Anda di halaman 1dari 13

Agricultural and Forest Meteorology 213 (2015) 304316

Contents lists available at ScienceDirect

Agricultural and Forest Meteorology


journal homepage: www.elsevier.com/locate/agrformet

Spatial and temporal variability in spectral-based surface energy


evapotranspiration measured from Landsat 5TM across
two mangrove ecotones
David Lagomasino a,b, , Ren M. Price a,b , Dean Whitman a ,
Assefa Melesse a , Steven F. Oberbauer c
a
Florida International University, Department of Earth and Environment, Miami, FL, United States
b
Florida International University, Southeast Environmental Research Center, Miami, FL, United States
c
Florida International University, Department of Biological Sciences, Miami, FL, United States

a r t i c l e i n f o a b s t r a c t

Article history: Coastal mangroves lose large amounts of water through evapotranspiration (ET) that can be equivalent
Received 1 April 2014 to the amount of annual rainfall in certain years. Satellite remote sensing can play a crucial role in identi-
Received in revised form fying regional ET trends and surface energy changes after disturbances in isolated and inaccessible areas
17 November 2014
of coastal mangrove wetlands, like those found in Everglades National Park in southern Florida. Using
Accepted 18 November 2014
a combination of long-term datasets acquired from the NASA Landsat 5TM satellite database and the
Available online 4 December 2014
Florida Coastal Everglades Long-Term Ecological Research project, the present study investigates how
ET as well as radiation and other energy balance parameters in the Everglades mangrove ecotone have
Keywords:
Everglades
responded to multiple hurricane events and restoration projects over the past two decades. An energy
Hydrology balance model using satellite data was used to estimate latent heat (E) in tall and scrub mangrove
Disturbance environments. An eddy-covariance tower and weather tower supplied long-term data of multiple envi-
Latent heat ronmental and meteorological parameters that were used in calibrating and testing the modeled results
Soil heat ux from the Landsat images. Results identied signicant differences in E and soil heat ux measurements
Surface energy balance between the tall and scrub, and fringe and basin mangrove environments. The scrub mangrove site had
the lowest E rates, highest soil heat ux and lowest biophysical index (i.e., Fractional Vegetation Cover
(FVC), Normalized Difference Vegetation Index (NDVI), and Soil-Adjusted Vegetation Index (SAVI) values.
Mangrove damage and mortality associated with two strong hurricanes decreased FVC, NDVI, and SAVI,
and increased soil heat ux at the tall mangrove site located in a basin-type environment. Recovery of the
spectral characteristics, energy balance parameters and E following hurricane disturbance was quicker
in fringe mangroves than in basin mangroves. Latent heat uxes (E) were also relatively high after each
storm and may have increased as a result of increasing vapor pressure decits. Remote sensing of the
surface energy balance and E of mangrove forests can help our understanding of how these environ-
ments respond to disturbances to the landscape and the effect that these changes can have on the energy
and water budget. Moreover, relationships between energy and water balance components developed
for the coastal mangroves of southern Florida could be extrapolated to other mangrove systems in the
Caribbean to measure changes caused by natural events and human modications.
2014 Elsevier B.V. All rights reserved.

1. Introduction

Evapotranspiration (ET) is a large component of the water


cycle in subtropical and tropical wetlands (Rockstrom et al.,
1999) and can amount to 70110% of the annual rainfall in these
regions (Sumner, 1996). The strong coupling between rainfall
Corresponding author. Current address: Universities Space Research Association,
and ET in these regions can be attributed to high levels of rainfall
NASA Goddard Space Flight Center, Greenbelt, MD, United States.
Tel.: +1 301 614 6666. recycling from high rates of ET (Eltahir and Bras, 1994; Price et al.,
E-mail address: david.lagomasino@nasa.gov (D. Lagomasino). 2008). Many of the coastal subtropical and tropical wetlands are

http://dx.doi.org/10.1016/j.agrformet.2014.11.017
0168-1923/ 2014 Elsevier B.V. All rights reserved.
D. Lagomasino et al. / Agricultural and Forest Meteorology 213 (2015) 304316 305

dominated by mangrove forests, which have some of the highest has now been drained, channelized and converted to agricultural,
carbon storage among forest ecosystems, yet are disappearing at urban, and water conservation areas to support the growing popu-
a rate of 2% per year. Changes in ET induced by natural processes lation and the subsequent water demand (Light and Dineen, 1994;
such as climate change as well as anthropogenic alterations such Chimney and Goforth, 2001). Water ow into Everglades National
as land-use change can lead to regional variability and signicant Park (ENP) has been reduced to 1525% of the original inow,
changes to the water budget (Pielke et al., 2002). allowing seawater to propagate upstream in the surface water
Remote sensing has proved to be an effective tool for estimat- and intrude landward into the aquifer (Fennema et al., 1994; Price
ing ET rates and various other energy balance parameters across a et al., 2006). More than two-thirds of the 1.2 m annual rainfall in
variety of different ecosystems including: agricultural lands (Boegh south Florida occurs during the wet season between late May and
et al., 2002), terrestrial forests (Chen et al., 2005), grasslands (Jiang November, with peak rainfall occurring in June and August (Duever
and Islam, 2001) and wetlands (Chen et al., 2002; Jiang et al., et al., 1994; Childers et al., 2006). During the wet season, the Florida
2009; Anderson et al., 2011). There have been few studies ded- peninsula is prone to daily thunderstorms and tropical cyclones,
icated to modeling ET and surface energy balance in mangrove whereas in the dry season infrequent frontal systems account for
environments (Monji et al., 2002; Barr et al., 2013, 2014), despite much of the precipitation (Obeysekera et al., 1999).
a number of studies that have measured or calculated various An extensive coastal mangrove environment covers a large
biophysical parameters (i.e., leaf area index, water use efciency, portion of the southern Everglades (1445 km2 ) (Simard et al.,
spectral indices) that are useful in estimating ET (see reviews by 2006; Rivera-Monroy et al., 2011) and recent models suggest a
Green et al., 1998; Kuenzer et al., 2011). Even fewer studies have landward transgression of mangrove communities over the next
investigated the role of satellite remote sensing in modeling sur- century under various regimes of sea-level rise (Krauss et al., 2011).
face energy balance and ET in mangroves (Jiang et al., 2009). In In fact, in some southeastern areas of the Everglades, mangroves
that study, Jiang et al. (2009) used coarse resolution imagery to have migrated 1.5 km landward over the past ve decades (Ross
estimate ET across south Florida with little emphasis on vegeta- et al., 2000). The landward transgression of mangroves has been
tion community. There appears to be a gap in the literature in attributed to reduced freshwater ow caused by the compartmen-
regard to detailed surface energy and ET mapping specically in and talization of the surface water upstream that is used to support
among mangrove communities using higher resolution satellite the urban growth and development of South Florida. These anthro-
remote sensing techniques. Measurements of ET over large areas pogenic modications to the hydrology, compounded with climatic
are needed to support efforts aimed at modeling climate variability factors (e.g., sea-level rise, storm events) may ultimately alter the
at global scales, particularly in mangroves where carbon accumu- ecohydrology and water cycle dynamics of the mangrove ecotone
lation, ET, and water-use-efciency are high (Ball, 1988; Barr et al., (Rivera-Monroy et al., 2011).
2013). Two main conduits deliver water from the Everglades into
Exhaustive research has investigated energy balance and ET in Florida Bay and the Gulf of Mexico, Taylor River and Shark River
south Florida (German, 2000; Douglas et al., 2009; Schedlbauer (Fig. 1). Taylor River, the smaller of the two rivers (247 km2 ), is
et al., 2011; Barr et al., 2013, 2014). Many of the referenced located on the southeastern boundary of ENP and transports fresh-
studies utilized expensive meteorological equipment in the form water through a series of small, interconnected creek channels,
of eddy-covariance towers to measure changes to solar irradi- and discharges into the northeastern parts of Florida Bay (Olmsted
ance, temperature, wind velocity, and water vapor. Using the et al., 1980; Sutula et al., 2001) (Fig. 1). Tidal uctuations in Tay-
sophisticated equipment, surface energy balance models and ET lor River are primarily controlled by wind, as the eastern portion
models, the south Florida studies identied site-specic seasonal of Florida Bay is restricted from tidal exchange with the Gulf of
patterns with the greatest monthly average ET (56 mm day1 ) Mexico and the Atlantic Ocean (Wang et al., 1994). During the dry
occurring during the early summer (MayJuly) and the lowest ET season and the associated diminished freshwater ows, salinities
(23 mm day1 ) in early winter (German, 2000; Barr et al., 2013). in Taylor River can increase to 50 psu (Brinceno et al., 2013). Shark
Excluding the recent work by Barr et al. (2013, 2014), the major- River is the largest river (991 km2 ) and delivers water from the
ity of the measurements originate from vegetated environments northern ENP boundary to the Gulf of Mexico after moving through
outside the mangrove zone. A regional-scale evaluation of the spa- relatively large and wide tidal channels along the southwestern
tial variability of ET across a coastal mangrove community using Florida coast. Shark River experiences diurnal tidal uctuations
satellite remote sensing will provide a better understanding of the (1 m) with progressively less tidal inuence upstream of the Gulf
variability in energy and water budgets associated with mangrove of Mexico (Wdowinski et al., 2013). The largest tidal uctuations
productivity and their response to damage and recovery from trop- occur during the wet season when water levels are high because of
ical storms. increased freshwater ows (Provost, 1973; Twilley and Chen, 1998;
The present study sought to investigate the spatial and tempo- Krauss et al., 2006).
ral variability of ET along the Florida coastal Everglades mangrove Mangrove forests line the coastal areas and tidal channels of
ecotone region (EMER) using a multi-year dataset of optical satel- both Taylor and Shark Rivers. Taylor River is predominately inha-
lite images. More specically, the study estimates the variability bited by the scrub, or stunted Rhizophora mangle (red mangrove)
in ET and energy balance components between tall, riverine man- that stand 12 m tall and have relatively low above ground biomass
grove and scrub mangrove forests over time and in response to accumulation rates when compared to the much taller mangroves
strong tropical storm events. This study provides some of the rst in Shark River (Coronado-Molina et al., 2004; Ewe et al., 2006).
satellite-derived surface energy balance estimates in mangrove Along Shark River, the mangroves are a mix of three separate
ecotones and expands upon earlier site-specic mangrove studies species; Avicennia germinans (black mangrove), Laguncularia race-
by introducing spatial and temporal variability in the mangrove mosa (white mangrove), and R. mangle (red mangrove) that also
forests. occur along many coastlines of the Caribbean and South America.
Near the mouth of Shark River, all three mangrove species occur
2. Study area together and the forests contain some of the largest and tallest man-
groves in the region (Chen and Twilley, 1999; Simard et al., 2006).
The Florida Everglades once covered most of South Florida, Upstream along Shark River, the mangroves become progressively
extending from Lake Okeechobee south to Florida Bay and south- smaller and shorter and the community structure changes to pre-
west to the Gulf of Mexico (Fig. 1). Most of the original Everglades dominately red mangroves and Conocarpus erectus (buttonwood)
306 D. Lagomasino et al. / Agricultural and Forest Meteorology 213 (2015) 304316

