You are on page 1of 26

ANNUAL

REVIEWS Further
Quick links to online content

Copyright 1970. All rights reserved

THE ATMOSPHERIC BOUNDARY LAYER


A. S. MONIN
Institute oj Oceanology, Academy oj Sciences, Moscow, USSR

Definition of atmospheric boundary layer.-In large-scale air currents, the


Annu. Rev. Fluid Mech. 1970.2:225-250. Downloaded from www.annualreviews.org

combined action of turbulent friction and Coriolis force results in the forma­
tion, near the surface of a planet, of the atmospheric boundary layer. Un­
like most boundary layers dealt with in engineering, the atmospheric boun­
by Ryerson University on 02/25/13. For personal use only.

dary layer is characterized by the i nfluence of Coriolis force (i.e., the planet's
rotation) and the density stratification of air (affecting turbulence through
buoyancy forces), thus the atmospheric boundary layer is a turbulent
boundary layer in rotating heavy stratified fluid.
In the equatorial zone and in typhoons, the thickness of the atmospheric
boundary layer can change in the wind directioJ?, growing with its fetch in
a manner similar to the thickness of a boundary layer near a plate. Beyond
these specific regions, the thickness of the layer can be determined by the
equation h=c(G/j), where G is the wind velocity at the upper boundary of
the layer, j= 2w sin cjJ is the vertical projection of the vorticity of the planet's
rotation, called the Coriolis parameter (w being the angular velocity of the
planet's rotation and cp the latitude), and c is a nondimensional factor. Such
an atmospheric boundary layer is called an Ekman layer in honour of
Ekman (1), who was the first to construct a theoretical model of the boun­
dary layer in a rotating fluid. Below we shall confine ourselves to the con­
sideration of the Ekman layer alone.
According to Charney (2), an atmospheric boundary layer will be hydro­
dynamically stable (and may be steady) only so long as its Reynolds number
R= (Gh/K) (where K is the effective value of turbulent viscosity) does not
exceed some critical value Rer• On putting K""-'h2j, we obtain the criterion
h"2:. (G/�f) for stability of an Ekman layer. In their experiments on the
stability of the laminar Ekman layer Faller (3), Tatro & Mollo-Christensen
(4), and Green (5) obtained �100. If this estimate is extended into the
turbulent Ekman layer the minimum thickness of a stable layer for the
Earth, with G= lO m/sec and j= lO-4sec-l, appears to be 1 km, i.e. , one
order l ess than the effective thickness of the Earth's atmosphere, ha�lO km.
In other words, the Ekman layer on the Earth is thin, and from this point of
view the Earth should be admitted to be a rapidly rotating planet.
The scales of inhomogeneities of meteorological fields much larger and
much smaller than the effective thickness of the atmosphere ha should
reasonably be called "large" and "small" respectively [ Kolesnikova &
225
226 MONIN

Monin (6, 7)]. The large-scale inhomogeneities are evidently quasi-two­


dimensional (quasi-horizontal), whereas the small-scale ones, on the con­
trary, are essentially three-dimensional (quasi-isotropic); the latter are the
object of micrometeorology. Thus the atmospheric boundary layer on the
Earth is a micrometeorological formation.

A veraging. Turbulence is the most important of the micrometeorological


-

phenomena; its presence is an essential peculiarity of the atmospheric


boundary layer. Therefore a hydromechanical description of the layer re­
quires the use of methods from the theory of turbulence. This description
should be first of all statistical, based on the use of a certain operation of
Annu. Rev. Fluid Mech. 1970.2:225-250. Downloaded from www.annualreviews.org

averaging and a consideration of the average hydromechanical parameters


obtained.
Practically, averaging in time is used. And since the spectrum of oscilla­
tion periods of meteorological fields is continuous, the danger of a "level
by Ryerson University on 02/25/13. For personal use only.

evolution" arises, i.e., of a statistical instability of the obtained average


parameters and their dependence on the value of the period of averaging.
Yet, this difficulty is overcome because in the spectrum of periods between
the large-scale and micrometeorological regions (in the range of periods from
minutes to hours, with the center in the vicinity of the period of the order
of ha/�20 min), there is a deep mesometeorological minimum (discovered
for the first time in the wind-velocity spectrum by Van der Hoven (8); in
(6, 7) its presence is proposed to be explained by a strong damping action
of turbulent viscosity in the mesometeorological region). Therefore the averag­
ing by periods of the mesometeorological interval yields statistically per­
sistent results.

Significance of the atmospheric boundary layer . Large scale processes or


- -

"weather conditions" are "external parameters" determining the internal


structure of the �tmospheric boundary layer and, first of all, the regime of
turbulence. The latter, in its turn, exerts a reverse influence on the large­
scale processes, being a dissipation factor for them and generating, in the
regions of turbulent drag of the large-scale currents, upwelling fluxes (with a
velocity determined in a simple model of the layer by Dyubyuk's (9) equa­
tion w",h6.pljPa, where 6.p is the horizontal pressure Laplacian and Pa the air
density), and finally realizing an exchange between the underlying surface
and the atmospheric boundary layer in momentum, heat, humidity, and
other admixtures that are further transferred from the layer to the free
atmosphere (hence to large-scale processes) in the regions of cumulus con­
vection and in the regions of separation of the layer-at atmospheric fronts
and in typhoons [vide Priestley (10) ] .
Thus a knowledge o f the atmospheric boundary layer is necessary in
computing the evolution of large-scale atmospheric processes for long-term
weather forecasts and for the creation of a theory of climate. Besides, turbu­
lence in the· layer is of determining significance in the vibration of land con-
THE ATMOSPHERIC BOUNDARY LAYER 227
structions (buildings, bridges, masts, electro-transmission lines) under the
action of wind, in the generation of wind waves and drift currents in the
ocean, in the bumping of airplanes, lifting rockets, and other aircraft, in the
diffusion of atmospheric contamination, in the amplitude and phase fluctua­
tions of light and radiowaves propagating in the atmosphere from land and
space sources, in the scattering of short radiowaves creating conditions for
long-range ultra short wave and telecommunication, and in many other
phenomena of practical importance.

Model of the atmospheric boundary layer. M easurements show that the


-

structure of the atmospheric boundary layer and, in particular, the dis­


Annu. Rev. Fluid Mech. 1970.2:225-250. Downloaded from www.annualreviews.org

tribution of wind with height are by no means always of a similar character


and are often very irregular. For instance, sometimes a wind-velocity maxi­
mum is observed at relatively small heights of some 100 to 300 m, which in
the most pronounced cases assumes the form of the so-called "low level jet
by Ryerson University on 02/25/13. For personal use only.

stream." Irregular changes of wind direction with height having no resem­


blance whatsoever to the "Ekman spiral" are also observed rather often.
Such irregularities may be a consequence of a number of complicating fac­
tors: density stratification of the air, horizontal inhomogeneity of the tem­
perature field [creating the so-called thermic wind-vide original data, for
instance, in the paper by Utina (11)] and of the underlying surface [vide
Laikhtman (12) and (13-17)], irregular density and curvature of isobars
[cyclonicity increases and anticYclonicity decreases wind turning in the
layer-see, for instance (18)], orographic effects on air currents, nonsta­
tionarity effects, etc. Yet even in the absence of such factors that make the
studies much more difficult, the structure of the layer proves to be rather
complicated. Therefore we shall confine ourselves mainly to discussion of the
simplest case, of a stationary atmospheric boundary layer above a plane and
homogeneous underlying surface with rectilinear and evenly distributed
isobars, i.e., with the pressure field
p(x, y, z) = P(z) + pofG(x sin a y cos a)
- 1.
Here x, y, z are Cartesian coordinates (where z is vertical, the direction of
the x-axis coincides with that of the surface wind, and a is the angle between
the isobar and the surface wind), Po is the standard average value of air den­
sity in the layer, and G in this case is the geostrophic wind velocity. Under
the pressure field of Equation 1 the average equations of motion (Reynolds

( )
equations) have the form
a au
f(v - G sin 0:) + - Tzz + v- = 0
az az
a ( av)
-f(u-Gcosa)+- TII.+V- = 0
2.

az az
Here 12, ii are the components of the mean wind velocity, II is the coefficient
228 MONIN

of molecular viscosity of air, and T",.= -u'w' and T".= - v'w' are the com­
ponents of vertical turbulent momentum flux (Reynolds stresses) divided
by air density (w' being the fluctuation in vertical velocity; here and below,
a bar over a letter means averaging and primes are used to mark fluctuations,
i.e., deviations from average values).
According to our choice of direction for the x-axis, T".+v(du/dZ) vanishes
as the underlying surface is approached, and the value of T",.+v(dU;dZ) tends
to some positive limit denoted by U*2j the value u* is called the friction
velocity.
To describe the density stratification in the atmospheric boundary layer
determined by the stratification of temperature and air humidity, let us take
Annu. Rev. Fluid Mech. 1970.2:225-250. Downloaded from www.annualreviews.org

the average equations of heat influx and humidity transfer, which in the
present case of statistical stationarity and horizontal homogeneity (and in
the absence of phase transformations of humidity in the air) are reduced in
by Ryerson University on 02/25/13. For personal use only.

