ArticlePDF Available

Surface Heterogeneity and Vertical Structure of the Boundary Layer

Authors:
  • NorthWest Research Associates

Abstract and Figures

The concepts of vertical structure of the boundary layer over homogeneous surfaces are discussed, including the roughness sublayer, surface layer (inertial layer) and outer layer. As an interpretive literature survey, the vertical depth of the influence of surface heterogeneity, relative to such vertical structure, is examined in terms of blending height theory, convective boundary-layer scaling and internal boundary-layer theory. These concepts are examined with data over different surface types. The rich variety of types of surface heterogeneity and background flow preclude description of their influence on the boundary layer by one single approach. Nonetheless, new scaling arguments offer promise for convective conditions. Internal boundary layers are found to form only in certain situationsand even then may exhibit more diffuse vertical structure comparedto the textbook internal boundary layer. New scaling arguments aredeveloped to describe the potential for internal boundary-layerdevelopment in flow of warm air over a cooler surface. These argumentsexplain some aspects of the observations but remain incomplete.
Content may be subject to copyright.
SURFACE HETEROGENEITY AND VERTICAL STRUCTURE OF THE
BOUNDARY LAYER
L. MAHRT?
College of Oceanic and Atmospheric Sciences, Oregon State University, Corvallis, OR 97331,
U.S.A.
(Received in final form 29 October 1999)
Abstract. The concepts of vertical structure of the boundary layer over homogeneous surfaces are
discussed, including the roughness sublayer, surface layer (inertial layer) and outer layer. As an
interpretive literature survey, the vertical depth of the influence of surface heterogeneity, relative to
such vertical structure, is examined in terms of blending height theory, convective boundary-layer
scaling and internal boundary-layer theory.
These concepts are examined with data over different surface types. The rich variety of types of
surface heterogeneity and background flow preclude description of their influence on the boundary
layer by one single approach. Nonetheless, new scaling arguments offer promise for convective
conditions.
Internal boundary layers are found to form only in certain situations and even then may exhibit
more diffuse vertical structure compared to the textbook internal boundary layer. New scaling ar-
guments are developed to describe the potential for internal boundary-layer development in flow of
warm air over a cooler surface. These arguments explain some aspects of the observations but remain
incomplete.
Keywords: Heterogeneity, Surface fluxes, Convective boundary layer, Blending height, Internal
boundary layer.
1. Introduction
The atmospheric boundary layer is traditionally partitioned into different idealized
layers including the canopy layer, roughness sublayer, surface layer, outer layer
and the entrainment zone (Figure 1). This division has served the community well
and provides a framework for organizing our investigation of the boundary layer.
However, mixing in the convective boundary layer is generated by eddies on both
small and large scales and large eddy motions cross the boundaries of these layers
without any special recognition of a sharp change in physics. Although existing
concepts of the vertical structure of the boundary layer are even less applicable
over heterogeneous surfaces, such concepts provide a useful andnecessary starting
point.
To some extent, the earth’s surface is always heterogeneous. When is surface
heterogeneity important and when must the above concepts of vertical structure
?E-mail: mahrt@oce.orst.edu
Boundary-Layer Meteorology 96: 33–62, 2000.
© 2000 Kluwer Academic Publishers. Printed in the Netherlands.
34 L. MAHRT
Figure 1. Idealized layering of the boundary layer.
be modified? The answer is partly a matter of both horizontal scale and amplitude
of the surface heterogeneity. To address this query, the vertical structure of the
boundary layer over homogeneous surfaces is reviewed in Section 2. Section 3
reviews blending height theory and convective boundary-layer scaling in terms of
the impact of surface heterogeneity. Section 4 studies the influence of the scale
of the surface heterogeneity on the vertical structure of the boundary layer, and
Section 5 describes four data sets that are analyzed in Sections 6–7. Section 6
assesses the applicability of the blending height and convective boundary-layer
SURFACE HETEROGENEITY AND VERTICAL STRUCTURE OF THE BOUNDARY LAYER 35
scaling. Section 7 evaluates existing concepts for the internal boundary layer over
heated and cooled surfaces.
2. Vertical Structure over Homogeneous Surfaces
2.1. ROUGHNESS SUBLAYER
Two definitions of the roughness sublayer are used in the literature. In the first
definition, the roughness sublayer is the layer immediately above the surface where
the individual surface roughness elements induce horizontal variability of the time-
averaged flow. This horizontal variation is due to spatial preference of wakes and
thermals semi-attached to vegetation elements (Figure 1). In this respect, the top
of the roughness sublayer (z) is the ‘blending height’ for the surface roughness
elements. However, in this study, the term blending height will be reserved for de-
scribing the height-dependence of the influence of surface heterogeneity on a scale
that is large compared to the scale of individual roughness elements. Horizontally-
averaged fluxes are required in the roughness sublayer since observations for a
fixed point, such as on a tower, would not be representative and would be altered if
the sensor is displaced horizontally even a few metres.
The roughness sublayer is sometimes also defined as the region below the ‘sur-
face layer’, where Monin–Obukhov (MO) similarity theory (ST) no longer applies.
This definition emphasizes that the flux-gradient relationship in the roughness sub-
layer is different from that predicted by MO similarity theory. For example, the
eddy diffusivity for momentum may be enhanced beyond that predicted by MO
similarity due to mixing associated with wakes behind the roughness elements. As
a result, a modified similarity theory is formulated for the roughness sublayer (Gar-
ratt, 1980; Cellier and Brunet 1992; Raupach, 1992; Physick and Garratt, 1995). In
principle, similarity theory in the roughness sublayer should be constructed from,
and applied to, horizontally-averaged flow.
The above two definitions are not necessarily equivalent. For example, Wieringa
(1993) defines an intermediate layer that is above the influence of individual rough-
ness elements on the time-averaged flow yet MO similarity does not apply. This
speculation is referred to as a horizontally homogeneous part of the roughness sub-
layer in Figure 1. The existence of the homogeneous part of the roughness sublayer
has not been clearly shown from observations over land. However, in the analogous
wave boundary layer over lakes and oceans, the time-averaged flow is horizontally
homogeneous close to the surface yet MO similarity does not apply because the
flux-gradient relationship depends on the surface wavelength. MO similarity theory
does apply in the overlying surface layer, located above this wave boundary layer.
Generally, the depth of the roughness sublayer is considered to be proportional
to the mechanical properties of the surface such as the momentum roughness length
(Wieringa, 1993; Parlange and Brutsaert, 1993), spacing between the roughness
36 L. MAHRT
elements (Raupach et al., 1980), height of the roughness elements (Garratt, 1980;
Wieringa, 1993; Raupach, 1994), and silhouette area of the roughness elements
(Lettau, 1969). In Raupach (1994), the depth of the roughness sublayer extends
a vertical distance of 2(hcd) above the displacement height, d,wherehcis
the height of the roughness elements or canopy. Some of these formulations are
surveyed in Hess (1994) while a detailed treatment is provided by Raupach et al.
(1991). However, with heated weak wind situations, thermals may occur at pre-
ferred locations with respect to the individual roughness elements, as speculated in
Figure 1. Preferred locations for thermals emanating from the top of an individual
roughness element may extend the depth of the roughness sublayer beyond the
above predictions, which is consistent with the findings of Garratt (1978). Garratt
(1978) points out the potentially important role of thedifferent vertical distributions
of heat and momentum sources and sinks, which in turn vary between different
types of canopies.
2.2. SURFACE LAYER AND OUTER LAYER
MO similarity theory is applied in the surface layer (Figure 1) and assumes that the
underlying roughness sublayer is sufficiently thin, that the flux in the surface layer
is close to the spatially-averaged surface flux. In the surface layer, MOST is used
to predict the flux-gradient relationship in terms of the height above ground (or
displacement height) and the Obukhov length. The bulk formula based on vertically
integrated similarity theory is the primary method for predicting surface fluxes in
models. Sun et al. (1999) emphasize that the roughness length and aerodynamic
surface variables represent the influence of the surface characteristics on the flux-
gradient relationship only in the surface layer.
Above the surface layer, in the outer layer, the flux-gradient relationship is influ-
enced by the boundary-layer depth and additional scaling arguments are required
(Sorbjan, 1989). With weak winds and strong surface heating, the surface layer
may be separated from the outer mixed layer by a free convection regime (e.g.,
Holtslag and Nieuwstadt, 1986). If the boundary layer is shallow and/or the rough-
ness sublayer is deep, the surface layer may be squeezed from existence (Mahrt et
al., 1998). Then there is no layer where MOST applies since the influence of the
boundary-layer depth essentially extends downward to the roughness sublayer.
These concepts are based on homogeneous surfaces. Over heterogeneous sur-
faces, the surface layer may be eliminated, depending on the horizontal scale of
the surface heterogeneity. Then MOST is expected to break down. The vertical
extent of the influence of surface heterogeneity and the breakdown of surface layer
similarity theory are considered in the next two sections.
SURFACE HETEROGENEITY AND VERTICAL STRUCTURE OF THE BOUNDARY LAYER 37
3. Scaling Approaches
If the surface heterogeneity is due only to the roughness elements, and therefore
influences only the roughness sublayer, the flow is considered to be homogeneous.
Studies of surface heterogeneity have implicitly assumed a distinct separation of
scales between the surface heterogeneity and the roughness elements, as will be
assumed here. This condition is often not met. For example, some forests are char-
acterized by clumping of trees. Is a clump of trees a surface roughness element or
should it be considered as very small-scale heterogeneity?
Surface spatial variations often occur simultaneously on a continuum of ho-
rizontal scales or consist of gradual changes without a well-defined scale, as
sometimes occur with natural changes of vegetation or soil moisture gradients
associated with recent precipitation patterns. In the next two sections, we assume
that surface heterogeneity occurs on a single well-defined scale. In Section 6, we
consider observed atmospheric flows where the surface heterogeneity may occur
simultaneously on a variety of horizontal scales.
3.1. BLENDING HEIGHT LOCAL DIFFUSION
The blending height is viewed here as a scaling depth that describes the decrease
of the influence of surface heterogeneity with height. The blending height is not a
level where the influence of surface heterogeneity suddenly and completely van-
ishes but instead is the level where the influence of surface heterogeneity gradually
decreases below some threshold value. The blending height can be re-formulated in
terms of internal boundary layers and footprint theory (Mahrt, 1996). The blending
height has been studied mainly from numerical models, although the existence of
the blending height has been inferred from tethered balloon measurements (Grant,
1991). The various blending height derivations were intended more as qualitative
scaling arguments based in part on linear theory. Therefore, quantitative compar-
isons of the blending height formulations with more complex atmospheric flows
in this study may exceed the intended use by the original authors. One role of the
nonlinearity is evident in the response of the atmosphere to spatial variations of
the surface roughness. Schmid and Bünzli (1995) show that the effective rough-
ness length applied to the flow above the blending height is augmented due to the
fact that the effects of the rough-to-smooth and smooth-to-rough transitions on the
area-averaged momentum flux do not cancel, leading to significant ‘second order
roughness’.
