SCALE VARIABILITY OF SURFACE FLUXES OF ENERGY AND CARBON OVER A TROPICAL RAIN FOREST IN SOUTH-WEST...

Post on 28-Mar-2023

1 views 0 download

Transcript of SCALE VARIABILITY OF SURFACE FLUXES OF ENERGY AND CARBON OVER A TROPICAL RAIN FOREST IN SOUTH-WEST...

1

SCALE VARIABILITY OF SURFACE FLUXES OF ENERGY AND

CARBON OVER A TROPICAL RAIN FOREST IN SOUTH-WEST

AMAZONIA. I. DIURNAL CONDITIONS

Celso von Randow, Leonardo D. A. Sá, Prasad S. S. D. Gannabathula and

Antonio O. Manzi

Laboratório Associado de Meteorologia e Oceanografia (LMO), Centro de Previsão de

Tempo e Estudos Climáticos (CPTEC), Instituto Nacional de Pesquisas Espaciais (INPE),

São José dos Campos, Brazil

2

Abstract

The aim of this study is to investigate the low-frequency characteristics of diurnal

turbulent scalar spectra and cospectra near the Amazonian rain forest. This is because the

available turbulent data are often non stationary and there is no clear spectral gap to

separate data in 'mean' and 'turbulent' parts. Daubechies-8 orthogonal wavelet is used to

scale project turbulent signals in order to provide scale variance and covariance estimations.

Based on the characteristics of the scale dependence of the scalar fluxes, the total scalar

covariance of each 4 hours data run was partitioned into 'turbulent', 'large eddy', and

'mesoscale' covariance contributions according to the method proposed by Sun et al. [1996].

This permits to study some of the statistical characteristics of the scalar turbulent fields in

each one of the classes above mentioned and thus, to have a insight to explain the origin of

the variability of the scalar fields close to Amazonian forest. The data were measured in a

60m height micrometeorological tower built in Rebio-Jaru Biological Reserve (10o 04' S;

61o 56' W), in brazilian north-western state of Rondonia during the LBA (Large Scale

Biosphere-Atmosphere Experiment in Amazonia) 1999 wet season campaign. 10.4 Hz fast

response turbulent measurements were performed at 62.7 m above the ground (27 m above

the top canopy mean height). The largest amount of the sensible heat, latent heat and CO2

covariance contributions occurred on 'turbulent' length scales. Larger scale eddy motions,

however, can contribute up to 30 % to the total covariances under weak wind conditions.

Analysis of scale correlation coefficient (r(Tq)) between temperature (T) and humidity (q)

signals show that the scale pattern of T and q variability are not similar, and r(Tq) < 1 for

all analysed scales.

3

1. Introduction

The discussions about the accuracy on turbulent fluxes estimations using the eddy-

correlation method are many times forgotten in scientific works, although their importance

are evident, as discussed by [Wyngaard, 1983; Shuttleworth et al., 1984; Mahrt, 1991a;

Vickers and Mahrt, 1997; Mahrt, 1998; Affre et al., 2000; McNaughton and Laubach,

2000], among others. It is an extremely complex problem and contains several uncertainties

associated with the nature of turbulence [Lumley and Panofsky, 1964; Chimonas, 1985;

Kaimal et al., 1989; Wyngaard, 1992; McNaughton and Laubach, 2000], which can be

aggravated when measurements are made under certain peculiar conditions, such as on

towers [Mahrt, 1998], on aircrafts [Mann and Lenschow; 1994; Réchou and Durand, 1997],

over complex surfaces [Mahrt, 1991a], or under transient conditions [Howell and Mahrt,

1994a; Mahrt, 1998], in the context of disturbed surface layers [McNaughton and Laubach,

2000] or disturbed wet tropical boundary layer [Garstang and Fitzjarrald, 1999]. Two main

sources of imprecision on turbulent fluxes estimations could be stressed: the first one is

result of all the errors associated with the measurements themselves, and the second comes

from the way in which eddy-correlation method is used. On this work we will discuss the

second aspect of the problem, by means of investigating questions related to the

measurement of turbulent fluxes under non-stationary conditions or over non-homogeneous

surfaces and the dependence of flux calculations to the choice of cutoff frequency (or

averaging period) used on these calculations [Réchou and Durand, 1997; Mahrt, 1998].

This imprecision on flux estimations appears when the power spectra of turbulent

variables, such as w (vertical wind velocity) do not show a clear spectral gap (Figure 1),

4

when sampled at one hour scale. This problem was discussed by several authors: Lumley

and Panofsky [1964, pg. 45], on their classical study about atmospheric turbulence

structure, identified scenarios where the autocorrelation function does not tend to zero with

increase on time (or on length) in such a way that the determination of integral scales

sometimes becomes particularly difficult. It is evident that turbulent signals under such

conditions led to difficulties on definition of averaging periods and fluctuations [Mahrt,

1989; Hildebrand, 1991; Mahrt,1991a, 1998].

According to Pasquill [1974], two major objections could be raised against the

current statement concerning the validity of the stationarity and homogeneity hypotheses on

atmospheric boundary layer (ABL) turbulent fields. The first one results from the fact that

the power spectra of turbulent fluctuations extend to larger scales of motion, such as

mesoscale ones, which existence is connected to perturbations introduced by low-frequency

motions [McNaughton and Laubach, 2000]. Consequently, turbulent variables calculated

statistically are strongly dependent on sampling period and often their variances do not

show a stable mean value along the data sampling, as detected by authors such as

Hildebrand [1991] and Mahrt [1991a], on studies about flux estimation over complex

terrain. According to Tennekes and Wyngaard [1972] this problem tends to be aggravated

in higher order moment evaluations. The second complication is associated to the factors

regarding the variability imposed to the flow, by external forcing such as the roughness of

the terrain [McNaughton and Laubacch, 2000], the existence of complex distributions of

sources or sinks of heat and moisture at the surface [Brutsaert, 1998; Desjardins et al.,

1997], the interaction of the ABL flow with gravity waves exterior to it [Nai-Ping et al.,

5

1983; Chimonas, 1985; Stull, 1988], the effects originated from the top of the ABL [Mahrt,

1991b], etc.

Trying to deal with the determination of scale dependent fluxes, Sun et al [1996],

based on the characteristics of this dependence, proposed to partition the total flux into

turbulent, large-eddy, and mesoscale fluxes due to motions on scales smaller than 1 km,

between 1 and 5 km, and bigger than 5 km, respectively. In order to explain these

characteristics of flux-scale dependence, McNaughton and Laubach [2000] have proposed a

different three class scheme based on the earlier ideas of Kader and Yaglom [1990] and

their attempt to find a generalization of the classical Monin-Obukhov Similarity Theory

(MOST). After McNaughton and Laubach [2000], three different scaling regimes could act

on the surface layer turbulent exchanges: an inner-layer scaling (ILS), an outer-layer scaling

(OLS) and a combined scaling (CS). Each one of these regimes would explain the

characteristics of turbulent power spectra and cospectra on specific regions and their

classification depends on height and frequency. They have associated OLS components

with "inactive" turbulence and ILS components with "active" turbulence in the sense

proposed by Townsend's hypothesis about turbulence atmospheric structure [Townsend,

1976; Högström, 1990]. Still after McNaughton and Laubach, if the power spectra of active

and inactive components of turbulence are separated by a spectral gap, there would be no

interactions between them. But, on the other hand, if there is no obvious spectral gaps, there

would be a matching region with -1 and no more -5/3 power law (for u, v, and scalar

variables, but not for w-wind velocity component) corresponding to the CS region.

Therefore, each one of these classes should have a specific parameterization to represent the

subgrid fluxes on models of simulations of ABL, what is particularly difficult for large

6

eddies and mesoscale motions situations. Other class-separation criteria have been

proposed, still. Howell and Mahrt [1994b] also studied the scale dependence of surface

fluxes using the Haar wavelet transform to decompose the turbulent signals in 4 classes,

including the three cited before, and the 'fine scale' class, corresponding to very small scale

motions, with nearly isotropic characteristics.