the layer to the form


aT
PoToS'w' - CpPoX - + H R const 3.
az
=

HL iJq
- - Po5)- = E = const 4.
.£ az
Here S is the entropy, T is the temperature (To being its standard average
value in the layer), q is the specific humidity, Cp "" 1 J/g deg and .e "" 2500 J/g
are the specific heat capacity of air under constant pressure and latent heat
of evaporation, cppoX and Po5) are the coefficients of molecular conductivity
and water vapor diffusivity, HR is the vertical radiative heat flux, HL=
£poq'w' is the vertical latent heat flux, and E is the vertical humidity flux
which, since it is constant, equals the rate of evaporation on the underlying
surface. According to (19), the first term in Equation 3 should be written in
the form HT+ [(Sv-Sd)/Cp]' (cpTo/ .£)HL, where HT=CpPO T 'w' is the vertical
turbulent heat flux, and Sv and Sd are the specific entropies of water vapor and
dry air. If vertical changes of radiative heat flux HR (usually resulting in
unsteadiness) are neglected, Equation 3 is reduced to the form
aT
HT - C,ppox -- = H = const 3a.
dZ

Energy equation.-The influence of density stratification on turbulence is


clearly explained with the aid of the turbulent energy equation. Its form
for a steady and horizontally homogeneous atmospheric boundary layer
[vide §§6.2-6.6 of the book by Monin & Yaglom (20)] is
( iJu
Tz.- + T".-
iJfJ) - -p'w'
g _ 1 aQ
- - -- = E 5.
iJz 8z Po Po ih
THE ATMOSPHERIC BOUNDARY LAYER 229

where g is the acceleration of gravity, Q is the vertical flux of turbulent


energy, and E is the rate of energy dissipation (per unit mass). The term in
parentheses on the left side of Equation 5 describes the generation of turbu­
lent energy caused by the work of Reynolds stresses. The next term is the
gain or loss of turbulent energy caused by the work of buoyancy forces under
lapse or stable stratification. According to (19), it reduces to
g _ (j(HT + bHL)
- -p'w' = ------ 6.
Po CpPo
where (3 is the buoyancy parameter, i.e., the product of g by the thermal
Annu. Rev. Fluid Mech. 1970.2:225-250. Downloaded from www.annualreviews.org

expansion coefficient, equal for air to liTo, and b= (RvlRq,-l)cpTo/£,


Rv=0.461 Jig deg and Rq,=O.287 Jig deg being the gas constants for water
vapor and dry air. (The value Bo=Hr/HL is called the Bowen ratio). Thus
if HT+bHL;CO the presence of heat and humidity fluxes HT and E affects
by Ryerson University on 02/25/13. For personal use only.

turbulent energy so that neither heat nor humidity can be considered pas­
sive substances. The relative contribution of density stratification to the
generation of turbulent energy may be characterized by the ratio of the work
of buoyancy forces to the work of Reynolds stresses

Rf=
( au av) 7.
CpPo
az az
7".- + 7".-

called the flux Richardson number.

Quasi-steadiness.-The requirement of steadiness seems to be the most


limiting simplification accepted in models (Equations 2, 3a, 4). Under na­
tural conditions it is often disturbed both because of the synoptic evolution
of the pressure field (though usually slow) and as a result of diurnal changes
of radiative heat influx leading to changes of thermal stratification, then of
turbulence, and finally of wind. In the model formulated such changes can
be described by accepting the hypothesis of quasi-steadiness, i.e., by assum­
ing alI parameters of the atmospheric boundary layer to be dependent on
time only through the internal parameters u*, H, and E or their determining
external parameters G, oT=Th-T. and Oq=qh-q. (here and below sand h
denote values at the underlying surface and at the upper boundary of the
layer). The results of quasi-steady and unsteady models of the diurnal
changes of the atmospheric boundary layer wiII be compared below.

Structure of the atmospheric boundary layer (surface layer, dynamic sublayer,


viscous su blayer) . The lower part of the atmospheric boundary layer, in
-

which only slight changes with height are observed in not only the vertical
heat and humidity fluxes (Equations 3a and 4) but also the components of
vertical momentum flux
230 MONIN

au af)
Tz• + V-:::o< u.2, T�. + v-:::o< 0 8.
i)z i)z

is called the surface layer of air. In other words, the surface layer is the lower
part of the atmospheric boundary layer in which the action of the Coriolis
force can be neglected (and the surface layer should resemble the wall region
of boundary layers in nonrotating stratified fluid). The thickness of the sur­
face layer is tens of meters. In fact, we may show [see, for instance, (20)
§6.6] that the thickness of the layer in which the conditions of Eq. 8 are ful­
filled with an accuracy no lower than a percent does not exceed au.2/fG;
usually u*/G,-..,O.OS, and for G= 10 m/sec,f= 10-' sec -1, and a = 20 percent,
Annu. Rev. Fluid Mech. 1970.2:225-250. Downloaded from www.annualreviews.org

the thickness of the surface layer equals 50 m. A similar estimate confirms


rather well the quasi-stationarity of the layer with respect to diurnal and
synoptic changes of the external parameters.
Below, it is shown that the effect of density stratification on turbulence
by Ryerson University on 02/25/13. For personal use only.

in the atmospheric boundary layer decreases as the underlying surface is


approached, and there is a sublayer near the latter in which the influence
of stratification can be neglected (and it should resemble the wall region of a
boundary layer in homogeneous fluid). Consequently, turbulence in this
sublayer is determined only by dynamic factors, which is why it is called
the dynamic sublayer. Below we shall see that the thickness of the dynamic
sublayer is of the order of CppQU*3/(B! Hp+bHL!); in the atmospheric boun­
dary layer on the Earth it changes from several meters in the cases of very
strong hydrostatic instability or stability to very large values under neutral
stratification when Hp+bHL is close to zero (in which case the whole of
the surface layer can be dynamic).
All the dynamic parameters of the dynamic sublayer are well determined
by the two constant parameters v and u. and the roughness parameters of
the underlying surface-first of all, by the mean height of roughness h•. If
h8 <' (v/U.), roughness does not affect the structure of the dynamic sublayer;
in this case the underlying surface is called dynamically smooth. In its
vicinity, a viscous sublayer is formed in which Reynolds stresses are small
compared with viscous stresses. By omitting in the viscous sublayer the term
T",. in the first Equation 8 we obtain a linear velocity profile a(z) = (u.2z/v);
this law proves to apply to the layer z �5(v/u.) (in the atmosphere, about
one millimeter thick) . If h.»(v/u.) the current near the underlying surface
consists of vortices formed by flow past roughness elements, and no viscous
sublayer exists; in this case the surface is called dynamically rough. The
land surface is always such.

DYNAMIC SUBLAYER AND AIR-SURFACE INTERACTION


Dynamic interaction.-For z»(v/u.), h. all dynamic parameters of the
dynamic sublayer determined by the components of turbulence of not too
small scales (whose regime depends very little on molecular air viscosity)
THE ATMOSPHERIC BOUNDARY LAYER 231

can depend only on the one dimensional parameter u'" (and of course on the
height z). Such parameters include, evidently, the mean velocity gradient,
for which from dimensional considerations the following equation is ob­
tained: oU/OZ=U*/KZ (where K is the so-called von Karman constant for
which measurements in the dynamic sublayer yield the value 0.4). The well­
known logarithmic law follows from this:
U* Zl
U(Zl) - U(Z2) = - 19 - 9.
K Z2
The upper part of the dynamic sublayer, in which this law is fulfilled, is
Annu. Rev. Fluid Mech. 1970.2:225-250. Downloaded from www.annualreviews.org

sometimes called the logarithmic sublayer. If the drag coefficient Cf(z) is


defined as the ratio of turbulent stress POU*2 to pou2, and U=u*/ vCj is put
into Equation 9, the resulting equation for Cf will show that in the logarith­
mic sublayer the valuezo=z exp { -K/ VCf(Z)} does not depend on the height
by Ryerson University on 02/25/13. For personal use only.