Fig. 1. Location of Florida (A), South Florida, Lake Okeechobee and Everglades National Park (B) and mangrove study sites in Everglades National Park (C). Dashed gray boxes
indicated the locations of evapotranspiration maps in Figs. 9 and 10.

(Lugo and Snedaker, 1974; Chen and Twilley, 1999). Red mangroves used for calibration and verication of the modeled surface energy
comprise 20100% of coastal mangrove forests in the Everglades parameters.
and plant biomass decreases from 162 Mg ha1 along the coast
to 100 Mg ha1 at the inland extent of the EMER (Castaneda-Moya
3.1. Landsat 5 TM images
et al., 2013).
Landsat 5 Thematic Mapper (Landsat) data were obtained from
3. Methods the US Geological Survey-Earth Explorer database and from the
Florida International University (FIU) Geographic Information Sys-
Four sites that corresponded to long standing research sites tems database. Landsat data consist of seven bands in the visible
in the Florida Coastal Everglades: Long-Term Ecological Research [Band1 (blue: 450520 nm), Band2 (green: 520600), Band3 (red:
(FCE LTER) program were monitored. Three mangrove sites were 630690)], near-infrared (NIR) [Band4 (760900 nm), mid-infrared
selected in Shark River; SRS4 (upstream), SRS6 (downstream), and [Band5 (15501750 nm), and Band7 (20802350 nm)] and ther-
SRS-coast (coastal) (Fig. 1). The upstream site, SRS4, is located mal infrared wavelengths [Band6 (10,40012,500 nm)]. The spatial
18.2 km from the mouth of Shark River, while SRS6 is located about resolution of each band is 30 m 30 m, with the exception of the
4.1 km from the mouth (Fig. 1). These two sites, SRS4 and SRS6, thermal band, Band6, which has a resolution of 60 m 60 m.
represent the lowest and highest salinities, respectively, along the Sixteen satellite images (path: 15; row: 42) were used in the
Shark River transect. The third SRS site, SRS-coast, is located 1 km study and covered a 16-year period from 19932009. The pri-
inland from the coast and located within a basin mangrove com- mary selection criteria for the images were based on minimal cloud
munity, as characterized by Lugo and Snedaker (1974). Only one cover and the availability from the FIU databases. Satellite images
site, TS6, was monitored in Taylor River which represented the were conned to the dry season (NovemberMay) because of a
typical scrub mangrove environment located in the Taylor River lack of cloud-free images during the wet season (MayNovember)
estuary (Fig. 1). Weather towers located near TS6 and SRS6 were when convective clouds begin to form. The distribution of images
D. Lagomasino et al. / Agricultural and Forest Meteorology 213 (2015) 304316 307

throughout the dry season provided a detailed inter-seasonal vari- interactions with the canopy (Choudhury et al., 1987). Therefore,
ability and span reported minimum (NovemberDecember) and the FVC was used as a scaling factor in the equation
maximum (MayJune) surface energy balance uxes (Barr et al.,
G G
  G
 
2014). Under those circumstances, only cloud-free images from the = FVC + (1 FVC) (5)
Rn Rn veg Rn soil
dry season were used in this study and were not necessarily from
the same year or spaced at regular intervals. where (G/Rn )veg and (G/Rn )soil represent vegetation and soil
Each of the satellite images was corrected for atmospheric end-member values and equal 0.1 and 0.5, respectively (Boegh
effects using the ATCOR2 model (ReSe Applications Schlpfer, Wil, et al., 2002). Daily ET rates (mm day1 ) were calculated using the
Switzerland) in ERDAS Imagine (Hexagon Geospatial, Madison, AL). equation:
Atmospheric corrections were determined using solar geometry, a
E
visibility of 35 km and solar and thermal regional models of rural ET = (6)

and tropical conditions, respectively. After the satellite images were
corrected, spectral reectance data from each band were extracted where  equals the latent heat of vaporization of water
from a 2 2 pixel square (3600 m2 ) from each of the Shark River (2404 kJ kg1 for water with salinity of 20 at 20 C; Sharqawy
sites, SRS4, SRS6, and SRS-coast and the Taylor River site, TS6. The et al., 2010). Sensible heat ux (H) was determined from the
spectral response exhibited by each of the sites was an indicator satellite-modeled Rn values using the equation:
of the stability, biophysical status, and density of the respective G
mangrove forest. The wavelength reectance intensity from each H = Rn (7)
(1 + 1/)
Landsat band and the combinations of ratios of NIR to visible wave-
lengths were used in the satellite-based surface energy balance where is the ratio between H and E, known as Bowens ratio.
model. The extracted data from each site were averaged from each A static value of = 0.603 was used to calculate H at each site.
Landsat band and used to estimate the vegetation indices: normal- The value of 0.603 was the 5-year average (20062011) that
ized difference vegetation index (NDVI), soil-adjusted vegetation was calculated from midday values of H and E measured directly
index (SAVI) and fractional vegetation cover (FVC), as well as energy at SRS6. The numerator of Eq. (7) was also treated as a constant
balance components. The standard deviation (S.D) of the 2 2 pixel where Rn G = Rn *0.9, with the assumption that G is approximately
square was less than one reectance percentage point. 10% of Rn .
The satellite energy balance model was designed using the
model maker tool in ERDAS IMAGINE. Non-mangrove areas (e.g.,
3.2. Satellite-based energy budget and ET model
sawgrass, hardwoods, and water) were masked from each of the
satellite images using the 1999 Vegetation Map of the Everglades
Several energy balance parameters and ET rates were deter-
generated by the University of Georgia (Welch et al., 1999) and
mined using Landsat images. Instantaneous uxes were modeled
available at the FCE LTER website (http://fcelter.u.edu). Once the
for midday (10:0016:00) values. Latent heat of vaporization (E)
mask was overlaid on each image, Eqs. (1)(7) were applied to each
estimates were calculated as the residual in the surface energy bal-
atmospherically-corrected mangrove pixel which generated model
ance equation (Soegaard, 1989; Kustas et al., 1994; Boegh et al.,
estimates of FVC, and instantaneous Rn , H, G and E across Shark and
2002):
Taylor Rivers.
E = Rn H G (1)
3.3. Meteorological data
where Rn is net radiation, H is sensible heat ux, and G is soil heat
ux. The Rn component was estimated from the equation: A suite of atmospheric and meteorological measurements
was collected continuously from 2006 to 2011 from an eddy-
Rn = Rs (1 ) + RL (1 s ) s Tb4 (2)
covariance tower installed within the tall mangroves at SRS6 from
where Rs is the incoming shortwave radiation, RL is the incoming which all data was made available from the FCE LTER website
long wave radiation, is the surface albedo, s is surface emis- (http://fcelter.u.edu). Explicit details of the instrument and tower
sivity (constant at s = 0.98),  is the Stefan Boltzmans constant deployment, specications and corrections are outlined in Barr
( = 5.6703 108 J m2 s1 K4 ) and Tb is surface temperature. et al. (2010) and Barr et al. (2012). Incoming and outgoing radi-
Both Rs and RL were measured directly from the eddy-covariance ation were measured above the canopy at half hour intervals using
tower located at SRS6. Surface albedo, , was estimated using a a Kipp and Zonen CNR1 net radiometer (Bohemia, NY) at both the
combination of Landsat bands, TM4 and TM7 as outlined in Brest Shark River (height of 27 m) and Taylor River (height of 4 m) sites
(1987) and Tb was determined from the thermal band, TM6, of every half hour. Measurements of incoming solar radiation (i.e., Rs
Landsat via the ATCOR2 model (Richter and Schlapfer, 2011). and Rl ) at the SRS6 tower were used in Eq. (2) to calculate Rn at each
The soil heat ux, G, was estimated by remote sensing tech- of the Shark River study sites, SRS4, SRS6 and SRS-coast, and at the
niques described by Clothier et al. (1986) and Kustas and Daughtry Taylor River site, TS6. Air temperature, G, H, and E measured at
(1990). A soil-adjusted vegetation index (SAVI) was calculated the eddy-covariance tower at SRS6 were used to test and compare
using the following equation (Huete, 1988): the results from the satellite-derived energy balance parameters.
Latent heat (E) and H were calculated using high-frequency mea-
TM4 TM3 surements from a three-dimensional sonic anemometer (RS-50,
SAVI = (3)
TM4 + TM3 + 0.5 Gill Co., Lymington, England) and an open path infrared water vapor
gas analyzer (LI-7500, LI-COR, Inc., Lincoln, Nebraska). Soil heat ux
The SAVI value was then used to determine the fractional vege-
(G) was determined using heat ux plates (model HFT 31.1 Heat
tation cover, FVC (Ormsby et al., 1987; Purevdorj et al., 1998) using
Flux Plates, Campbell Scientic Inc. Logan, Utah) at a depth of 0.1 m
the linear equation:
below the soil surface.
FVC = 1.62 SAVI 0.37 (4) Meteorological data, such as Rn and air temperature, were also
collected at a weather tower south of TS6 (Fig. 1). Environmental
A linear relationship between G and Rn has been noted for bare data, tower and instrument specications, and deployment meth-
soils, but the (G/Rn ) for vegetation can be different because of ods for the Taylor River weather tower are described in Zapata-Rios
308 D. Lagomasino et al. / Agricultural and Forest Meteorology 213 (2015) 304316

