You are on page 1of 9

Forest Ecology and Management 223 (2006) 45–53

www.elsevier.com/locate/foreco

Estimation of tree canopy cover in evergreen oak


woodlands using remote sensing
João M.B. Carreiras *, José M.C. Pereira, João S. Pereira
Department of Forestry, Instituto Superior de Agronomia, Tapada da Ajuda, 1349-017 Lisboa, Portugal
Received 1 April 2005; received in revised form 21 October 2005; accepted 24 October 2005

Abstract
The montado/dehesa landscapes of the Iberian Peninsula are savannah-type open woodlands dominated by evergreen oak species (Quercus
suber L. and Q. ilex ssp. rotundifolia). Scattered trees stand over an undergrowth of shrubs or herbaceous plants. To partition leaf area index
between trees and the herbaceous/shrubby understorey requires good estimates of tree canopy cover and is of key importance to understand the
ecology and the changes in land cover. The two vegetation components differ in phenology as well as in radiation and rainfall interception, water
and CO2 fluxes. The main goal of this study was to estimate tree canopy cover in a montado/dehesa region of southern Portugal (Alentejo) using
remote sensed data. For this purpose we developed empirical models combining measurements obtained through the analysis of aerial photos and
reflectance from Landsat Thematic Mapper (TM) individual channels, vegetation indices, and the components of the Kauth–Thomas (K–T)
transformation. A set of 142 plots was designed, both in the aerial photos and in the satellite data. Several simple and multiple linear regression
models were adjusted and validated. A subset of 75% of the data (n = 106) was used for model fitting, and the remainder (n = 36) was used for
model assessment. The best linear equation includes Landsat TM channels 3, 4, 5 and 7 (r2 = 0.74), but the Normalised Difference Vegetation Index
(NDVI), the components of the K–T transformation, and the Atmospherically Resistant Vegetation Index (ARVI) also performed well (r2 = 0.72,
0.70, and 0.69, respectively). The statistics of prediction residuals and tests of model validation indicates that these were also the models with better
predictive capability. These results show that detection of low/medium tree canopy cover in this type of land cover (i.e. evergreen oak woodlands)
can be accomplished with the help of high and medium spatial resolution satellite imagery.
# 2005 Elsevier B.V. All rights reserved.

Keywords: Landsat Thematic Mapper (TM); Aerial photo; Linear regression; Evergreen oak woodlands; Tree canopy cover

1. Introduction are exploited for cork, wood and extensive agriculture or grazing,
in proportions that vary with local conditions and history. The
Savannah-type Mediterranean evergreen oak woodlands are landscapes are typically heterogeneous and include three distinct
widely distributed in the Iberian Peninsula, as well as in other variants, which differ in terms of land use intensity and land cover
areas with the Mediterranean type of climate, e.g., in parts of structure (Blanco et al., 1997):
California, Chile, South Africa and Australia. These are complex
ecosystems with open, heterogeneous canopies with shrub or - Relatively denser oak woodlands are more common in steep
annual herbaceous understories (Joffre et al., 1999; Baldocchi areas and poor soils, unsuitable for agricultural use. Land use
et al., 2004). In the Iberian Peninsula, large areas of these intensity is low, and a diverse shrub community species
woodlands depend upon human action, forming a multiple use dominates the understorey. In Portugal dense cork oak stands
agroforestry system, called montado in Portugal and dehesa in were planted for cork production in better soils.
Spain. For example, in Portugal, they cover an area of - Pastures represent an intermediate level of land use intensity.
approximately 1.2 Mha (DGF, 2001). The most common oak Tree cover is sparser, and the understorey is dominated by a
species are holm oak (Quercus rotundifolia Lam. syn. Quercus great variety of annual/biennial herbaceous plant commu-
ilex L. ssp. rotundifolia) and cork oak (Quercus suber L.). They nities. The productivity of these communities is dominated by
climatic factors, and typically they display large interannual
variability in plant cover.
* Corresponding author. Tel.: +351 213653387; fax: +351 213645000. - Dryland farming of cereal crops in areas with deeper soils,
E-mail address: jmbcarreiras@isa.utl.pt (João M.B. Carreiras). such as the bottom of valleys. During fallow years, these areas
0378-1127/$ – see front matter # 2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.foreco.2005.10.056
46 J.M.B. Carreiras et al. / Forest Ecology and Management 223 (2006) 45–53