z. This parameter of the logarithmic sublayer can depend only on the size
and form of mean roughness heights of the underlying surface and is there ­

fore called the roughness parameter (since the logarithmic sublayer is part
of the dynamic sublayer, Zo cannot depend on surface-layer stratification).
From the definition of Zo it follows that C,=K2/(Ig Z/20)2; by returning to
u=u",/VC, we obtain
U", Z
u(z) = -lg- 9a.
K Zo
Above land, values of Zo comprise only a small fraction of the height h.
and are equal to 0.001 (in centimeters) above a smooth snow surface, 0.03
above the sandy surface of deserts, 0.2 to 0.7 above mowed grass (h.= 1.5 to
3 em), 9.0 to 3.7 above high grass (h.=60 to 70 em) (for wind velocity of 1.5
to 6 m/sec), 10 above shrubs and trees, SO to 100 above forests (h.,,-,10 m),
and some 100 above cities [vide, for instance, Laikhtman (12) and Priestley
(21) and one of the latest papers by Lettau (22)]. The above-mentioned
dependence of Zo on wind velocity above high grass is explained by the bend­
ing of grass stems to the ground under the action of wind.
The logarithmic law (Eq. 9a) has so far been substantiated only on con­
dition that z»h If z/h. is not too large, h. can affect the wind velocity pro­
•.

file and then we have ou/oZ=(U*/KZ)CPl(h./z), where CPICn is some function


characterized by .pl(O) 1. Yet, if heights are counted from the level
=

Zl =.p/(O)h.. with an accuracy to the second order in the small quantity


h./(Z-ZI) we shall have au/a�U*/K(Z-Zl), the logarithmic law (Eq. 9a)
re maining true after z is replaced by Z-ZI' Such a reference level ZI can be
called the displacement height (by analogy with the displacement height
from boundary-layer theory). For high vegetation Zl usually lies between
h./2 and h•.

Specific character of the sea surfa ce. - Above the sea surface, Zo depends on
232 MONIN

a number of factors, and first of all on the local wind velocity. Heights of
roughness of the sea surface can be measured by the scale h. = U*2jg. When
the wind is weak [u* <' (gv)1/3 and h. <' (vj�)] the sea surface proves to be
dynamically smooth and zo=mo(IJju*), where, according to the experimental
data, m�O.l. Under moderate or strong wind the sea-surface resistance
changes depending on wind duration or fetch. For h.»(vju*) and fully
developed waves (when the average wave height exceeds 100 h.) we may
assume that zO=ml(u*2jg) [ Charnock (23)], where, according to the experi­
mental data, m{"0.035. The drag coefficient C, appears to depend on wind
velocity linearly at moderate winds, as was found by Munk (24), which
Annu. Rev. Fluid Mech. 1970.2:225-250. Downloaded from www.annualreviews.org

is confirmed by some direct measurements of momentum flux U*2= -u'w'


and wind profile u(z) [ Zubkovsky et al. (25, 26)]. See an approach to the
theoretical computation of the sea-surface resistance in the book by Phillips
(27), and reviews of the experimental data in Roll's book (28) and in the
by Ryerson University on 02/25/13. For personal use only.

papers by Kitaigorodsky and Volkov (29, 30).


Wind action on the sea surface results in the formation in the upper sea
layer of the so-called drift current, the direction of which, generally speak­
ing, differs from the direction of the tangential wind stress [i.e., according
to (Eq. 8), from the x-axis] ; the velocity of this current is usually one order
less than the friction velocity u*. Let u., v. be the components of this current
so that it forms with the x-axis an angle a. =tan-1(v./u.). The angle between
the wind direction and the x-axis near the sea surface should be the same,
yet it rapidly decreases with height to zero. According to Faller's (31) mea­
surements, in the temperate latitudes of the Northern Hemisphere a.�13°.
Laikhtman's (32) theoretical computations show that in the Northern
Hemisphere a�5° to 150 (and in the Southern Hemisphere a.<O).
Heat and humidity exchange.-In the dynamic sublayer, heat and humid­
ity are passive substances, ie ., they do not exert a dynamic influence on
.

turbulence and the buoyancy parameter (3 is not essential here. At the same
time, the hydrodynamic parameters associated with heat and humidity
transfer through the dynamic sublayer should depend respectively on the
parameters Hand E determined by Equations (3a) and 4. With the aid of
these parameters and the friction velocity u* we may determine the scales
of vertical change of temperature and specific humidity
H
T* = -
---
10.

where the numerical factor K has been introduced into the denominators for
convenience. Signs were chosen so that with temperature and humidity de­
creasing with height we have T*<0 and q* <0 (because then H>0 and
E>O), and, conversely, with temperature and humidity increasing with
height we have T*>0 and q*>0 (because in this case H <0 and E <0).
In the logarithmic sublayer, all statistical parameters of the temperature
and humidity fields determined by the components of turbulence of not too
THE ATMOSPHERIC BOUNDARY LAYER 233
small scale can depend only on the three constant dimensional parameters
u*, T*, and q* (and on the height z). Hence for the mean temperature and
humidity gradients the following equations are obtained:
aT T* aq q*
-- = --
, -=--,
f)z aOHZ f)z aOqz
where aOH and aOq are numerical constants (if the exchange coefficients for
momentum, heat, and humidity

�2 -H -E
11.
Annu. Rev. Fluid Mech. 1970.2:225-250. Downloaded from www.annualreviews.org

KM = -- , KH = --::::--, K = ---
f)u/iJz cppoiJTjih q poiJq/az
are introduced, the constants aOH and aOq will be the values of the ratios
aH=KH/KM and aq=Kq/KM in the logarithmic sublayer). Consequently,
by Ryerson University on 02/25/13. For personal use only.

for the mean temperature and humidity profiles in the logarithmic sublayer,
logarithmic equations similar to Equation 9 are obtained that can be written
in the form

T(z) - T. =
T*
aT, + - lg
Z
aOH Zo
- , q(z) - q. = aq. +
q*
-
aOq
z
19 -
20
12.
where oT. and oq. are parameters not depending on z, the determination of
which is equivalent to finding the coefficients of heat and humidity transfer
H E
-----
CH = , Cq = 13.
cpPou(T. - T) pou(q. - q)
(sometimes called the Stanton and Dalton numbers). The quantities aT./T*
and oq./q* should be regarded as functions, first of all, of the roughness
Reynolds number u*h./v. They are apparently close to each other; above
land, according to laboratory measurements, they are close to (1/5) (u*h./v)1/2
[ Owen & Thomson (33)].
Above the sea, the saturation humidity at temperature T. is taken as q..
and u2*/g can be used, as above, for h The available experimental data on
•.

values of CH and Cq (obtained under weak and moderate winds) show that
these coefficients change within two orders depending on a number of fac­
tors, and that the observed scatter can be significantly decreased by con­

1
sidering these coefficients as functions of u*h./" or of u*zo/" [vide Kitaigo­
rodsky & Volkov (34) and Fig. taken from the latter paper, as well as the
theoretical work by Bortkovsky & Byutner (35)].
Some authors support the possibility of using standard values of the
coefficients CH and Cq to compute turbulent heat and humidity fluxes above
the oceans [for instance, Robinson (36)]. Data in Fig. 1 show that the error
in such computations may be as large as many hundred percent. These errors
are particularly dangerous in attempts to use Equations 13 for the estima-
234 MONIN


Annu. Rev. Fluid Mech. 1970.2:225-250. Downloaded from www.annualreviews.org
by Ryerson University on 02/25/13. For personal use only.

FIG. 1

tion of the climatic mean values II and E from climatic maps of u, T, and
q, as was done by J akobs (37) and his followers: if CH and Cq increase rapidly
with wind velocity, storms should contribute significantly to the climatic
mean values Hand E, which cannot be taken into consideration in computa­
tions based on climatic maps.

SIMILARITY THEORY FOR THE SURFACE LAYER

Wind, temperature, and humidity profiles.-Above the dynamic sublayer,


heat and humidity cannot be regarded as passive substances and the number
of parameters determining the turbulent regime should be enlarged to in­
clude the buoyancy parameter {3. Then we can accept the following similar­
ity hypothesis [developed by Obukhov and Monin (38-41), whose pertinent
papers are listed in Chapter 4 of their book (20), as a generalization of the
similarity hypotheses for the logarithmic sublayer discussed above]: in the
surface layer with z»(v/u*), h. the laws governing vertical changes of the
statistical parameters of hydrodynamic fields determined by the components
of turbulence of not too small scale can depend only on the four constant
dimensional parameters u*, H/cppo, E/po, and {j.
As in the logarithmic sublayer, we can use the parameter u* as a velocity

Equations 10.
scale and determine the temperature and humidity scales T* and q* by
But whereas in the logarithmic sublayer the only reference
length. was the height z, in the surface layer the available dimensional pa­
rameters can be used to construct the length scale
THE ATMOSPHERIC BOUNDARY LAYER 235

14.