The classes established by Sun et al. [1996] refer to oceanic tropical boundary layer

which in some situations, in the vicinity of large cloud cluster systems, do not show a clear

gap in the power spectra [Williams et al, 1996] and present many affinities with the ABL

over Amazon rain forest on wet-season [Garstang and Fitzjarrald, 1999]. According to these

authors, the tropical wet season ABL could be characterized as a disturbed state of the

boundary layer, which presents peculiar inter-scale links properties. It has qualitatively

different characteristics than that observed on undisturbed boundary layers and exhibits

peculiar tropospheric phenomena, such as outflows, among others, which play important

roles on vertical transports and on defining vertical exchange processes time-scales. Under

these conditions, is hard to characterize a superior border for the ABL, whose

thermodynamic characteristics are strictly associated to the nature of the convection on the

humid troposphere. It is important to mention that, although not hardly studied, the

exchanges of heat and moisture on the top of ABL also appear to influence the variability of

turbulent scalar variables measured at surface [Mahrt, 1991b; Holtslag and Moeng 1991;

Smedman et al., 1995; Wiliams et al., 1996; Garstang and Fitzjarrald, 1999]. Mahrt [1991b]

showed that, depending on the ABL moisture regime, different characteristic scales of

potential virtual temperature and humidity are present on some close to surface turbulent

processes.

7

Even disregarding the top of ABL influence, at least four physically distinct factors

may contribute to the variability of measurements at surface: (i) local circulation induced by

horizontal heterogeneity of the terrain [Mahrt and Ek, 1993; Mahrt et al., 1994, Mahrt et al.,

1998], which is expected to exist in Rebio Jaru site (this study) due the existence of "fish-

bone pattern" trips of alternating forested/deforested areas surrounding the reserve [Dias et

al., 2001] and also due to the wavy pattern of the vegetated crown top; (ii) contribution

from updrafts and downdrafts to surface fluxes [Wyngaard and Moeng, 1992], associated

with the very strong convective activity present in Amazonia [Garstang and Fitzjarrald,

1999]; (iii) contributions due to the existence of coherent structures on turbulent flow

[Mahrt and Gibson, 1992; Mahrt and Howell, 1994; Hogström and Bergström, 1996], what

presents specific characteristics over vegetated covers [Paw U et al., 1992; Gao and Li,

1993; Collineau and Brunet, 1993; Lu and Fitzjarrald, 1994; Raupach et al., 1996; Brunet

and Irvine, 2000]; (iv) the effects of slow wind direction and wind strength variation on

scalar transport processes near the ground [McNaughton and Laubach, 2000]. Such

variations are identified with the concept of "inactive" turbulence [Townsend, 1976].

These factors can drive important contributions to the total fluxes by low frequency

motions, as reported by Sakai [personal communication, 2000]. The author found for a

summer-time signal, obtained above a midlatitude deciduos forest, that large eddies

presenting periods ranging from 4 to 30 minutes might contribute with about 17 % to total

surface fluxes of heat, water vapor and CO2. However, it is likely that these flux fractions

would not be taken into account if they were calculated by the current popular averaging-

periods procedures. This is because these usual sampling time intervals are too short to

resolve the larger eddies present in the flow [Mahrt, 1998]. Sakai [2000] has also found that

8

short-averaging periods might underestimate daytime CO2 fluxes at standard towers by 10-

40 %, depending on wind speed conditions. Most of research groups from the flux

measurement community use frequency response corrections suggested by Moore [1986],

and slightly corrected by Moncrieff et al. [1987]. However, Culf et al. [2001] highlight that

this correction scheme might generate large unrealistic low-frequency corrections under

certain circumstances (for instance, under unstable conditions and low wind speed).

In this work we use turbulent data measured at Amazon rain forest during LBA

1999 wet-season campaign. The measuring heights, 62m and 67 m, are likely to be in a

surface transition sub-layer. During daytime conditions, low-frequency contributions to

eddy-correlation turbulent fluxes were often present. We apply wavelet transform to project

the data on scales and calculate by scale covariances in a similar way to the one used by

Howell and Mahrt [1994b] and Katul and Parlange [1994]. To accomplish such

calculations, we use the Daubechies-8 wavelet transform [Daubechies, 1992] to decompose

turbulent signals of wind velocity (u,v,w), temperature (T), humidity (q) and CO2

concentration (c). In spite of some restrictions concerning the feasibility of its applications

[Treviño and Andreas, 1996], wavelet analysis is a powerful tool to analyze turbulent

signals [Farge, 1992; Katul and Parlange, 1994; Abry, 1997]. Using the scale projected

signals, we then calculate the fluxes partitioned into turbulent, large-eddy and mesoscale

classes.

2. Site description and instruments deployed

9

During the 1999 Wet Season in Amazonia, several activities of the LBA Project

(Large Scale Biosphere-Atmosphere Experiment in Amazonia) took place at the biological

reserve of Jaru (Rebio Jaru), located about 100 km North of Ji-Paraná, Rondonia, Brazil

[Dias et al., 2001]. Rebio Jaru is a terra firme forest reserve owned by the Brazilian

Environmental Protection Agency (IBAMA). As reported by Culf et al. [1996], in the

predominant wind direction (sector from the north west clockwise to south-south east), the

fetch condition is mainly of an undisturbed forest for tens of kilometers. However, in the

other directions, it is much less (about 800 m), where the Ji-Paraná river forms the western

boundary of the reserve. On the other side of the river the rain forest has been progressively

cleared during the last 25 years, following a very peculiar fish-bone pattern, which

alternates patches of forest and of degraded land are observed [Andreae et al., 2001].

The Rebio Jaru reserve canopy has a mean height of 35m; however, some of the

higher tree branches have heights up to 45m. Details of vegetation at this site are given by

McWilliam et al. [1996]. At the end of 1998, within the collaboration of LBA/AMC and

LBA/EUSTACH projects, a new 60 m tall micrometeorological tower was built at this site

(10o 4.706' S; 61o 56.027' W, at the height of 145 m A.S.L.). Details of instrumentation and

measurements at this tower are given by Dias et al. [2001] and Andreae et al. [2001].

Two datasets were used on this work. The first one (dataset A) is composed by

measurements made between days 93 (Apr. 3rd) and 99 (Apr. 9th) of 1999, with a 3-D sonic

anemometer (Solent A1012R, Gill Instruments), together with a infrared gas analyzer (LI-

6262, LICOR Inc.), both recording data at a sampling rate of 10.4 Hz.. , This data were

obtained at the height of 62.7 m, and is part of long term measurements of surface fluxes

10

supported by the LBA/EUSTACH project (Andreae et al., 2001). The second dataset (B) is

composed by turbulent data collected from day 31 (Jan. 31st) to 60 (Mar. 1st) of 1999,

during WETAMC campaign (Dias et al, 2001), with a different 3-D sonic anemometer

installed at the height of 67 m above the forest floor (CSAT3, Campbell Scientific Inc.), at

a sampling rate of 16 Hz. The sonic anemometers measure the three wind components

(u,v,w) and virtual air temperature (Tv), and the IRGA measures the concentration of water

vapor (q) and CO2 (c) on the air. Quality control was done using the QC pack software of

Vickers and Mahrt (1997). Records with bad data points, instrument dropouts, poor

resolution and abrupt changes are flagged and then examined visually. Bad data records

were not used in the analysis.

3. Methodology

a. Determination of power spectra and cospectra low-frequency ends of the turbulent

variables

The spectral gap, usually appears on power spectra of turbulent variables measured

at middle and high latitude sites over uniform terrain, and provide the necessary

information to a good choice of the sampling segment size (time of duration of

measurements to flux determination). This also supply information on the temporal scale

where the variables should be divided into mean and fluctuation parts [Stull, 1988].