Below the blending height, the turbulence may not be in equilibrium with the
local vertical gradient, which excludes application of MO similarity. MO similar-
ity can be applied above the blending height only if the blending height is low
compared to the boundary-layer depth. This condition is required in order that
the flux immediately above the blending height is close to the spatially-averaged
38 L. MAHRT
Figure 2. Scaling regimes based on blending height concepts (lower main part) and convective
boundary-layer scaling (upper part). The diagram is not to scale and the boundaries between different
regimes depend on stability and other boundary-layer variables. Here the surface layer is chosen
arbitrarily to be equal to 0.05 h so that the flux in the surface layer would be within 5% of the surface
value for a linear decrease of the flux with height to near zero at the boundary-layer top. The hatched
area denotes the scale regime where MO similarity is not valid. The vertical dashed line is defined
by Equation (11).
surface value and that the boundary-layer depth does not influence the flux-gradient
relationship.
Different formulations of the blending-height approach are discussed in Mason
(1988), Wood and Mason (1991), Claussen (1990), and Schmid and Bünzli(1995).
The general form of the blending height, zblend, can be expressed as
zblend =Cu
Up
Lhetero,(1)
where uis the friction velocity based on the horizontally-averaged momentum
flux, Uis the speed of the spatially averaged wind vector based on the flow at
the blending height, Lhetero is the horizontal scale of the surface heterogeneity, pis
usually 2 and Cis a nondimensional coefficient usually taken as unity. With p=2,
zblend becomes the drag coefficient times the scale of surface heterogeneity. The
diffusion height corresponds to p=1 (Wood and Mason, 1991; Claussen, 1990).
Observations do not normally provide winds at the time-dependent blending height
and winds from a fixed observational level(s) must be substituted. The influence of
SURFACE HETEROGENEITY AND VERTICAL STRUCTURE OF THE BOUNDARY LAYER 39
surface heterogeneity is small above the blending height (Equation (1), region B in
Figure 2) and theoretically vanishes at the diffusion height.
The blending height increases linearly with horizontal scale of the surface het-
erogeneity so that the influence of larger scale surface features extends upwards
farther into the boundary layer. Consider a reference level, z, that might be the
first model level or the observational level. Equation (1) can be used to predict the
minimum horizontal scale of the surface feature, Lblend, required to significantly
influence the flow at level zby equating zwith zblend. Then, the minimum horizontal
scale for important influence of the surface heterogeneity becomes
Lblend =CblendzU
u2
.(2)
If the horizontal length scale of the surface feature, Lhetero, is less than Lblend,then
the surface feature does not induce significant horizontal heterogeneity at level z;
that is, level zis above the blending height (region B in Figure 2). From another
point-of-view, the impact of surface heterogeneity at height zis an increasing func-
tion of Lhetero/Lblend. This dependence will be incorporated in Section 3.5. Based
on observations in Section 6.1, Cblend is chosen to be roughly 0.6.
3.2. THERMAL BLENDING HEIGHT
The horizontal length scale, Lblend, decreases slowly with increasing instability
since u/U is predicted to increase slowly with increasing instability. However,
when surface heating is important, this weak dependence on instability under-
estimates the importance of instability (Section 6) and a more explicit stability-
dependence is required. Wood and Mason (1991) derive a thermal version of the
blending height based on a linearized thermodynamic budget where the perturb-
ation flow is induced by the surface heterogeneity. The thermodynamic budget
reduces to an approximate balance between the temperature advection and ver-
tical divergence of the perturbation heat flux where the perturbation heat flux is
due to surface heterogeneity. The resulting thermal blending height for unstable
conditions is
zwm
Lheterow0θ0
sf c
U2
o
,(3)
where w0θ0
sf c is the absolute value of the spatially-averaged surface heat flux and
2ois the averaged potential temperature. In the derivation of this relationship, the
influence of the spatial variation of the heat flux cancels and the resulting blending
height for heat depends only on the background surface heat flux. This cancellation
occurs due to linearization; presumably the spatial variation of the heat flux would
be introduced in higher order analysis.
40 L. MAHRT
Equating the reference level zwith zwm, the minimum horizontal scale of the
surface heterogeneity that significantly influences the flow at level zgives
Lwm =CwmzU2
o
w0θ0
sf c
.(4)
Scales of surface heterogeneity larger than Lwm are predicted to significantly influ-
ence the flux at level z.With decreasing wind speed and increasing surface heating,
Lwm decreases and small scale surface features exert a stronger influence on the
flow at level z.ThevalueofCwm is estimated from observations in Section 6.1 to
be roughly 3.1 ×103.
3.3. BOUNDARY-LAYER CONVECTIVE SCALING
The blending height concept (Section 3.1) is based on shear-driven mixing where
the transport is more diffusive and local. In contrast, in the convective boundary
layer, large boundary-layer eddies can transfer information from the surface to the
top of the deep boundary layer on the time scale of a few tens of minutes and the
concept of a blending height is less applicable. For this case, Raupach and Finnigan
(1995) pose the impact of surface heterogeneity in terms of the basic time scales
of the convective boundary layer. The mixing time scale is defined as zi/w,where
ziis the depth of the convective boundary layer and wis the Deardorff convective
velocity scale. During one mixing time scale, the flow travels
LRau CRau Uzi
w
,(5)
where CRau is a nondimensional coefficient estimated from observations in Section
6.1 to be roughly 0.8. For horizontal scales of surface heterogeneity smaller than
LRau, the influence of surface heterogeneity is confined to depths which are small
compared to the boundary-layer depth. The large convective boundary-layer eddies
horizontally mix the influence of the surface heterogeneity to the extent that it does
not significantly influence the bulk of the boundary layer.
This convective scaling does not address the height-dependence of the influence
of surface heterogeneity in that it predicts the influence of the surface heterogen-
eity on the boundary layer as a whole. Since the eddy length scale increases with
height in the lower part of the boundary layer, even in convective conditions, one
might expect that the influence of smaller scales of heterogeneity decreases with
height. Then, near the top of the boundary layer, only the largest scales of surface
heterogeneity are important, as with blending height arguments. We return to this
problem in Section 3.5.
An additional larger time scale is introduced into the convective scaling ap-
proach of Raupach and Finnigan (1995). The entrainment time scale,Te, describes
the period of growth of the convective boundary layer since its initial development
SURFACE HETEROGENEITY AND VERTICAL STRUCTURE OF THE BOUNDARY LAYER 41
earlier in the morning. The corresponding length scale, TeU, defines the largest
horizontal scale where the surface heterogeneity influences the boundary layer. For
surface heterogeneity on scales significantly larger than TeU, the entire boundary
layer reaches approximate equilibrium with the local surface conditions.
3.4. THE THREE HORIZONTAL LENGTH SCALES
We can summarize the three horizontal length scales as follows:
1. Lblend is based on local diffusive mixing by shear-generated turbulence and
seems most applicable to near-neutral or stable conditions.
2. Lwm allows buoyancy-modification of the turbulence.
3. LRau represents the case of nonlocal bulk mixing due to buoyancy-driven
boundary-layer eddies on the scale of the boundary layer and seems most
applicable to the convective boundary layer.
All three length scales estimate the minimum horizontal scale of surface hetero-
geneity that significantly influences the flow at level z, or in the case of LRau,the
entire boundary layer. All three length scales are proportional to the wind speed
and inversely proportional to some measure of the strength of the turbulence.
3.5. OVERALL MODEL FORMAT
Toformulate a measure of the spatial variability of the flux, we define the standard
deviation of the flux, σF, divided by the spatially-averaged flux, F, both evaluated
at level z. Based on the above arguments, one can postulate
σF
F=fLhetero
LXGσTsf c
θo,(6)
where LXrefers to one of the horizontal length scales in the previous subsection.
The function, f(L
hetero/LX), is expected to increase with Lhetero/LXsince the tur-
bulence substantially reduces the influence of surface heterogeneity on horizontal
scales of LXand smaller. The function fcould consist of multiple terms where
LXdepends on both Lblend and LRau, in order to include the limiting cases of
near-neutral flows and the convective boundary layer. The dependence of σF/F
on the amplitude of surface heterogeneity, G, will be posed later in terms of the
surface radiation temperature. Over surfaces with complex vegetation, the surface
radiation temperature needs to be replaced with a more reliable indicator of surface
heterogeneity. Simplification of Equation (6) is pursued in Section 6 by analyzing
aircraft data. Examination of Equation (6) with data allows consideration of only
horizontal scales of heterogeneity that are smaller than the observational domain.
42 L. MAHRT
4. Scale Regimes
Based on the horizontal length scales derived in the previous section, we now for-
mulate different scale regimes for the surface heterogeneity based on the vertical
extent of their influence. The various scaling regimes are schematically illustrated
in Figure 2. The terminology at the bottom of the figure is modified from the survey
of Mahrt (1996) while the notation at the top follows Raupach and Finnigan (1995).
The influence of heterogeneity on the scale of the individual roughness elements is
confined to the roughness sublayer (Section 2), noted by “E” in Figure 2. We now
discuss the vertical depth of the influence of surface heterogeneity on the micro-,
meso- and macro-scales.
4.1. MICROSCALE HETEROGENEITY
If the blending height is low compared to the depth of the boundary layer,
zblend << h, (7)
then the flux immediately above the blending height is close to the spatially-
averaged surface value and Monin–Obukhov similarity theory can be used to
predict the surface flux. This case is referred to as microscale surface heterogeneity
(left hatched region in Figure 2) and from Equations (1) and (7) corresponds to
Lhetero << Cblend(U/u)2h. (8)
For the thermal blending height (Equation (4)), the corresponding condition is
Lhetero << Cwm
U2
o
w0θ0
sf c
h. (9)
Convective scaling does not explicitly address the influence of surface heterogen-
eity in the surface layer, although one would anticipate that Lhetero <L
Rau is a
necessary condition in order that MO similarity theory is applicable.
The above inequalities can be used to assess the suitability of the tile approach
used to partition subgrid surfaces in numerical models, also known as the mosaic or
flux aggregation approach (see Avissar and Pielke, 1989, or additional references
in Mahrt, 1996). Such an approach was formulated for momentum transfer over
heterogeneous surfaces in terms of an effective roughness length (Taylor, 1987).
Each grid surface area is divided into smaller subgrid areas corresponding to dif-
ferent surface types while the variables at the first model level, presumably within
the surface layer, are assumed to equal their grid-averaged values, unaffected by
subgrid surface heterogeneity. Such an assumption can be justified as a rough ap-
proximation only if the blending height is below the first model level, which in turn
must be below the top of the surface layer. One could derive restrictions on the scale
SURFACE HETEROGENEITY AND VERTICAL STRUCTURE OF THE BOUNDARY LAYER 43
of the surface heterogeneity for the tile approach by replacing the boundary-layer
depth in Equations (7)–(9) with the height of the first model level. For larger scale
surface heterogeneity (higher blending height), the tile approach formally breaks
down. Then one must compare the errors resulting from this breakdown against
errors resulting from complete neglect of subgrid surface heterogeneity.
4.2. MESOSCALE OR INTERMEDIATE-SCALE HETEROGENEITY
If the equalities in Section 4.1 are not met, then the influence of the surface het-
erogeneity extends upward to a significant fraction of the boundary-layer depth
(region C in Figure 2). Then the flow immediately above the blending height is too
high to estimate the surface flux (region B, Figure 2). For convective scaling, the
mesoscale regime is defined as LRau <L
hetero <T
eU. In terms of blending height,
the condition is h < Lblend <hwhere may be chosen as 0.05, for example, to
approximate the depth of the surface layer.