(For HL=O this is determined uniquely to within a numerical factor, and


the dependence on HL is accepted here in accordance with Equation 6 for
the work of buoyancy forces; the sign of L is chosen in such a way that L <0
for HT+bHL> 0, when the density stratification is lapse, and conversely
L>O when HT+bHL<O and the stratification is stable; the numerical factor
K is introduced into the denominator for convenience.) According to the ac­

cepted similarity hypothesis, the nondimensional one-point statistical pa­


Annu. Rev. Fluid Mech. 1970.2:225-250. Downloaded from www.annualreviews.org

rameters obtained when using scales L, u*, T*, and q* should be universal
functions of the nondimensional height f=z/L. In particular,

z aT cJ>(r)
--= , -- 15.
by Ryerson University on 02/25/13. For personal use only.

T* az aHG')
where cJ>(f) is some universal function, and aH and aq, as above, are the ratios
of exchange coefficients KH/ KM and Kq/KM' By substituting the first of
Equations 15 into 7 we obtain the relationship Rf=r/cJ>(f) for the flux
Richardson number, which shows that f and Rf are equivalent parameters
for the density stratification. With the aid of Equation 11 the following
equation is derived from Equation 7:

Rf = aHRi, Ri = (3
(aT
- + b-
.£,
-aHall -ailaz) (au)-z
-az 16.
az Cp

where Ri is the gradient Richardson number. By integrating Equations 15


we obtain

U(Zl) - U(Z2) :* �(�) - f(�)]


=

17.
T(Zl) - T(Z2) = T* �T(�) - iT (�)]

and a similar equation for the humidity profile q(z). (The available data
enable one to suppose that aq=aH and fq=Jr; therefore below we shall not
discuss humidity profiles.)
As neutral stratification is approached, i.e., for HT+bHL-,>O, when
ILI-'>oo and t=z/L-'>O, Equations 17 should reduce to the equations of
the logarithmic law (Eqs. 9, 12), and it must be that cJ>(O) = 1; then we have

f(r) � 19 t + const, 18.

In this case Rpr and Ri:::r::=. !aoH. It should be noted that the condition
236 MONIN
I rl «1 is reached also for fixed HT+bHL if z�O; this proves both the exist­

I
ence of the dynamic sublayer and the estimate of the order of LI for its
thickness. Somewhat higher than the dynamic sublayer, where rl <1, we
expand the right-hand sides of Equation 15 into power series in r and retain
only the linear terms of these series to obtain, after integration,

1(0 � 19 r + f3..l + const,

(I r I < 1)
Annu. Rev. Fluid Mech. 1970.2:225-250. Downloaded from www.annualreviews.org

The numbers f3u, {3T are different for r >0 and r <0 (but all of them should be
positive to provide the flattening of wind and temperature profiles caused by
intensification of turbulent mixing under an increasing lapse stratification).
by Ryerson University on 02/25/13. For personal use only.

Equations 18a constitute the "log-plus-linear law" for wind and temperature
profiles suggested in (40, 41).
Under very strong instability, i.e., at large positive HT+bHL, the values
of L prove to be small and negative, and r=zjL is large and negative. Such
values of r are reached for fixed HT+bHL >0 as well, if u*-40, and the
limiting state of strong instability is free convection. For the latter, the
second Equation 17 should not contain u*, which is possible only for
., -1/3. Note, however, that exclusion of the parameter u* does not yet
fT(S),-...S
mean the absence of wind; it is only required that momentum become, so to
say, a passive substance. Therefore the limita-OOH of the value ofaH=KHjKM
as r� - 00 can be finite. But in this case f(t) "'t -l/a as well, and we have

fer) � cr-i + const, 19.

where C is a numerical constant. (Then Rf=orooHRi= - 3r4/3j C ) This .

"minus-one-third law" was advanced in (38, 39); it was later suggested by


Priestley (42) for h(t) from other considerations (Priestley, in addition,
established empirically that it is applicable already for t < -0.03); see also
the paper by Kazansky & Monin (43).
Under very strong stability, i.e., at large negative HT+bHL, the values of
L prove to be small and positive and r=zjL is large and positive. Buoyancy
forces interfere with the development of turbulence, and the energy equa­
tion (Eq. 5) shows that turbulence can be maintained only at not too large
Rf. Since Rf evidently increases as r grows, the presence of a finite limit R
on Rf for large r>O follows from the restriction of Rf. Then cp(t)�(r/R).
As stability grows, the value of aH apparently decreases. [As a model of
limiting stability we can use a free surface of fluid through which turbulent
heat transfer is impossible yet momentum can be transferred by pressure
fluctuations, and a H =O, vide Stewart (44).] If aH at large positive t has a
nonzero limit aooH, it follows from Equation 15 that
THE ATMOSPHERIC BOUNDARY LAYER 237

-I-
Annu. Rev. Fluid Mech. 1970.2:225-250. Downloaded from www.annualreviews.org
by Ryerson University on 02/25/13. For personal use only.

-3

-4 �--�--�----�
-J -2. -I 0 (

?=�/L
FIG. 2

t
f(t) � - + const, (r» 1) 20.
R
If aooH=O, fT(S) will grow with the growth of its argument more rapidly
than according to the linear law of Equation 20.
The theoretical predictions contained in Equations 17 to 20 agree well
with experimental data [summed up in Chapter 4 of (20) and in the book by
Lumley & Panofsky (45)]. As an example, Figure 2 presents an empirical
graph of the function f(t), constructed for the first time in (40) and clearly
showing the validity of the asymptotic laws, Equations 18 to 20. The nu­
merical constants in these laws were most thoroughly evaluated by Zilitinke­
vich & Chalikov (46), who used the abundant measurements in the range
-1.2 <t<0.4 and obtained the values" ",,0.43, aOH ",,0.83, (3u ",,9.9 under
stable and 1.4 under lapse stratification, (3T "" 10.4 under stable and 2.0 under
lapse stratification, C.., 1.25, and a-ooH ""0.87; if the additive constants in
238 MONIN
Equation 18a are taken as equal to zero, those in Equation 19 become 0.24
and 0.11. For rough estimates it is recommended in (46) to assume that
both! and iT are given by (1+5r-1/3)/4 for r< -0.07, by 19 1 rl for O?:r?:
-0.07, and by 19 r+l0r for t>O.
It should be noted, however, that according to measurements in Aus­
tralia that covered cases of stronger instability (up to !'= -4.5), O',-""H
appears to be much larger-about 3.5 [vide Charnock (47)]. Moreover, ac­
cording to some data [summed up in § 8.2 of (20); see also (48)], for r<-1
the "minus-one-third law" of Equation 19 for temperature profiles is dis­
turbed, being replaced by a still more rapid flattening of temperature with
height, as if aH(t) grows according to a power law (the so-called "windless
Annu. Rev. Fluid Mech. 1970.2:225-250. Downloaded from www.annualreviews.org

convection"). A discrepancy is also found in the case of very strong stability:


if for R( = l/fJu) most authors obtained values of the order of 0.1, for a""H
(=fJ.JfJT) in some measurements [summed up in § 8.2 of (20) and in (49) J
very small values, of the order of 0.01, were obtained, and Ri=RflaH ap­
by Ryerson University on 02/25/13. For personal use only.

pears to be of the order of 10.

Turbulent fluctuations.-The similarity hypothesis formulated in the


preceding section can be applied to the statistical parameters of turbulent
fluctuations of velocity, temperature, and humidity in the surface layer­
their one-point moments, correlation and structure functions, and spectra
(5Q.-53). Nondimensional root-mean-square values of these fluctuations
u../u*, u.lu*, uw/u*, uTIIT* I and uqllq*1 should be functions of t=z/L.
According to measurements [vide § 5.3 and § 8.5 of (20)], the first three of
these are equal in the logarithmic sublayer to 2.3,1.7,and 0.9 (some authors
have obtained somewhat larger values for uwlu* in the atmosphere). As the
stratification changes from stable to lapse, these three functions tend to
increase; for strong instability they should grow like 1 til/a, tending to hori­
zontral isotropy and an excess of Uw over u.. "'"0'••
Temperature fluctuations are found to be dependent on stratification in
a peculiar way. Under adiabatic temperature stratification they are very
small, since mixing of an adiabatically stratified layer cannot result in
temperature fluctuations; therefore in this case the heat flux tends to zero
(54). However, in both lapse and stable stratification the temperature
fluctuations grow-in the latter case, despite the fact that stable stratifica­
tion suppresses turbulence. Under strong instability uTI I T* 1 should decrease
like Itl-I/3; under strong stability it approaches a constant.
The u' and w' correlation coefficient is negative; according to measure­
ments, under neutral stratification it is close to -0.5, its absolute value de­
creasing as instability grows. The T' and w' correlation coefficient is posi­
tive; under neutral stratification it is close to 0.5 and apparently grows with
increasing instability. Hence it can be seen that u' and T' should mostly have
opposite signs so that there should exist a horizontal turbulent heat flux
Hl = cpPou'T' opposite in sign to the vertical heat flux HT and tending to
zero only under neutral stratification and strong instability (the same should
THE ATMOSPHERIC BOUNDARY LAYER 239

be true for humidity fluxes). The direct measurements of HI made by


Zubkovsky & Zvang (55) yielded values for HI/HT ranging from -3 for
weak instability to -1 for moderate instab i lity .