Fig. 2. Half-hourly, daily and instantaneous (midday) net radiation (Rn ) recorded at
the tower at in Shark River versus Rn recorded at the tower in Taylor River. The gray
dashed line represents the 1:1 line.

and Price (2012). When plotted together, corresponding daily and


midday (10:0016:00) daily average Rn , and 3-year midday aver-
ages (20082010) Rn on the DOY of each Landsat image depict
slightly higher Rn at Shark River relative to Taylor River (Fig. 2).
Given these similar Rn conditions between sites, and the limited
ground-truthing available in the Everglades, data measured at the
Taylor River tower were used to validate the satellite-modeled
results of Rn from Eq. (2) at TS6. Average daily values for each year
were derived by averaging midday (10:0015:00 h) measurements
of each parameter for each day available of the 20062011 record.
Fig. 3. Reectance values of all visible and infrared Landsat bands and corrected
The daily data were then converted to a day-of-year (DOY) value to
surface temperature (Landsat Band6) through time at the (A) scrub mangrove site
generate a 5-year daily average record. The latter dataset was then TS6, and tall mangrove sites (B) SRSS-coast, (C) SRS4, and (D) SRS6.
ltered through a 13-point moving average.

level (Klimberg et al., 2010). Highly accurate models tend to have


3.4. Data analysis PFE values less than 20 (Lawrence et al., 2008).

An analysis of variance (ANOVA) was conducted to identify the 4. Results


spatial variability of the mangrove spectral response between each
of the mangrove sites. Landsat reectance values were organized 4.1. Spectral response
according to site to identify relationships and differences between
the various mangrove environments (i.e., fringe and basin) and Wavelengths in the near and mid infrared measured by Landsat
mangrove types (i.e., tall and scrub). In order to test for any sig- (Band4 and Band5) exhibited the greatest variability when com-
nicant changes over time, several t-tests were employed for each pared to bands in the visual spectrum (Band1, Band2, and Band3).
individual site comparing the means between the 1990s (n = 8) and At the Shark River sites, reectance values in Band4 ranged from a
2000s (n = 8). low of 17% at SRS-coast to a high of 44% at SRS4 (Fig. 3). The TS6
The accuracy of the energy balance models was determined site had less variability and magnitude in Band4 and only ranged
by a number of error statistics; root-mean-square-error (RMSE), between 9% and 20% (Fig. 3). Reectance data from the visible bands
normalized-root-mean-square (NRMSE), Nash-Sutcliffe, mean- (Band1, Band2, and Band3) exhibited much less variability over the
absolute-percent error (MAPE), percent forecast error (PFE) and 16-year period of the Landsat acquisitions with the most variable
bias analyses. The RMSE was used to determine the deviation visible reectance at TS6 and SRS-coast. The reectance in Band7
between the measured and model results, but the NRMSE was used reached a maximum of 2425% in May 1993 and May 2006 satel-
to normalize the RMSE to the range of energy balance parameters. lite acquisitions at SRS-coast; however, similar but smaller maxima
The NashSutcliffe index is widely used in hydrologic studies to occurred during the same periods at the other sites (Fig. 3). The ther-
calculate model efciency but can be sensitive to sample size, out- mal band, Band6, generally ranged between 10 and 33 C for each
liers, and bias (McCuen et al., 2006). Therefore a bias analysis was of the sites. Additionally, the highest temperature readings from
also recorded with the Nash-Sutcliffe index. Lastly, MAPE and PFE Band 6 tended to occur during acquisitions that had the lowest
analyses were also conducted. The MAPE provided a percent model Band4 reectance (Fig. 3).
error but could be skewed by higher percent errors when modeled At the riverine mangrove sites in Shark River, SRS4 and SRS6, the
values were larger than the actual values (Makridakis and Hibon, overall average (n = 16) spectral index values (i.e., FVC, NDVI, and
1995). The PFE has been shown to be a good model for accuracy SAVI) were nearly the same, with averages ranging between 0.81
which can estimate the next time period with a 95% condence and 0.91 for all indices (Table 1; Fig. 4). The indices remained fairly
D. Lagomasino et al. / Agricultural and Forest Meteorology 213 (2015) 304316 309

Table 1
Average values and standard deviations of fractional vegetation cover (FVC), net radiation (Rn ), soil heat ux (G), sensible heat ux (H), and latent heat ux (E) for each site.
The units for Rn , G, H and E are in W m2 .

Site FVC * 100 Rn G H E

SRS4 90.4 2.8 425.4 27.5 129.2 8.8 172.3 13 160.5 14.3
Shark River SRS-coast 64.4 7.7 512.8 25 174.6 14.8 173.5 13.1 175.5 14.5
SRS6 83.9 3.8 523.9 26.2 139.9 8 174.2 13.3 218 19.6

Taylor River TS6 16.4 5.5 487.3 33.8 227.4 17.3 177.4 13.9 104.4 9.9

consistent over the course of the 16-year Landsat dataset when S.D. away from the other sites (Table 1). Sensible heat ux, H, mea-
compared to values at SRS-coast and TS6 (Fig. 4). There were sharp surements were within one S.D of E uxes at SRS4 and SRS6. In
decreases in FVC, NDVI and SAVI (over one S.D) in March 1993 and contrast, E was signicantly larger than H at SRS6 and signicantly
May 2006 for all three Shark River sites and in particular, at the lower than H at TS6 (Table 1). The Taylor River site, TS6 experienced
SRS-coast site (Fig. 4). The spectral ratios at the SRS-coast site were the highest average G values (227 W m2 ) throughout the 16-year
about one-fth of those from SRS4 and SRS6 in March 1993 and time period and the lowest average G was measured at SRS4 and
slowly increased back to average values after 1996. In the May 2006 SRS6 (129140 W m2 , respectively) (Table 1). Sensible heat uxes
Landsat image, there was a relatively uniform drop in spectral index (H) closely mirrored Rn , accounting for slightly over 30% of the total
values across all three SRS sites (Fig. 4). The scrub mangrove site Rn at each site. The FVC values that were used to calculate G were
at TS6 exhibited a different pattern from the mangroves in Shark signicantly lower (p < 0.05) at the TS6 site when compared to the
River. Values across each of the spectral indices (e.g., NDVI, SAVI, tall, Shark River mangrove sites (Table 1). Moreover, the two river-
and FVC) steadily increased over time from 1993 to 2009 (Fig. 4A). ine sites, SRS4 and SRS6, exhibited signicantly higher FVC values
than the basin mangroves, SRS-coast and TS6.
During the 2000s, average G values were one S.D. lower than
4.2. Energy balance
those from the 1990s at all mangrove sites, with the exception at
SRS6 where the values remained constant (Fig. 5). The two largest
The satellite-modeled Rn values were similar (within one S.D.)
drops in G were found at TS6 and SRS-coast. The scrub mangrove
across most sites, except SRS4, with decadal averages ranging
site, TS6, decreased from an average of 256 W m2 in the 1990s
between 487 and 524 W m2 . The northern most site, SRS4, tended
to 199 W m2 in the 2000s. The Shark River site, SRS-coast, also
to have relatively lower Rn values (425 W m2 ) and greater than one
experienced a similar drop in G starting at 191 W m2 in the 1990s
and decreasing to 151 W m2 and SRS4 dropped from 147 W m2
to 111 W m2 (Fig. 5). There was a decrease in average G values
observed across three of the four mangrove sites from the 1990s to
the 2000s, but only the reduction in G at SRS4 was signicant at the
5% level (p = 0.05) according to t-tests (Fig. 5). The other two sites,
TS6 and SRS-coast, had p-values (t-tests) of 0.08 and 0.10, respec-
tively. Sensible heat uxes (H) remained relatively consistent over
time with modeled averages ranging between 173 and 177 W m2
(Table 1).
Net radiation (Rn ) modeled for Shark River and Taylor River sites
using Eq. (2) and Rs and RL data from the SRS6 tower provided Rn val-
ues that compared well with the actual eld-measured Rn recorded
at weather towers at SRS6 and TS6 (Fig. 6). Sensible heat ux (H)
was calculated as a function of Rn (Eq. (7)) and therefore, H values
are correlated with modeled Rn . There was also a greater devia-
tion between measured and modeled H values when compared to
Rn and E (Fig. 6). Satellite-derived G measurements were overes-
timated and were nearly an order of magnitude higher for every

Fig. 4. Landsat 5TM-derived spectral indices, Normalized Difference Vegetation


Index, Fractional Vegetation Cover and Soil-Adjusted Vegetation Index at (A) TS6, Fig. 5. Multi-year average of net radiation (Rn ), soil heat ux (G), sensible heat ux
(B) SRS-coast, (C) SRS4, and (D) SRS6. The gray and black dashed lines indicated the (H), latent heat of evaporation (E) and FVC*100 for the 1990s and the 2000s for
arithmetic mean (n = 16) for SAVI and FVC, respectively. each site. Error bars denote one standard deviation (n = 8).
310 D. Lagomasino et al. / Agricultural and Forest Meteorology 213 (2015) 304316

Fig. 6. Measured versus modeled scatter plots of (A) net radiation (Rn ), (B) soil
heat ux (G), (C) sensible heat ux (H) and (D) latent heat (E). Gray dashed line
represents a 1:1 ratio.