In this scaling regime, MO similarity theory cannot be used to predict the
spatially-averaged surface fluxes. The blending height is sufficiently high that there
is no longer a level above the blending height where the flux-gradient relationship
depends on z/L alone and where the spatially-averaged flux is close to the surface
value. This failure of similarity theory is probably responsible for the failure of the
emergence of a unifying approach for predicting surface fluxesovermesoscale het-
erogeneity in terms of effective roughness lengths or transfer coefficients. Different
studies yield different predictions, depending on the details of the surface condi-
tions and assumptions invoked. For mesoscale heterogeneity, no rigorous approach
is available.
4.3. MACROSCALE OR LARGE-SCALE HETEROGENEITY
If the predicted blending height is greater than the boundary-layer top,
zblend >h, (10)
then the local surface feature controls the entire boundary layer. This condition
occurs when the horizontal scale of surface heterogeneity exceeds
Lhetero >C
blendhU
u2
,(11)
corresponding to region D in Figure 2. In this region, the boundary layer establishes
equilibrium with the local surface type and similarity theory can be applied without
concern for the change of surface conditions far upstream.
This boundary-layer equilibrium regime occurs with convective boundary-layer
scaling if
Lhetero TeU. (12)
44 L. MAHRT
This regime is referred to as macroscale heterogeneity. For small wind speed in
Equation (12) or small boundary-layer depth in Equation (11), macroscale hetero-
geneity may occur on scales that are normally referred to as mesoscale. Note that
the above discussions refer to mesoscale and macroscale organization of the tur-
bulent fluxes and do not address the direct vertical transport by mesoscale vertical
motions.
5. Data Sets
To assess the impact of surface heterogeneity on fluxes in the atmospheric bound-
ary layer, we will analyze data collected by the Twin Otter of the Canadian National
Research Council in three different field programs. The measurements were made
at approximately 35 m above the surface, which roughly corresponds to the top of
the surface layer.
5.1. SGP
The primary data set for this study is repeated aircraft passes at approximately
35 m over the El Reno track during 10 flight days of the Southern Great Plains
Experiment (SGP). The flight on 2 July 1997 is selected as a case study since it
contains the most complete data set for this field program. On this day, the winds
were from the southwest andcontained asignificant wind component parallel to the
east-west track, allowing examination of the impact of surface heterogeneity and
advection. On most of the days, the wind was mainly perpendicular to the track.
The 13-km El Reno track consists of mixed farming in the west, senescent wheat
and bare fields in the central part where evapotranspiration was smaller and green
grasslands in the eastern part where the evapotranspiration was larger. The time
series from repeated aircraft passes have been stretched to account for variable air-
craft ground speed and have been aligned using ground markers based on remotely
sensed NDVI (Normalized Difference of Vegetation Index) and surface radiation
temperature. Fluxes, variances and means are computed from aircraft data for
windows of 1-km width using unweighted means. The 1-km window generally cap-
tured almost all of the turbulent flux. The window is sequentially marched through
the record by increments of 250 m. For some calculations, variables are first aver-
aged over all of the aircraft passes for each window position. The number of passes
within each flight ranged from 3 to 12. Boundary-layer depth was estimated from
aircraft soundings that were executed at variable intervals during the flight periods.
For two of the flights, the boundary-layer depth was not well-defined, leaving eight
flights for the analysis below.
SURFACE HETEROGENEITY AND VERTICAL STRUCTURE OF THE BOUNDARY LAYER 45
5.2. CODE, BOREAS AND RASEX
We will also analyze data from two flights in the California Ozone Deposition
Experiment (CODE), each consisting of eight repeated passes at 35 m by the NRC
TwinOtter over irrigated and nonirrigated surfaces (Pedersen et al., 1995; Mahrt et
al., 1994). Land-lake contrasts will be examined with NRC Twin Otter data taken
in the Boreal Ecological and Atmospheric Study (BOREAS) described in Sun et
al. (1997). We choose the case where the wind was most aligned with the aircraft
track.
Internal boundary layers are also studied using data collected during the Risø
Air Sea Experiment (RASEX) with an instrumented mast, 2 km off the Danish
coast (Mahrt et al., 1998) and an instrumented tower at the shore. As a case study,
we have chosen 24 October 1994 where a stationary stable boundary layer formed
for a 5-hour period of offshore flow of warm air over cooler water. Fluxes, variances
and means are computed from the tower data in terms of deviations from a 10-
minute mean. The fluxes and variances are then averaged over a one-hour period.
6. Evaluation of Length Scales
Determination of the blending height from data is problematic, partly because
the influence of surface features gradually fades with height and other mesoscale
features gradually become dominant. For example, in the uppermost part of the
boundary layer in the Southern Great Plains Experiment, transient spatial variations
of the boundary-layer depth and entrainment led to substantial spatial variation of
measured fluxes along the aircraft track in the upper part of the boundary layer.
These variations partially mask any remaining influence of surface heterogeneity
on the upper part of the boundary layer. Therefore this study concentrates on the
low level passes.
The influence of transient motions and large random flux sampling errors can
be reduced by first averaging over all of the passes for a given flight for each 1-km
segment (Section 5). Unfortunately, this averaging incorporates different boundary-
layer conditions within the averaging period associated with diurnal variation
during the flight period of several hours. Nonetheless, the relationship between
the atmospheric response σF/F and the horizontal length scales (Section 4) is
improved by first averaging variables over the passes. These relationships probably
would improve further had more repeated passes been available. The following
reports results only for flight-averaged values.
6.1. FLUX CORRELATION WITH SURFACE TEMPERATURE
The influence of surface heterogeneity on the boundary layer can be assessed in
terms of the correlation between the heat flux at level zand the surface radiation
temperature. Here we have computed the correlation between the spatial variation
46 L. MAHRT
of the flight-averaged heat flux and the flight-averaged surface radiation temper-
ature for the 1-km overlapping windows. Considering all of the measurement
difficulties and complexity of real air-surface interactions, the relationship between
the heat flux-surface temperature correlation and the various length scales (Lblend,
Lwm and LRau) is promising (Figure 3), even though the scatter is large. When
these length scales are large, the impact of smaller scale surface heterogeneity on
the heat flux at level zis reduced and the correlation between the heat flux and the
surface radiation temperature decreases. Within the accuracy of the data, a clear
distinction between the predictive value of the length scales cannot be determined.
To estimate the nondimensional coefficients for the horizontal length scales
(LX=Lblend,L
wm or LRau) from the observations, we assume that the surface
heterogeneity exerts a significant influence on the fluxes at the aircraft level (near
the top of the surface layer) when the correlation between the flux and the sur-
face radiation temperature exceed 0.7 (50% or more variance-explained). This
threshold value was estimated from Figure 3 by subjectively drawing a curved line
through the data points. The value of the horizontal length scale corresponding to
this threshold, L
X, is then equated to the estimated scale of the dominant surface
heterogeneity, so that
L
X=Lhetero.(13)
The length scale for the surface heterogeneity, Lhetero, is not easily determined since
surface heterogeneity occurs simultaneously on a variety of scales. Based on sur-
face radiation temperature and NDVI, the principal scale of surface heterogeneity
appears to be about 4 km although significant variations occur on smaller scales.
Here, we chose 3 km for the horizontal scale of the surface heterogeneity and then
estimate the nondimensional coefficients from Equation (13) and the definitions of
LX. The numerical values of the coefficients were reported in Section 3 after the
defining relationships for LX. Because of the various uncertainties, the values of
the nondimensional coefficients are only order-of-magnitude estimates. Note that
the calibrated values of LXestimate the minimum horizontal scale of the surface
heterogeneity required for the most important influences on the fluxes at level
z.Thatis,LXis a scaling variable, not a cutoff value, and smaller scale surface
heterogeneity will still exert some influence on the flow at level z. To include all
of the significant influences on the fluxes at level z, one should decrease the values
of the nondimensional coefficients for the length scales by an order of magnitude.
The above estimates are prejudiced against smaller scales since the fluxes had to
be averaged over a one-kilometre window. Averaging over significantly smaller
windows would be vulnerable to random flux errors.
6.2. RELATIVE FLUX VARIATION
The prediction of the magnitude of the spatial variation of the flux at a given level,
z, is of considerable importance for a number of practical applications, including
SURFACE HETEROGENEITY AND VERTICAL STRUCTURE OF THE BOUNDARY LAYER 47
Figure 3. The correlation between the heat flux and surface radiation temperature as a function of the
horizontal length scale: (a) Lblend,(b)Lwm,and(c)LRau.
48 L. MAHRT
scaling up of area-averaged fluxes. The spatial variation of the heat flux will be
represented by the spatial standard deviation of the flux computed from the over-
lapping 1-km windows that are averaged over all of the passes in the flight (σF).
The relative standard deviation is computed by dividing this standard deviation by
the track-averaged flux for the flight (F).
The relative standard deviation of the surface heat flux (σH/H ) shows no ob-
vious relationship with the blending height predictions, Lblend (Equation (2)) and
Lwm (Equation (3)). The failure of these horizontal length scales for the present
data set is due to: (a) the importance of the depth of the convective boundary layer,
(b) lack of information on the amplitude of the surface heterogeneity, and (c) a
number of additional complications noted below. The influence of the value of the
boundary-layer depth on σH/H is strong for boundary-layer depths greater than
about 400 m. Smaller boundary-layer depths correspond to transient cases earlier
in the morning when the convective eddies are not well-developed.
The relative standard deviation of theheat flux tends to decrease with increasing
LRau although the scatter is quite large (Figure 4a). With large values of LRau,the
influence of small-scale surface heterogeneity is more effectively eliminated by
large boundary-layer eddies. However, the large scatter in Figure 4a suggests that
other factors are important. To include a measure of the amplitude of the surface
heterogeneity, we pursue Equation (6) using LRau as the principal horizontal length
scale. Assuming linear dependencies between the relative standard deviation of the
heat flux and the scaling variables, Equation (6) is written as
σH
HLhetero
LRau
σTsf c
θo
.(14)
For bookkeeping convenience, we define
LRT CRT
LRauθo
σTsf c
,(15)
in which case
σH
HLhetero
LRT
.(16)
The value of the nondimensional coefficient, CRT , is estimated from the method
in Section 6.1 to be 4.3 ×103. For the present data, the spatial variation of the
heat flux seems to be better related to LRT (Figure 4b) than LRau (Figure 4a) al-
though large scatter remains. The relative spatial variation of the heat flux depends
significantly on the amplitude of the surface heterogeneity. The amplitude of the
surface heterogeneity varies from day to day partly due to the precipitation history
along the flight track. After recent rains, the surface heterogeneity is less since the
moisture flux over the bare soil/senescent wheat region is almost as large as that
over the grassland.
SURFACE HETEROGENEITY AND VERTICAL STRUCTURE OF THE BOUNDARY LAYER 49
Figure 4. Relative standard deviation of the spatial variation of the heat flux for the eight flights in
SGP as a function of: (a) LRau and (b) LRT and (c) the relative variation of the momentum flux as a
function of LRT .