The vertical turbulent fluxes of momentum Pou*2= -Pou' w', heat


HT= cpPoT'w', and humidity E= Poq'w' can be measured directly from
synchronous recordings of fluctuations in u', w', T', an!;l q' or may be com­
puted from measured profiles of u(z), T(z), and q(z). Let, for instance, the
wind velocity u at a fixed height z and the temperature and humidity dif­
ferences aT and aq at heights z/2 and 2z be measured. In this case u*/u,
HT!cpPouoT, and E/pouoq (the latter two apparently coinciding) are func­
tions of the empirical Richardson number (3z(oT+b£oq/cp)u-2 and non­
Annu. Rev. Fluid Mech. 1970.2:225-250. Downloaded from www.annualreviews.org

dimensional roughness parameter zo/z. These functions can be expressed by


means of the functionsj(r) and!T('\) of Equations 17, and knowing the lat­
ter, one may construct corresponding nomograms. Such nomograms were
constructed by Kazansky & Monin (43, 56,57); they were recently improved
by Ryerson University on 02/25/13. For personal use only.

by Zilitinkevich & Chalikov (58).


The nomograms for a height of measurement z = 1m show that as the
stratification changes from strong stability (Ri,......,0.3) to strong instability
(Ri......,-0.05) the value of u*lu increases by a factor of 1.5 to 2 (from
0.06 or 0.07 to 0.11 or 0.12), and HT/CppouoT becomes 25 times as great.
Since strong stability (oT <0) is observed, as a rule, under weak wind and
since, on the contrary, no calm weather is usually observed under strong
instability, the product 1uoTI also increases (by several times), and HT
changes by two orders: stratification influences turbulent heat and humidity
transfer much more strongly than it affects momentum transfer. Velocity
fluctuations of the order of u* are influenced by only a factor of several times,
whereas temperature fluctuations of the order of T* change by factors of ten.
Let us point out an empirical estimate of the contribution of humidity
to density stratification given by blBo, where Bo=HTIHL is the Bowen
ratio [see Equations 6 and 14] j this contribution is essential above the sea,
where in two-thirds of the cases it proves to be positive, of the order of 0.2
to 1.0, yet in 10 to 15 percent of the cases b/Bo is negative, being sometimes
even less than -1 (in the latter case the humidity contribution is dominant).
The energy dissipation E in the dynamic sublayer is U*8/KZ, and under
strong instability it approaches a constant proportional to the work of
buoyancy forces (Equation 6). Having a similar meaning, the rate of decay
of temperature inhomogeneities ET=X(VT)2 in the dynamic sublayer is
Ku*T*2/aOHz, and under strong instability it attenuates like 1.\1-413 [see § 7.5
of (20)]. The contribution of turbulent energy diffusion to the energy equa­
tion (Eq. 5) in the dynamic sublayer is small, and under strong instability
it becomes comparable with that of the work of buoyancy forces [see (45)].
When turbulent fluctuations are measured by the scales u*, T*, q*, and
wave numbers k by the scale liz, the nondimensional space spectra of hori­
zontal fields of fluctuation in the region of not too large k will be functions
only of the nondimensional wave number kz and the stratification parameter
240 MONIN
r=z/L or Ri. From the space spectra of fluctuation fields on straight lines
parallel to the wind direction, one may pass to frequency spectra of fluctua­
tions at a fixed point of space with the aid of the frozen turbulence hypothe­
sis of G. I. Taylor, i.e., by assuming k =w/ u. For instance, the spectral den­
sity of vertical velocity fluctuations will have the form
U.2z (WZ )
SIb(W) = Sib --:: I Ri 21.
u %�
-_-

where Sw is some universal function. The results of measurements of spec­


tra and co-spectra of the fluctuations of u', v', w', T', and q' are discussed
Annu. Rev. Fluid Mech. 1970.2:225-250. Downloaded from www.annualreviews.org

in (52), (45) and in § 23 of (20) [vide also a recent paper by Zubkovsky &
Koprov (59) on the spectra of momentum and heat fluxes]. Measurements
show that within a wide range of nondimensional wave numbers the spectra
prove to be proportional to (wz/ u)-5/S, i.e., the Kolmogorov-Obukhov "minus­
by Ryerson University on 02/25/13. For personal use only.

fiv e -thi rds law" is fulfilled. The lower boundary of this inertial range ap­
2.5 at Ri = -0.76;
pears to be dependent on stratification; for Sw(w) it iswz/ u=
4.5 at Ri=O, and 12 at Ri=0.28. With a further decrease of w, sometimes
at sufficiently large heights under lapse stratification, velocity fluctuation
spectra grow slower and those of temperature faster than according to
the "minus-five-thirds law." Under stable stratification, on the contrary,
velocity-fluctuation spectra grow more rapidly and those of temperature
slower [see, for instance, the theoretical computation made by Monin (60)
and § 2 1.7 of (20)]. Further, the growth of spectra ceases, their maximum
being observed at periods of the order of a minute.

Semi-empirical theories of the surface layer.-Attempt s at theore tic al


determination of the universal functions, the existence of which follows from
the similarity theory [a review of such attempts being given in §§ 7.4 to 7.5
of (20)], have been based on the use, first of all, of the energy equation (Eq.
5). In this equation it is convenient, instead of E, to introduce the turbulence
scale f = (KM8/E)1/4, which in the logarithmic sublayer is KZ, and at greater
heights assumes the form KZ·A<t), where A(t) is some universal function.
Let (u- 1) be the ratio of the diffusivity term in Equation 5 to the work of
buoyancy forces; in this case Equation 5 will be
cP' - uN3 - X-' = 0 Sa.

In the simple semi-empirical theory, it is assumed that X= 1 and u =const


(and aH=const); thencp = (1-uRf)-1/4, and u is the inverse of the constant R
in Equations 20; for u�12 to 14 such a theory agrees well with measure­
ments. In a more detailed theory, the vertical flux of turbulent energy was
taken as equal to Q= -PoKQ(iJB2/iJZ), where B2 is the kinetic energy of
turbulence per unit mass, related to KM by KM""B( [ Monin (39)]. More
detai led equations were also suggested for the turbulence scale (; for instance,
by generalizing th e well-known von Karman equation (= -K(iJfl/iJZ)/(iJ2fl/iJz2)
THE ATMOSPHERIC BOUNDARY LAYER 24 1
,
Zilitinkevich & Laikhtman (61) arrived at the relationship ( = -K'If/(a'If/a1.)
where 'If=Bj( (considered together with Equation 5, and in the dynamic
sublayer turning to von Karman's equation, and for r« -1 and r»1 giving
the asymptotes (=3/2KZ and 1/2Kz).
The most detailed semi-empirical theory of the surface layer published
so far is that developed by Monin (62). Instead of one energy equation
(Eq. 5), he examined dynamic equations for all one-point second moments
of the fluctuations u', v', w', and T' (neglecting the third moments of these
fluctuations, i.e. diffusivity terms, and using relationships of the type

(
p' au.' au/ ) -
Annu. Rev. Fluid Mech. 1970.2:225-250. Downloaded from www.annualreviews.org

B __ __

- -+ - = Cl-t
Po ax; ax.
22.
by Ryerson University on 02/25/13. For personal use only.

where ni are the components of the unit vertical vector and CIt C2 are nu­
merical constants). It became possible to express all the universal functions,
including aH(r). in terms of X(r). [It turned out that X( - co ) should be
finite, and X,.",l/r ast--tco.]
DYNAMICS OF THE ATMOSl'HERIC BOUNDARY LAYER

-
Similarity theory for the atmospheric boundary layer. Above the surface
layer, the influence of the Coriolis force becomes apparent; therefore the
Coriolis parameter should be added to the four parameters u*, H/cppo,
E/po, and {3 taken into consideration in the similarity theory for the surface
layer. Then in addition to L, one more length scale h=Ku*/f can be formed
(which can be interpreted as the thickness of the atmospheric boundary
layer, and the factor c in the equation h=c(G/f) in our first section will
be c=Ku*/f; it has values of the order of 10-2). Consequently, besides the
nondimensional parameter of stratification t=z/L (either Rf or Ri) which
changes with height we have the constant parameter of stratification

Jl = -
h

L
= -
K2{3(HT + bH L)
cpPou.2j
23.