may be used as pastures. There is a tendency for abandonment quantified against a very noisy background of bare soils,
of this type of agriculture and the montado/dehesa reverts to agricultural crops, pastures and shrubs.
one of the former land cover types (Lourenço et al., 1998). The main goal of this study was to estimate tree canopy
cover in a montado/dehesa region of southern Portugal
Tree density in montado ecosystems frequently ranges from (Alentejo) using remote sensed data. Several studies have
30 to 60 trees ha1 (Blanco et al., 1997), but higher and lower produced good and promising results. Joffre and Lacaze (1993)
densities are common. In Portugal, about half of the area of Q. used panchromatic SPOT High Resolution Visible (HRV) (10 m
suber montados has tree canopy cover in the 10–30% class, one spatial resolution) and Landsat Thematic Mapper (TM) (30 m
quarter of the area in the 30–50% class, and the remaining spatial resolution) satellite imagery to estimate tree density in a
quarter in the >50% tree canopy cover class, whereas holm oak savannah-type ecosystem (holm and cork oaks), in southern
woodlands are even sparser, with about 85% of the area in the Spain; tree density was estimated with reasonable accuracy
10–30% canopy cover class (DGF, 2001). These low tree through the application of spatial filters to the panchromatic
densities result from specific land use objectives and manage- SPOT HRV band (r = 0.94), from the original panchromatic
ment practices, which transformed the relatively denser SPOT HRV (r = 0.85), and using a combination of Landsat
primordial woodlands and forests into landscapes structurally TM channels 2 (TM2), 3 (TM3) and 4 (TM4) (r = 0.61).
similar to savannas. Salvador and Pons (1998) used Landsat TM imagery combined
The Mediterranean evergreen oak woodlands differ from with fieldwork to estimate dendrometric variables (including
tropical savannahs in the timing of rainfall, which in the tree canopy cover) used in forest inventories in the
former occur during the relatively cool autumn–winter Mediterranean region, relying on regression models; these
months. This leads to severe water deficits during the hot were found to be consistent with the expected vegetation
Mediterranean summer. As a consequence large intra-annual spectral response; most of the multiple linear regression models
variations in plant cover occur (Baldocchi et al., 2004). The fit well the data, allowing quantitative predictions for several
understanding of ecosystem structure and function (e.g., field variables from remote sensing data; the best linear
radiation and rainfall interception, water and CO2 fluxes) regression (r2 = 0.64) for predicting tree canopy cover in Q. ilex
requires the partitioning of total leaf area index between the stands included Landsat TM channel 7 (TM7). Pereira et al.
tree component and the herbaceous/shrubby understorey. The (1995) used Landsat TM imagery and fieldwork measurements
two vegetation layers differ in phenology and functioning. For to determine biomass, percent canopy cover and canopy volume
example, whereas the sparse tree cover remains green and may in Mediterranean shrublands; the best results were obtained
have access to perennial water sources for year-long metabolic with the Normalized Difference Vegetation Index (NDVI) to
activity (David et al., 2004), the herbaceous layer dries out in estimate percent canopy cover (r2 = 0.65). Oliveira (1998) used
the dry summer. In addition to the spatial and temporal field radiometry and Landsat TM imagery to estimate several
heterogeneity, the montado/dehesa landscapes are today shrub parameters in the Alto-Dão e Lafões region (Portugal);
subject to rapid changes resulting from land abandonment, field radiometry results showed that the Landsat TM3 channel
changing fire regimes and tree decline (Brasier, 1992, 1996; was the one better correlated with canopy cover (r = 0.91) and
Terradas, 1999; Stoate et al., 2001; Romero-Calcerrada and the NDVI proved to be the vegetation index (VI) with better
Perry, 2004). performance (r = 0.75); regarding the Landsat TM imagery, it
Given the economical, ecological, and historical signifi- was found that TM3, channel 5 (TM5), and TM7 yielded the
cance of these ecosystems, it is important to monitor the better correlation with canopy cover (r = 0.87) and the
changes that affect them, in a timely and cost-effective manner. Atmospherically Resistant Vegetation Index (ARVI) was the
Satellite remote sensing provides a potentially useful tool for best VI (r = 0.86). Calvão and Palmeirim (2004) collected field
such monitoring, but the problem is complex because of the data over Mediterranean shrublands and developed correlations
intrinsic diversity, structural complexity, and seasonal varia- between several biophysical parameters and spectral variables
bility of the montado/dehesa landscapes. The designation of a (single channel reflectance and NDVI) from Landsat TM data;
relatively broad range of land cover types, under a common the higher correlation for canopy cover was obtained with TM3
land use name, leads to a problem of mismatch between (r = 0.91) and NDVI (r = 0.91).
informational classes (i.e. those meaningful to a user), and
spectral classes (those which share similar spectral properties) 2. Study area and methods
(Swain and Davies, 1978). In the case of the montado/dehesa
systems, informational land use classes tend to be very The study area is located near the city of Évora (University
heterogeneous from the spectral standpoint, reflecting their of Évora/Mitra Campus, 388320 N and 88000 W), comprising an
internal ecological variability. This is a severe problem for area equivalent to a Landsat TM mini-scene (approximately
qualitative land use mapping, where the goal is to discriminate 50 km  50 km). The data used in this study came from two
between distinct land cover types. Quantitative characterisation distinct sources: 1-m spatial resolution ortho-rectified digital
of a biophysical attribute such as tree canopy cover is also quite infrared aerial photography from a flight in the summer of 1995
challenging in these landscapes, because of the very broad (acquired between 28 August and 13 September) (Table 1) and
range in tree density, and of the wide variety of understorey 30-m Landsat TM imagery from the same year (acquired on 15
cover types. The signal (tree canopy) has to be detected and August). Proximity between the dates of the two data sources is
J.M.B. Carreiras et al. / Forest Ecology and Management 223 (2006) 45–53 47