50 L. MAHRT
The remaining scatter in Figure 4b could be due partly to the influence of non-
stationarity, errors in estimation of the surface fluxes and boundary-layer depth,
and the influence of transient mesoscale disturbances not removed by compositing
over the passes. The dependence on wind direction with respect to the surface
heterogeneity was found to be small. The influence of partial cloud cover was small
for flight-averaged values.
Consider the practical problem of estimating the required horizontal resolution
for remotely sensed surface variables used to infer spatial variation of surface
fluxes. Such inferences can be based on empirical models which are calibrated
in terms of aircraft flux data. The value of LRT determines the smallest scale of
surface heterogeneity, which significantly influences the flow at the aircraft level
(approximate top of the surface layer). Finer spatial resolution of surface character-
istics would not be needed, especially considering the approximate nature of such
models.
6.3. MOISTURE FLUX
The surface moisture flux seems to vary in a less organized fashion compared to the
surface heat flux. The relative standard deviation of the moisture flux for the flight-
averaged flow is 0.19, as compared to 0.34 for the surface heat flux. Over the drier
regions, the moisture flux decreases less than the heat flux increases so that the sum
of H+LE is greater. Here, Hand LE are the sensible and latent heat fluxes in
Wm
2. We speculate that this difference is due to horizontal convergence of heat
and moisture below the aircraft level over warmer surfaces. However, the horizontal
convergence could not be evaluated from the data with sufficient accuracy.
The relationship of the surface moisture flux to LRau and the other length scales
is weaker than that for the surface heat flux. The heat flux may respond to the
surface heterogeneity in a more organized fashion because the temperature is more
dynamically active compared to moisture and, therefore, more strongly coupled
with updrafts and downdrafts. Furthermore, the moisture flux even near the surface
is often significantly influenced by entrainment.
6.4. MOMENTUM FLUX
The relative spatial variation of the momentum flux, σF/F ,whereFis the mag-
nitude of the stress vector, decreases rapidly with increasing values of all of the
heterogeneity length scales LX. As an example, this decrease for LRT is shown in
Figure 4c. The relative flux variation, σF/F , is much larger for momentum than
that for heat and moisture at small values of LRT . Apparently bluff roughness
elements such as occasional clumps of trees or buildings and edges of fields, exert
a strong influence on the momentum transfer even when such obstacles do not
significantly influence the heat and moisture flux.
SURFACE HETEROGENEITY AND VERTICAL STRUCTURE OF THE BOUNDARY LAYER 51
7. Internal Boundary Layers
How does the formation of internal boundary layers affect the applicability of the
blending height arguments and convective boundary-layer scaling? To address this
question, it is necessary to distinguish between mesoscale and local internal bound-
ary layers (Garratt, 1990). Mesoscale internal boundary layers are associated with
large-scale surface features and eventually engulf the entire pre-existing boundary
layer. Mesoscale convective internal boundary layers due to flow of cool air over
a warm surface are often well-defined with a capping inversion, which slopes up-
ward in the downstream direction. Flow of marine air over a heated land surface
is a classic example of the mesoscale internal boundary layer (see references in
Garratt, 1990). Flow of warm air over cooler water on the scale of 100 km can lead
to a stable internal boundary layer with well-defined vertical structure (Garratt and
Ryan, 1989).
On smaller scales, this well-defined state is not necessarily achieved, as will
be discussed below. These small-scale local internal boundary layers grow to a
maximum depth which is small compared to the upstream boundary-layer depth,
hIBL h, (17)
and then encounter a new surface type and lose surface support. Here hIBL is the
top of the internal boundary layer, assuming that it is definable. Using traditional
scaling arguments, the maximum value of the internal boundary-layer depth is
estimated in Mahrt (1996) to be
max(hIBL)=CIBL σw
ULhetero,(18)
where CIBL is thought to be O(0.1) as also found in Section 7.3, although more ob-
servations are needed. This is a scaling estimate only and does not account for the
decrease of the growth rate of the internal boundary layer in the downstream direc-
tion. Local internal boundary layers are defined to be those cases where max(hIBL)
is small compared to the boundary-layer depth, which in turn requires that the scale
of the surface heterogeneity is sufficiently small that
Lhetero U
CIBLσw
h. (19)
If σwis formulated in terms of the convective velocity scale w, then the right hand
side of Equation (19) is proportional to LRau and scales of surface heterogeneity
that are small compared to LRau correspond to local internal boundary layers.
52 L. MAHRT
7.1. COOL TO WARMER
Consider an example of flow of cool air from a partially irrigated cool region
over a hot unirrigated surface, as observed during CODE Flight 19 (Section 5).
In this example, the heat flux was near zero over the partially irrigated surface
and quite large over the dry surface, about 0.2 K m s1. Aircraft passes through
the sloping top of the internal boundary layer over the dry surface indicate that
the instantaneous top of the internal boundary layer is sometimes well-defined in
terms of a jump in potential temperature (Figure 5a). This example corresponds to
a slope of the top of the internal boundary layer of about 14%. However most of
the passes encountered a more diffuse zone of temperature change.
Since the position of the top varies in time, apparently in response to individual
thermals and fluctuations in wind speed, the time-averaged flow is characterized
by more gradual temperature changes compared to the instantaneous temperature
distribution. During periods of weak winds, the thermals over the heated surface
rise more vertically while during wind gusts the thermals are more tilted in the
downstream direction. As a result, the time-averaged horizontal structure is smooth
and no sharp changes are encountered. In this sense, the top of the internal bound-
ary layer represents the upper limit of the measurable influence of thermals during
the averaging period. For the present case, this upper limit rises steeply for the flow
composited over all of the passes (Figure 5b). We will consider this composite
to be an estimate of the time-averaged flow. Pockets of slightly warmer air are
even encountered just upstream from the surface discontinuity on three of the eight
passes (also evident in the composited temperature in Figure 5b), perhaps resulting
from eddies which temporarily reversed the horizontal flow. Ignoring this upstream
anticipation of the surface discontinuity, the heated internal boundary layer grows
with a slope of roughly 45% although uncertainties in the location of the top of
the internal boundary layer and alignment errors become significant for such steep
slopes.
We conclude that the top of the internal boundary layer for the time-averaged
flow: (1) is not associated with a sharp temperature change, (2) may rise with any
slope including nearly vertical orientation in weak-wind, heated conditions, and (3)
rises more vertically with greater averaging time which increases the probability of
capturing vertically oriented thermals.
The internal boundary layer forming in the flow of cool air from Candle Lake
over the heated land surface also initiates an internal boundary layer which rises
steeply (Figure 6). On other days when the ambient wind is weak, a divergent land
breeze leads to unstable internal boundary layers over land on both sides of the
lake, both of which rise steeply (Figure 5 in Sun et al., 1997).
Traditionally, the time-averaged internal boundary layer consists of a thin
thermal equilibrium layer at the surface (Brutsaert, 1982; Garratt, 1990; de Bruin
et al., 1991), where the potential temperature horizontally varies only slowly (Fig-
ure 5b), and a much thicker transition layer between the equilibrium layer and the
SURFACE HETEROGENEITY AND VERTICAL STRUCTURE OF THE BOUNDARY LAYER 53
Figure 5. (a) An example of an instantaneous sharp horizontal change of temperature due to a single
aircraft crossing the sloped top of the internal boundary layer during Flight 19 of CODE. (b) Averages
over the eight passes.
54 L. MAHRT
Figure 6. Instantaneous air and surface temperature and vertical velocity in a stable internal boundary
layer over Candle Lake, where the wind is easterly and along the flight track at 6 m s1. The lake is
slightly warmer at the west end, where the lake is shallow.
top of the internal boundary layer. The top of the instantaneous internal boundary
layer varies within the transition layer for the time-averaged flow. Based on po-
tential temperature for the time-averaged flow (Figure 5), the inferred depth of the
equilibrium layer for this data increases downstream with a slope of only about
2%. The concept of an equilibrium layer is not precise since the turbulence in the
equilibrium layer interacts with non-equilibrium turbulence above the equilibrium
layer.
7.2. WARM TO COOLER
Now consider flow from a warm surface to a cooler surface of limited width where a
stable internal boundary layer may or may not develop. McNaughton and Laubach
(1998) show that convective eddies advected over a cooler surface modulate the
cool stable air near the surface. Presumably, if the surface temperature change is
sufficiently small, mixing by the advected large eddies may prevent formation of
a stable internal boundary layer over the cool surface. As an instructive scaling
argument, consider cooler air which attempts to form in a layer near the surface
SURFACE HETEROGENEITY AND VERTICAL STRUCTURE OF THE BOUNDARY LAYER 55
TABLE I
Evaluation of R. Field program (described in Section 5), change in surface radiation tem-
perature, upstream air temperature – downstream surface radiation temperature 1Tsf c
(C), upstream boundary-layer depth h(m), upstream standard deviation of the vertical
velocity σw, the ratio R, and ratio of downstream to upstream σw.
Field prog. 1Tsf c (C) (C) h(m) σw(m s1)
wratio
Candle Lake 9.0 8.5 1500 0.6 0.7 0.22
CODE flt 13 28.0 2 900 0.5 3.75 0.35
RASEX Unavail. 1.5 500 0.4 2.0 1.0
SGP flt 13 9.5 0 550 0.75 Large 0.90
with a temperature deficit of ,whereis estimated as the upstream air temper-
ature minus the surface radiation temperature of the new surface. How far can the
advected large eddies with vertical velocities σwlift cool surface air upward? Here,
σwis chosen to be the standard deviation of the vertical velocity in the boundary
layer upstream. Based on solutions of the vertical equation of motion, Mahrt (1979)
found that the depth of lifting by large eddy updrafts can be estimated in terms of
the vertical scale ασ 2
w/g(1θ /2o). The coefficient αis probably large compared to
unity since the vertical motions of the most energetic updrafts and downdrafts are
significantly larger than σw.
Furthermore, the temperature deficit of air lifted from near the surface is less
than since the near-surface air will be warmer than the surface radiation tem-
perature. If the predicted depth of mixing is comparable to, or greater than, the
depth of the upstream boundary layer, then the cool air will be mixed throughout
the boundary layer and a new internal boundary layer does not form.
The ratio of the depth of upward transport of cool air to the upstream boundary-
layer depth is
Rασ2
w
hg 1θ /2 o
.(20)
If this ratio is large, then the advected large eddies from the boundary-layer up-
stream have enough kinetic energy to mix the new cool air throughout the boundary
layer. If this ratio is small, then a new cool internal boundary layer is more likely
to form. This approach ignores changes in the roughness of the surface and the
difficulty of defining the surface radiation temperature over complex vegetation. To
make Rof O(1) with typical development of internal boundary layers, αis chosen
to be 103. With this choice, how large must R(Equation (20)) become before a
stable boundary layer fails to develop? The ratio Rwas evaluated from the four
data sets described in Section 5 and then recorded in Table I.
56 L. MAHRT
As one quantitative measure of the strength of the stable internal boundary layer,
we define the ratio of the upstream standard deviation of the vertical motion to the
downstream standard deviation of vertical velocity within the modified flow. This
measure was better defined than those based on temperature. The σwratio is close
to unity for two of the three cases where Ris greater than unity. Other factors are
important as discussed below.