[introduced by Kazansky & Monin (63)]. Thus all nondimensional func­


tions of t discussed on pages 234 to 241 will depend in the atmospheric
boundary layer also on the parameter Jl. (It is more convenient to regard
them as functions of7J= z/h=rlJl and Jl, which at small7J, i.e., in the surface
layer, depend only on the product of their arguments 7JJl=r) Such a simi­
larity theory for the atmospheric boundary layer is formulated in (63); for
neutral stratification (Jl=O) the theory was used by Monin (64) as far back
as 1950.
By writing equations of the type of Equation 17 (with the nondimen-
242 MONIN

sional functions depending also on p.) for the velocity components Uh G cosa,
=

and 'lJh G sina at the upper boundary of the atmospheric boundary layer and
=

the differences oTo= Th- T(zo) and oqo qh-q(ZO) we shall see that G/u*,
= ,

oTo/T*, and oqo/ q* are functions of zo/h and JL. These three relationships
can be used to express the internal parameters u*, H/cpPo, and E/po in
terms of the external parameters G, oTo, oqo, Zo, and the constant parameters
(3 and! [the true external parameters are oT=Th-T. and Oq=qh-q., but
the differences oT. = T(zo) T. and og. q(zo) -g. should be determined from
- =

Equations 12 and 13]. The latter six parameters can be used to construct
two nondimensional combinations, by choosing for the purpose the Rossby
number Ro and the external parameter of stratification S determined by the
Annu. Rev. Fluid Mech. 1970.2:225-250. Downloaded from www.annualreviews.org

equations
G
Ro = -, 24.
by Ryerson University on 02/25/13. For personal use only.

zo!

Thus for the atmospheric boundary layer the following similarity hypothe­
sis can be accepted: for z»v/u*, h. the nondimensional statistical parameters
of the hydrodynamical fields determined by the components of turbulence
of not too small scales and obtained using the scales of velocity G, length
G/f, temperature oTo, and humidity ogo, can depend only on the two con­
stant nondimensional parameters Ro and S [Zilitinkevich & Monin (65, 66)].
The relationships connecting the internal and external parameters (the
laws of resistance and heat and humidity exchange for the atmospheric
boundary layer) are derived with the aid of logarithmic asymptotics of the
dynamic sublayer. They have the form

- A 2(p.);

sin a =
oTo
- = -
1 [ (u.Ro)
19 - - C(IL)
] 25.

T* aOH G

where A, B, C are functions of p. whose form cannot be predicted by logarith­


mic asymptotics. The equation forogo/g*is entirely analogous to Equation 25
for oTo/T*. The first of these equations (for p.= 0) was obtained for the first
time by Kazansky & Monin (67), the case JL=O being discussed also in the
papers by Gill (68), Charnock & Ellison (69), and Blackadar (70). (The latter
author considers also a stratified atmospheric boundary layer but uses as
the stability parameter zo/L, though according to Equation 17, Zo should
not enter the equations for velocity and temperature differences.)
Reviews of empirical data on the dependence of the geostrophic drag
coefficient u./G and wind turning angle a upon Ro (without taking into ac­
count stratification) are given by Lettau (71, 72) and Blackadar (70, 73).
According to these data, when Ro increases frpm 105 to 1010, u*/G decreases
THE ATMOSPHERIC BOUNDARY LAYER 243

S >0

------r----+--� p,�

_ .

9 ��������_____
Annu. Rev. Fluid Mech. 1970.2:225-250. Downloaded from www.annualreviews.org

----�-------+------� ��
;,
�I $ I
by Ryerson University on 02/25/13. For personal use only.

FIG. 3

roughly from 0.05 to 0.02, and a from 35° to 13°. According to the empirical
data obtained by Byzova & Mashkova (74) on u*/G and those of Kurpakova
& Orlenko (75) on a , as stratification changes from lapse to stable, u*/G
decreases (and at fixed G the thickness of the atmospheric boundary layer
decreases as well), and a grows.
In (68) are given estimates for A (O) ",, 4.2 to 4.7 and B(O) ",, 1 to 2 ; close
estimates [2 <A (0) <3 and B(O) "" 2 ] are given in (69). Empirical depen­
dencies of A , B, C on JL were derived by Zilitinkevich & Chalikov (76)
[they are presented also in (65) ] ; according to their data, as stability grows,
A increases and B and C decrease to large negative values. Proceeding from
these results Chalikov (77) constructed nomograms for the determination
of u*/G, a, and H/cppoG�To "" E/poG�qo as functions of IgRo and 19l sl , which
are contained in Fi gs. 3 to 5. Values of IgRo in these nomograms are written
on the curves. The above-mentioned nomograms can be used for describing
the atmospheric boundary layer in mathematical models of the general
atmospheric circulation.
The nondimensional wind velocity profiles of VU2+V2/U* as a function
of T] =z/h at different JL were constructed by Kazansky & M onin (63) and
Byzova & Mashkova (74). Values of VU2+V2/U* at each fixed T] proved to
be regularly increasing as JL grows, and profiles (63) under stable stratifica­
tion (JL >O) had well pronounced maxima at heights of 0.2 to 0.3 1] ; since at
large JL > O the atmospheric boundary layer is thin, these maxima look like
"low-level jet streams." The empirical dependencies of the values of
(G cos a - u)/u*, (G sin a - v)/u* on T] at JL =O were constructed by Gill (68) ,
and the nondimensional hodographs (spirals) of the wind vector (u/ G, v/G)
as a function of T] under lapse , neutral, and stable stratification were con­
structed by Kurpakova & Orlenko (75) . Examples of the nondimensional
temperature profiles ( T - T1) / 1 T* I as a function of T] at different JL are pre­
sented by Mashkova (78).
Annu. Rev. Fluid Mech. 1970.2:225-250. Downloaded from www.annualreviews.org
by Ryerson University on 02/25/13. For personal use only.

!�

244

.10
-1-0

10
20
50

2
S� O

!J
.,
6

��

FIG. 5
(/J

FIG. 4
MONIN

� Is I
1
S �O

t2
THE ATMOSPHERIC BOUNDARY LAYER 245

Most of the above-mentioned empirical data can be found in a review by


Zilitinkevich, Laikhtman & Monin (79) who summarized also the abundant
data available on the values of turbulent viscosity KM and energy dissipa­
tion (;.

Semi-empirical theories of the atmospheric boundary layer.-For a long


time the theory of the atmospheric boundary layer consisted only in the
solution of Reynolds equations (Eqs. 2), in which (T",. , Til') were taken as
proportional to «(Jajaz, (Jfj/(Jz) , and the proportionality coefficient KM was
given as some simple function of s [a review of such papers can be found,
Annu. Rev. Fluid Mech. 1970.2:225-250. Downloaded from www.annualreviews.org

for instance, in (79» ) . However, it would be more correct if KM were not


given but found simultaneously with a(s) and ti(s). Monin (64) used for this
purpose the energy equation (Eq. 5) in the case of neutral stratification, yet
taking into consideration the diffusivity term Q = -POKM«(}B2/(}S), where
by Ryerson University on 02/25/13. For personal use only.

KM �B(, and solved Equations 2 and 5 numerically, the turbulence scale


( = "s alone being given. For the first time, the dependencies of u*/G and 0:
upon Ro were computed.
During recent years, this approach has attracted the attention of various
authors who have made similar computations with more detailed equations for
the turbulence length scale (yet without consideration of the diffusivity term) :
Blackadar (73) for ( = Ks( l + Cfs/G)-l, Lettau (72) for ( = Ks [1 + C(s/W/4)-1,
Appleby & Ohmsted (80) for (=(", [ l - exp( -Kz!f",» ) , and Zilitinkevich,
Laikhtman & Tseitin (81) for t= - KW/«(}W/(}z), where W = B/f. Finally,
Bobyleva, Zilitinkevich & Laikhtman (82) used the latter equation for
computing a stratified atmospheric boundary layer (with the diffusivity
term) . Their computation of the layer appears to be the most complete one
of all those made so far. It contains the computation of �/G and 0: as func­
tions of Ro and p., as well as the nondimensional coefficients of turbulent
viscosity KM/"u*h and energy dissipation KhE/u*8, and the "velocity de­
ficiency" components (K/U*) (u - G cos 0:) and (K/U*) (fJ - G sin o:) as functions
of z/h at different J.I. [ these results are presented also in the review (79) ] ;
note the clearly defined "low-level jet streams" under stable stratification
in the two latter graphs.
Zilitinkevich & Chalikov (83) found that results close to (82) can be ob­
tained if KM/KU* I L I is taken under lapse stratification in the form I r l for
I r l Sm and m l tlm l 418 for I r l � m, and under stable stratification in the
form r for r sn and n for r �n, where r = z/L, L being the scale of the dy­
namic sublayer (Equation 14) , and m = 0.064 and n = O. l are numerical
constants. All the results, including the functions A (p.), B(p.), C(p.) of Equa­
tions 25 can be obtained here in analytic form without any difficulty.
The method of computing wind and temperature profiles and one-point
second moments of turbulent fluctuations in the surface layer suggested in
(62) was extended by Monin in (84) to the neutral and in (85) to the strati­
fied atmospheric boundary layer. In so doing, there were established, in
particular, the relations
246 MONIN

-- = --

iJuliJz Mias
and

(say, = tan if;),

and for the coefficients of velocity fluctuation anisotropy the following equa­
tions were derived:
Annu. Rev. Fluid Mech. 1970.2:225-250. Downloaded from www.annualreviews.org

<T�2 <T�o 2 M sin2 if;


- = -+ ,
B2 B02 1 - Rf
26.
by Ryerson University on 02/25/13. For personal use only.