Table 1 each digital aerial photo; this value was considered adequate to
Characteristics of the aerial photos used in this study
characterise the spatial distribution of tree canopy cover; (ii)
Aerial photograph # Tree density Acquisition date square plots of 120 m  120 m were delineated, centred in the
3535 Low 28 August 1995 points previously chosen; this plot size was chosen considering
3643 Low 28 August 1995 the spatial resolution of the satellite imagery, i.e. 4  4 Landsat
3685 High 28 August 1995 TM pixels; (iii) in each plot a systematic grid of 196 points was
5153 Low 13 September 1995 overlaid (14  14); (iv) the number of points falling in tree
5195 High 13 September 1995
crowns was counted in each plot (both cork and holm oaks);
tree canopy cover (TCC) in each plot was obtained dividing the
important in this study area because main phenological changes number of points overlaid in the crowns by the total in the plot,
in land cover are highly detectable, due to low tree density. This according to TCC = (number of tree crown points/total number
is more evident in croplands and pastures, which occupy a large grid points)  100%.
proportion of the understorey of Q. suber and Q. ilex ssp. Fig. 1 displays a plot overlaid on an aerial photo, and the grid
rotundifolia in this region (Gaspar, 1993). At the time of the used to determine tree canopy cover. It shows the points that
year of image acquisition (late summer) most of the croplands were considered to fall on a tree crown (white plus signs) and
had been harvested and the pastures were dry. This is a those falling on the understorey (black circles).
fundamental issue because it allows a better spectral contrast
between the overstorey and the understorey. 2.2. Satellite image analysis
The five aerial photos were selected to cover a broad range of
tree densities. Landsat TM mini-scene was calibrated to Several spectral measures related to the presence of
radiance and apparent reflectance using the time-dependent vegetation were calculated. Vegetation spectral indices (VIs)
coefficients proposed by Teillet and Fedosejevs (1995). Image combine information from two or more spectral channels to
georeferencing was performed with a linear polynomial enhance vegetation signal, while minimising soil, atmospheric,
function, using 21 control points extracted from 1:25 000 and solar irradiance effects (Jackson and Huete, 1991).
scale topographic maps from the Portuguese Army Geographic Several VIs were calculated, based on Landsat TM
Institute. A residual mean square error of 9.6 m was obtained. individual channels, namely the NDVI (Rouse et al., 1974),
Subsequently, the Landsat TM mini-scene and the digital aerial the Soil Adjusted Vegetation Index (SAVI) (Huete, 1988), the
photos were co-registered, to guarantee spatial correspondence modified soil adjusted vegetation index (MSAVI) (Qi et al.,
between the two data sources. 1994), the Green Normalised Difference Vegetation Index
The methodology used comprised three stages: (a) analysis of (GNDVI) (Gitelson et al., 1996), the ARVI (Kaufman and
digital aerial photos to estimate tree canopy cover; (b) calculation Tanré, 1992), the VI5 (Marchetti and Ricotta, 1993), and the
of spectral measures of vegetation cover, derived from satellite VI7 (López and Caselles, 1991). The equations for these indices
imagery; (c) regression modelling of tree canopy cover (aerial are shown in Table 2.
photos), as a function of spectral reflectance data (satellite). The NDVI, GNDVI, VI5 and VI7 all are normalized
difference VIs, and they differ only in the channels used.
2.1. Analysis of digital aerial photos Classical VIs (e.g., the NDVI) exploit the fact that vegetation is
highly reflective in the near infrared (NIR) region and strongly
Estimation of tree canopy cover using the aerial photos was absorbing in the visible (due to chlorophyll absorption)
accomplished as follows: (i) 30 points were randomly chosen in and short wave infrared (SWIR) (due to water absorption).

Fig. 1. (a) Plot insertion in the aerial photo, and (b) detail of the same plot and respective grid (aerial photo 3685, plot 10). The white plus signs represent the grid
points overlaid in tree crowns, and the black circles those overlaid in the understorey.
48 J.M.B. Carreiras et al. / Forest Ecology and Management 223 (2006) 45–53