Although is broadly accepted that the spectral gap exists at least on a statistical sense on

11

temporal scales close to 1 hour [Lumley and Panofsky, 1964; Stull, 1988], there are

situations when is hard to observe a clear gap, on intervals up to hundred of minutes [Sun et

al., 1996; Williams et al., 1996; Mahrt, 1998; Mc Naughton and Laubach, 2000]. Among

the several factors that may determine a non clear existence of the spectral gap, could be

highlighted: (a) the modulation of the turbulent fluxes by mesoscale motions [Sun et al,

1996; Williams et al, 1996]; (b) the existence of complex and horizontally heterogeneous

surfaces [Mahrt et al., 1994; McNaughton and Laubach, 2000]; (c) the existence of

mesoscale transient motions [Mahrt et al., 1994]; and (d) multiscale processes with a strong

non-linear interaction between the motions on several scales, what is typical of the humid

tropical boundary layer [Williams et al., 1996; Garstang and Fitzjarrald, 1999].

Most of the power spectra observed over Rebio Jaru during the wet season with data

measured during morning and afternoon times show variance spread over a wide range of

frequencies, without any obvious spectral gaps, even when the sampling sizes were

increased to several hundreds of minutes. This difficulty in determining a clear cutoff

frequency motivated the authors to use the methodology based on wavelet transforms, as

suggested by Howell and Mahrt [1994b]. Wavelet transforms allow the decomposition of

turbulent scalar fields into space and scale, and information related to the motions on a

particular scale can be extracted, even the spectra of signal of different scales. The Fourier

energy spectrum [Kaimal and Finnigan, 1994] has been one of the most popular techniques

for analysis of signals. Indeed, many of the traditional methods work in the Fourier space

for most of the time.

The Fourier energy spectrum E(k) of the real function f(x) is defined by

12

E(k) = | f(k) |2 for k ≥ 0 ( 1 )

where f(k) signifies Fourier transform.

The wavelet transform extends the concept of energy spectrum so that one can

define a local energy spectrum E(k, x) using the L2 norm wavelet [Farge, 1992]:

E(x, k) = C | F(x, k0/k) |2 for k ≥ 0 ( 2 )

where k0 is the peak wave number of the analysing wavelet, and C is a constant.

By measuring E(x, k) at different places in a flow field one can determine what parts

of the field contribute most to the overall Fourier energy spectrum and how the energy

spectrum depends on local conditions. For example, one can determine the type of energy

spectrum contributed by coherent structures, such as isolated vortices, and the type of

energy spectrum contributed by the unorganized part of the flow [Farge, 1992; Gao and Li,

1993].

Since the wavelet transform analyses the signal into wavelets rather than sine waves,

it is possible that the mean wavelet energy spectrum may not always have the same slope as

the Fourier energy spectrum. Perrier et al. [1995] have demonstrated the conditions on

which is possible to obtain spectral power law estimation by means of WT.

Howell and Mahrt [1994b] used also wavelet transform to develop an adaptive

method to divide the high frequency turbulent motions into turbulent and fine scale classes.

However, as our main goal is this study is to assess the low frequency contributions to the

13

total fluxes, we did not separate this fine scale class using Howell and Mahrt [1994b]

methodology.

b. Wavelet decomposition of turbulent data

The wavelet transform (WT) is a powerful analysis tool, that permits an

evolutionary spectral study of turbulent atmospheric signals [Daubechies, 1992; Farge,

1992]. WT is similar to, but an extension of Fourier analysis. WT is computationally

similar in principle to Fast Fourier Transform (FFT). The FFT uses cosines, sines and

exponentials to represent a signal, and is most useful for representing stationary functions.

Since many 1-D and 2-D signals display non-linear, chaotic, intermittent or fractal

behaviour, Fourier analysis is less suitable for analysing such signals. Wavelets offer a more

adequate method to analyse complex signals. The aim of the WT is to 'express' an input

signal into a series of coefficients of specified energy. The discrete numbers associated with

each coefficient contain all the information needed to completely describe the signal,

provided one knows which analysing wavelet was used for the decomposition. The WT

partitions a signal with respect to spatial frequency. This is achieved by filtering the signal

with a pair of dyadic orthogonal filters called a quadrature mirror filter (QMF). A QMF

comes in a pair: termed a 'father' wavelet and a 'mother' wavelet. The father wavelet

provides an approximate or blurred version of the signal at sucessive resolutions, while the

mother wavelet captures the detail at each resolution. In the wavelet transform certain

coefficients are large while others are negligible, and practically zero. This property can be

14

used to separate out the required information by suppressing out the coefficients close to

zero.

In this work we use the Daubechies-8 WT to project the turbulent signals on 16

scales. This WT is a discrete orthogonal wavelet and the cospectra based on this type of

wavelet can be interpreted as fluxes decomposed into values computed from moving

averages [Howell and Mahrt, 1997]. For analysis of turbulence measurements, discrete WT

are preferable since it is suitable to provide non redundant decomposition information and it

permits to obtain a inverse WT. Howell and Mahrt [1994b] also show that cospectra based

on the orthogonal Haar (multiresolution) decomposition are directly linked to varying the

Reynolds averaging length procedure.

The turbulent signals of horizontal (u) and vertical (w) wind velocity components,

virtual air temperature (Tv), specific humidity (q) and CO2 concentration (c) were divided

into overlapping intervals of 4 hours, and each of these intervals was decomposed into 16

scales (15 scales, in case of dataset B), using the wavelets subroutines of the software

package Matlab (The MathWorks, Inc.). The package performs the wavelet decomposition

of a turbulent signal similar to the methodology described by Katul and Parlange [1994].

The choice of the Daubechies-8 WT was, to some extent, arbitrary. This is because various

wavelets were tested, but the differences between scale variances or covariances calculated

with several wavelets were very small, result similar to the one obtained by Katul and

Parlange [1994], for turbulent data measured at atmospheric surface layer under unstable

and stable stability conditions.

15

3. Results and discussion

One of the main goals of this work is to investigate the scale contributions of

turbulent fluctuations to variances and covariances calculations. As was stressed by

Lenschow and Stankov [1986] an important consideration in observational studies of the

ABL is the length or time over which, on the average, a variable maintains some degree of

correlation with itself or with other physically significant variables. For fixed-point tower

measurements, accepting the Taylor's "frozen turbulence" assumption (Stull, 1988) a time

scale is specified in such a way that it can be related to a length scale by multiplying the

mean wind speed by the time scale. But to obtain flux measurements in tower-based

measurements is not an easy task. Mahrt [1998] has carried out a review of the main

problems which introduce errors in ABL turbulent flux measurements. One of these

problems is nonstationarity. As was discussed by him, nonstationarity of surface fluxes is

caused by diurnal trend, mesoscale motions, and passage of clouds. All these problems

probably are present in the tropical wet-season ABL [Garstang and Fitzjarrald, 1999].

Nonstationary mesoscale motions or mesoscale motions associated with heterogeneous

surfaces modulate the turbulent flux and sometimes lead to computed flux on scales larger

than turbulent motions [Mahrt et al., 1994; Mahrt, 1998]. After a variance analysis

performed by Mahrt et al. [1994] the momentum and ozone fluxes are more influenced by

transient mesoscale motions while fluxes of heat, moisture and carbon dioxide are more

influenced by surface heterogeneity. Updrafts and downdrafts sequence events sometimes

associated to well-organized convective cloud systems or to individual eddies motion also

contribute to introduce scale dependence on ABL fluxes [Sun et al., 1996]. As a tool for

16

displaying the scale dependence, we have computed the variances and covariances by

means of wavelet transform (WT) [Farge, 1992; Abry, 1997]. WT is suited to be a more

adequate tool to scale project nonstationary turbulent signals which present irregularly

spaced sharp transitions [Farge, 1992; Gao and Li, 1993; Collineau and Brunet, 1993;

Mahrt et al., 1994].

In order to project transient signals on selected scales, we apply the Daubechies-8

Wavelet transform to the turbulence time series of dataset A (B), divided on overlapping

records of 131072 (262144) data points, that correspond to approximately 3.5 (4.5) hours,

recovered from the datasets following a time step of one hour. These records lengths were

chosen to provide better analyses of low frequency motions, and because under diurnal

conditions, turbulent signals seldom presented spectra or cospectra low-frequency end at

frequencies higher than the ones corresponding to time scales of 2 to 3 hours. Variances and

covariances of the variables were then calculated for each projected time scale. Scale

variances represent the energy of the turbulent scalars (or wind component) contained on

each specific scale of variability [Gao and Li, 1993, Katul and Parlange, 1994].