The Candle Lake case (Figure 6) is characterized by the smallest value of R,the
smallest value of the σwratio, and the best-defined internal boundary layer in terms
of the horizontal variation of mean variables observed by the aircraft. As warm air
from land flows over the cooler lake surface, the standard deviation of the vertical
velocity eventually decreases by almost a factor of five. Note that the change of
surface radiation temperature is not large compared to the other experiments partly
because the surface radiation temperature over the forested land is reduced by the
influence of the cool ground surface and shaded canopy elements and is therefore
not representative of the surface heating (Sun and Mahrt, 1995). However, the dif-
ference between the upstream air temperature and downstream surface temperature
is substantially larger than in the other experiments.
As anticipated from the study of McNaughton and Laubach (1998), large
boundary-layer eddies advected from the heated land surface influence the flow
at the upstream east end of the lake (Figure 6), but this influence decays rapidly in
the downstream direction. Ignoring the locally warmer air over the land near the
shore, the temperature decreases gradually in the downstream direction until about
7 km offshore when the temperature suddenly drops. This drop is probably due to
penetration of the aircraft through the slowly thickening stable internal boundary
layer. Considering that the flight level is about 30 m above the water surface, the
corresponding upward slope of the internal boundary layer averages only about
0.4%. This corresponds to CIBL =0.15 in Equation (18). The other aircraft pass
on this day shows intersection of the internal boundary layer at approximately the
same location as in Figure 6 although the horizontal temperature gradient upstream
from this intersection is greater.
Repeated flights were flown over Candle Lake on four other days when the lake
surface was considerably cooler than the air. Occasionally, the aircraft appeared to
penetrate a well-defined top of the cool stable internal boundary layer. However,
on most of the aircraft crossings, the temperature decrease was more gradual.
In CODE, the change of surface radiation temperature is much larger but the
upwind air-ground temperature difference is also large. As a result, the difference
between the upwind air temperature over the hot dry surface and downwind surface
temperature of the irrigated surface is significantly less than that for the Candle
Lake case. Nonetheless, this difference is significant and Ris relatively small.
This leads to a reasonably well-defined internal boundary layer and a substantial
decrease of σw(Table I).
The towers in RASEX capture flow of warm air from land over cooler water.
The internal boundary layer is complicated by flow acceleration from 6 m s1,at
SURFACE HETEROGENEITY AND VERTICAL STRUCTURE OF THE BOUNDARY LAYER 57
the land mast on the coast, to about 9 m s1at the sea mast, 2 km downstream from
the coast. This acceleration is probably at least partly due to the decreased rough-
ness of the water surface. As a result of this acceleration, σwdoes not decrease in
the downwind direction in spite of increasing stability.
To various degrees, the above flow cases are complicated by:
(a) A roughness change that would act to accelerate the flow in the two cases of
flow over the water;
(b) An adverse component of the hydrostatic pressure gradient associated with
decreasing temperature in the downstream direction (Doran et al., 1995; Sun
et al., 1998);
(c) Acceleration of the flow immediately above the cool internal boundary layer,
that was originally within the upstream boundary layer, but is now decoupled
from the surface stress (Smedman et al, 1995; Mahrt, 1999).
These three influences are not included in R. Therefore, Ris, by itself,
incomplete.
As a final example, consider flow from a warm region of nontranspiring senes-
cent wheat and bare fields to a cooler region of transpiring green grass examined
with aircraft passes in SGP (Section 5). Although, the surface radiation temper-
ature decreases significantly between the senescent wheat and the grassland, the
air temperature upstream is approximately the same as the downstream surface
radiation temperature, making Rundeterminably large. An internal boundary layer
does not form in terms of the mean flow. Therefore, this case is considered to be
an “adjusting boundary layer” instead of an internal boundary layer, as discussed
in the next subsection.
7.3. ADJUSTING BOUNDARY LAYERS
An internal boundary layer may not develop in flow over small amplitude surface
heterogeneity. Instead, the original boundary layer gradually adjusts in the down-
stream direction without decoupling of the upstream boundary layer from the new
surface or modification of the mean vertical structure associated with formation of
internal boundary layers. To contrast the adjusting boundary layer with formation
of a new internal boundary layer, we consult the approximate stationary budget of
an arbitrary variable, φ, downstream from the surface change
uφ
∂x ≈−
w0φ0
∂z ,(21)
where the horizontal turbulent flux divergence has been neglected. The vertical
flux divergence in an internal boundary layer is large because the flux generally
decreases from its surface value to a small value at hIBL such that the right hand
side of Equation (21) scales as w0φ0
sf c /hIBL. This large vertical flux divergence
corresponds to large horizontal gradients in the mean flow. However in the adjust-
ing boundary layer, the vertical flux divergence is governed by the boundary-layer
58 L. MAHRT
depth and scales as w0φ0
sf c /h. This flux divergence is much weaker compared
to the internal boundary layer and therefore must correspond to weak horizontal
gradients. This distinction becomes more obscure in the nonstationary case.
Consider the flow from a strongly heated surface to a weakly heated surface,
as observed in SGP where air from senescent wheat and bare fields is advected
over green grasslands with large transpiration and weaker surface heat flux. The
surface heat flux decreases from about 0.15 to about 0.02 K m s1. Here a well-
defined internal boundary layer is not observed in terms of mean variables such
as temperature and moisture. Temperature decreases by a few tenths of a kelvin,
mainly due to the absence of thermals over the grassland (Figure 7). That is, the
mean air temperature over the grassland is about the same as the air between the
thermals over the upstream region of senescent wheat. The weak surface heating
over the grassland is sufficient to maintain mixing between the air near the sur-
face and overlying boundary layer advected from upstream, preventing significant
vertical flux divergence and formation of an internal boundary layer.
However, the temperature variance decreases rapidly in the downstream direc-
tion due to the absence of strong thermals (Figure 7). Such thermals are absent
because the potential temperature of the advected air is not significantly different
from the surface radiation temperature of the grassland. Therefore, the adjusting
boundary layer does exhibit horizontal changes of some of the higher moments
but is missing significant horizontal change of the mean variables. Such boundary-
layer adjustments are probably common but have received little attention in the
literature.
8. Conclusions
Through interpretation of the existing literature and analysis of new data, this
study has examined the vertical extent of the influence of surface heterogeneity
on different scales. If the influence of mesoscale heterogeneity extends vertically
to levels where the spatially-averaged flux is significantly different from the surface
flux, then MO similarity theory and bulk methods cannot be used to estimate the
area-averaged surface flux. As a possible result, a consensus for parameterization
of effective roughness lengths and transfer coefficients for heterogeneous surfaces
has not emerged for these scales of surface heterogeneity. The analysis in Sec-
tion 4.1 indicates that the flux aggregation method (tile or mosaic approach) for
semi-explicit inclusion of subgrid surface heterogeneity in numerical models is
formally valid only with small-scale heterogeneity and not valid for larger scales
of heterogeneity.
The response of boundary-layer fluxes to surface heterogeneity increases with
the scale of the heterogeneity and decreases with observational height, wind speed,
boundary-layer depth and stability. In addition, ubiquitous background transient
motions influence the spatial variation of the measured flux. Actual surface het-
SURFACE HETEROGENEITY AND VERTICAL STRUCTURE OF THE BOUNDARY LAYER 59
a)
b)
Figure 7. (a) The instantaneous spatial variation of potential temperature and (b) standard deviations
of temperature and vertical velocity computed over moving 100-m windows. An internal boundary
layer is not well-defined in terms of mean variables or change of σw.
erogeneity normally consists of simultaneous spatial variations on more than one
scale. Because of these multiple influences, simple scaling arguments can only
explain a limited fraction of the spatial variation of fluxes. Nonetheless, traditional
blending height formulations are able to roughly predict the degree of relationship
between the heat flux at the top of the surface layer and the surface radiation tem-
perature (Section 6.2). However, the blending height formulation cannot predict
the amplitude of the atmospheric response to the surface heterogeneity because:
(a) it does not include information on the amplitude of the surface heterogeneity,
and (b) the amplitude of the atmospheric response is reduced by horizontal mixing
due to large boundary-layer eddies that scale with the boundary-layer depth. Tra-
60 L. MAHRT
ditional blending height formulations do not include information on the height of
the boundary layer.
The scaling arguments of Raupach and Finnigan (1995) include information on
the boundary-layer depth and consequently successfully explain part of the spatial
variation of the heat flux when the flow is unstable. The scaling arguments of
Raupach and Finnigan (1995) are used to derive a new formulation which includes
the role of the amplitude of the surface heterogeneity. The resulting predictions
can be used to estimate the range of scales of surface heterogeneity that influence
the boundary layer at a given height (Section 6.2) and the required resolution of the
remotely-sensed surface heterogeneity for inferring the spatial variation of heat and
moisture fluxes. The spatial variation of the moisture flux is much less organized
than that for the heat flux and more data are required before a formulation can be
derived for the moisture flux. The above simple scaling arguments did not appear
to be as successful when applied to additional data sets with less passes per flight
(not discussed above). Data are required for near-neutral conditions to combine
with the present data sets. This would allow evaluation of combining the blending
height and convective scaling approaches.
With flows over surface discontinuities, local internal boundary layers often
fail to develop. When the change of surface properties is not sharp and/or of
small-amplitude, the boundary layer adjusts without formation of a new internal
boundary layer within the existing boundary layer. Suchadjusting boundary layers
exhibit significant horizontal variation of some of the higher moments, such as heat
flux and temperature variance, but do not show significant horizontal variation of
mean quantities. Even when local internal boundary layers do form, they normally
include a vertically thick diffuse transition layer without a sharp top. The top of
the internal boundary layer is only occasionally observed as a sharp interface from
instantaneous data. The top fluctuates in time so that the top of the time-averaged
internal boundary layer is diffuse. Over heated surfaces, the top of the internal
boundary layer represents the upper limit of the thermals over the new surface and
can rise quite steeply in weak wind conditions.
In contrast, internal boundary layers forming due to marked surface changes
affecting a large area, such as flow of marine air over a heated land surface (meso-
scale internal boundary layers), are characterized by well-defined tops associated
with capping inversions (Garratt, 1990). The literature has concentrated on flow
over sharp discontinuities. Little is known about flow over surface heterogeneity
where the boundary layer adjusts in a less dramatic fashion and traditional internal
boundary-layer concepts are of limited use.
Acknowledgements
I gratefully acknowledge Ian McPherson for the NRC Twin Otter data and Sesha
Sadagopan for computational assistance. The extensive comments of the reviewers,
SURFACE HETEROGENEITY AND VERTICAL STRUCTURE OF THE BOUNDARY LAYER 61
Dean Vickers, Nigel Wood and Jielun Sun are greatly appreciated. This material is
based upon work supported by Grants DAAH04-96-10037 and DAAD19-9910249
from the U.S. Army Research Office and Grant NAG5-8601 from the NASA
Hydrology Program.
References
Avissar, R. and Pielke, R. A.: 1989, ‘A Parameterization of Heterogeneous Land Surfaces for At-
mospheric Numerical Models and its Impact on Regional Meteorology’, Mon. Wea. Rev. 117,
2113–2136.