Table 2
Spectral indices calculated with the Landsat data. TM1, TM2, TM3, TM4, TM5, TM7 – reflectances of Landsat TM channels. L: soil adjustment factor; a value of 0.5
is reasonable for a broad range of soil types (Huete, 1988). g: atmospheric self-correcting factor, dependent on aerosols type; when the canopy is sparse and
atmospheric data are unknown, a value of g = 1 is recommended for most remote sensing applications (Kaufman and Tanré, 1992)
Spectral indices/K–T components Equation
NDVI (TM4  TM3)/(TM4 + TM3)
SAVI ((TM4  TM3)/(TM4 + TM3))(1 + L)
MSAVI (2TM4 + 1  ((2TM4 + 1)2  8(TM4  TM3))1/2)/2
GNDVI (TM4  TM2)/(TM4 + TM2)
ARVI (TM4  RB)/(TM4 + RB), RB = TM3  g(TM1  TM3)
VI5 (TM4  TM5)/(TM4 + TM5)
VI7 (TM4  TM7)/(TM4 + TM7)
K–T brightness 0.2043TM1 + 0.4158TM2 + 0.5524TM3 + 0.5741TM4 + 0.3124TM5 + 0.2303TM7
K–T greenness 0.1603TM1  0.2819TM2  0.4934TM3 + 0.7940TM4-0.0002TM5  0.1446TM7
K–T wetness 0.0315TM1 + 0.2021TM2 + 0.3102TM3 + 0.1594TM4  0.6806TM5  0.6109TM7

This spectral behaviour contrasts with that of most other land spectral variables. Simple linear regression models were
surfaces and atmospheric features and is the key to spectral developed to estimate tree canopy cover as a function of each
detection of vegetation and quantification of its abundance. The VI. Multiple linear regression was used for the models based on
VI5 and VI7 replace the visible channel of classical vegetation the components of the K–T transformation and for the model
indices with SWIR channels. These are sensitive to vegetation based on individual Landsat TM channels. For the latter model,
moisture, enhancing the contrast between vegetation and drier variables were chosen using a stepwise forward selection
areas, and are less affected by atmospheric effects. The GNDVI procedure, with penter = 0.05 and premove = 0.40. Model
uses the green spectral region instead of the red region, to evaluation was performed using goodness-of-fit statistics
increase sensitivity to the presence of chlorophyll (Gitelson (coefficient of determination (r2) and adjusted r2).
et al., 1996). Huete (1988) developed a VI (SAVI) to minimise Model validation was based on the analysis of the difference
confounding effects induced by the soil background, especially between observed and predicted values (residuals) for the
in areas with relatively sparse vegetation cover. Qi et al. (1994) validation subset (n = 36), using some of the tests for model
proposed a modification of SAVI (MSAVI); it uses a variable assessment recommended by Soares and Tomé (1993).
soil adjustment factor that depends on the amount of vegetation Residual analysis was performed using the mean prediction
present. Kaufman and Tanré (1992) developed the ARVI, meant residuals, mean absolute prediction residuals, and sum of
to be resistant toward atmospheric effects; it uses the blue squared residuals. The statistical tests were performed to
channel to quantify, and compensate for, the magnitude of the evaluate model bias. First, the Shapiro-Wilk W test (Shapiro
atmospheric effect. et al., 1968) was used to test for normality of the distribution of
Other spectral indicators of the presence of vegetation were prediction residuals. Student’s t-test and regression tests were
also calculated, namely the components of the Kauth–Thomas performed for models with normally distributed residuals.
(K–T) transformation, also know as the tasselled cap Student’s t-test evaluates model accuracy, based on the null
transformation (Crist and Cicone, 1984). The K–T transforma- hypothesis of a mean prediction residual equal to zero. The
tion converts the six reflective channels of the Landsat TM into regression tests are based on the null hypothesis that when
three orthogonal components, which are linear combinations of observed values are regressed over the corresponding predicted
the original channels. The first of these component, designated values, a regression line with unit slope and zero intercept
brightness, expresses differences in soil properties, like particle should be obtained, if the model is unbiased. For models with
size and organic matter content; the second component, non-normally distributed residuals the non-parametric Wilcox-
greenness, is well correlated with tree canopy cover, leaf area on’s signed-rank test was used. This test is based on the sign and
index and live biomass; the third component, wetness, is magnitude of the rank of the differences between pairs of
sensitive to soil and plant moisture (Crist and Cicone, 1984). measurements. The formal null hypothesis for Wilcoxon’s test
The Landsat TM satellite data were converted to apparent is that the population distribution of differences is symmetrical
reflectance, so we used the K–T equivalent transformation for about zero. Details about all these tests can be found in standard
reflectance factor data (Crist, 1985). statistics textbooks (e.g., Conover, 1980; Steel and Torrie,
1980; Zar, 1999).
2.3. Model adjustment and validation
3. Results
The 142 tree canopy cover estimates were randomly split
into two subsets, one for model fitting, with 75% of the data 3.1. Determination of tree canopy cover
(n = 106), and another for validation, with the remainder 25%
(n = 36). Ordinary least squares linear regression modelling Table 3 shows a summary of the results of the estimation of
was employed to estimate tree canopy cover as a function of tree canopy cover in the five aerial photos. Some of the aerial
J.M.B. Carreiras et al. / Forest Ecology and Management 223 (2006) 45–53 49