On the following section we show the contribution of each scale to the total

variances and fluxes and assess the partition of these fluxes between turbulent, large eddies

and mesoscale classes, as proposed by Sun et al [1996]. In order to obtain useful

information about the scaling properties of the turbulent signals on the various selected

scales on which they have been projected, we also performed scale calculations of the

correlation coefficients between temperature (T) and specific humidity (q). This is because

De Bruin et al. [1999], based on the conclusions of Hill [1989] have shown that if the

correlation coefficient between T and q (r(Tq)) is equal to one, temperature and specific

17

humidity both will obey the Monin-Obukhov Similarity Theory (MOST). Further, under

such conditions, it is possible to determine the Bowen ratio from r(Tq). As examples of

situations in which r(Tq) ≠ 1, De Bruin et al. [1999] have mentioned: conditions of local

advection and conditions in which the entrainment of dry air at the top of the ABL is

stronger than surface evaporation, when large eddies are able to transport drier air from

ABL top towards the surface [Mahrt, 1991b]. In these situations, the MOST is expected to

be violated.

Scale variances and covariances partitioning

Once the wavelet coefficients are available, one can assess the energy contained in

each scale (scale variance). With this information, it is possible to investigate in which eddy

motion classes are partitioned the contributions from different physical processes to total

fluxes. One starting point to perform this analysis is by simply calculating the variances and

covariances of these coefficients and verifying whether they agree or not with some

classification proposed. In this work, we will analyse the feasibility of applying the

classification proposed by Sun et al. [1996] to explain our results. This is carried out in

Figure 2. It shows the scale variances of w, Tv, q and c, averaged during diurnal periods

(9:00 to 17:00 hs), from day 93 to 98. The scales are presented as spatial scales of eddies,

which were estimated assuming the Taylor’s hypothesis [Stull, 1988]. The vertical dashed

lines indicate the scales that divide the three classes of motion:

Turbulent scales (TS) – main scales of vertical turbulent transport of mass and energy in the

atmospheric surface layer;

18

Large eddies (LE) – motions involving eddies of scales on the order of the height of the

ABL;

Mesoscale (MS) – motions on horizontal scales much larger than the height of the ABL,

involving processes not expected to be controlled only by ABL parameters.

Nevertheless, based only on the information provided by Figure 2, it is not easy to

assess specific scales that should separate the above classes. To have some more insight

about the end of each one of these classes, we depict also informations regarding: (i) the

scale covariance values, as a function of the scale itself, which are useful to better

understand the regions in which some properties of the eddy motions are present (Figure 3);

(ii) the scale correlation coefficients between Tv and q along a large range of scales (Figure

4); (iii) the mean values of H, λE, FCO2, var(w), var(Tv), var(q), and H/ λE, which are

presented in Table 1.

Figure 2 shows a broad peak for var(w) values ranging from scales around 50m to

500m (the peak corresponds to a 4.0min time-scale), while scalar variances grow up slowly

continuously in this region (without any local maximum). Figure 3 shows, still, broader

peaks than the former var(w) one, for scalar cospectrum peaks. Their local

maximum/minimum values are also inside the interval located between 50m to 500m: while

scalar H and λE show maximum values, FCO2 presentes a minimum value there. On scales

higher than 800 m, we see a continuous drop on <w'T>' and <w'q'> scale values and a

increase on <w'c'> scale values (Figure 3), as well as a continuous drop on the scale

correlation coefficient between Tv and q (Figure 4). Due to this factors, the scales around

300m, which corresponds to temporal scales around 3 minutes, present the more energetic

19

surface layer eddies and are associated with physical events whose role in tropical wet ABL

is essential to determine some important features of the structure of turbulence there.

Hence, this scale actually separetes two different classes of turbulent eddy patterns, as

discussed by Williams et al. [1996], and not three, as proposed by Sun et al. [1996] for the

tropical marine boundary layer. So, it appears that all the scales greater than the ones

associated with the spectral and cospectral peaks present similar physical characteristics.

However, the variances of Tv, q, and c increase considerably on scales higher than 3 km

(figure 2) and, despite the facts discussed above, the low frequency motions seems to

behave slightly different on scales ranging from 800 m to 3 km, and higher than 3 km (see

figure 5 discussed later, where we can see that these two classes present opposite signals).

Due to this fact, we choose the 3 km scale to separate the large eddy and mesoscale classes.

However, this and other criteria to split these low frequency scales in two different classes

should be investigated later.

It is interesting to note that most of the variance of w occur on turbulent scales (TS),

what explains the fact that most of the mass and energy exchange happen on these scales. It

shows a maximum on scales close to 300m, similar to results obtained by Howell and

Mahrt [1994b]. Due to the fact that large-scale motions are predominantly horizontal at

levels close to the surface, variances of w are low at LE and MS classes. McNaughton and

Laubach (2000) reported, however, that although w-spectrum is well scaled by Monin-

Obukhov Similarity, it still might present enhancement at low frequencies under disturbed

conditions, influenced by outside ABL parameters, such as updrafts and downdrafts.

Variances of Tv, q and c show the higher values at mesoscale, but is important to notice

that q still presents pronounced energy on turbulent processes, compared to higher scales.

20

The presence of the maximum variance at low frequency motions (except for w, which is

surface limited, as explained above) probably indicate the high influence of mesoscale

motions and convective systems that act on amazonian ABL. Several authors have also

reported high variability of turbulent scalars at low frequencies, such as Wiliams et al

[1996], Sun et al. [1996] and Réchou and Durand [1997].

In spite of the fact that the scalar variance curves present similar shapes, the var(q)

curve grows more quickly inside the TS range comparatively to the var(T) and var(c)

curves, which present enhancement in their scale variation rate at the separation border

between TS and LE regions, result which seems to be related with the concept of 'scale

surface heterogeneity' [Mahrt, 1996]. As discussed by Stull [1988], for 'classical' situations,

there are experimental and modeling evidences that var(T) is expected to be greater than

var(q) at the lower levels of the ABL. This is because close to the surface, there would be a

more intensive variation of sources and sinks of sensible heat, which generate more

intensive fluctuations of T in this region comparatively to humidity field. But our results do

not present such a difference. One first possible explanation for these results is that the

measurements were performed below the 'blending height', loosely defined as the level

above which the flow becomes horizontally homogeneous in the absence of other

influences [Mahrt, 1996]. One other possible explanation for these results is that in a rain

forest as the Rebio-Jaru reserve, the scale thermal and humidity heterogeneity strength

patterns are not the same. Besides, the former grows more slowly than the last when the

spatial scale increases. As stressed by McNaughton and Laubach [2000], 'advected variance'

has little effect on this kind of results, and since all scalars over the experimental field are

transported by the same velocity field, the main differences in scalar spectra must have had

21

their origin in differences in scalar sources at the ground. At the lower frequency range we

observe a clear variation on the values of d[var(S)]/ds (where S is a generic scalar and s, is

the spatial scale), except for d[var(q)]/ds.