Brutsaert, W. H.: 1982, Evaporation into the Atmosphere. D. Reidel, Dordrecht, 299 pp.
Cellier, P. and Brunet, Y.: 1992, ‘Flux-Gradient Relationships above Tall Plant Canopies’, Agric. For.
Meteorol. 58, 93–117.
Claussen, M.: 1990, ‘Area-Averaging of Surface Fluxes in a Neutrally Stratified, Horizontally
Inhomogeneous Atmospheric Boundary Layer’, Atmos. Environ. 24a, 1349–1360.
de Bruin, H. A. R., Bink, N. J., and Kroon, L. J. M.: 1991, ‘Fluxes in the Surface Layer under
Advective Conditions’, in T. J. Schmugge and J.-C. Andre (eds.), Land Surface Evaporation;
Measurement and Parameterization, Springer Verlag, pp. 157–170.
Doran, J. C., Shaw, W. J., and Hubbe, J. M.: 1995, ‘Boundary Layer Characteristics over Areas of
Inhomogeneous Surface Fluxes’, J. Appl. Meteorol. 34, 559–571.
Garratt, J. R.: 1978, Flux-Profile Relations above Tall Vegetation’, Quart. J. Roy. Meteorol. Soc.
104, 199–212.
Garratt, J. R.: 1980, ‘Surface Influence upon Vertical Profiles in the Atmospheric Near-Surface
Layer’, Quart. J. Roy. Meteorol. Soc. 106, 803–819.
Garratt, J. R.: 1990, ‘The Internal Boundary Layer – A Review’, Boundary-Layer Meteorol. 50,
171–203.
Garratt, J. R. and Ryan, B. F.: 1989, ‘The Structure of the Stably Stratified Internal Boundary Layer
in Offshore Flow over the Sea’, Boundary-Layer Meteorol. 47, 17–40.
Grant, A. L. M.: 1991, ‘Surface Drag and Turbulence over an Inhomogeneous Land Surface’,
Boundary-Layer Meteorol. 56, 309–339.
Hess, G. D.: 1994, ‘Recent Ideas on Roughness, Inhomogeneities and Mixing in the Atmospheric
Boundary Layer’, Aust.Met.Mag.43, 239–253.
Holtslag, A. A. M. and Nieuwstadt, F. T. M.: 1986, ‘Scaling the Atmospheric Boundary Layer’,
Boundary-Layer Meteorol. 36, 201–209.
Lettau, H. H.: 1969, ‘Note on Aerodynamic Roughness-Parameter Estimation on the Basis of
Roughness-Element Description’, J. Appl. Meteorol. 8, 823–832.
Mahrt, L.: 1979, ‘Penetrative Convection at the Top of a Growing Boundary Layer’, Quart. J. Roy.
Meteorol. Soc. 105, 469–485.
Mahrt, L.: 1996, ‘The Bulk Aerodynamic Formulation over Heterogeneous Surfaces’, Boundary-
Layer Meteorol. 78, 87–119.
Mahrt, L.: 1999, ‘Stratified Atmospheric Boundary Layers’, Boundary-Layer Meteorol. 90, 375–396.
Mahrt, L., Sun, J., Vickers, D., MacPherson, J. I., Pederson, J. R., and Desjardins, R. L.: 1994,
‘Observations of Fluxes and Inland Breezes over a Heterogeneous Surface’, J. Atmos. Sci. 51,
2165–2178.
Mahrt, L., Vickers, D., Edson, J., Sun, J., Højstrup, J., Hare, J., and Wilczak, J.: 1998, ‘Heat Flux in
the Coastal Zone’, Boundary-Layer Meteorol. 86, 421–446.
Mason, P. J.: 1988, ‘The Formation of Areally Averaged Roughness Lengths’, Quart. J. Roy.
Meteorol. Soc. 114, 399–420.
62 L. MAHRT
McNaughton, K. G. and Laubach, J.: 1998, ‘Unsteadiness as a Cause on Non-Equality of Eddy Dif-
fusivities for Heat and Vapour at the Base of an Advective Inversion’, Boundary-Layer Meteorol.
88, 479–504.
Parlange, M. and Brutsaert, W.: 1993, ‘Regional Shear Stress of Broken Forest from Radiosonde
Wind Profiles in the Unstable Surface Layer’, Boundary-Layer Meteorol. 64, 355–368.
Pederson, J. R., Mahrt, L., Delaby, A., Oncley, S., Den Hartog, G., Newmann, H. H., Mickle, R.
E., Shaw, R. H., Paw U, K. T., Grantz, D. A., MacPherson, J. I., Desjardins, R., Schuepp, P. H.,
Pearson Jr., R., and Arcado, T. E.: 1995, ‘California Ozone Deposition Experiment: Methods,
Results, and Opportunities’, Atmos. Environ. 29, 3115–3132.
Physick, W. and Garratt, J. R.: 1995, ‘Incorporation of a High-Roughness Lower Boundary into
a Mesoscale Model for Studies of Dry Deposition over Complex Terrain’, Boundary-Layer
Meteorol. 74, 55–71.
Raupach, M. R.: 1992, ‘Drag and Drag Partition on Rough Surfaces’, Boundary-Layer Meteorol. 60,
375–395.
Raupach, M. R.: 1994, ‘Simplified Expressions for Vegetation Roughness Length and Zero-Plane
Displacement as Functions of Canopy Height and Area Index’, Boundary-Layer Meteorol. 71,
211–216.
Raupach, R. R. and Finnigan, J. J.: 1995, ‘Scale Issues in Boundary-Layer Meteorology: Surface
Energy Balances in Heterogeneous Terrain’, Hydro. Proc. 9, 589–612.
Raupach, M. R., Antonia, R. A., and Rajagopalan, S.: 1991, ‘Rough-Wall Turbulent Boundary
Layers’, Appl. Mechanics Revs. 44, 1–25.
Raupach, M. R., Thom, A. S., and Edwards, I.: 1980, ‘A Wind Tunnel Study of Turbulent Flow Close
to Regularity Arrayed Rough Surface’, Boundary-Layer Meteorol. 18, 373–397.
Schmid, H. P. and Bünzli, D.: 1995, ‘The Influence of Surface Texture on the Effective Roughness
Length’, Quart. J. Roy. Meteorol. Soc. 121A, 1–21.
Smedman, A.-S., Bergström, H., and Högström, U.: 1995, ‘Spectra, Variances and Length Scales in
a Marine Stable Boundary Layer Dominated by a Low Level Jet’, Boundary-Layer Meteorol. 76,
211–232.
Sorbjan, Z.: 1989, Structure of the Atmospheric Boundary Layer, Prentice Hall, Englewood Cliffs,
NJ, 317 pp.
Sun, J. and Mahrt, L.: 1995, ‘Relationship of the Heat Flux to Microscale Temperature Variations:
Application to BOREAS’, Boundary-Layer Meteorol. 76, 291–301.
Sun, J., Desjardins, R., Mahrt, L., and MacPherson, J. I.: 1998, ‘Transport of Carbon Dioxide, Water
Vapor and Ozone by Turbulence and Local Circulations’, J. Geophys. Res. 103, 25,873–25,885.
Sun, J., Lenschow, D., Mahrt, L., Crawford, T., Davis, K., Oncley, S., MacPherson, J. I., Wang,
Q., Dobosy, R., and Desjardins, R. L.: 1997, ‘Lake-Induced Atmospheric Circulations during
BOREAS’, J. Geophys. Res. 102, 29,155–29,166.
Sun, J., Massman, W., and Grantz, D.: 1999, ‘Aerodynamic Variables in the Bulk Formula’,
Boundary-Layer Meteorol. 91, 109–125.
Taylor, P. A.: 1987, ‘Comments and Further Analysis on Effective Roughness Lengths for Use in
Numerical Three-Dimensional Models’, Boundary-Layer Meteorol. 39, 403–418.
Wieringa, J.: 1993, ‘Representative Roughness Parameters for Homogeneous Terrain’, Boundary-
Layer Meteorol. 63, 323–364.
Wood, N. and Mason, P. J.: 1991, ‘The Influence of Stability on the Effective Roughness Lengths for
Momentum and Heat Flux’, Quart. J. Roy. Meteorol. Soc. 117, 1025–1056.
... Wang et al., 2010). The land surface is usually heterogeneous over a wide range of spatial scales due to variability in (among other parameters) vegetation, terrain, soil texture and wetness, cloud cover, and urban areas (Desai et al., 2005;Desai, Paleri, et al., 2022;Mahrt, 2000). However, measurements at a single location, such as eddy correlation flux towers, are often used to represent the properties of a larger region. ...
... The lack of understanding of the small-scale dynamics of water vapor throughout the PBL leads to strong limitations in predicting localized phenomena in weather models (Couvreux et al., 2005;Hill et al., 2008Hill et al., , 2011Steinfeld et al., 2007). As such, the multiscale nature of water vapor has continued to defy a generalized approach or theory for "characterizing" its impact on the PBL (Mahrt, 2000). Heat and moisture exchange measurements between the land surface and the atmosphere are critical to understanding the causes of variability in the PBL. ...
Article
Full-text available
The water vapor transport associated with latent heat flux (LE) in the planetary boundary layer (PBL) is critical for the atmospheric hydrological cycle, radiation balance, and cloud formation. The spatiotemporal variability of LE and water vapor mixing ratio (rv) are poorly understood due to the scale‐dependent and nonlinear atmospheric transport responses to land surface heterogeneity. Here, airborne in situ measurements with the wavelet technique are utilized to investigate scale‐dependent relationships among LE, vertical velocity (w) variance (σw2 ${\sigma }_{w}^{2}$), and rv variance (σH2O2 ${\sigma }_{\mathrm{H}2\mathrm{O}}^{2}$) over a heterogeneous surface during the Chequamegon Heterogeneous Ecosystem Energy‐balance Study Enabled by a High‐density Extensive Array of Detectors 2019 (CHEESEHEAD19) field campaign. Our findings reveal distinct scale distributions of LE, σw2 ${\sigma }_{w}^{2}$, and σH2O2 ${\sigma }_{\mathrm{H}2\mathrm{O}}^{2}$ at 100 m height, with a majority scale range of 120 m–4 km in LE, 32 m–2 km in σw2 ${\sigma }_{w}^{2}$, and 200 m–8 km in σH2O2 ${\sigma }_{\mathrm{H}2\mathrm{O}}^{2}$. The scales are classified into three scale ranges, the turbulent scale (8–200 m), large‐eddy scale (200 m–2 km), and mesoscale (2–8 km) to evaluate scale‐resolved LE contributed by σw2 ${\sigma }_{w}^{2}$ and σH2O2 ${\sigma }_{\mathrm{H}2\mathrm{O}}^{2}$. The large‐eddy scale in PBL contributes over 70% of the monthly mean total LE with equal parts (50%) of contributions from σw2 ${\sigma }_{w}^{2}$ and σH2O2 ${\sigma }_{\mathrm{H}2\mathrm{O}}^{2}$. The monthly temporal variations mainly come from the first two major contributing classified scales in LE, σw2 ${\sigma }_{w}^{2}$, and σH2O2 ${\sigma }_{\mathrm{H}2\mathrm{O}}^{2}$. These results confirm the dominant role of the large‐eddy scale in the PBL in the vertical moisture transport from the surface to the PBL, while the mesoscale is shown to contribute an additional ∼20%. This analysis complements published scale‐dependent LE variations, which lack detailed scale‐dependent vertical velocity and moisture information.