Landsat acquisition date than the average, eld-measured values


at SRS6 (Fig. 6).
The modeled average values for Rn , H and E were in agreement
with the eld averages despite the variability between particular
dates (Fig. 6). The accuracy of the satellite-derived energy balance
model was tested using the measured parameters at the eddy-
covariance tower at SRS6. The modeled data had a RMSE of ranging
between 11.8 W m2 for G and 53.0 W m2 for Rn at SRS6 (Table 2). Fig. 7. Satellite-based latent heat ux (E) in W m2 for each image acquisition at
The NashSutcliffe index was highest (0.710.88) for modeled Rn (A) TS6, (B) SRS-coast, (C) SRS4, and (D) SRS6.
and E values and lowest for G (12.6), while the MAPE was lowest
for Rn and E and highest for G but occurred with an overestimate but had a lesser effect on E at the Taylor River and SRS-coast sites
bias of 113 (Table 2). The PFE indicated that most of the models (R2 < 0.37) (Fig. 8A). The strong relationship between Rn and E at
were highly accurate with PFE values below 20% for all the models the Shark River sites was most likely related to the spatial vege-
except for G. tation patterning between the Shark and Taylor River mangrove
environments. The Shark River sites, SRS4 and SRS6, were predo-
4.3. Latent heating variability minately and homogenously covered by mangroves at the Landsat
pixel scale (30 30 m grid), which was suggested by the high (>0.8)
Latent heat ux (E) modeled from Landsat images ranged spectral index values and the small range (0.751.1) of FVC values
between 120 and 350 W m2 at the Shark River sites and between (Fig. 8B). At the Taylor River site, FVC values were much lower (<0.6)
50 and 160 W m2 at the Taylor River site (Fig. 7). The averaged E and much more variable exhibiting a wider range (0.3 to 0.6) of
were highest at SRS6 (218 W m2 ) and lowest at TS6 (104 W m2 ) FVC values suggesting that each Landsat pixel was a combination
(Table 1). Over the 16-year period of satellite acquisitions, E of both vegetation and open water because of the smaller man-
remained relatively consistent through the 1990s and became more grove basal area (2.36.3 m2 ha1 ; Coronado-Molina et al., 2004)
variable in the 2000s (Fig. 6). In contrast, E at SRS6 exhibited a relative to the larger basal area at SRS4 and SRS6 (10 m2 ha1 and
different decadal variability with the highest ux occurring from 40 m2 ha1 respectively; Chen and Twilley, 1999). The poor rela-
1992 to 1995 and the again in 2006 and 2008 (Fig. 7). Modeled E tionships between Rn and E at SRS-coast and TS6 may be caused by
at the Taylor River site was always lower than E at any of the Shark the prolonged inundation that could facilitate energy removal from
River sites (e.g., SRS4, SRS6, and SRS-coast). When the Landsat data the system via water ow. Mixed pixels may also occur at SRS-coast
were separated into decades (i.e., 1990s and 2000s) there was lit- directly after strong storm events, which were suggested by the
tle change in the averaged model E between SRS4 and SRS6 over lower FVC as a result of the mixing of background reectance from
time. In contrast, there was an increase in average E at SRS-coast soils or reectance from damaged, defoliated mangroves (Fig. 8B).
and TS6; increasing from 110 to 160 W m2 at SRS-coast and from Anthoni et al. (2000) reported similar ndings and indicated that
65 to 150 W m2 at TS6. upwelling longwave radiation is higher within sparse canopies than
closed canopies and therefore, Rn , and the subsequent available
5. Discussion energy tends to be lower in sparse canopy cover.
The satellite-derived energy balance model estimates were in
5.1. Surface energy balance variability: tall versus scrub agreement with the measured values in Taylor and Shark River,
mangroves except for G values (Table 2). Net radiation (Rn ), H, and E estimates
each had MAPE of 10 or lower, and PFE values all under 20 suggest-
Net radiation (Rn ), G, H and E varied between each of the man- ing the satellite model was a good predictor of the average observed
grove sites. Particular patterns were evident when comparing the conditions. Soil heat uxes (G) were always overestimated by an
tall and scrub mangroves. Net radiation (Rn ) was strongly corre- order of magnitude, which was one of the major limitations of the
lated with E at the Shark River sites SRS4 and SRS6 (R2 > 0.76), satellite-derived model. Nonetheless, despite the errors in G values
D. Lagomasino et al. / Agricultural and Forest Meteorology 213 (2015) 304316 311

Table 2
Error analyses table for measured versus modeled values energy balance parameters at TS6 and SRS6.

Error analysis Formula TS Rn SRS Rn SRS G SRS H SRS E SRS6/TS6

PFE (%) (2 se /Yt+1 ) 100 13.0 11.6 21.6 15.6 19.5 11.6
MAPE (%) ((|Yt Ft |)/Yt )/n 5.84 10.2 80.3 19.8 8.21 43.5
Nash-Sutcliffe (Y Ft )2 /(Yt Yavg )2 0.79 0.71 12.6 0.21 0.88 0.24
t
RMSE (W m2 ) (Yt Ft )2 /n 55.0 53.0 11.8 46.4 29.4 64.1
NRMSE RMSE/(Ymax Ymin ) 11.5 13.8 8.55 29.6 12.2 18.5
Bias (W m2 ) (Yt Ft )/n 21.9 26.0 113 -0.11 1.31 155

PFE = Percent Forecast Error; MAPE = Mean Absolute Percent Error; RMSE = Root Mean Square Error; NRMSE = Normalized RMSE; CVRMSE = Coefcient of Variation RMSE. Yt
are the measured values, Ft are the modeled values, Yt+1 , is the next forecasted value, Yavg is the mean of the measured values, Ymax and Ymin are the maximum and minimum
of the measured values, respectively, n is the number of samples and se is the standard error of the measured values.

the estimations of E were still highly accurate. The over estimation the numerator in Eq. (7) were kept constant throughout each of
of G in combination with accurate residual E calculations suggests the Landsat images and could cause over or underestimates of H,
that energy was also dissipated by other factors, most likely from depending on the time of year. Despite the static values used in
water and tidal actions, as well as changes in water vapor pressure the calculations, there was still relatively good agreement between
decits (Barr et al., 2014). measured and modeled H values (Table 2; Fig. 6).
The Bowens ratio value of 0.603 used in the satellite-based sur- The spatial differences between the Shark and Taylor River
face energy balance equations implies that sensible heat should study areas were manifested in the FVC, EVI, and NDVI spectral
always be below the uxes from E. However, this was not the indices calculated from Landsat images (Fig. 4). Satyanarayana et al.
case at all the sites. Meanwhile, average E was always greater (2011) measured similar responses in NDVI, using multispectral
than H at the two riverine mangrove sites, SRS4 and SRS6. Sen- data from the Quickbird satellite, from varying mangrove commu-
sible heat ux (H) values were lower (>1S.D) in the 2000s when nities; larger, healthier sites generally had higher NDVI values and
compared to the 1990s at SRS-coast and TS6. The lower H values at environmentally stressed sites typically had lower NDVI values. The
SRS-coast and TS6 coincided with a similar decrease in G over the highest variability in FVC was exhibited at SRS-coast and could be a
same time period (Fig. 5). The different relationships between H function of storm related damage and the location of the mangroves
and E at the riverine-type and basin-type mangroves suggest that in a basin-type environment (Table 1). In combination with the
there is less energy available for E in the basin-type mangroves lower FVC values, TS6 also exhibited higher G values than the Shark
because more energy may be transferred to soil and water. Yet E River mangrove sites (Fig. 5). Soil heat ux (G) has been shown to
modeled at SRS-coast resembles E rates at SRS4 and SRS6. Bowens be sensitive to changes in vegetation cover because as vegetation
ratio values around SRS6 can range between 0.3 in the wet season covers and shades more ground, it dampens the incoming solar
and 1.0 in the dry season (Barr et al., 2012). The eddy-covariance energy, and effectively attenuates the temperature transfer to the
tower measures surface energy balance parameters without energy soil and, thereby, reduces G (Yang et al., 1999). Therefore, the higher
balance closure, whereas, the satellite-based models assumes an G estimated at the Taylor River site during the entire 16-year study
energy closure to estimate E. period suggest that there was more energy transferred to the soils
Modeled H values were greater than one S.D higher than E val- and most likely, the water than at the taller and heavily vegetated
ues at TS6 and greater than one S.D. lower than E at SRS6 (Table 1). sites of Shark River. The limited tidal ushing, continual inundation,
The magnitude of H values (100250 W m2 ; Fig. 6) across all the and the low FVC in the Taylor River scrub mangrove forests play a
study sites tend to be higher than typical H values found in subtrop- major role in surface energy processes, and perhaps, in conjunction
ical and tropical forests (1030 W m2 ; da Rocha et al., 2004) and with the higher water use efciency, limit ET.
subtropical wetlands (2070 W m2 ; Schedlbauer et al., 2011). The Latent heat (E) calculated for the current study derived from
higher H values reported in the current study are an average of mid- spectral data collected from Landsat with minimal inputs from eld
day values, while the other studies reported daily values (da Rocha data (i.e., Rs and RL ) measured at SRS6 agreed well with the eld
et al., 2004; Schedlbauer et al., 2011). The discrepancy between measured Rn values at SRS6. In addition, Rn values calculated at
the modeled H uxes in the present study could arise from the TS6 (using Rs and RL inputs from the weather tower near TS6)
calculation used to determine H at each site. Bowens ratio and also coincided with eld-measured Rn (Fig. 6A). Modeled E values

Fig. 8. (A) Net radiation (Rn ) versus latent heat (E) and (B) fractional vegetation cover (FVC) versus latent heat plots for each site.
312 D. Lagomasino et al. / Agricultural and Forest Meteorology 213 (2015) 304316

Fig. 10. Modeled latent heat (E) values in the scrub mangroves in southern Florida
near Taylor River in (A) post-Hurricane Andrew conditions (March 1993 acquisi-
tion) and in (B) steady-state conditions (November 2008). Mangrove study sites are
marked with black diamonds. The location of the mapped area is dened in Fig. 1.