Figure 3 shows the scale dependence of fluxes of sensible and latent heat, and CO2,

averaged for the same period as in figure 2. On the smallest scales, that should involve

mainly isotropic motions [Howell and Mahrt, 1994], the fluxes are very small, but different

from zero. The fact that the fluxes are not always null in the inertial subrange, raises some

interesting questions discussed by Katul et al. [1997] concerning the inertial subrange eddy

motion contamination by the larger scale organized eddy motion. As they have stressed,

irrespective of the anisotropy source, if interactions between large scales and small scales

are significant, then it is possible that the anisotropy of the large scale persists in the inertial

subrange. In such situations, basic atmospheric surface layer assumptions as Kolmogorov's

hypotheses could be affected. From scale of 10m towards higher scales, the fluxes increase

significantly, characterizing well the high transports on the region of turbulent scales. Both

sensible and latent heat fluxes, as well as CO2 flux, reach a maximum at scales close to 400

m, corresponding to time scales of approximately 3 minutes, and then decrease to small

values at higher scales. It is interesting to guess about the physical origins for these 3min

maximum-energy eddies in the wet tropical boundary-layer. As was discussed by Gartang

and Fitzjarrald [1999, pg.16], horizontal gradients in temperature and humidity tend to

disappear in the lower tropical atmosphere. As they have stressed, in this context locally

produced horizontal gradients in temperature, moisture, and pressure are product of

perturbations of the velocity field and not a result of pre-existing gradients in those

quantities. Thus, we might expect that these local perturbations of the velocity field are the

22

main source for generation of turbulent kinectic energy (TKE) in the atmospheric surface

layer of tropical regions. Gu et al. [2001] have recently presented evidences of cloud

modulation of solar irradiance in a Amazonia pasture (located not far from our

experimental site) associated with cloud gap patterns which seems to be responsible for a

convective regime and a turbulence diffusion regime whose long time period fluctuations

are of the order of 3 min, the same time-scale order as we have obtained in our results.

Hence, we suggest that these cloud gap effects represent the main source of w-TKE in the

Amazon forest wet boundary layer and they could explain the physical origin of our 3min

energy-peaks.

Although the largest amount of the energy and mass fluxes occur on turbulent scales

below or of the order of the w-spectral peak, larger scale eddy motions could generate

important contributions to the total fluxes. In addition, according to our results, these low-

frequency contributions to surface fluxes also show large variation among the various

investigated data records, as we can see on Figure 3. This sohows up by the estimated large

sampling errors.

Two explanations for these results could be proposed: (i) one presented by

McNaughton and Laubach [1998, 2000] in their investigation about the consequences of the

unsteadiness of the wind field on the scalar fields. After them, the surface fluxes of

temperature and humidity do indeed vary in step with low-frequency variations in the wind

under certain non homogeneous surface conditions. By means of a theoretical treatment,

they showed that this would affect the values of the eddy-diffusivities for temperature and

humidity in different ways. It is interesting to note that this kind of low-frequency wind

velocity variability and that dissimilarity between T and q fluctuations were observed in

23

Rebio-Jaru turbulent data, too. It is important also to keep in mind that our experimental

site is in a rain forest strip which is surrounded by deforested areas in a very peculiar "fish-

bone" pattern. Such a non homogeneous boundary condition could generate mesoscale

circulations as the discussed by Sun et al. [1996] and Mahrt and Ek [1993], that would

explain our low-frequency scalar fluxes. (ii) one second explanation concerns, not the

heterogeneity surface conditions, but the peculiar complex characteristics of the 'disturbed

state' wet tropical ABL. According to Garstang and Fitzjarrald [1999, pg. 13], when deep

convective clouds are present in a moist convective tropical boundary layer, the links to the

surface in the vertical column of the atmosphere are complex. Besides, mean upward

motions may be present over a substantial area occupied by these systems, which

concentrate the vertical transport of heat, trace gases, and matter there. In such situations,

there is a failure of the familiar concept of planetary boundary layer, in which there is a

simple connection between the surface and some shallow horizontally homogeneous layer.

This is because the deep clouds operate as very effective mixers, leaving the atmosphere

much more nearly neutrally stratified than in the undisturbed ABL familiar state [Garstang

and Fitzjarrald, 1999, pg. 285]. We suggest that this peculiar behaviour of the wet tropical

boundary layer is responsible for our results concerning the low-frequency features of the

scalar covariances.

To provide quantitative information about the amount of flux contained in each

turbulence-pattern class earlier defined, we present Table 1. In this table we present the

mean by-class and total values of sensible heat flux (H, in W/m2), latent heat flux (λE, in

W/m2), the bowen ratio (H/λE), CO2 flux (FCO2, in µmol/m2/s), variance of w (in m2/s2),

variance of Tv (in K2), variance of q (in g2/kg2), and variance of c (in µmol2/mol) averaged

24

during diurnal periods, from day 93 to 98. These calculations do not provide a true

climatology of fluxes, which would require averaging over a much larger dataset including

a wider range of conditions, but provide a simple summary of the scale dependence of

fluxes over Rebio Jaru forest. Our purpose here is to warn about the numerous effects that

can influence the exchange processes on this disturbed ABL, as described by Garstang and

Fitzjarrald (1999).

Starting from these results, some interesting conclusions are obtained: (i) About

24.5% of sensible heat and 21 % of latent heat is transported by large eddies. Mesoscale

motions are responsible for approximately 5 % of both sensible and latent heat fluxes. For

CO2 fluxes, the results are similar for large eddies, but quite higher for mesoscale

processes: about 24.5% occur on large eddies scales and 11 % on mesoscale. (ii) The scale

H/λE ratio value remains almost constant around 0.4, for each one of the investigated

classes. This reflects the fact that our measurements were carried out in a tropical rain forest

ABL, during the wet season. Therefore, a Bowen ratio value of 0.4 expresses the fact that

most of the available energy in the forest environment is being used to generate water vapor

transport from the canopy to the atmosphere. (iii) The w-variance is almost totally

contained in the TS class. This is not a surprising result. As it has been demonstrated by

authors as McNaughton and Laubach [2000], the w-power spectra are associated mainly

with inner-layer scaling regimes and so, are insensitive to large-scale eddy motions which

are predominantly horizontal. After them, such large-scale motions are presumably driven

by outer-layer scaling regime motions. This would explain the fact that our data

contributions for var(Tv) and var(c) enhance with the increasing of length-scale motion. But

this was not true for var(q) whose values integrated over turbulent scales are even higher

25

than on MS class. These results show that the turbulent fluctuations pattern of Tv and q are

not similar. For this reason we present figure 4 in which scale correlation coefficients r(Tq)

are related with the scale on which each signal was projected. Observing these results, it is

clear that r(Tq) < 1 for all analyzed scales: r(Tq) = 0 at the length scales of the order of 1 m

(corresponding to the inertial subrange) and increases until reaches the 100m length scale

where r(Tq) presents a maximum value of 0.6. For the length scales greater than 100 m the

r(Tq) value curve falls down until it crosses the zero-axis at a length scale of the order of 5

km and becomes negative. Hence, r(Tq) in the LE class region is still positive but ranges

between 0.4 and 0.1. In the MS class, r(Tq) values range between 0.1 and -0.2. Starting

from these results and based on Hill [1989] and De Bruin et al. [1999] conclusions, one first

finding is that MOST does not hold for all investigated scale-ranges. We could formulate

various tentative explanations for the above results: (i) For the smaller scales,

corresponding to the transition subrange layer, we actually expect that the MOST is not

useful to fully explain the turbulence structure there [Thom et al., 1975; Garratt, 1980;

1983; Thom and Raupach, 1981?; Fitzjarrald et al., 1990]. So, it is not surprising to obtain

r(Tq) ≠1 in such region. (ii) For the scales ranging between 10m and 100m, r(Tq)>0.5, but

is lower than 1. Therefore, we expected to be in a disturbed surface layer context. Starting

from the verification that there are topographically-induced eddying as well as

convectivelly-induced eddying over the Rebio-Jaru experimental site, it would be useful for

us to apply McNaughton and Laubach [2000] arguments to explain the breakdown of

MOST in this region. As they have shown, such situations could introduce low-frequency

wind speed fluctuations which generate dissimilarities between the eddy diffusivities for

temperature and humidity. (iii) As stressed by Mahrt (1991b) different boundary layer

26

moisture regimes may present different patterns of similarity or dissimilarity between T and

q fluctuations. He has shown that in some peculiar wet boundary layers, the T and q

fluctuations are not similar, at least in part, because some dry air patches are transported

from the boundary layer top to the surface. This could explain our results in a very disturbed

tropical boundary layer.