Table 3 Table 4
Descriptive statistics of tree canopy cover estimation in the aerial photos Performance of the nine tested models, showing coefficients estimates, coeffi-
cient of determination (r2) and adjusted r2 (radj
2
) (n = 106); TCC: tree canopy
Aerial Number Tree canopy cover statistics
cover; BRIGHT, GREEN, WET – K–T brightness, greenness, and wetness
photograph # of plots
components, respectively
Mean Minimum Maximum S.D.
Model r2 2
radj
3535 30 31.6 2.0 61.2 15.2
3643 30 7.4 0.0 42.3 13.0 1. TCC = 36.298 + 278.354NDVI 0.72 0.71
3685 27 27.5 0.0 56.1 16.7 2. TCC = 44.472 + 552.388SAVI 0.66 0.66
5153 27 11.6 0.0 48.2 15.4 3. TCC = 42.502 + 628.068MSAVI 0.64 0.64
5195 28 28.5 0.0 100.0 30.6 4. TCC = 87.977 + 397.043GNDVI 0.64 0.64
5. TCC = 12.491 + 152.326ARVI 0.69 0.69
6. TCC = 57.223 + 231.738VI5 0.59 0.59
7. TCC = 8.569 + 182.749VI7 0.63 0.63
8. TCC = 54.723  36.617BRIGHT 0.70 0.69
+ 746.741GREEN + 157.412WET
9. TCC = 63.626  447.222TM5 + 623.837TM4 0.74 0.73
 714.626TM3 + 281.354TM7

data is concentrated in the classes with lower canopy cover,


which is to be expected in these relatively open woodlands.

3.2. Model adjustment and validation

The results of the adjustment of the different models can be


seen in Table 4. The model with the best performance is the one
resulting from stepwise regression (with individual channels as
Fig. 2. Histogram of tree canopy cover (%) for all the plots (n = 142). independent variables), followed by the model with the NDVI
as independent variable, and then those using the three K–T
photos have fewer than 30 plots because the corresponding grid components or the ARVI.
point was not totally inside the photo, so those plots were To get more insight into the data, a 2D perspective, i.e.
rejected. Fig. 2 shows the histogram for tree canopy cover, displaying several independent variables versus the dependent
considering all plots (n = 142). As expected, the majority of the variable, is presented. In Fig. 3, the different scatterplots, using

Fig. 3. Scatterplots of tree canopy cover (n = 142) vs. Landsat TM individual channels: (a) TM3; (b) TM4; (c) TM5; (d) TM7.
50 J.M.B. Carreiras et al. / Forest Ecology and Management 223 (2006) 45–53

the individual Landsat TM channels are represented, and in validation. The four models with better goodness-of-fit are also
Fig. 4 the same representation is shown for the NDVI, ARVI, those with lower arp and ssr; model 9 (with Landsat channels
and K–T components. Regarding Fig. 3, there is a non-linear TM3, TM4, TM5 and TM7) has the lowest arp and ssr, followed
inverse relation with tree canopy cover, with the exception of by model 8 (with K–T components), model 1 (with NDVI), and
TM4 reflectance, where there is no correlation. In Fig. 4, as model 5 (with ARVI). The W value of the Shapiro–Wilk test is
would be expected for the VIs, the highest tree canopy cover significant (for a = 0.05) for models 1–3, 6, and 7. The results
corresponds to the highest VIs values. Moreover, the highest K– of the t-test indicate that there is insufficient information to
T brightness component occurs in those plots with lowest tree reject the null hypothesis for all the tested models (for
canopy cover; the opposite happens for the K–T greenness and a = 0.05). The statistic of the intercept = 0 regression t-test is
wetness components. significant for model 3 (for a = 0.05), and for models 6 and 7
Table 5 shows model validation results. The models with (for a = 0.001). The slope = 1 regression t-test is significant for
lower mean prediction residuals (rp) (models 2, 3, and 7) have, models 1, 2, and 9 (for a = 0.05), for model 3 (for a = 0.01), and
however, higher mean absolute prediction residuals (arp) and for models 6 and 7 (for a = 0.001). These results indicate that
sum of squared residuals (ssr). This could result from both models 6 and 7 are highly biased (for a = 0.001), and that
higher negative and positive residuals that sum to a lower value models 1, 2, 9, and 3 are slightly biased. However, we should
(lower rp) but resulting in higher arp and ssr; therefore the arp note that the previous t-tests only have meaningful interpreta-
and ssr measures seem to be more appropriate to assess model tion for those models displaying normally distributed residuals.