MRf
- = - - --- ,
B2 Bo2 1 - Rf

where M (qvo2 - qvo2)/Bo2, and subscript zero denotes values under neutral
=

stratification. Finally, for KM and the nondimensional momentum flux


r = (r",.+ir".)/u*2 were obtained the equations

27.

Unsteadiness.-If the hypothesis of quasistationarity of the atmospheric


boundary layer formulated in an earlier section is not used, then to describe
changes of the layer in time t it is necessary to replace the zeros on the right­
hand sides of the Reynolds equations (Eqs. 2) by the terms auNt and
ail/at, and the equations of heat influx and humidity transfer should be writ­
ten as poTo(as/at) -aQ./az, po(agjiJt) = -iJQq/az, where Q. and Qq are
=

vertical entropy and humidity fluxes. In papers devoted to diurnal changes


of the parameters of the atmospheric boundary layer, such equations have
been used for a long time in a very simple form: turbulent fluxes of momen­
tum, heat, and humidity were determined from equations of the type of
Equations 11, with the coefficients KM, KH, Kg given in the form of products
of simple functions of z by periodic functions of t.
In the last few years, some authors dealing with diurnal changes of the
atmospheric boundary layer began to express the exchange coefficients in
terms of wind velocity and temperature gradients through relationships
allied to the stationary energy equation (Eq. 5) [Estoque (86), Sharon (87),
Krishna (88)]. Finally, Yager & Zilitinkevich (89) used the unsteady form
of Equation S, adding the term iJB2/iJt to its right side and solving it together
with the unsteady equations of motion and heat transfer, the values of
temperature r(zo, t) at the lower boundary of the atmospheric boundary
layer being given as periodic functions of time (with amplitude Tm) .
THE ATMOSPHERIC BOUNDARY LAYER 247
In particular, they computed diurnal changes of the nondimensional
parameters It*/G, a, and H/cpPoGTm and found from the latter the diurnal
changes of the parameter JI. = JI. (t) . By excluding the time t, Yager & Zilitin­
kevich could construct curves of the dependencies of u*/G, a , and H/cpPoGTm
on JI. in the form of " hysteresis loops" which have, at each value of JI., two
ordinates, one of which corresponds to an i ncrease, and the other to a
decrease of JI.. For u*/G and particularly for H/cpPoGTm these loops turned
out to be rather narrow, i . e . , nonstationarity affects but little the results of
computing these i mportant parameters. This can j ustify the use of the
hypothesis of quasisteadiness of the atmospheric boundary layer.
Annu. Rev. Fluid Mech. 1970.2:225-250. Downloaded from www.annualreviews.org
by Ryerson University on 02/25/13. For personal use only.
248 MONIN

LITERATURE CITED
1. Ekman, V. W., A rkiv. Mat. Astron. 23. Charnock, H., Quart. J. Roy. Meteorol.
Fiz., 2, No. 1 1, 1-52 (1905-1906) Soc., 81, No. 350, 639-40 (1955)
2 . Charney, J. G., Okeanologiya, 9, No. 24. Munk, W. H., Quart. J. Roy. Meteorol.
1, 143-45 (1969) Soc., 8 1 , No. 349, 320-32 (1955)
3. Faller, A. J., J. Fluid Mech. , 15, No. 4, 25. Zubkovsky, S. L., Timanovsky, D. F.,
560-76 (1963) IZ1I. A kad. Nauk SSSR, A tmo-
4. Tatro, P. R., Mollo-Christensen, E. spheric and Oceanic Physics, 1, No.
L., J. Fluid Mech., 28, No. 3, 531- 10, 1005-13 (1965)
45 (1967) 26. Zubkovsky, S. L., Kravchenko, T. K.,
5 . Green, A. W., An experimenlal study of IZ1I. Akad. Nauk SSSR, A tmo-
the interactions between non-steady spheric and Oceanic Physics, 3, No.
Ekman layers and an annular vor- 2, 1 2 7-35 (1967)
Annu. Rev. Fluid Mech. 1970.2:225-250. Downloaded from www.annualreviews.org

tex (Doctoral Thesis, M IT, Cam- 27. Phillips, O. M., The Dynamics of the
bridge, Mass., 1968) Upper Ocean (Cambridge Univ.
6. Kolesnikova, V. N., Monin, A. S., Press, London, 1966)
IZ1I. A kad. Nauk SSSR, A tmo- 28. Roll, H. U., Physics of the Marine
spheric Oceanic Physics, 1, No. 7, A tmosphere (Academic, New York-
653-69 (1965) London, 1965)
by Ryerson University on 02/25/13. For personal use only.

7. Kolesnikova, V. N., Monin, A. S., 29. Kitaigorodsky, S. A., Volkov, Yu. A.,
Meteorol. Res., 16, 30-56 (1968) Izv. A kad. Nauk SSSR, A lmo-
8. Van der Hoven, J., J. Meteorol., 14, spheric and Oceanic Physics, 1 , No.
No. 2, 1 60-64 (1957) 9, 973-88 (1965)
9. Dyubyuk, A. F., TT. Res. Inst. GUGMS, 30. Kitaigorodsky, S. A., Izv. A kad. Nauk
2, No. 4, 18-51 (1947) SSSR, A tmospheric and Oceanic
10. Priestley, C. H. B., Phys. Fluids, Physics, 4, No. 8, 870-78 (1968)
Suppl., 38-46 (1967) 31. Faller, A. J., Tel/us, 16, No. 3, 363-70
1 1 . Utina, Z. M., Tr. GI. Geojiz. Observ., (1964)
No. 187, 146-48 (1966) 32. Laikhtman, D. L., Izv. A kad. Nauk
1 2 . Laikhtman, D. L., Physics ofthe A tmos- SSSR, A tmospheric and Oceanic
pheric Boundary Layer, 142-94 Physics, 3, No. 10, 1 0 1 7-25 (1966)
(Gidrometizdat, 1961) 33. Owen, P. R., Thomson, W. H., J.
13. Panofsky, H . A., Townsend, A. A., Fluid Mech., 15, No. 3, 32 1-34
Quart. J. Roy. Meteorol. Soc., 90, (1963)
No. 384, 147-55 (1964)
34. Kitaigorodsky, S. A., Volkov, Yu. A.,
14. Oliphant, J. E., Panofsky, H. A., Final
Izv. A kad. Nauk SSSR, A tmo-
Rep. Contract No. AF (604)-6641,
spheric and Oceanic Physics, 1 , No.
Pennsylvania State Univ. (1965) 12, 1 3 1 7-36 (1965)1
15. Townsend, A. A., J. Fluid Mech., 22,
35 . Bortkovsky, R . S ., Byutner, E . K.
No. 4, 799-822 (1965)
16. Townsend, A. A., J. Fluid Mech., 23,
Iz'll. A kad. Nauk SSSR, A lmo-
spheric and Oceanic Physics, S, No.
No. 4, 767-78 (1965)
5, 494-503 (1969)
1 7. Blackadar, A. A., Panofsky, H. A.,
36. Robinson, G. D., Quart. J. Roy.
Glass, P. E., Boogard, J. F., Phys.
McleoroI. Soc., 92 , N o. 394, 451-65
Fluids, Suppl., . 209-1 1 (1967)
( 1966)
1 8. Monin, A. S ., Izv. A kad. Nauk SSSR,
37. Jakobs, W. C., Compendium Meteorol.,
Ser. Geograph. Geojiz., 3, No. 3,
220-37 (1949) 1057-70 (Am. Meteorol. Soc., Bos-
19. Monin, A. S., Dokl. A kad. Nauk SSSR, ton, Mass., 1951)
175, No. 4, 819-22 (1967) 38. Obukhov, A. M., Tr. Inst. Tear.
20. Monin, A. S., Yaglom, A. M., Statistical Geojiz. A kad. Nauk SSSR, 1, 95-
Hydromechanics (Nauka Publishers, 1 1 5 (1946)
Moscow, Part 1, 1965; Part 2, 1967) 39. Monin, A. S ., Inform. Bull. Chief A d-
2 1 . Priestley, C. H. B., Turbulent Transfer ministr. Hydrometeorol. Serv., No.
in the Lower A tmosphere (Chicago 1, 13-27 (1950)
Univ. Press, 1959) 40. Monin, A. S., Obukhov, A. M., Dakl.
22. Lettau, H. H., In The Collection and A kad. Nauk SSSR, 93, No. 2,
Processing of Field Data, 3-40 (In- 223-26 (1953)
terscience, New York, 1967) 41. Monin, A. S., Obukhov, A. M., Tr.
THE ATMOSPHERIC BOUNDARY LAYER 249
Geoftz. Inst. A kad. Nauk SSSR, 6 1 . Zilitinkevich, S.S., Laikhtman, D. L.
No. 24 ( 1 5 1 ) , 1 63-87 (1954) Tr. GI. Geoftz. Observ. , No. 167, 44-
42. Priestley, C. H. B., Quart. J. Roy. Met­ 8 (1965)
eorol. Soc., 81, No. 348, 1 39-43 62. Monin, A. S., Izv. A kad. Nauk SSSR,
(1955) A tmospheric and Oceanic Physics,
43 . Kazansky, A. B., Monin, A. S., IZ1/. 1, No . 1 , 45-54 (1965)
A kad. Nauk SSSR, Ser. geoftz., 63. Kazansky, A. B., Monin, A. S., IZ1/.
No. 6, 741-51 (1958) A kad. Nauk SSSR , Ser. Geoftz., No.
44. Stewart, R. W., Advan. Geophys., 6, I, 165-8 (1960)
303-1 1 (1959) 64. Monin, A. 5., Izv. A kad. Nauk SSSR ,
45. Lumley, J. L., Panofsky, H. A., The Ser. Geograph. Geoftz., 14, No. 3,
Structure of A tmospheric Turbulence 232-54 (1950)
(Interscience, New York, 1 964) 65. Zilitinkevich, S. S., Monin, A. S .,
46. Zilitinkevich, S. S., Chalikov, D. V., "Global A tmospheric Research Pro­
gramme, " Rep. Study Conf. Stock­
Annu. Rev. Fluid Mech. 1970.2:225-250. Downloaded from www.annualreviews.org