In Figure 5 we present the ratio of scale wT covariance contributions of each eddy-

pattern class to total sensible heat flux as a function of horizontal wind speed. It is

interesting to observ that the turbulent eddy (TS) contribution becomes more important to

total sensible heat flux as mean horizontal wind speed enhances. For horizontal mean

speeds greater than 1.5 m/s, the large eddy (LE) w'T' contributions for heat fluxes shows a

constant decrease, and the mesoscale (MS) are even negative.This seems to confirm one our

guess that LE contributions to the total heat flux are associated with processes induced by

deep convective clouds such as updrafts, and MS processes are more associated with

macrobursts, large downbursts with 4km or larger outflow diameter, with damaging winds

lasting 5 to 20 min [Garstang and Fitzjarrald, 1999, pg.303]. The discussions of Mahrt

[1998] might support other possible arguments to this results: (i) with weak large-scale flow

and significant surface heating, the velocity fluctuations may more closely approach that of

pure updrafts and downdrafts and the special flow distorcion effects under these situations

enhance low frequency motions; (ii) the existence of stationary eddies, that might be

attached to the surface heterogeneity and might be slow moving with weak winds. These

stationary eddies can also be important factors that conttribute to low frequency transport.

27

4. Summary and conclusions

High frequency measurements (10.4 Hz, or 16 Hz) of vertical wind velocity (w),

virtual temperature (Tv), specific humidity (q) and CO2 concentration (c) of air, obtained

over Rebio Jaru tropical rain forest reservation, in south west Amazonia, were projected

into 15 scales using the Daubechies-8 wavelet transform (WT). The relative contributions

of each scale to the total variances and covariances were then assessed, and, based on the

general characteristics of the scale-dependence of fluxes, they were partitioned into three

classes:

Turbulent scales (TS) – main scales of vertical turbulent transport of mass and energy in the

atmospheric surface layer, with range from very small scales with nearly isotropic

motions to scales on the order of 800 m (or 6.5 minutes);

Large eddies (LE) – motions involving eddies of scales on the order of the height of the

ABL, ranging from 800 m to 3 km (corresponding to time scales of approximately 6.5

and 25 minutes, respectively);

Mesoscale (MS) – motions on horizontal scales much larger than the height of the ABL,

involving processes not expected to be controlled only by ABL parameters. These scales

are on the order of 3 km or higher.

The variance of w shows most of turbulent kinetic energy occurring on turbulent

scales (TS) reaching a maximum on scales close to 300m and decreasing to very low values

on higher scales. Due to the fact that large-scale motions are predominantly horizontal at

levels close to the surface, variances of w are low at LE and MS classes.Variances of Tv, q

28

and c show the higher values at mesoscale, indicating a likely high influence of mesoscale

motions and convective systems that act on amazonian ABL on these data.

The largest amount of the sensible heat, latent heat and CO2 fluxes occur on

turbulent length scales below or of the order of the w-spectral peak scale. Larger scale eddy

motions could, however, generate important contributions to the total fluxes. In addition,

these low-frequency contributions to surface fluxes show large variation among the various

investigated data records, and can be either positive or negative irrespectively of mean ABL

gradient conditions. About 24.5% of sensible heat and 21 % of latent heat averaged during

diurnal periods from day 93 to 98 were transported by large eddies. Mesoscale motions

during the same period were responsible for approximately 5 % of both sensible and latent

heat fluxes. For CO2 fluxes, the results are similar for large eddies, but quite higher for

mesoscale processes: about 24.5% occur on large eddies scales and 11 % on mesoscale.

These calculations, however, do not provide a true climatology of fluxes, which would

require averaging over a much larger dataset including a wider range of conditions.

We also performed scale calculations of the correlation coefficients, r(Tq), between

temperature (T) and specific humidity (q). The results show that the turbulent fluctuations

pattern of T and q are not similar, and r(Tq) < 1 for all analysed scales.

29

5. Acknowledgments

This work is part of The Large Scale Biosphere-Atmosphere Experiment in Amazonia

(LBA) and was supported by the Fundação do Amparo à Pesquisa do Estado de São Paulo

(FAPESP)/Brazil-process 1997/9926-9. Thanks are due to Dr. Maria Assunção Faus da

Silva Dias who is the coordinator of this research project. The authors are also grateful to

INCRA/Ji-Paraná and to IBAMA/Ji-Paraná. C. von Randow, P. Gannabathula are

supported by a fellowship from Conselho Nacional de Pesquisa e Desenvolvimento

Tecnológico (CNPq).

6. References

Abry, P., Ondelettes et Turbulences, Diderot-Editeur, 292 pp., Paris, 1997.

Affre, C., A. Lopez, A. Carrara, A. Druilhet and J. Fontan, The analysis of energy and

ozone flux data from the LANDES 94 experiment, Atmospheric Environment, 32, 803-821,

2000.

Andreae, M. O. et al., Towards an understanding of the biogeochemical cycling of carbon,

water, energy, trace gases and aerosols in Amazonia: An overview of the LBA-EUSTACH

experiments, J. Geophys. Res., this issue, 2001.

30

Brunet, Y., and M. R. Irvine, The Control of Coherent Eddies in Vegetation Canopies:

Streamwise Structure Spacing, Canopy Shear Scale and Atmospheric Stability, Boundary

Layer Meteorol., 94 (1), 139-163, 2000.

Brutsaert, W., Land-surface water vapor and sensible heat flux: Spatial variability,

homogeneity, and measurement scales, Water Resources Res., 34 (10), 2433-2442, 1998.

Chimonas, G., Apparent Counter-gradient Heat Fluxes Generated by Atmospheric Waves,

Boundary Layer Meteorol., 31 (1), 1-12, 1985.

Collineau, S., and Y. Brunet, Detection of Turbulent Coherent Motions in a Forest Canopy.

Part II: Time-scales and Conditional Averages, Boundary Layer Meteorol., 66 (1-2), 49-73,

1993.

Culf, A. D., J. L. Esteves, A. O. Marques Filho and H. R. Rocha, Radiation, temperature

and humidity over forest and pasture in Amazonia, in Amazonian Deforestation and

Climate, edited by C. A. Nobre, J.H.C. Gash, J.M. Roberts and R.L. Victoria, pp. 175-191,

Wiley, Chichester, 1996.

Culf, A. D. et al., Errors arising from the correction of eddy correlation data under unstable

conditions, J. Geophys. Res., this issue, 2001.

Daubechies, I., Ten Lectures on Wavelets, 357 pp., SIAM, Philadelphia, 1992.

31

De Bruin, H. A. R., B. J. J. M. Van Den Hurk and L. J. M. Kroon, On the Temperature-

Humidity Correlation and Similarity, Boundary Layer Meteorol., 93, 453-468, 1999.

Desjardins, R. L., J. I. MacPherson, L. Mahrt, P. H. Schuepp, E. Pattey, H. Neumann, D.

Baldocchi, S. Wofsy, D. Fitzjarrald, H. McCaughey and D. W. Joiner, Scaling up flux

measurements for the boreal forest using aircraft-tower combinations, J. Geophys. Res., 102

(D24), 29125-29133, 1997.

Dias, M. A. F. S., S. Rutledge, P. Kabat, P. Silva Dias, C. Nobre, G. Fisch, H. Dolman, E.

Zipser, M. Garstang, A. Manzi, J. Fuentes, H. Rocha, J. Marengo, A. Plana-Fattori, L. D. A.

Sá, R. C. S. Avalá, M. Andreae, P. Artaxo, R. Gielow and L. Gatti, Clouds and rain

processes in a biosphere atmosphere interaction context in the Amazon Region, J. Geophys.

Res., this issue, 2001.

Farge, M., The Wavelet Transform and its Applications to Turbulence, Annu. Rev. Fluid

Mech., 24, 395-457, 1992.

Fitzjarrald, D. R., K. E. Moore, O. M. R. Cabral, J. Scolar, A. O. Manzi and L. D. A. Sá,

Daytime Turbulent Exchange Between the Amazon Forest and the Atmosphere, J. Geophys.

Res., 95 (D10), 16825-16838, 1990.

32

Gao, W., and B. L. Li, Wavelet Analysis of Coherent Structures at the Atmosphere-Forest

Interface, J. Appl. Meteorol., 32 (11), 1717-1725, 1993.