Fig. 4. Scatterplots of tree canopy cover (n = 142) vs. NDVI, ARVI, and K–T components: (a) NDVI; (b) ARVI; (c) K–T brightness; (d) K–T greenness; (e) K–T
wetness.
J.M.B. Carreiras et al. / Forest Ecology and Management 223 (2006) 45–53 51

Table 5
Summary statistics of prediction residuals and results of statistical tests of model validation (n = 36); rp: mean prediction residuals, arp: mean absolute prediction
residuals, ssr: sum of squared residuals; the Wilcoxon’s signed-rank test is not shown for models 4, 5, 8, and 9, as there is no evidence to reject the null hypothesis of
the Shapiro–Wilk test (for a a = 0.05)
Model Summary statistics Statistical tests
rp arp ssr Shapiro–Wilk test (W) t-test (t) Regression t-test Wilcoxon test (z)
Intercept = 0 Slope = 1
1 0.66 7.82 3717.43 0.930 0.385 1.442 2.399 0.220
2 0.38 8.76 4448.15 0.938 0.201 2.002 2.639 0.377
3 0.52 9.21 4967.28 0.937 0.261 2.155 2.812 0.518

4 0.79 8.77 4304.25 0.973 0.427 1.606 1.813
5 1.83 7.90 3374.09 0.990 1.118 0.292 1.352 –
6 0.67 24.43 41117.04 0.902 0.118 4.945 6.794 0.896
7 0.64 24.28 41317.05 0.900 0.112 4.822 6.814 0.676
8 0.91 7.20 2988.73 0.982 0.591 0.989 1.917 –
9 0.73 6.59 2713.60 0.953 0.498 1.512 2.638 –

The Wilcoxon z-test, appropriate for residuals with a non- greater than approximately 60%, although this could be due to
normal distribution, indicates that there is not enough an insufficient number of observations in this range.
information to reject the null hypothesis for the tested models
(for a = 0.05). Therefore, the previously mentioned four 4. Discussion
models with better goodness-of-fit, and lower prediction
residuals statistics, also passed on the appropriate statistical At this time of the year (late summer) the understorey has a
tests, the exception being model 9 for the slope = 1 regression t- very light colour, which is due to the presence of dry vegetation
test. and soil. The tree canopy is darker and when tree canopy cover
Fig. 5 shows the scatterplots of observed versus predicted increases the reflectance of the plot decreases (reflective TM3),
tree canopy cover (in the validation subset) for these four because the soil reflects solar radiation and the chlorophyll
models. Models 8 and 9 yielded a better agreement between absorbs it (Fig. 3a). In the middle infrared (TM5 and TM7) the
observed and predicted data. Another relevant result in these situation is the same, but the absorption is due to the presence of
models is the tree canopy cover overestimation for values water in the living tissues (Fig. 3c and d). In the NIR (TM4)

Fig. 5. Observed vs. predicted tree canopy cover (%) for the (a) model 1, (b) model 5, (c) model 8, and (d) model 9. The data refers to the validation subset (n = 36).
The 1:1 line represents perfect agreement.
52 J.M.B. Carreiras et al. / Forest Ecology and Management 223 (2006) 45–53

crowns are highly reflective, as well as the understorey of bright simplest, followed by that with the K–T components. All the
soils and dry grass, thus resulting in the weak correlation shown parameters in model 9 are interpretable, except the one
in the scatterplot (Fig. 3b). associated with the TM7; which should be negative, like the one
Satellite imagery acquired in the late summer maximize the associated with TM5, reflecting the presence of water in the
spectral contrast between the evergreen tree crowns and the dry living tissues, typical of the SWIR spectral region; one possible
herbaceous background. However, the study area is still explanation is the observed high correlation between TM5 and
characterized by an extremely variable understorey, containing TM7 (r = 0.96); the signal is positive for TM4 because a higher
bare soil, dry grass, and a few evergreen shrubs. The differences tree canopy cover is associated with a high value of reflectance
between VIs values of plots with the same tree canopy cover are in the NIR; the opposite happens with TM3, the negative signal
due to the presence of different types of understorey, which indicating that a higher absorption (less reflectance) is a sign of
have different spectral properties, and/or different vegetation the presence of high values of chlorophyll, and so more
structure. Huete (1989) refers that in areas with significant soil- vegetation. As expected, the equation with the NDVI or the
brightness changes that arise from moisture differences, ARVI present the parameter with a positive signal, indicating a
roughness variations, shadow, or organic-matter differences, direct relation between the increase of these VIs and tree
there are soil-induced influences on the vegetation indices canopy cover. The equation with the K–T components has the
values, and that these effects are predominant in partially parameters with the appropriate signals: areas with dense
vegetated canopies, as in this study. canopy cover are dark (negative brightness coefficient), green
To distinguish different plots according to tree canopy cover, (positive greenness sign), and contain abundant leaf tissue
a diagram representing Landsat channels TM3 and TM4 was moisture (positive wetness sign).
built (Fig. 6). The plots were grouped into five classes,
according to tree canopy cover estimated from the aerial 5. Conclusions
photos. Richardson and Wiegand (1977) observed that bare soil
pixels were generally aligned according to a straight line The low tree density characteristic of these Mediterranean
(beginning in the origin), which approximately bisects the red- evergreen oak woodlands has a large influence on the estimation
NIR space (soil line). They also observed that points of tree canopy cover with remote sensing data. The spectral
representing pixels of water were located close to the origin contribution of the understorey for the signal received by the
of the soil line, while pixels of vegetation were always located satellite sensor increases with decreasing tree density. Therefore,
to the left of the same line. It is evident that plots with lower tree one should select a period when there is maximum spectral
canopy cover (i.e. with higher proportion of understorey) are contrast between the overstorey and the understorey.
located closer to the soil line. The higher or lower reflectance of The model with best predictive capability (model 9) includes
these plots in these channels is related with the type of the one channel in the visible domain, one in the NIR, and two in
background. An increase in tree canopy cover, within the same the SWIR, and produces good estimates of tree canopy cover.
type of understorey, shows an increase of the reflectance in the The second best model (model 8) includes the components of
NIR (TM4) and a decrease in the red (TM3). the K–T transformation, yielding a robust and easily
Considering the four best models, model 9 is the harder to interpretable model. Two other models present good predictive
interpret, because it is a multiple linear regression model. Those capability, namely those including the NDVI (model 1) and the
with the NDVI (model 1) or with the ARVI (model 5) are the ARVI (model 5).
It is also important to stress that the type of relation
established can be helpful in understanding the dynamics of
tree canopy cover in this region of the country. The use of high
and medium spatial resolution satellite imagery could be an
instrument to monitor changes that occur in this type of forest,
with lower tree density. Another important key aspect is the
presence of several types of understory associated with
different crown cover percentage, which favours the robustness
of the model being built.