Izv. Akad. Nauk SSSR, A tmo­


spheric (:J> Oceanic Physics, 4, No. holm 28 June-11 July, 1 967, 5, 1-37
3, 294-302 (1968) (Stockholm, 1967)
47. Charnock, H., Quart. J. Roy. Meteorol. 66. Monin. A. S. Turbulence in the atmos­
Soc., 93, No. 395, 97-100 (1967) pheric boundary layer, Phys . .
48. Deardorff, J. W., Willis, G. E ., Quart. Fluids, 10, No. 9, Part I I , Suppl.,
by Ryerson University on 02/25/13. For personal use only.

J. Roy. Meteorol. Soc., 93, No. 39 6, Boundary Layers and Turbulence,


166-75 (1967) pp. 831-837 (1967)
49. Monin, A. S., A tmospheric Turbulence 67. Kazansky, A . B., Monin, A. S ., IZ1/.
and Radiowave Propagation, 1 1 3- A kad. Nauk SSSR, Ser. Geoftz., No.
20 (Nauka Publishers, Moscow, 5, 786-8 (1961)
1967) 68. Gill, A. E., The Turbulent Ekman Layer
50. Monin, A. S., Theory of probability and (Dept. ApDl. Math. Theor. Phys.,
its application, 3, No. 3, 2 85-31 7 Cambridge Univ., 1967)
(1958) 69. Charnock, H., Ellison, T. H., Global
51. Monin, A. 5., J. Geophys. Res., 64, No. A tmospheric Research Programme,
12, 2 196-97 (1959) Rept. Study Conf. Stockholm 28
52. Monin, A. S., J. Geophys. Res., 67, No. June - 11 July 1 967, 3, 1-16 (Stock­
8, 3103-09 ( 1962) holm, 1967)
53. Monin, A. 5., Tr. Inst. A tmospheric 70. Blackadar, A. K., Global A tmospheric
Physics, A kad. Nauk SSSR, No. 4, Research Programme, Rept. Study
5-20 (1962) Conf. Stockholm 28 June-ll July
54. Monin, A. S., IZ1I. A kad. Nauk SSSR, 1967. 4, 1-1 1 (Stockholm, 1967)
Mechanics of Fluid and Gas, No. I , 71. Lettau , H. H., A dvan. Geophys., 6, 241-
3 7-43 (1966) 57 (1959)
55. Zubkovsky, S. L., Zvang, L. R., Izv. 72. Lettau, H. H., Beilr. Phys. A tmosph. ,
A kad. Nauk SSSR, A tmospheric and
35, No. 3-4, 1 95-2 12 (1962)
Oceanic Physics, 2, No. 12, 1307-10
73. Blackadar, A. K., J. Geophys. Res., 67,
(1965)
56. Kazansky, 5., Izv.
No. 8, 3095-102 (1962)
74. Byzova, N. L., Mashkova, G. B.,
A. B., Monin, A.
A kad. Nauk SSSR, Ser. Geoftz. ,
No. I, 79-86 (1956)
Izv. A kad. Nauk SSSR, A tmos­
pheric and Oceanic Physics, I , No.
5 7. Kazansky, A. B., Monin, A. S.,
1 1 , 1209- 1 1 ( 1964) ; 2, No. 7, 681-87
Meteorol. Hydrol, No. 12, 3-8
(1966)
(1962)
58. Zilitinkevich, S. 5., Chalikov, D. V., 75. KurDakova, T. A., Orlenko, L. R.,
Izv. A kad. Nauk SSSR, A tmo­ Tr. GI. Geoftz. Observ., No. 205,
spheric and Oceanic Physics, 4, No. 1 3-24 (1967)
9, 9 1 5-29 (1968) 76. Zilitinkevich, S. 5., Chalikov, D. V.,
59. Zubkovsky, S. L., Koprov, B. M., Izv. A kad. Nauk SSSR, A tmo­
IZTJ. A k ad. Nauk SSSR, A tmo­ spheric and Oceanic Physics, 4, No.
spheric and Oceanic Physics, 5, No. 7, 765-72 (1968)
4, 323-31 (1 969) 7 7. Chalikov, D. V., Meleorol. Hydrol. , No.
60. Monin, A. S. , IZ1/. A kad. Nauk SSSR, 8, 10-19 (1968)
A tmospheric and Oceanic Physics, 78. Mashkova, G. B., Tr. Inst. Appl. Geo­
No. 3, 397-407 (t 962) phys., No. 2, 44-56 (1955)
250 MONIN
79. Zilitinkevich, S. S., Laikhtman, D. L., Meleorol. Hydrol., No. 2, 1 1-26
Monin, A. S . , Iz. A kad. Nauk (1968)
SSSR, A tmospheric and Oceanic 84. Monin, A. S., Izv. A kad. Nauk SSSR,
Physics, 3, No. 3, 297-333 (1967) A tmospheric and Oceanic Physics,
80. Appleby, J. F., Ohmsted, W. D., Proc. I, No, 3, 258-65 (1965)
Army Sci. Con!. (1964) , Washing­ 85. Monin, A. S., Iz'IJ. A kad. Nauk SSSR,
ton, 1 , 85-99 (1 965) A tmospheric and Oceanic Physics,
8 1 . Zilitinkevich, S. S., Laikhtman, D. L., 1, No. 5, 490-500 (1965)
Tseitin, G. H., A ir-Ocean Interac­ 86. Estoque, M. A., J. Geophys. Res., 868,
tion, 154-63 (Naukova Dumka, 1 1 03-14, No. 4, ( 1963)
Kiev, 1966) 87. Sharon, S. W., J. Geophys. Res., 70, No.
82. Bobyieva, 1. M., Zilitinkevich, S. S . , 8, 1801-7 (1 965)
Laikhtman, D. L., A tmospheric 88. Krishna, K, Monthly Weather Rev., 96,
Turbulence and Radiowave Propaga­ No. 5, 269-76 (1968)
tion, 1 79-90 (N auka Publishers, 89. Yager, B. G., Zilitinkevich, S. S.
Annu. Rev. Fluid Mech. 1970.2:225-250. Downloaded from www.annualreviews.org

Moscow, 1967) Meteorol. Hydrol., No. 7, 3-111


83. Zilitinkevich, S. S., Chalikov, D . V., (1968)
by Ryerson University on 02/25/13. For personal use only.

You might also like