Garratt, J. R., Surface Influence upon Vertical Profiles in the Atmospheric near-surface

layer, Q. J. R. Meteorol. Soc., 106 (450), 803-819, 1980.

Garratt, J. R., Surface Influence upon Vertical Profiles in the Nocturnal Boundary Layer,

Boundary Layer Meteorol., 26 (1) 69-80, 1983.

Garstang, M., and D. R. Fitzjarrald, Observations of Surface to Atmosphere Interactions in

the Tropics, 405 pp., Oxford-University-Press, New York, 1999.

Gu, L., J. D. Fuentes, M. Garstang, J. Tota da Silva, R. Heitz, J. Sigler and H. H. Shugart,

Cloud Modulation of surface solar irradiance at a pasture site in southern Brazil, Agric.

Forest Meteorol., 106, 117-129, 2001.

Hildebrand, P. H., Error in Eddy Correlation Turbulence Measurements from Aircraft:

Application to HAPEX-MOBILHY, in Land Surface Evaporation - Measurement and

Parameterization, edited by T.J.Schmugge and J.-C. André., pp. 231-243, Springer-Verlag,

New York, 1991.

Hill, R. J., Implications of Monin-Obukhov Similarity Theory for Scalar Quantities, J.

Atmos. Sci., 46 (14), 2236-2244, 1989.

33

Holtslag, A. A. M., and C.-H. Moeng, Eddy Diffusivity and Countergradient Transport in

the Convective Atmospheric Boundary Layer, J. Atmos. Sci., 48 (14), 1690-1698, 1991.

Howell J. F., and L. Mahrt, An Adaptative Multiresolution Data Filter: Applications to

Turbulence and Climatic Time Series, J. Atmos. Sci., 51 (14), 2165-2178, 1994a.

Howell, J. F., and L. Mahrt, An Adaptative Decomposition: Application to Turbulence, in

Wavelets in Geophysics, edited by P. Kumar and Efi Foufoula-Georgiou, pp. 107-128,

Academic-Press, San Diego, 1994b.

Howell J. F., and L. Mahrt, Multiresolution Flux Decomposition, Boundary Layer

Meteorol., 83, 117-137, 1997.

Högström, U., Analysis of Turbulence Structure in the Surface Layer with a Modified

Similarity Formulation for Near Neutral Conditions, J. Atmos. Sci., 47 (16), 1949-1972,

1990.

Högström, U., and H. Bergström, Organised Turbulence in the Near-Neutral Atmospheric

Surface Layer, J. Atmos. Sci., 53 (17), 2452-2464, 1996.

Kader, B. A., and A. M. Yaglom, Mean fields and fluctuation moments in unstable

stratified turbulent boundary layers, J. Fluid Mech., 212, 637-662, 1990.

34

Kaimal, J. C., S. F. Clifford and R. J. Lataitis, Effect of Finite Sampling on Atmospheric

Spectra, Boundary Layer Meteorol., 47 (1-4), 337-347, 1989.

Katul, G. G., and M. B. Parlange, On the Active Role of Temperature in Surface-Layer

Turbulence, J. Atmos. Sci., 51 (15), 2181-2195, 1994.

Katul, G., C.-I. Hsieh and J. Sigmon, Energy-Inertial Scale interactions for Velocity and

Temperature in the Unstable Atmospheric Surface Layer, Boundary Layer Meteorol., 82

(1), 49-80, 1997.

Lenschow, D. H., and B. B. Stankov, Length scales in the convective boundary layer, J.

Atmos. Sci., 43, 1198-1209, 1986.

Lu, C.-H., and D. R. Fitzjarrald, Seasonal and Diurnal Variations of Coherent Structures

over a Deciduous Forest, Boundary Layer Meteorol., 69 (1-2), 43-69, 1994.

Lumley, J. L., and H. A. Panofsky, The Structure of Atmospheric Turbulence, 239 pp.,

Wiley, New York, 1964.

Mahrt, L., Intermittency of Atmospheric Turbulence, J. Atmos. Sci., 46 (1), 79-95, 1989.

35

Mahrt, L., Heat and Moisture Fluxes over the Pine Forest in HAPEX, in Land Surface

Evaporation - Measurement and Parameterization, edited by T. J. Schmugge, pp. 261-273,

Springer-Verlag, New York, 1991a.

Mahrt, L., Boundary-layer moisture regimes, Q. J. R. Meteorol. Soc., 117 (A), 497, 151-

176, 1991b.

Mahrt, L., and W. Gibson, Flux Decomposition into Coherent Structures, Boundary Layer

Meteorol., 60 (1-2), 143-168, 1992.

Mahrt, L., and M. Ek, Spatial Variability of Turbulent Fluxes and Roughness Lengths in

HAPEX-MOBILHY, Boundary Layer Meteorol., 65 (4), 381-400, 1993.

Mahrt, L., J. I. Macpherson and R. Desjardins, Observations of Fluxes over Heterogeneous

Surfaces, Boundary Layer Meteorol., 67, 345-367, 1994.

Mahrt, L., and J. F. Howell, The influence of coherent structures and microfronts on scaling

laws using global and local transforms, J. Fluid Mech., 260, 247-270, 1994.

Mahrt, L., The Bulk Aerodynamic Formulation for Surface Heterogeneity, Boundary Layer

Meteor., 78, 87-119, 1996.

36

Mahrt, L., Flux Sampling Errors for Aircraft and Towers, J. Atmos. Oceanic Tech., 15, 416-

429, 1998.

Mahrt, L., D. Vickers, J. Edson, J. Sun, J. Hojstrup, J. Hare and J. M. Wilczak, Heat Flux in

the Coastal Zone, Boundary Layer Meteorol., 86 (3), 421-446, 1998.

Mann, J., and D. H. Lenschow, Errors in airborne flux measurements, J. Geophys. Res., 99

(D7), 14519-14526, 1994.

McNaughton, K.G., and J. Laubach, Unsteadiness as a Cause of Non-Equality of Eddy

Diffusivities for Heat and Vapour at the Base of an Advective Invertion, Boundary Layer

Meteorol., 88 (3), 479-504, 1998.

McNaughton, K. G., and J. Laubach, Power Spectra and Cospectra for Wind and Scalars in

a Disturbed Surface Layer at the Base of an Advective Inversion, Boundary Layer

Meteorol., 96, 143-185, 2000.

Moncrieff, J. B., J. M. Massheder, H. A. R. de Bruin, J. Elbers, T. Friborg, B. Huesinkveld,

P. Kabat, S. Scott, H. Soegaard and A. Verhoef, A System to Measure Surface Fluxes of

Momentum, Sensible Heat, Water Vapour and Carbon Dioxide, J. Hydrol., 188-189, 589-

611, 1997.

37

Moore, C. J., Frequency Response Corrections for Eddy Correlation Systems, Boundary

Layer Meteorol., 37, 17-35, 1986.

Nai-Ping, L., W. D. Neff and J. C. Kaimal, Wave and Turbulence Structure in a Disturbed

Nocturnal Inversion, Boundary Layer Meteorol., 26 (2), 141-155, 1983.

Pasquill, F., Atmospheric Diffusion, Wiley, New York, 1974.

Paw U, K. T., Y. Brunet, S. Collineau, R. H. Shaw, T. Maitani, J. Qiu and L. Hipps, On

coherent structures in turbulence above and within agricultural plant canopies, Agric.

Forest Meteorol., 61 (1-2), 55-68, 1992.

Perrier, V., T. Philipovitch, C. Basdevant, Wavelet spectra compared to fourier spectra, J.

Math. Phys., 36, 1506-1519, 1995.

Raupach, M. R., and A. S. Thom, Turbulence in and above Plant Canopies, Annu. Rev.

Fluid Mech., 13, 97-129, 1981.

Raupach, M. R., J. J. Finnigan and Y. Brunet, Coherent Eddies and Turbulence in

Vegetation Canopies: The Mixing-layer Analogy, Boundary Layer Meteorol., 78 (3-4), 351-

382, 1996.