Acknowledgements

The authors gratefully acknowledge the two anonymous


reviewers for their helpful comments and suggestions. This
study was funded by the Portuguese ‘‘Subprograma Ciência e
Tecnologia do 28 Quadro Comunitário de Apoio’’ within the
grant Ref. GGP XXI/BTI/4458/96. This grant was integrated in
a broader project called ‘‘Monitorização da Resposta ao Stress
Fig. 6. Representation of the dataset (142 plots), based in tree canopy cover Ambiental em Povoamentos de Sobreiro’’, Ref. PRAXIS XXI/
(TCC) classes, in the TM3–TM4 reflectance plane. 3/3.2/FLOR/2122/95.
J.M.B. Carreiras et al. / Forest Ecology and Management 223 (2006) 45–53 53

References López, G.M.J., Caselles, V., 1991. Mapping burns and natural reforestation
using thematic mapper data. Geocarto Int. 1, 31–37.
Baldocchi, D.D., Xu, L., Kiang, N., 2004. How plant functional type, weather, Lourenço, N., Pinto-Correia, T., Jorge, M.R., Machado, C.R., 1998. Farming
seasonal drought, and soil physical properties alter water and energy fluxes strategies and land use changes in southern Portugal: land abandonment
of an oak-grass savanna and an annual grassland. Agric. For. Meteorol. 123, or extensification of the traditional systems? Mediterrâneo 12/13, 191–
13–39. 208.
Blanco, E., Casado, M.A., Costa, M., Escribano, R., Garcia, M., Genova, M., Marchetti, M., Ricotta, C., 1993. L’impiego di dati LANDSAT TM per il
Gomez, A., Gómez, F., Moreno, J.C., Morla, C., Regato, P., Sainz, H., 1997. monitoraggio della ripresa vegetativa in aree incendiate. Monti e Boschi 3,
Los Bosques Ibéricos: Una Interpretación Geobotânica. Editorial Planeta, 22–26.
Barcelona, Spain. Oliveira, T.M., 1998. Cartografia Quantitativa de Formações Arbustivas Empre-
Brasier, C.M., 1992. Oak tree mortality in Iberia. Nature 360, 539. gando Dados de Detecção Remota. Tese de Mestrado em Gestão dos
Brasier, C.M., 1996. Phytophthora cinnamomi and oak decline in southern Recursos Naturais. Universidade Técnica de Lisboa, Instituto Superior
Europe Environmental constraints including climate change. Annales des de Agronomia, Lisbon, Portugal.
Sciences Forestières 53, 347–358. Pereira, J.M.C., Oliveira, T.M., Paul, J.C.P., 1995. Satellite-based estimation of
Calvão, T., Palmeirim, J.M., 2004. Mapping Mediterranean scrub with satellite Mediterranean shrubland structural parameters. EARSeL Adv. Remote
imagery: biomass estimation and spectral behaviour. Int. J. Remote Sens. 25 Sens. 4 (3), 14–20.
(16), 3113–3126. Qi, J., Chehbouni, A., Huete, A.R., Kerr, Y.H., Sorooshian, S., 1994. A modified
Conover, W.J., 1980. Practical Non-parametric Statistics, 2nd ed. John Wiley soil adjusted vegetation index. Remote Sens. Environ. 47, 1–25.
and Sons, New York. Richardson, A.J., Wiegand, C.L., 1977. Distinguishing vegetation from soil
Crist, E.P., 1985. A TM tasselled cap equivalent transformation for reflectance background information. Photogrammet. Eng. Remote Sens. 43 (12), 115–
factor data. Remote Sens. Environ. 17, 301–306. 125.
Crist, E.P., Cicone, R.C., 1984. A physically-based transformation of thematic Romero-Calcerrada, R., Perry, G.L.W., 2004. The role of land abandonment in
mapper data – the TM tasseled cap. IEEE Trans. Geosci. Remote Sens. 22 landscape dynamics in the SPA ‘Encinares del rı́o Arberche y Cofio’ Central
(3), 256–263. Spain1984–1999. Landsc. Urban Plan. 66, 217–232.