Fig. 9. Modeled latent heat (E) values in the tall mangroves in southwestern Florida
near Shark River in (A) post-Hurricane Andrew conditions (March 1993 acquisition)
and in (B) steady-state conditions (November 2008). Trajectories of tropical storms 5.2. Forest latent heat comparisons
are shown in Fig. 10A. Mangrove study sites are marked with black diamonds. The
location of the mapped area is dened in Fig. 1.
The present study has provided some of the rst satellite-
based surface energy balance models for mangrove forests. As the
results indicate, E values can vary signicantly (104218 W m2 ;
Table 1) in mangrove forests based upon canopy cover and man-
generally exhibited more spatial variability on the southwestern grove height. Complications in estimating E in mangrove forests
coast near Shark River than on the southern coast near Taylor River arise from added physical forcings such as inundation, tidal ushing
(Figs. 9 and 10). The southwest coast is primarily composed of and storm activity. Nevertheless, the results from the present study
homogenous stands of mangroves over 8 m tall with large basal provide E estimates that are comparable among other forested
areas, whereas the mangroves in around Taylor River are predom- ecosystems (Table 3). The E estimated in the present study were
inantly less than 4 m and spatially discontinuous (Simard et al., made as instantaneous calculations with respect to midday val-
2006). The relatively homogenous modeled E in the scrub man- ues (10:0016:00), and therefore, reect near-maximum rates. In
groves around Taylor River was most likely related to the structure addition, the satellite images were acquired on cloud free days,
and low stature of the mangroves and the relatively stable shal- which in turn, represented ideal energy conditions. Estimated
low groundwater geochemical conditions as compared to the more E were within the range of instantaneous midday E measured
variable conditions along Shark River (Zapata-Rios and Price, 2012; from eddy-covariance data (Barr et al., 2014) and satellite data
Lagomasino et al., 2014) (Fig. 10). In addition, Lagomasino et al. (Zhaira et al., 2009) in mangrove forests in the Everglades and
(2014)) reported that the spectral signature of the scrub man- India, respectively (Table 3). Field measurements from mangrove
grove leaves remain consistent throughout the year, and suggested forests in Thailand were slightly higher (200425 W m2 ; Monji
that the stable, shallow groundwater chemistry may lead to per- et al., 2002). When compared to other forest types, mangroves are
sistent biophysical stress. Similarly, Hao et al. (2009) suggested on par with satellite-based E estimates from coniferous forests
the lowered photosynthetic activity in mangroves was caused by (100180 W m2 ; Venalainen et al., 1999) and within the wide
higher interstitial soil salinity and lack of tidal ushing while higher range of temperature deciduous forests (40400 W m2 ; Baldocchi
carbon-13 isotope values (Lin and Sternberg, 1992) and the increase and Wilson, 2001) (Table 3). Lastly, mangrove forests in Florida fall
in water use efciency (Lovelock and Feller, 2003) in scrub man- just short of Mediterranean evergreens (223393 W m2 ; Zahira
groves also support the notion that scrub mangroves are almost et al., 2009) and tropical rain forests (260310 W m2 ; Malhi et al.,
continuously under stressed conditions. 2002).

Table 3
Comparison of midday values of satellite-modeled and eld-measured latent heat ux (E; W m2 day1 ) for various types of forested vegetation communities.

Biome Location Satellite models Reference Field data Reference

Florida 104218 Present study 115231 Barr et al. (2014)


Mangrove Forests Thailand 200425 Monji et al. (2002)
India 125190 Ganguly et al. (2008)
Mediterranean Evergreen Algeria 223393 Zahira et al. (2009)
Tropical Rain Forest Brazil 260310 Malhi et al. (2002)
Coniferous Forest Sweden 100180 Venalainen et al. (1999)
Temperate Deciduous Forest Tennesse 40400 Baldocchie and Wilson (2001)
D. Lagomasino et al. / Agricultural and Forest Meteorology 213 (2015) 304316 313

5.3. Change over time damage was caused to island and fringe-type mangroves after the
passage of Hurricane Andrew (Smith et al., 2009). The local mini-
Hydrogeomorphic and biophysical features along with restora- mums in FVC, NDVI, and SAVI measured from the Landsat images
tion efforts between Shark River and Taylor River area may provide March 1993 and May 2006 suggest signicant damage to the man-
insight into the long-term changes to the energy balance parame- groves at each of the Shark River sites following hurricanes Andrew
ters and ET in the Everglades mangrove region. The scrub mangrove and Wilma with the greatest detriment occurring at SRS-coast, as
site, TS6, had an increase in averaged, satellite-based E from the indicated by the much lower index values (Fig. 4C). Barr et al. (2012)
1990s to the 2000s. The increase in E at TS6 was in conjunction presented a similar spectral index (i.e., EVI) response after Hurri-
with an increase in FVC and a decrease in G (Figs. 4A and 7A). The cane Wilma using MODIS. The relatively quick recovery (1 year)
cause of these trends over the recent decades may be linked to the suggested by the Landsat data was also mirrored in the MODIS EVI
growth and expansion of the scrub mangrove patches in the Taylor data where EVI values return to pre-Wilma conditions approxi-
River area, as previously reported in the mid-1990s by Ross et al. mately 1 year after the storm (Barr et al., 2012). The drop in spectral
(2000) and more recently, Rivera-Monroy (personal communica- vegetation indices, FVC, NDVI and SAVI, recorded in acquisitions
tion 2013). following hurricanes Andrew (1992) and Wilma (2005) was most
The increase in FVC observed at TS6 between the 1990s and the likely caused by a combination of defoliation and tree mortality
2000s suggests more vegetation shading of the water and soil, and from strong winds during the hurricane and the subsequent storm
more homogenization of the Landsat pixels, and in turn less energy surge.
was transferred to water and soils in the recent decade. Variability Despite the surveys conducted after Hurricane Andrew identify-
in E over time is closely associated with sharp decreases to FVC ing most of the damage areas in fringe and island mangroves (Smith
and G in satellite images collected after strong tropical storm events et al., 2009), there was very little change in spectral values at SRS4
(Fig. 7). Increased E after tropical storms were above 250 W m2 and SRS6, both fringe mangrove sites, as opposed to the damage
at SRS6 and above 150 W m2 at SRS5. The storms, however, did done in the basin mangroves at SRS-coast (Fig. 4). The eye wall of
not play a clear role at the SRS-coast and TS6 sites. This observa- Hurricane Andrew passed just north of the SRS4 and SRS6. Storm
tion was most likely a result of different environmental impacts surge (14 m) related to the southern eyewall ooded the man-
to these specic regions. At SRS-coast there was more tree mor- grove wetlands up to 10 km inland and deposited up to 2050 cm of
tality/defoliation relative to SRS4 and SRS6 as suggested by low carbonate-rich muds (Risi et al., 1995; Swiadek, 1997). The result-
FVC values (<0.5) and documented by surveys (Smith et al., 1994). ing surge and reduced tidal ushing in the basin mangroves around
Meanwhile, the change in E at TS6 was incremental and increased SRS-coast may be the cause of the increased mangrove mortality
with increasing FVC (Fig. 7). at SRS-coast relative to SRS4 and SRS6 and the reason there was a
more signicant drop in FVC and other spectral indices at SRS-coast.
5.4. Physical forcings on energy balance In addition to the spectral response, the modeled G values from
SRS-coast and TS6 exhibited a decrease from the 1990s to the 2000s
South Florida is prone to tropical cyclone activity and has had (Fig. 5). Occurring in tandem with the lower average G values in
a number of tropical storms and hurricanes pass over or near the the 2000s, average E increased over the same time period. The
study area during the 16-year time span of the Landsat data. Three decrease in G at SRS-coast and TS6 can be attributed to an increase
hurricanes during that time period of are particular interest because in vegetation growth and mangrove recovery following Hurricane
they were signicant storms and coincide closely with Landsat Andrew. Reectance values in TM1 and TM3 also decrease from
image acquisitions: hurricanes Andrew (1992), Katrina (2005) and 1993 to 2009 at the TS6 site suggesting increased absorption from
Wilma (2005). Weaker tropical storms including Harvey (1999), photosynthetic pigments caused by new leaf growth in the scrub
Irene (1999), Ernesto (2006) and Fay (2008) also affected the study mangroves at Taylor River (Fig. 3A). Recently, researchers have
area. measured an increase in above ground net primary production of
Mangroves are well adapted and resilient against storm condi- mangroves from 1800 g C m2 yr1 in 2001 to 5800 g C m2 yr1
tions to some degree. After strong storms, larger mangroves with in 2004 at a site near TS6 (Ewe et al., 2006) as well as an increase
a diameter-breast-height (DBH) greater than 5 cm have a signi- in scrub red mangrove patch sizes from 2003 to recent (Rivera-
cantly higher chance of mortality than smaller mangroves (Smith Monroy, personal communication, 2013). The combination of a
et al., 1994; Risi et al., 1995). Additionally, Smith et al. (1994) steady increase in FVC and steady decrease in G at TS6 can be
documented a nearly 50% increase in mangrove mortality rates directly linked to increased scrub mangrove production as individ-
months and years after the passing of Hurricane Andrew (category ual mangrove patches have grown larger. River discharge has not
4) through South Florida in August 1992. The observation of ini- shown any signicant increase since the inception of the Taylor
tial tree mortality directly after the storm followed by a continued River Slough restoration despite recent studies indicating a strong
increase in post-storm mangrove mortality, associated with storm link between water ushing times and freshwater input (e.g., local
surge damage and salinization, may be the cause of the higher mid- management practices) (Sandoval, 2013). Therefore, the increase
NIR response (Fig. 3) and lower vegetation index response (Fig. 4) in E and FVC at TS6 may be related to short-term changes in sea-
noted in this study at SRS-coast. Fractional vegetation cover (FVC), level that have been more conducive to the growth and expansion
NDVI, and SAVI values were at their lowest values at SRS-coast of mangroves (e.g., nutrient inputs).
in the 3 years following Hurricane Andrew. Likewise, the spec- Sharp increases in satellite-based E were identied after storm
tral indices were also lowest at the TS6 directly after Hurricane events (1993 and 2005) at the SRS6 site (Fig. 7). Latent heat ux (E)
Andrew. Below average FVC, NDVI and SAVI values also occur across estimated at SRS4, SRS-coast, and SRS6 sites were relatively equal
all three Shark River sites in the May 2006 Landsat image, which throughout the study period, except for time periods directly after
was acquired a few months after the passage of hurricanes Katrina strong storm events when satellite-based E at SRS6 were nearly
and Wilma in 2005. double the rates at SRS4 and SRS-coast. A possible explanation for
Previous studies have indicated that the structure of mangroves the higher E at SRS6 directly following hurricanes, even though
stands can inuence the disturbance effects caused by storms our results show no direct evidence of this, could be from increased
(Ward et al., 2006). Surveys conducted after hurricane Andrew nutrients that were brought onshore by the storm surge associated
and hurricane Wilma identied higher mangrove mortality rates in with the passage of hurricanes Andrew and Wilma. Phosphorus-
basin-type mangroves after Hurricane Wilma, but relatively more limited mangrove environments, like in the Shark River and Taylor
314 D. Lagomasino et al. / Agricultural and Forest Meteorology 213 (2015) 304316