... The land surface has a well-documented ability to influence planetary boundary layer (PBL) characteristics (Avissar & Pielke, 1989;Brunsell et al., 2011;H. Huang & Margulis, 2013;Hubbe et al., 1997;Kustas & Albertson, 2003;Mahrt, 2000), clouds and precipitation (Berg & Stull, 2005;Kang, 2016;Pielke Sr, 2001;Rieck et al., 2014;Schrieber et al., 1996), and hydrometeorological extremes Fischer et al., 2007;Hirschi et al., 2011;Mocko et al., 2021;Santanello et al., 2015). Even relatively small scale surface features create important temperature and moisture gradients, such as between rivers/lakes and adjacent land (Ramos da Silva et al., 2011;Y. ...
Article
Full-text available
Land‐atmosphere interactions are central to the evolution of the atmospheric boundary layer and the subsequent formation of clouds and precipitation. Existing global climate models represent these connections with bulk approximations on coarse spatial scales, but observations suggest that small‐scale variations in surface characteristics and co‐located turbulent and momentum fluxes can significantly impact the atmosphere. Recent model development efforts have attempted to capture this phenomenon by coupling existing representations of subgrid‐scale (SGS) heterogeneity between land and atmosphere models. Such approaches are in their infancy and it is not yet clear if they can produce a realistic atmospheric response to surface heterogeneity. Here, we implement a parameterization to capture the effects of SGS heterogeneity in the Community Earth System Model (CESM2), and compare single‐column simulations against high‐resolution Weather Research and Forecasting (WRF) large‐eddy simulations (LESs), which we use as a proxy for observations. The CESM2 experiments increase the temperature and humidity variances in the lowest atmospheric levels, but the response is weaker than in WRF‐LES. In part, this is attributed to an underestimate of surface heterogeneity in the land model due to a lack of SGS meteorology, a separation between deep and shallow convection schemes in the atmosphere, and a lack of explicitly represented mesoscale secondary circulations. These results highlight the complex processes involved in capturing the effects of SGS heterogeneity and suggest the need for parameterizations that communicate their influence not only at the surface but also vertically.
... In such situations, flux averaging should be preferred (e.g. Wood and Mason 1991;Raupach and Finnigan 1995;Mahrt 2000). However, the observational constraints do not allow accurate separation of the flux contributions from water, sea-ice and form drag in heterogeneous icewater surfaces. ...
Article
Full-text available
Surface turbulent exchanges play a key role on sea ice dynamics, on ocean and sea ice heat budgets and on the polar atmosphere. Uncertainties in parameterizations of surface turbulent fluxes are mostly held by the transfer coefficients and estimates of those transfer coefficients from field data are required for parameterization development. Measurement errors propagate through the computation of transfer coefficients and contribute to its total error together with the uncertainties in the empirical stability functions used to correct for stability effects. Here we propose a methodology to assess their contributions individually to each coefficient estimate as well as the total drag coefficient uncertainty and we apply this methodology on the example of the SHEBA campaign. We conclude that for most common drag coefficient values (between $$1.0\times 10^{-3}$$ 1.0 × 10 - 3 and $$2.5\times 10^{-3}$$ 2.5 × 10 - 3 ), the relative total uncertainty ranges from 25 and 50 $$\%$$ % . For stable or unstable conditions with a stability parameter $$|\zeta |>1$$ | ζ | > 1 on average, the total uncertainty in the neutral drag coefficient exceeds the neutral drag coefficient value itself, while for $$|\zeta |<1$$ | ζ | < 1 the total uncertainty is around 25 $$\%$$ % of the drag coefficient. For closer-to-neutral conditions, this uncertainty is dominated by measurement uncertainties in surface turbulent momentum fluxes which should therefore be the target of efforts in uncertainty reduction. We also propose an objective data-screening procedure for field data, which consists of retaining data for which the relative error on neutral drag coefficient does not exceed a given threshold. This method, in addition to the commonly used flux quality control procedure, allows for a reduction of the drag coefficient dispersion compared to other data-screening methods, which we take as an indication of better dataset quality.
... It is also notable that, although lateral mixing with the environment generally causes the updraft properties to converge with height, updrafts from the various surface types retain distinct temperature distributions up to at least 800 m in this instance, at which point the forest updraft velocities are no longer positive and they detrain. The level at which updraft properties converge may be considered an "updraft blending height," analogous to the blending height at which the atmosphere over a heterogeneous surface approaches homogeneity (Mahrt, 2000). In this case, the height is a reflection of updraft properties 225 rather than variability at the scale of surface heterogeneity. ...
Preprint
Full-text available
Earth System Models (ESMs) typically represent surface heterogeneity on scales smaller than the atmospheric grid, while land-atmosphere coupling is based on grid mean values. Here we present a general approach allowing subgrid surface heterogeneity to influence the updraft thermodynamic properties in a multi-plume mass flux parameterization. The approach is demonstrated in single column experiments with an Eddy Diffusivity Mass Flux (EDMF) boundary layer scheme. Instead of triggering based on grid mean surface values, updrafts are explicitly assigned to individual surface tiles with positive buoyancy flux. Joint distributions of near-surface vertical velocities and thermodynamic variables are defined over individual surface tiles, and updraft properties are drawn from the positive tails of the distributions. The approach allows updraft properties to covary with surface heterogeneity, and updrafts from different tiles maintain distinct properties to heights of several hundred metres. Mass flux contributions to subgrid variances are increased near the surface, but impacts on mean state variables are relatively small. We suggest that larger impacts might be obtained by adding a specialized plume to represent the effects of secondary circulations.
... This phenomenon means that atmospheric boundary layer (ABL) flow 25 can be perturbed not just at one specific length or time scale but at multiple scales, which excites physical processes such as advection, secondary circulations, and non-equilibrium turbulence (Bou-Zeid et al., 2020). Further, the issue of land surface heterogeneity is fundamentally an issue of scale (Mahrt, 2000;Bou-Zeid et al., 2020) coupled to local biogeophysical processes, as well as larger scale synoptic circulations. Observing and quantifying these interactions and their role in ABL development calls for space and scale intensive deployments seldom seen in micrometeorological field experiments (Wulfmeyer et al., 2018). ...
Preprint
Full-text available
Observational studies and large eddy simulations (LES) have reported secondary circulations in the turbulent atmospheric boundary layer (ABL). These circulations form as coherent turbulent structures or mesoscale circulations induced by gradients of land surface properties. However, simulations have been limited in their ability to represent these events and their diurnal evolution over realistic and heterogeneous land surfaces. In this study, we present a LES framework combining the high-resolution observational data collected as part of the CHEESEHEAD19 field campaign to overcome this gap and 5 test how heterogeneity influences the ABL response. We simulated diurnal cycles for four days chosen from late summer to early autumn over a large (49×52 km) heterogeneous domain. To investigate surface atmospheric feedbacks such as self-reinforcement of mesoscale circulations over the heterogeneous domain, the simulations were forced with an interactive land surface model with coupled soil, radiative transfer and plant canopy model. The lateral and model top boundary conditions were prepared from National Oceanic and Atmospheric Administration High-Resolution Rapid Refresh (HRRR) meteorologi-10 cal analysis fields. Comparing the simulated profile and near surface data with field measured radiosonde and eddy covariance station data showed a realistic evolution of the near-surface meteorological fields, heat and moisture fluxes and the ABL. The LES had limitations in simulating the night-time cooling in the nocturnal boundary layer. The simulated fields were strongly modulated by the imposed HRRR derived mesoscale boundary conditions, resulting in a slightly warmer and drier ABL. The simulations were run without clouds which resulted in higher daytime sensible heat fluxes for some scenarios. 15 Our findings demonstrate the capability of the PALM model system to realistically represent the daytime evolution of the ABL response over unstructured heterogeneity and the limitations involved therein with respect to the role of boundary conditions and the representation of the nocturnal boundary layer. The simulation setup and dataset described in this manuscript build the baseline to tackle specific research questions associated with the CHEESEHEAD19 campaign, particularly to address questions about the role of heterogeneous ecosystems in modulating surface-atmosphere fluxes and near surface meteorological 20 1 https://doi.
Article
Full-text available
The study presents the first results from the multi-platform observational campaign carried out at the Mukhrino peatland in June 2022. The focus of the study is the quantification of spatial contrasts of the surface heat budget terms and methane emissions across the peatland, which arise due to the presence of microlandscape heterogeneities. It is found that surface temperature contrasts across the peatland exceeded 10 ◦C for clear-sky conditions both during day and night. Diurnal variation of surface temperature was strongest over ridges and drier hollows and was smallest over the waterlogged hollows and shallow lakes. This resulted in strong spatial variations of sensible heat flux (H) and Bowen ratio, while the latent heat varied much less. During the clear-sky days, H over ryam exceeded the one over the waterlogged hollow by more than a factor of two. The Bowen ratio amounted to about unity over ryam, which is similar to values over forests. Methane emissions estimated using the static-chamber method also strongly varied between various microlandscapes, being largest at a hollow within a ridge-hollow complex and smallest at a ridge. A strong nocturnal increase in methane mixing ratio was observed and was used in the framework of the atmospheric boundary layer budget method to estimate nocturnal methane emissions, which were found to be in the same order of magnitude as daytime emissions. Finally, the directions for further research are outlined, including the verification of flux-aggregation techniques, parameterizations of surface roughness and turbulent exchange, and land-surface model evaluation and development.
Article
Full-text available
A large number of surface features (e.g., frost, gullies, slope streaks, recurring slope lineae) are observed on Martian slopes. Their activity is often associated with the specific microclimates on these slopes, which have been mostly studied with one‐dimensional radiative balance models to date. We develop here a parameterization to simulate these microclimates in 3D Global Climate Models. We first demonstrate that any Martian slope can be thermally represented by a poleward or equatorward slope, that is, the daily average, minimum, and maximum surface temperatures depend on the North‐South component of the slope. Based on this observation, we implement here a subgrid‐scale parameterization to represent slope microclimates (radiative fluxes, volatile condensation, ignoring slope winds for now) in the Mars Planetary Climate Model and validate it through comparisons with surface temperature measurements and frost detections on sloped terrains. With this new model, we show that slope microclimates do not have a significant impact on the seasonal CO2 and H2O cycles on a global scale. Furthermore, short‐scale slopes (i.e., less than ∼1 km in length) do not significantly impact the thermal state of the atmosphere. Ninety‐one percent of the active gullies are found where our model predicts CO2 frost, suggesting that their activity is related to processes involving CO2 ice. However, the low thicknesses (≤tens of cm) predicted at mid‐latitudes rule out mechanisms involving large amounts (∼meters) of ice. This model opens the way to new studies on surface‐atmosphere interactions in present and past climates.
