m
and γθ = 4K/km. Discuss the time evolution of the potential temperature
profile and the relevant physical processes.
b) Show that the differential equations given by (6ab,c) can be combined to
give:
∂∆ 1 + β ∆ ∂h ∂h
+ = γθ (7)
∂t β h ∂t ∂t

89
c) Note that equation (7) can be written as

∂∆ 1 + β ∆
+ = γθ (8)
∂h β h

Solve equation (8) and obtain ∆(t) as a function of h(t).

Exercise 23
Consider the mixed-layer growth of the horizontally homogeneous boundary
layer driven by a constant heat flux wθ0 . Without subsidence the mixed layer
equations can be summarised as:
∂h
= we
∂t
∂θm
h = wθ0 − wθh
∂t
∂∆ ∂θm
= γθ we −
∂t ∂t
where the entrainment velocity is defined as we ≡ −(wθ)h /∆, θm is the potential
temperature in the mixed layer, ∆(t) is the potential temperature jump at the
inversion, h(t) the boundary layer depth, γθ the potential temperature lapse
rate above the inversion and β a constant (about 0.2). In order to close the set
of equations we assume

wθh = −βwθ0

a) Give a physical interpretation of the terms in the equations.


b) Derive a differential equation in terms of the unknowns h and ∆ (eq. 6.16 of
Garratt), and solve this equation (there exists a more general solution than eq.
6.17 of Garratt).
c) The differential equation has a homogeneous and a particular solution The
homogeneous solution has an unknown constant (for instance denoted by A).
Determine A from the initial value of ∆, γθ and h. Under which conditions is
A = 0? Show these results graphically.
d) Determine for A = 0 the evolution of h and ∆ as a function of time.

Exercise 24
Describe the structure of the CBL. Identify the scaling parameters and the
region in the CBL where they apply to. Express σu , σw and σθ in these scales
and find the explicit relationship in the intermediate region.

Exercise 25

90
Derive the equations of motion for a pure convective boundary layer, where the
buoyancy flux is the only source of turbulence production (no mean flow).

Exercise 26
We consider an atmospheric boundary layer.
a) Which parameter tells us whether the boundary layer is convective?
b) Give an expression for σθ in the surface layer.
c) Do the same in the for the mixed layer.
c) And again for the intermediate layer, just above the surface layer.

Exercise 27
Consider a dry horizontally homogeneous stable boundary layer (SBL) driven
by a constant (in time and space) geostrophic wind (Ug , Vg ). If we neglect the
radiative flux divergence and molecular diffusion terms in the SBL the Reynolds
averaged thermodynamic equation becomes:

∂Θ ∂wθ
=−
∂t ∂z
In steady state conditions and in case we can neglect the turbulent transport
term the TKE equation reduces to:
∂U ∂V g
0 = −uw − vw + wθ − 
∂z ∂z θ0
a) How does the heat flux change with height if we assume a quasi-stationary
boundary layer (∂/∂t(∂Θ/∂z) = 0 )? Denote the surface heat flux as wθ (< 0)
and take the heat flux at the boundary layer depth (h) equal to zero.
b) How is the flux Richardson number defined?
c) Write down an equation for the net work against gravity (conversion of kinetic
energy into potential energy) and shear production averaged over the whole
boundary layer. Use the stationary horizontal momentum equations to rewrite
the shear production term.
d) If we assume the flux Richardson number constant throughout the SBL and
equal to its critical value (about 0.25), then derive a relationship between h, the
friction velocity at the surface (u?0 ), (wθ)0 and the geostrophic wind from the
results obtained above. Discuss the relationship between h and the surface heat
flux for a given u?0 and geostrophic wind.

Exercise 28
a) What could be the cause for the occurrence of the so-called ’low-level jet’
during a cloudless night?
b) Formulate the equations of motion in this case and solve these.

91
c) Make a sketch of wind speed as a function of time for a 24 h period (cloudless
conditions and over land) at a height of 200 and 10 m.

CHAPTER 5

Exercise 29
a) Discuss the various terms in the surface energy balance.
b) Make a sketch of the diurnal cycle of these terms on a cloudless day over
land.
c) Define the Bowen ratio
d) Assume that we know the net radiation and the soil heat flux. Which addi-
tional measurements do we need to determine the Bowen ratio?

Exercise 30
On 30 November 1996 most components of the surface energy balance were
measured near the 200 m high meteorological tower at Cabauw. The surface in
Cabauw is relatively flat and at that time of the year covered with short grass.
Conservation of energy at the surface requires that:

Rn = G + H + LE

where Rn is the net radiation (positive downwards), G is the heat flux into
the soil, H is the sensible heat flux and LE is the latent heat flux. The heat
flux (measured with a sonic anemometer) and net radiation are measured at
3m height. The latent heat flux is estimated by the Bowen ratio method. The
soil heat flux is calculated from the volumetric heat capacity of the soil and
temperature measurements at different depths in the soil. The surface energy
balance on 30 November is shown in the figure.
a) Discuss the diurnal variation of the different components of the energy balance
shown in the left figure.
b) Discuss the relationship between the heat flux as a function of height and the
temporal evolution of the temperature profile after sunset (shown in the figure
on the right).
c) Which other physical process(es), besides turbulence, can cool the air in the
upper part of the boundary layer?
d) What is the Bowen ratio method?
e) Describe three other (empirical) methods from which the latent heat flux can
be estimated.

Exercise 31
We design an experiment in horizontally homogeneous terrain in order to de-
termine the turbulent fluxes of heat and momentum in the surface layer. The

92
Figure 7.2: The hourly averaged components of the (left) surface energy bal-
ance and (right) temperature profiles after 1500 UTC on 30 November 1996 as
measured in Cabauw. Local time is one hour plus UTC. The arrows indicate
time of sunrise and sunset, respectively.

aerodynamic roughness of the terrain z0 is known. We measure mean wind and


temperature only. Give a minimum configuration of the measurement. Make a
sketch and discuss it.

Exercise 32
Consider the top-layer of the ocean with thickness d and temperature Ts . The
temperature respons due to solar heating is predicted by

dTs Ts − Te Q(t)
=− + .
dt τ ρ0 cp d

The equilibrium temperature is Te , the response time τ (∼ 105 s) and Q is a


periodic forcing with mean zero. we investigate the response by Fourier series
expansions of Q and Ts . Analyze the response in amplitude and phase shift as
a function of the frequency of the forcing. Take d= 20 m. discuss both the
diurnal and yearly cycle.

93

You might also like