38

Réchou, A., and P. Durand, Conditional Sampling and Scale Analysis of the Marine

Atmospheric Mixed Layer - Sofia Experiment, Boundary Layer Meteorol., 82, 81-104,

1997.

Shuttleworth, J. W., J. H. C. Gash, C. R. Lloyd, C. J. Moore, J. Roberts, A. O. Marques

Filho, G. F. Fisch, V. P. Silva Filho, M. N. G. Ribeiro, L. C. B. Molion, L. D. A. Sá, C. A.

Nobre, O. M. R. Cabral, S. R. Patel and J. C. Moraes, Eddy correlation measurements of

energy partition for Amazonian forest, Q. J. R. Meteorol. Soc., 110, 466: 1143-1162, 1984.

Smedman, A.-S., H. Bergström and U. Högström, Spectra, Variances and Length Scales in

a Marine Stable Boundary Layer Dominated by a Low Level Jet, Boundary Layer

Meteorol., 76 (3), 211-232, 1995.

Stull, R. B., An Introduction to Boundary Layer Meteorology, 666 pp., Kluwer Academic,

Dordrecht, 1988.

Sun, J., J. F. Howell, S. K. Esbensen, L. Mahrt, C. M. Greb, R. Grossman and M. A.

LeMone, Scale Dependence of Air-Sea Fluxes over the Western Equatorial Pacific, J.

Atmos. Sci., 53 (21), 2997-3012, 1996.

Tennekes, H., and J. C. Wyngaard, The Intermittent Small-Scale Structure of Turbulence:

Data-processing Hazards, J. Fluid Mech., 55, 93-103, 1972.

39

Thom, A. S., J. B. Stewart, H. R. Oliver and J. H. C. Gash, Comparison of aerodynamic and

energy budget estimates of fluxes over a pine forest, Q. J. R. Meteorol. Soc., 101, 93-105,

1975.

Townsend, A. A., The Structure of Turbulent Shear Flow, 429 pp., C. U. Press, Cambridge,

1976.

Treviño, G., and E. L. Andreas, On Wavelet Analysis of Nonstationary Turbulence,

Boundary Layer Meteorol., 81 (3-4), 271-288, 1996.

Vickers, D., and L. Mahrt, Quality Control and Flux Sampling Problems for Tower and

Aircraft Data, J. Atmos. Oceanic Tech., 14 (3, Part 1), 512-526, 1997.

Williams, A. G., H. Kraus and J. M. Hacker, Transport Processes in the Tropical Wram

Pool Boundary Layer. Part I: Spectral Composition of Fluxes, J. Atmos. Sci., 53 (8), 1187-

1202, 1996.

Wyngaard, J. C., Lectures on the Planetary Boundary Layer, in Mesoscale Meteorology -

Theory, Observations and Models, edited by D.Lilly and T.Gal-Chen, pp. 603-650, Reidel,

Hingham, 1983.

Wyngaard, J. C., Atmospheric Turbulence, Annu. Rev. Fluid Mech., 24, 205-233, 1992.

41

Table 1 Partition of the total surface flux contribution from each one of the three

turbulence-pattern classes.

T.S. L.E. M.S. Total

Sensible heat (H, in W/m2) 68.18 23.95 5.68 97.81

Latent heal (λλλλE, in W/m2) 201.96 57.24 13.58 272.78

Bowen ratio (=H/λλλλE) 0.34 0.42 0.42 0.36

CO2 flux (FCO2, in µmol/m2/s) -9.90 -3.57 -1.69 -15.17

Var(w) (in m2/s2) 0.24 0.03 0.009 0.28

Var(T) (in K2) 0.1087 0.1886 0.3059 0.6032

Var(q) (in g2/kg2) 0.1071 0.0639 0.084 0.255

Var(c) (in µmol2/mol2) 5.4341 11.7563 21.574 38.7644

42

Figure Captions.

Figure 1. Fourier power spectra of vertical wind velocity component (w, black dash-dotted

line) and virtual temperature (Tv, gray thick line), measured from 10:00 to 12:00 hs (local

time), on day 98, over Rebio Jaru forest.

Figure 2. Scale variances of vertical wind velocity (w), virtual temperature (Tv), specific

humidity (q) and CO2 concentration (c), averaged during diurnal periods (9:00 to 17:00 hs),

from day 93 to 98 over Rebio Jaru forest. The error bars represent the 95% confidence

level. The regions between the vertical dashed lines are denoted TS (Turbulent Scales), LE

(Large Eddies) and MS (Mesoscale).

Figure 3. Scale covariance contribution to fluxes of sensible heat (H), latent heat (λE) and

CO2 (FCO2), averaged for the same period as in figure 2. Details are the same as in figure

2.

Figure 4. As in figure 2, except for the scale correlation coefficient between virtual

temperature (Tv) and specific humidity (q).

Figure 5. Ratio of sensible heat fluxes of each eddy-pattern class to total sensible heat flux,

as a function of horizontal wind speed, calculated from day 31 to 60 over Rebio Jaru forest.

The values were calculated as averages of fluxes separated by wind speed classes. The error

bars represent the 95% confidence level.

43

log(f) in Hz0.001 0.01 0.1 1 10

f S2 αα

(f)

0.001

0.01

0.1

1

vertical velocity (w)virtual temperature (t)

Figure 1. Fourier power spectra of vertical wind velocity component (w, black dash-dotted

line) and virtual temperature (Tv, gray thick line), measured from 10:00 to 12:00 hs (local

time), on day 98, over Rebio Jaru forest.

44

Rebio Jaru - Days 93 to 98

Spatial scale (m)

0.1 1 10 100 1000 10000

varia

nce

of v

ertic

al w

ind

vel.

(m2 /s

2 )

0.00

0.01

0.02

0.03

0.04

0.05

varia

nce

of a

ir te

mp.

(K2 )

0.00

0.05

0.10

0.15

0.20

0.25va

rianc

e of

spe

cific

hum

idity

(g2 /k

g2 )

0.00

0.02

0.04

0.06

0.08

0.10

0.12

varia

nce

of C

O2 c

onc.

(µm

ol2 /m

ol2 )

0

2

4

6

8

10

12

14

16

var (w)var (t)var (q)var (c)

T.S. L.E. M.S.

Figure 2. Scale variances of vertical wind velocity (w), virtual temperature (Tv), specific

humidity (q) and CO2 concentration (c), averaged during diurnal periods (9:00 to 17:00 hs),

from day 93 to 98 over Rebio Jaru forest. The error bars represent the 95% confidence

level. The regions between the vertical dashed lines are denoted TS (Turbulent Scales), LE

(Large Eddies) and MS (Mesoscale).

45

Rebio Jaru - Days 93 to 98

Spatial scale (m)0.1 1 10 100 1000 10000

Sens

ible

and

late

nt h

eat f

luxe

s (W

/m2 )

-10

0

10

20

30

40

50

60

70

CO

2 Flu

x (µ

mol

/m2 s)

-4

-3

-2

-1

0

1

2

HλEFCO2

L.E. M.S.

Figure 3. Scale covariance contribution to fluxes of sensible heat (H), latent heat (λE) and

CO2 (FCO2), averaged for the same period as in figure 2. Details are the same as in figure

2.

46

Rebio Jaru - Days 93-98

Spatial scale (m)

0.1 1 10 100 1000 10000

T-q

corre

latio

n co

effic

ient

-0.4

-0.2

0.0

0.2

0.4

0.6

0.8T.S. L.E. M.S.

Figure 4. As in figure 2, except for the scale correlation coefficient between virtual

temperature (Tv) and specific humidity (q).

47

Horizontal wind speed (m/s)

1 2 3 4

Rat

io to

tota

l sen

sibl

e he

at fl

uxes

-0.25

0.00

0.25

0.50

0.75

1.00

1.25

TurbulentLarge eddiesMesoscale

Figure 5. Ratio of sensible heat fluxes of each eddy-pattern class to total sensible heat flux,

as a function of horizontal wind speed, calculated from day 31 to 60 over Rebio Jaru forest.

The values were calculated as averages of fluxes separated by wind speed classes. The error

bars represent the 95% confidence level.