David, T.S., Ferreira, M.I., Cohen, S., Pereira, J.S., David, J.S., 2004. Con- Rouse, J.W., Haas, R.W., Schell, J.A., Deering, D.W., Harlan, J.C., 1974.
straints on transpiration from an evergreen oak tree in southern Portugal. Monitoring the vernal advancement and retrogradation (greenwave effect)
Agric. For. Meteorol. 122, 193–205. of natural vegetation. NASA/GSFCT Type III Final Report, Greenbelt, MD,
DGF, 2001. Inventário Florestal Nacional. Portugal Continental. 3 Revisão, USA.
1995–1998. Relatório Final. Direcção-Geral das Florestas, Lisbon, Portu- Salvador, R., Pons, X., 1998. On the applicability of Landsat TM images to
gal. Mediterranean forest inventories. For. Ecol. Manage. 104, 193–208.
Gaspar, J., 1993. As Regiões Portuguesas. Ministério do Planeamento e da Shapiro, S.S., Wilk, M.B., Chen, H.J., 1968. A comparative study of various
Administração do Território, Secretaria de Estado do Planeamento e tests of normality. J. Am. Stat. Assoc. 63, 1343–1372.
Desenvolvimento Regional, Lisbon, Portugal. Soares, P., Tomé, M., 1993. Procedures for validation of forest management
Gitelson, A.A., Kaufman, Y.J., Merzlyak, M.N., 1996. Use of a green channel in growth models. An application to continuous forest inventory data. In:
remote sensing of global vegetation from EOS-MODIS. Remote Sens. Vanclay, K., Skousgaard, J.P., Gertner, G.T. (Eds.), Growth and Yield
Environ. 58, 289–298. Estimation from Successive Forest Inventories. Proceedings from IUFRO
Huete, A.R., 1988. A soil-adjusted vegetation index (SAVI). Remote Sens. Conference, Copenhagen. Forskningsserien 3, pp. 167–180.
Environ. 25, 295–309. Steel, R.G.D., Torrie, J.H., 1980. Principles and Procedures of Statistics, 2nd
Huete, A.R., 1989. Soil influences in remotely sensed vegetation-canopy ed. McGraw-Hill, Singapore.
spectra. In: Asrar, G. (Ed.), Theory and Applications of Optical Remote Stoate, C., Boatman, N.D., Borralho, R.J., Rio Carvalho, C., de Snoo, G.R.,
Sensing. John Wiley & Sons, New York, pp. 107–141. Éden, P., 2001. Ecological impacts of arable intensification in Europe. J.
Jackson, R.D., Huete, A.R., 1991. Interpreting vegetation indices. Prev. Vet. Environ. Manage. 63, 337–365.
Med. 11, 185–200. Swain, P.H., Davies, S.M., 1978. Remote Sensing. The Quantitative Approach.
Joffre, R., Lacaze, B., 1993. Estimating tree density in oak savanna-like dehesa McGraw-Hill, New York.
of southern Spain from SPOT data. Int. J. Remote Sens. 14 (4), 685–697. Teillet, P.M., Fedosejevs, G., 1995. On the dark target approach to atmospheric
Joffre, R., Rambal, S., Damesin, C., 1999. Functional attributes in Mediterra- correction of remotely sensed data. Can. J. Remote Sens. 21 (4), 374–386.
nean-type ecosystems. In: Pugnaire, F.I., Valladares, F. (Eds.), Handbook of Terradas, J., 1999. Holm oak and holm oak forests: an introduction. In: Roda,
Functional Plant Ecology. Marcel Dekker, New York, pp. 347–380. F., Retana, J., Gracia, C.A., Bellot, J. (Eds.), Ecology of Mediterranean
Kaufman, Y.J., Tanré, D., 1992. Atmospherically resistant vegetation index evergreen oak forests. Springer-Verlag, Berlin, pp. 3–14.
(ARVI) for EOS-MODIS. IEEE Trans. Geosci. Remote Sens. 30 (2), 261– Zar, J.H., 1999. Biostatistical Analysis, 4th ed. Prentice-Hall, Upper Saddle
270. River, NJ, USA.

You might also like