River portions of the Everglades, have been shown to have low leaf and most likely reect the increase in mangrove basal area (Ross
water potential, low stomatal conductance, photosynthetic carbon- et al., 2000; Rivera-Monroy, personal communication) (Fig. 10B).
assimilation rates, and less conductive xylem (Lovelock et al., 2006).
The water stress factors noted by Lovelock et al. (2006) were mit-
6. Conclusions
igated when more P was added to the system. The storm deposits
from Hurricane Wilma showed a relative increase in Ca-bound P
The Everglades have endured numerous modications from
(2025% of total P pool) compared to the composition from the
both human engineering (e.g., canals, levees) and natural events

top soil (Castaneda-Moya et al., 2010). A similar increase in E
(e.g., tropical storms, sea-level rise). Long-term datasets acquired
occurred in the April 2008 image (Fig. 7). The elevated E estimated
from the NASA Landsat mission and the FCE LTER project have
in the April 2008 images is associated with elevated Rn values that
provided insight into the spatial and temporal variability of E
were 1.5 times greater than the average measured values for that
and energy balance components in the Everglades, and more
particular day. This misestimate may be a result of modeling param-
specically, how the mangrove ecotone has responded to multi-
eters, as there was no signicant change in the spectral reectance
ple hurricane events and restoration projects during the recent
nor the spectral indices for April 2008 (Figs. 3 and 4).
decades. Decreases in NDVI, SAVI, and FVC measured from Landsat
Moreover, E has also been linked to changes in vapor pres-
images after hurricane events identied mangrove mortality that
sure decits (Barr et al., 2014). In that study, Barr et al. (2014)
was directly related to damage from strong winds and storm surge.
measured low VPD during the dry season relative to the doubled
Modeled G values in the scrub mangroves decreased from the 1990s
VPD during the wet season. The opening of the mangrove forest
to the 2000s indicating new growth and regrowth to areas around
canopy from defoliation and tree mortality after strong storms
Taylor River in the coastal Everglades. Latent heat (E) remained
(e.g., Hurricanes Andrew, Wilma, and Katrina) can lead to reduc-
relatively constant across all sites with the exception of SRS6 that
tions in H, enabling more energy to be partitioned into E, as
exhibited higher ET rates following strong hurricanes. Minimum E
seen in elevated wet and dry season E values directly after Hurri-
were found in the scrub mangroves while maximum uxes were
canes Katrina and Wilma in 2005 (Barr et al., 2012). Similar results
obtained from the tall mangroves. There was also a decrease in G
were identied in the present study (Fig. 7). More importantly,
and an increase in FVC in the scrub mangroves over time that was
using remote sensing techniques, an extrapolation of E values
also accompanied by a rise in E as a result of increases in biomass
across the mangrove ecotone has been presented and has provided
and the greater available energy over the same time period. Despite
detailed information about the spatial variability in surface energy
some limitations to the satellite surface energy model related to
processes in mangrove ecosystem beyond the single site measure-
physical forcings, the close comparisons among the site-specic
ments.
energy balance parameters and values from the literature suggest
The mangrove forests have been identied as having charac-
that there are other exports of energy from the environment from
teristics of semi-arid regions ( > 1.0) during the dry season and
water heat ux and tidal uctuations.
broadleaf deciduous forests ( < 0.5) during the wet season (Barr
Additional Everglades-related restoration projects that are
et al., 2014). In lieu of this, and in addition to the VPD variability
expected to affect both Shark and Taylor River have been completed
and tidal uctuations, the results presented here are preliminary
or planned in the years since the last Landsat image (2008) used in
estimates of surface energy balance modeling for mangrove forests.
the present study. Continued monitoring through remote sensing
These satellite-based surface balance components estimated by the
can provide better regional understanding of how the mangrove
present study provide insight to regional surface energy dynamics
environment responds to natural and anthropogenic intervention.
that can be overlooked at the eld site scale.
Moreover, linking changes to the hydrogeological, meteorological
The regional maps of E that were generated using Landsat
and biophysical conditions in the coastal mangroves in tropical and
imagery demonstrated the variability of E across the mangrove
subtropical coastal wetlands are important for global water, energy
ecotones (e.g., tall versus scrub mangroves) and each ecotones
and carbon budgets. Moreover, this study highlights the vari-
E response to strong tropical storms and steady-state or typical
ability in surface energy across various mangrove environments
conditions for the Shark River area (Fig. 9) and Taylor River area
and biomass productivity while at the same time illustrates the
(Fig. 10). The southwestern mangroves in the Shark River vicin-
accuracy, as well as the uncertainty, in estimating surface energy
ity exhibited high spatial variability in E values images collected
balance in complex, tidally inuenced ecosystems. The inclusion of
within 13 years after the storm (Fig. 10A). Latent heat values (E)
older images from the long-term Landsat archive (1970s2000s)
were highest in the islands around SRS6 and along Shark River,
and the new acquisitions from Landsat 8 as part of the Landsat
whereas to the north, E values were lower by 58 W m2 . Many
Continuity Mission (2013present) will undoubtedly provide cru-
of the mangroves north of Shark River are separated from tidal
cial spatial and temporal diagnostics of changes to E, the energy
channels and were also along the trajectory of the tropical storms
balance and biophysical conditions that may not be detected at
(Fig. 9A). Latent heat (E) maps from the area identied two man-
the eld-scale. Energy balance modeling in the coastal Everglades,
grove responses after storms; 1) mortality and defoliation from
though difcult because of many physical forcings, has the potential
storm surge and high winds (Smith et al., 1994) as suggested by
to be used in other similar environments throughout the Caribbean
the low FVC values (<0.5), and 2) and elevated E values in hot-spot
where data collection is minimal and could provide for better inter-
areas. Under more meteorologically stable conditions, for instance
pretations of the region.
the November 2008 map, indicates more homogeneity in E val-
ues across the mangroves most likely related to forest regrowth
and recovery (Fig. 9B). Acknowledgements
The E values for mangroves along the southern coast around
Taylor River remained generally consistent throughout the region This material was supported directly by the National Science
and for the most part were independent of the storm condi- Foundation through the Florida Coastal Everglades Long-Term
tions (Fig. 10). In the March 1993 image collected after Hurricane Ecological Research program (FCE-LTER) under Grant No. DBI-
Andrew, the modeled E values were low (180260 W m2 ) 0620409 and DEB-1237517 and from the National Aeronautics
across the scrub mangrove region (Fig. 10A). Fifteen years later and Space Administrations (NASA) Water Science of Coupled
in November 2008 the E values in the scrub mangroves have Aquatic Processes in Ecosystems from Space (WaterSCAPES) Uni-
increased, yet still have little spatial variability (300360 W m2 ) versity Research Center program under Grant No. NNX-10AQ13A.
D. Lagomasino et al. / Agricultural and Forest Meteorology 213 (2015) 304316 315