Article
The structure and impact of thermally induced secondary motions in stably stratified channel flows with two-dimensional surface temperature inhomogeneities is studied using direct numerical simulation (DNS). Starting from a configuration with only spanwise varying surface temperature, where the streamwise direction is homogeneous (Bon & Meyers, J. Fluid Mech., 2022, pp. 1–38), we study cases where the periodic temperature strip length lx/h (with h the half-channel height) assumes finite values. The patch width (ly/h={π/4,π/8}) and length are varied at fixed stability and two different Reynolds numbers. Results indicate that for the investigated patch widths, the streamwise development of the secondary flows depends on the patch aspect ratio (a=lx/ly), while they reach a fully developed state after approximately 25ly. The strength of the secondary motions, and their impact on momentum and heat transfer through the dispersive fluxes, is strongly reduced as the length of the temperature strips decreases, and becomes negligible when a≲1. We demonstrate that upward dispersive and turbulent heat transport in locally unstably stratified regions above the high-temperature patches lead to reduced overall downward heat transfer. Comparison to local Monin–Obukhov similarity theory (MOST) reveals that scaled velocity and temperature gradients in homogeneous stably stratified channel flow at Reτ=550 agree reasonably well with empirical correlations obtained from meteorological data. For thermally heterogeneous cases with strips of finite length, the similarity functions only collapse higher above the surface, where dispersive fluxes are negligible. Lastly, we show that mean profiles of all simulations collapse when using outer-layer scaling based on displacement thickness.
Article
Odor plume tracking is important for many organisms, and flying insects have served as popular model systems for studying this behavior both in field and laboratory settings. The shape and statistics of the airborne odor plumes that insects follow are largely governed by the wind that advects them. Prior atmospheric studies have investigated aspects of microscale wind patterns with an emphasis on characterizing pollution dispersion, enhancing weather prediction models, and for assessing wind energy potential. Here, we aim to characterize microscale wind dynamics through the lens of short-term ecological functions by focusing on spatial and temporal scales most relevant to insects actively searching for odor sources. We collected and compared near-surface wind data across three distinct environments (sage steppe, forest, and urban) in Northern Nevada. Our findings show that near-surface wind direction variability decreases with increasing wind speeds and increases in environments with greater surface complexity. Across environments, there is a strong correlation between the variability in the wind speed (i.e., turbulence intensity) and wind direction (i.e., standard deviation in wind direction). In some environments, the standard deviation in the wind direction varied as much as 15°–75° on time scales of 1–10 min. We draw insight between our findings and previous plume tracking experiments to provide a general intuition for future field research and guidance for wind tunnel design. Our analysis suggests a hypothesis that there may be an ideal range of wind speeds and environment complexity in which insects will be most successful when tracking odor plumes over long distances.
Preprint
Full-text available
The spatiotemporal variability of latent heat flux (LE) and water vapor mixing ratio (rv) variability are not well understood due to the scale-dependent and nonlinear atmospheric energy balance responses to land surface heterogeneity. Airborne in situ and profiling Raman lidar measurements with the wavelet technique are utilized to investigate scale-dependent relationships among LE, vertical velocity (w) variance (s2w), and rv variance (s2wv) over a heterogeneous surface in the Chequamegon Heterogeneous Ecosystem Energy-balance Study Enabled by a High-density Extensive Array of Detectors 2019 (CHEESEHEAD19) field campaign. Our findings reveal distinct scale distributions of LE, s2w, and s2wv at 100 m height, with a majority scale range of 120m-4km in LE, 32m-2km in s2w, and 200 m – 8 km in s2wv. The scales are classified into three scale ranges, the turbulent scale (8m–200m), large-eddy scale (200m–2km), and mesoscale (2 km–8km) to evaluate scale-resolved LE contributed by s2w and s2wv. In the large-eddy scale in Planetary Boundary Layer (PBL), 69-75% of total LE comes from 31-51% of the total sw and 39-59% of the total s2wv. Variations exist in LE, s2w, and s2wv, with a range of 1.7-11.1% of total values in monthly-mean variation, and 0.6–7.8% of total values in flight legs from July to September. These results confirm the dominant role of the large-eddy scale in the PBL in the vertical moisture transport from the surface to the PBL. This analysis complements published scale-dependent LE variations, which lack detailed scale-dependent vertical velocity and moisture information.
Article
Full-text available
Nocturnal land breezes and daytime lake breezes are studied using data collected by the Canadian Twin Otter aircraft and a deck boat which traversed Candle Lake during the Boreal Ecosystem-Atmosphere Study (BOREAS). The nocturnal vertical transport of CO2, water vapor, and ozone over the lake consists of two parts: (1) mesoscale rising motion associated with land breeze convergence and (2) significant turbulence and vertical mixing driven by buoyancy in the lower part of the internal boundary layer and shear generation in the top part of the internal boundary layer. For comparison, the role of the lake in the daytime is examined in terms of formation of a stable internal boundary layer due to advection of warm air from land with small CO2 concentration over the cooler lake surface. Analysis of the aircraft and boat data indicates that the nocturnal land breeze plays an important role in the regional CO2 budget in the lake region. In the present study, CO2 is advected horizontally by a nocturnal land breeze circulation and vented vertically over Candle Lake (“chimney effect”). Such near-surface horizontal transport implies that part of the respirated CO2 never reaches the tower observational level, particularly under light wind conditions. This study speculates that preferred locations of vertical venting of CO2 may also occur due to convergence of nocturnal drainage circulations or flow meandering, although probably weaker than that associated with the land breeze. These circulations partly explain recent findings that tower-measured nocturnal turbulent fluxes of CO2 above the canopy and the subcanopy storage of CO2 frequently sum to less than the total respiration of CO2, leading to “missing CO2.” Unfortunately, the present study does not allow evaluation of all of the terms in the carbon dioxide budget.
Article
The Monin-Obukhov similarity theory forms the basis of the methodology used to calculate surface fluxes of momentum, and sensible and latent heat in numerical models of the atmosphere. However, observed surface properties are often not homogeneous over the subgrid length scales used in models as assumed in the theory. Several simple models are presented to generalise the concept of roughness to apply in non-homogeneous conditions because of variations in terrain, vegetation, or the presence of buildings or other obstacles. Turbulent mixing within the atmospheric boundary layer (ABL) is another important element of ABL parameterisation. -from Author
Article
This review considers theoretical and experimental knowledge of rough-wall turbulent boundary layers, drawing from both laboratory and atmospheric data. The former apply mainly to the region above the roughness sublayer (in which the roughness has a direct dynamical influence) whereas the latter resolve the structure of the roughness sublayer in some detail. Topics considered include the drag properties of rough surfaces as functions of the roughness geometry, the mean and turbulent velocity fields above the roughness sublayer, the properties of the flow close to and within the roughness canopy, and the nature of the organized motion in rough-wall boundary layers. Overall, there is strong support for the hypothesis of wall similarity: At sufficiently high Reynolds numbers, rough-wall and smooth-wall boundary layers have the same turbulence structure above the roughness (or viscous) sublayer, scaling with height, boundary-layer thickness, and friction velocity.
Article
In each surface grid element of the numerical model similar homogeneous land patches located at different places within the element are regrouped into subgrid classes. Then, for each one of the subgrid classes, a sophisticated micrometeorological model of the soil-plant-atmosphere system is applied to assess the surface temperature, humidity, and fluxes to the atmosphere. This parameterization was incorporated into a mesoscale numerical model to test the impact of subgrid-scale land surface heterogeneities on the development of local circulations. Where strong contrasts in total sensible heat flux are generated by land surface heterogeneities, circulations as strong as sea breezes may develop. -from Authors
Article
To remove small-scale variance and noise, time series of data are generally filtered using a moving window with a specified distribution of weights. Such filters unfortunately smooth sharp changes associated with larger scale structures. In this study, an adaptive low-pass filter is developed that not only effectively removes random small-scale variations but also retains sudden changes or sharp edges that are part of the large-scale features. These sudden changes include fronts, abrupt shifts in climate, sharp changes associated with a heterogeneous surface, or any jump in conditions associated with change on a larger scale.To construct the filter, gradients on different scales and at different positions in the time series are computed using a multiresolution representation of the data. The low-pass filter adapts to include smaller-scale variations at positions in the time series where the small-scale gradient is steep and represents change on a larger scale. The action of the filter is to apply a more concentrated distribution of weights at locations in the original time series where the signal is rapidly varying.As application examples, the filter is applied to turbulence data observed under strong wind conditions and climate data corresponding to 52 years of a Southern Oscillation index.
Article
This paper describes results from a June 1992 field program to study the response of the boundary layer over a site with well-defined extreme differences in sensible and latent heat fluxes over clearly separated areas, each with characteristic length scales of 10 km or more. The experiment region consisted of semiarid grassland steppe and irrigated farmland. Sensible heat flux maxima over the steppe regularly reached values in excess of 300 W m-2 and were typically a factor of 4 or more greater than those over the farmland. Two days were selected for analysis: one with moderate winds of 7-10 m s⁻¹ and one with lighter winds of 4-7 m s⁻¹ over the steppe. In both cases the wind directions were nearly perpendicular to the boundary between the steppe and farm. An analysis of potential temperature soundings showed that mixed-layer characteristics over both the farm and the steppe were largely determined by heating over the steppe, with advection from the steppe to the farm playing a significant role. On the day with the lighter winds, a secondary circulation related to the thermal contrasts between the two areas was observed. A simple conceptual model is described that predicts the extent of the cooler area required to generate such circulations. The observations illustrate how predictions of boundary layer structure in terms of local surface sensible heat fluxes may be compromised by advective effects. Such difficulties complicate efforts to obtain accurate representations of surface fluxes over inhomogeneous surfaces even if parameterizations of mesoscale contributions to the heat flux are included. 17 refs., 12 figs., 1 tab.
Article
We review scaling regimes of the idealized Atmospheric Boundary Layer. The main emphasis is given on recent findings for stable conditions. We present diagrams in which the scaling regimes are illustrated as a function of the major boundary-layer parameters. A discussion is given on the different properties of the scaling regimes in unstable and stable conditions.
Article
This paper examines the effect of non-stationarity of the wind on similarity of the eddy diffusivities for heat and vapour within a stable layer at the bottom of an internal boundary layer formed downwind of a dry-to-wet transition. First, we present some experimental data taken above a rice crop downwind of very extensive dry range lands at Warrawidgee, NSW, Australia. These data establish that periods of higher wind speed were associated with periods of higher saturation deficit in the canopy of the rice crop, and lower Bowen ratio. It is shown that Bowen ratios calculated for 30-second sub-intervals varied three-fold within a single 20-minute averaging period. Thus periods of higher wind speed corresponded to periods of higher moisture flux and smaller sensible heat flux.An idealized situation is then analysed theoretically. It is assumed that the time scale of the slow variations of the wind is long compared with the surface-layer time scale and that fetch is sufficient that the air near the ground is in continuous equilibrium with the surface. Using a two-scale Reynolds decomposition of the fluctuating wind and scalar variables into active and inactive components, it is shown that unsteadiness can lead to an eddy diffusivity for saturation deficit, calculated as the ratio of average flux to average gradient, that is larger than that for total energy calculated in a similar way. Using this ratio to calculate the ratio of diffusivities for temperature and humidity, K-T/K-q, it is found that the latter can be much larger than one if the Bowen ratio is small and negative. Despite this, assuming K-T = K-q and using the Bowen ratio method to calculate surface energy fluxes will usually incur only minor errors.