Additional support was provided by Everglades National Park and Fennema, R.J., Neidrauer, C.J., Johnson, R.A., MacVicar, T.K., Perkins, W.A., 1994. A
the Florida Education Fund McKnight Dissertation Year Fellowship. computer model to simulate natural everglades hydrology. In: Davis, S., Ogden,
J.C. (Eds.), Everglades: The Ecosystem and its Restoration. St. Lucie Press, Boca
This is SERC contribution number 701. Raton, Florida, pp. 249289.
Ganguly, D., Dey, M., Mandai, S.K., De, T.K., Jana, T.K., 2008. Energy dynamics and
References its implication to biosphere-atmosphere exchange of CO2 , H2 O and CH4 in a
tropical mangrove forest canopy. Atmos. Environ. 42, 41724184.
Anderson, M.C., Kustas, W.P., Norman, J.M., Hain, C.R., Mecikalski, J.R., Schultz, German, E.R., 2000. Regional evaluation of evapotranspiration in the Everglades. US
L., Gonzalez-Dugo, M.P., Cammalleri, C., dUrso, G., Pimstein, A., et al., 2011. Department of the Interior, US Geological Survey, Water-Resources Investiga-
Mapping daily evapotranspiration at eld to continental scales using geo- tions Report 004217.
stationary and polar orbiting satellite imagery. Hydrol. Earth Syst. Sci. 15, Hao, G.Y., Jones, T.J., Luton, C., Zhange, Y.J., Manzane, E., Scholz, F.G., Bucci, S.J., Cao,
223239. K.F., Goldstein, G., 2009. Hydraulic redistribution in dwarf Rhizophora mangle
Anthoni, P.M., Law, B.E., Unsworth, M.H., Vong, R.J., 2000. Variation of net trees driven by interstitial soil water salinity gradients: impacts on hydraulic
radiation over heterogeneous surfaces: measurements and simulation in a architecture and gas exchange. Tree Physiol. 29, 697705.
junipersagebrush ecosystem. Agric. For. Meteorol. 102 (4), 275286. Huete, A.R., 1988. A soil-adjusted vegetation index (SAVI). Remote Sens. Environ. 25,
Baldocchi, D.D., Wilson, K.B., 2001. Modeling CO2 and water vapor exchange of a 295309.
temperate broadleaved forest across hourly to decadal time scales. Ecol. Model. Jiang, L., Islam, S., 2001. Estimation of surface evaporation map over Southern Great
1-2, 155184. Plains using remote sensing data. Water Resour Res. 37, 329340.
Ball, M.C., 1988. Ecophysiology of mangroves. Trees 2, 129142. Jiang, L., Islam, S., Guo, W., Singh, Jutla, A., Senarath, S.U.S., Ramsay, B.H., Eltahir,
Barr, J.G., Engel, V., Fuentes, J.D., Zieman, J.C., OHalloran, T.L., Smith, T.J., Anderson, E., 2009. A satellite-based daily actual evapotranspiration estimation algorithm
G.H., 2010. Controls on mangrove forest-atmosphere carbon dioxide exchanges over South Florida. Global Planet. Change 67, 6277.
in western Everglades National Park. J. Geophys. Res.: Biogeosci. 115 (G2), Klimberg, R.K., Sillup, G.P., Boyle, K.J., Tavva, V., 2010. Forecasting performance meas-
G02020. ures what are their practical meaning? Adv. Business Manage. Forecast. 7,
Barr, J.G., Engel, V., Smith, T.J., Fuentes, J.D., 2012. Hurricane disturbance and recovery 137147.
of energy balance, CO2 uxes and canopy structure in a mangrove forest of the Krauss, K.W., Doyle, T.W., Twilley, R.R., Rivera-Monroy, V.H., Sullivan, J.K., 2006. Eval-
Florida Everglades. Agric. For. Meteorol. 153, 5466. uating the relative contributions of hydroperiod and soil fertility on growth of
Barr, J.G., Fuentes, J.D., DeLonge, M.S., OHalloran, T.L., Barr, D., Zieman, J.C., 2013. south Florida mangroves. Hydrobiologia 536, 311324.
Summertime inuences of tidal energy advection on the surface energy balance Krauss, K.W., From, A.S., Doyle, T.W., Doyle, T.J., Barry, M.J., 2011. Sea-level rise
in a mangrove forest. Biogeosciences 10, 501511. and landscape change inuence mangrove encroachment onto marsh in the ten
Barr, J.G., DeLonge, M.S., Fuentes, J.D., 2014. Seasonal evapotranspiration patterns in thousand islands region of Florida, USA. J. Coast. Conserv. 15, 629638.
mangrove forests. J. Geophys. Res.: Atmos., 119. Kustas, W.P., Daughtry, C.S.T., 1990. Estimation of the soil heat ux/net radiation
Boegh, E., Soegaard, H., Broge, N., Hasager, C.B., Jensen, N.O., Schelde, K., Thomsen, A., ratio from spectral data. Agric. For. Meteorol. 49, 205223.
2002. Airborne multispectral data for quantifying leaf area index, nitrogen con- Kustas, W.P., Perry, E.M., Doraiswamy, P.C., Moran, M.S., 1994. Using satellite remote
centration, and photosynthetic efciency in agriculture. Remote Sens. Environ. sensing to extrapolate evapotranspiration estimates in time and space over a
81, 179193. semiarid Rangeland basin. Remote Sens. Environ. 49, 275286.
Brest, C.L., 1987. Seasonal Albedo of an urban/rural landscape from satellite obser- Lagomasino, D., Price, R., Whitman, D., Campbell, P., Melesse, A., 2014. Estimating
vations. J. Appl. Meterol. Climatol. 26, 11691187. estuarine hydrogeochemistry using leaf and satellite reectance in two coastal
Brinceno, H., Miller, G., Davis, S.E., 2013. Relating freshwater ow with estuarine mangrove communities. Remote Sens. Environ. 154, 202218.
water quality in the southern Everglades mangrove ecotone. Wetlands. Lawrence, K., Klimberg, R., Lawrence, S., 2008. Fundamentals of Forecasting using

Castaneda-Moya, E., Twilley, R.R., Rivera-Monroy, V.H., Zhang, K., Davis, S.E., Ross, Excel. Industrial Press, New York, NY, pp. 5961.
M., 2010. Sediment and nutrient deposition associated with Hurricane Wilma Light, S.S., Dineen, J.W., 1994. Water control in the everglades: a historical per-
in mangroves of the Florida coastal Everglades. Estuaries Coasts 33, 4558. spective. In: Davis, S., Ogden, J.C. (Eds.), Everglades: The Ecosystem and its

Castaneda-Moya, E., Twilley, R.R., Rivera-Monroy, V.H., 2013. Allocation of biomass Restoration. , pp. 4784.
and net primary productivity of mangrove forests along environmental Lovelock, C.E., Feller, I.C., 2003. Photosynthetic performance and resource utilization
gradients in the Florida Coastal Everglades, USA. For. Ecol. Manage. 307, of two mangrove species coexisting in a hypersaline scrub forest. Oecologia 134,
226241. 4, 455-462.
Chen, R., Twilley, R., 1999. Patterns of mangrove forest structure and soil nutri- Lovelock, C.E., Feller, I.C., Ball, M.C., Engelbrecht, B.M.J., Ewe, M.L., 2006. Differences
ent dynamics along the Shark River estuary, Florida. Estuaries Coasts 22 (4), in plant function in phosphorus and nitrogen limited mangrove systems. New
955970. Phytologist 172 (3), 514522.
Chen, J., Kan, C., Tan, C., Shih, S., 2002. Use of spectral information for wetland Lugo, A.E., Snedaker, S.C., 1974. The ecology of mangroves. Annu. Rev. Ecol. Syst. 5,
evapotranspiration assessment. Agric. Water Manage. 55, 239248. 3964.
Chen, J.M., Chen, X., Ju, W., Geng, X., 2005. Distributed hydrological model for map- Malhi, Y., Pegorano, E., Nobre, A.D., Pereira, M.G.P., Grace, J., Culf, A.D., Clement,
ping evapotranspiration using remote sensing inputs. J. Hydrol. 305, 1539. R., 2002. The energy and water dynamics of a central Amazonia rain forest. J.
Childers, D.L., Boyer, J.N., Davis, S.E., Madden, C.J., Rudnick, D.T., Sklar, F.H., 2006. Geophys. Res., 107.
Relating precipitation and water management to nutrient concentrations in Makridakis, S., Hibon, M., 1995. Evaluating accuracy (or error) measures. INSEAD
the oligotrophic upside-down estuaries of the Florida Everglades. Limnol. working paper. INSEAD, Fontainebleau, France, pp. 31.
Oceanogr. 51 (1), 602616. McCuen, R.H., Knight, Z., Cutter, A.G., 2006. Evaluation of Nash-Sutcliffe efciency
Chimney, M.J., Goforth, G., 2001. Environmental impacts to the Everglades ecosys- index. J. Hydrologic Eng. 11, 597602.
tem: a historical perspective and restoration strategies. Water Sci. Technol. 44, Monji, N., Hamotaru, Hirano, T., Fukagawa, T., Yabuki, K., Jintana, V., 2002. CO2
93100. and heat exchange of mangrove forest in Thailand. J. Agric. Meteorol. 52,
Choudhury, B.J., Idso, S.B., Reginato, R.J., 1987. Analysis of an empirical model for 489492.
soil heat ux under a growing wheat crop for estimating evaporation by an Obeysekera, J., Browder, J., Hornung, L., Harwell, M.A., 1999. The natural South
infrared-temperature based energy balance equation. Agric. For. Meteorol. 39, Florida system I: climate, geology, and hydrology. Urban Ecosyst. 3 (34),
283297. 223244.
Clothier, B.E., Clawson, K.L., Pinter, P.J., Moran, M.S., Reginato, R.J., Jackson, R.D., 1986. Olmsted, I.C., Loope, L.L., Rintz, R.E., 1980. A survey and baseline analysis of aspects
Estimation of soil heat ux from net radiation during the growth of alfalfa. Agric. of the vegetation of Taylor slough. Report T-586. South Florida Research Center,
For. Meteorol. 37, 319329. Everglades National Park, Homestead, Florida, 157162 p.
Coronado-Molina, C., Day, J.W., Reyes, E., Perez, B.C., 2004. Standing crop and above- Ormsby, J.P., Choudhury, B.J., Owe, M., 1987. Vegetation spatial variability and its
ground biomass partitioning of a dwarf mangrove forest in Taylor River Slough, effect on vegetation indices. Int. J. Remote Sens. 8, 13011306.
Florida. Wetlands Ecol. Manage. 12, 157164. Pielke, R.A., Marland, G., Betts, R.A., Chase, T.N., Eastman, J.L., Niles, J.O., Niyogi, D.D.S.,
da Rocha, H.R., Goulden, M.L., Miller, S.C., Menton, M., Pinto, L., Freitas, H., Figueira, Running, S.W., 2002. The inuence of land-use change and landscape dynamics
A.S., 2004. Seasonality of water and heat uxes over a tropical forest in eastern on the climate system: relevance to climate-change policy beyond the radiative
Amazonia. Ecol. Appl. 14 (4), 2232. effect of greenhouse gases. Philos. Trans. R. Soc. London, Series A: Math. Phys.
Douglas, E.M., Jacobs, J.M., Sumner, D.M., Ray, R.L., 2009. A comparison of models for Eng. Sci. 360, 17051719.
estimating potential evapotranspiration for Florida land cover types. J. Hydrol. Price, R.M., Swart, P.K., Fourqurean, J.W., 2006. Coastal groundwater discharge an
373, 366376, ISSN 0022-1694. additional source of phosphorus for the oligotrophic wetlands of the Everglades.
Duever, M.J., Meeder, J.F., Meeder, L.C., McCollom, J.M., 1994. The climate of south Hydrobiologia 569, 2326.
Florida and its role in shaping the everglades ecosystem. In: Davis, S., Ogden, Price, R.M., Swart, P.K., Willoughby, H.E., 2008. Seasonal and spatial variation in the
J.C. (Eds.), Everglades: The Ecosystem and its Restoration. St. Lucie Press, Boca stable isotopic composition (18O and D) of precipitation in south Florida. J.
Raton, Florida, pp. 225248. Hydrol. 358, 193205.
Eltahir, E.A.B., Bras, R.L., 1994. Precipitation recycling in the Amazon basin. Q. J. R. Provost, M.V., 1973. Mean high water mark and use of tide lands in Florida. Florida
Meteorol. Soc. 120, 861880. Scientist 36, 2066.
Ewe, S.M.L., Gaiser, E.E., Childers, D.L., Iwaniec, D., Rivera-Monroy, V.H., Twilley, Purevdorj, T.S., Tateishi, R., Ishiyama, T., Honda, Y., 1998. Relationships between per-
R.R., 2006. Spatial and temporal patterns of aboveground net primary produc- cent vegetation cover and vegetation indices. Int. J. Remote Sens. 19, 35193535.
tivity (ANPP) along two freshwater-estuarine transects in the Florida Coastal Richter, R., Schlpfer, D., 2011. Atmospheric/Topographic Correction for Satellite
Everglades. Hydrobiologia 569, 459474. Imagery; DLR Report DLR-IB 565-02/11. DLR, Wessling, Germany, pp. 202.
316 D. Lagomasino et al. / Agricultural and Forest Meteorology 213 (2015) 304316

Risi, J.A., Wanless, H.R., Tedesco, L.P., Gelsanliter, S., 1995. Catastrophic sedimenta- of 22nd International Symposium on Remote Sensing of Environment, 1, pp.
tion from Hurricane Andrew along the southwest Florida coast. J. Coastal Res. 349367.
81, 82102. Sumner, D.M., 1996. Evapotranspiration from successional vegetation in a defor-
Rivera-Monroy, V.H., Twilley, R.R., Davis, S.E., Childers, D.L., Simard, M., Chambers, R., ested area of the lake wales ridge, Florida United States Geological Survey.
Jaffe, R., Boyer, J.N., Rudnick, D.T., Zhang, K., et al., 2011. The role of the Everglades Water-Resources Investigations Report 96-4244.
mangrove ecotone region (EMER) in regulating nutrient cycling and wetland Sutula, M., Day, J.W., Cable, J., Rudnick, D.T., 2001. Hydrological and nutrient budgets
productivity in south Florida. Crit. Rev. Environ. Sci. Technol. 41, 633669. of freshwater and estuarine wetlands of Taylor Slough in southern Everglades,
Rockstrom, J., Gordon, L., Folke, C., Falkenmark, M., Engwall, M., 1999. Linkages Florida (U.S.A). Biogeochemistry 56, 287310.
among water vapor ows, food production, and terrestrial ecosystems. Conserv. Swiadek, J.W., 1997. The impacts of Hurricane Andrew on mangrove coasts in south-
Ecol. 3, 5. ern Florida: a review. J. Coastal Res. 13, 242245.
Ross, M.S., Meeder, J.F., Sah, J.P., Ruiz, P.L., Telesnicki, G.J., 2000. The southeast saline Twilley, R.R., Chen, R., 1998. A water budget and hydrology model of a
Everglades revisited: 50 years of coastal vegetation. J. Veget. Sci. 11, 101112. basin mangrove forest in Rookery Bat, Florida. Marine Freshwater Res. 49,
Sandoval, E., (Master Thesis) 2013. Ten year study on water ushing times and 309323.
water quality in southern Taylor Slough, Everglades National Park, FL. Florida Venalainen, A., Frech, M., Heikinheimo, M., Grelle, A., 1999. Comparison of
International University Digital Commons, pp. 834. latent and sensible heat uxes over boreal lakes with concurrent uxes over
Satyanarayana, B., Mohamad, H.A., Idris, I.F., Husain, M.L., Guebas, F.D., 2011. Assess- a forest: implications for regional averaging. Agric. For. Meteorol. 98-99,
ment of mangrove vegetation based on remote sensing and ground-truth 535546.
measurements at Tumpat, Kelantan Delta, East Coast of Peninsular Malaysia. Wang, J.D., Van de Kreeke, J., Krishnan, N., Smith, D., 1994. Wind and tide response
Int. J. Remote Sens. 32 (6), 16351650. in Florida Bay. Bull. Marine Sci. 54, 579601.
Schedlbauer, J.L., Oberbauer, S.F., Starr, G., Jimenez, K.L., 2011. Controls on sensi- Ward, G.A., Smith, T.J., Whelan, K.R.T., Doyle, T.W., 2006. Regional pro-
ble heat and latent energy uxes from a short-hydroperiod Florida Everglades cesses in mangrove ecosystems: spatial scaling relationships, biomass,
marsh. J. Hydrol. 411 (3), 331341. and turnover rates following catastrophic disturbance. Hydrobiologia 569,
Sharqawy, M.H., Lienhard, J.H., Zubari, S.M., 2010. Thermophysical properties of sea- 517527.
water: a review of existing correlations and data. Desal. Water Treatment 16, Wdowinski, S., Hong, S.H., Mulcan, A., Brisco, B., 2013. Remote-sensing monitoring
67. of tide propagation through coastal wetlands. Oceanography 26, 6469.
Simard, M., Zhanq, K., Rivera-Monroy, V.H., Ross, M.S., Ruiz, P.L., Castaneda-Moya, Welch, R., Madden, M., Doren, R.F., 1999. Mapping the everglades. Photogrammetric
E., Twilley, R.R., Rodriguez, E., 2006. Mapping height and biomass of mangrove Eng. Remote Sens. 65 (2), 163170.
forests in Everglades National Park with SRTM elevation data. Photogrammetric Yang, Z.L., Dai, Y., Dickinson, R.E., Shuttleworth, W.J., 1999. Sensitivity of ground
Eng. Remote Sens. 72, 299311. heat ux to vegetation cover fraction and leaf area index. J. Geophys. Res. 101,
Smith, T.J., Robblee, M.B., Wanless, H.R., Doyle, T.W., 1994. Mangroves, hurricanes, 16, 19505-19514.
and lightning strikes. BioScience 44, 256262. Zhaira, S., Abderrahmane, H., Mederbal, K., Frederic, D., 2009. Mapping latent heat
Smith, T.J., Anderson, G.H., Balentine, K., Tiling, G., Ward, G.A., Whelan, K.R., 2009. ux in western forest covered regions of Algeria using remote sensing data and
Cumulative impacts of hurricanes on Florida mangrove ecosystems: sediment a spatialized model. Remote Sens. 1, 795817.
deposition, storm surges and vegetation. Wetlands 29, 2234. Zapata-Rios, X., Price, R.M., 2012. Estimates of groundwater discharge to a coastal
Soegaard, H., 1989. A comparison between satellite derived evapotranspiration and wetland using multiple techniques: Taylor Slough, Everglades National Park,
normalized difference vegetation index in the Sahelian zone. In: Proceedings USA. Hydrogeol. J. 20, 16511668.

Anda mungkin juga menyukai