Calibration of Landsat Thermal Data

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/222511686

Calibration of Landsat thermal data and application to water resource studies

Article  in  Remote Sensing of Environment · October 2001


DOI: 10.1016/S0034-4257(01)00253-X

CITATIONS READS
86 1,357

5 authors, including:

Nina Raqueno Dilkushi A de Alwis


Rochester Institute of Technology University of Lincoln
48 PUBLICATIONS   929 CITATIONS    24 PUBLICATIONS   318 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Assessing the progress of river restoration in the UK View project

SHARE2012 View project

All content following this page was uploaded by Dilkushi A de Alwis on 02 November 2017.

The user has requested enhancement of the downloaded file.


Remote Sensing of Environment 78 (2001) 108 – 117
www.elsevier.com/locate/rse

Calibration of Landsat thermal data and application to water


resource studies
John R. Schott*, Julia A. Barsi, Bryce L. Nordgren, Nina Gibson Raqueño, Dilkushi de Alwis
Center for Imaging Science, Rochester Institute of Technology, 54 Lomb Memorial Drive, Rochester, NY 14623-5604, USA
Received 13 February 2000; received in revised form 2 January 2001; accepted 30 April 2001

Abstract

The newest in the Landsat series of satellites was launched April 15, 1999. The imagery collected by Landsat is used for a myriad of
applications, from coral reef studies to land management. In order to take advantage of Landsat 7 data, the Enhanced Thematic Mapper+
(ETM+) instrument must be calibrated. This study focuses on the immediate postlaunch calibration verification of the Landsat 7 thermal band
(Band 6), specifically so that it can be useful in water resource studies. Two year’s worth of thermal calibration results using a combination of
underflight data and ground truth show the ETM+ to be extremely stable, though the prelaunch calibration produces an offset of 0.261 W/m2
sr mm. This paper focuses on the details of the calibration process, including problems faced with ground truth instrumentation. While the
technical emphasis in this paper is the calibration of Landsat thermal data, it is presented in the context of the water resource studies for which
calibrated thermal data are required. At certain times in the year, water quality in large lakes, particularly the spatial structure of water quality,
is driven by temperature of lake waters. During the spring warming, a phenomena called the thermal bar drives the current and sedimentation
of large water bodies. A long-term goal of this study is to use thermally driven hydrodynamic models of lake processes to better understand
and monitor water quality in large lakes. This paper presents the hydrodynamic model and the relationship between temperature and water
quality in the Great Lakes as one example of why high-resolution, well-calibrated data are critical to earth observing. D 2001 Elsevier
Science Inc. All rights reserved.

1. Background International Field Year on the Great Lakes before much of


the current monitoring technology existed (Aubert &
1.1. The Great Lakes as a natural resource Richards, 1981). Even localized studies of shoreline pro-
cesses and river discharges are relatively scarce because of
The Laurentian Great Lakes hold 18% of the world’s problems associated with sampling the receiving waters and
fresh water. The US coast line of the Great Lakes exceeds the length of the coastline.
that of the Atlantic coast. Roughly 10% of the US popu- The Great Lakes are often referred to as inland oceans;
lation and 32% of the Canadian population live in the Great however, unlike oceans, they are small enough that given a
Lakes drainage basin, approximately 35 million people. thorough study, it may be possible to understand the primary
Pollution loads, both from within the region and introduced drivers and processes that impact lake conditions. Our
with precipitation, represent an enormous burden with both interest in studying the Great Lakes is twofold. The first is
local manifestations, as well as an ongoing regional threat. the obvious need to understand, monitor, and forecast water
Yet, despite all this, the Great Lakes are surprisingly quality conditions impacting this precious resource. The
understudied as a system and poorly understood. The second is our belief that the size of the lakes makes them a
enormous surface area, volume, and dynamic nature of the good metric for global process monitoring [i.e., they are large
lakes make it very difficult to study just one, much less all enough to damp out localized phenomena yet small enough
five of the Great Lakes. The first, and last, serious attempt to (compared to oceans) to generate observable responses].
study Lake Ontario as a whole was in 1972 during the
1.2. The Great Lakes as a thermal target
* Corresponding author. Tel.: +1-716-475-5170; fax: +1-716-475-
5988. The thermal structure in the Great Lakes has a tremend-
E-mail address: schott@cis.rit.edu (J.R. Schott). ous influence on the hydrodynamics and water quality in the

0034-4257/01/$ – see front matter D 2001 Elsevier Science Inc. All rights reserved.
PII: S 0 0 3 4 - 4 2 5 7 ( 0 1 ) 0 0 2 5 3 - X
J.R. Schott et al. / Remote Sensing of Environment 78 (2001) 108–117 109

lakes. This is best exemplified by the spring thermal bar.


The thermal bar occurs in dimictic temperate zone lakes,
resulting from the fact that the maximum density of water is
4C. To understand the thermal bar, consider lakes in the
winter. At this time, lakes are in winter stratification with
cold water on top and often with some degree of ice
coverage. As insolation increases in the spring, both direct
heat input and indirect input from runoff begin to thaw the
ice. Once the ice has melted, warm water runoff and
insolation effects dominate the shallow nearshore processes
and a localized summer stratification sets up with warm water
on top and cooler water underneath. A ring of water in
summer stratification forms around the lake. The central core,
still in winter stratification, slowly warms from the surface.
The surface water sinks, setting up a very slow mixing process
in the core of the lake. Where the two bodies (cold core and
warm shore waters) meet, they equilibrate towards 4C and the
water sinks. The result is the formation of the thermal bar, a
stable boundary between the warmer nearshore water and the
cold core, persisting for many weeks into early summer
(cf. Fig. 1). During this period, the thermal bar moves out
from shore and slowly constricts. Eventually, the thermal bar
breaks down and the whole lake goes into summer stratifica-
tion with relatively well-mixed water on the surface, a fairly
sharp thermocline and cooler stratified water underneath.
Because there is very little mixing across the thermal bar,
it is a controlling factor in water quality conditions. Nutri-
ent- and pollution-rich runoff is trapped in the nearshore
region. While the thermal bar persists, localized studies have
shown the nearshore water to be more turbid and nutrient-,
pollution-, and bacteria-rich (Menon, Dutka, & Jurkovic, Fig. 1. A selection of Landsat 5 images of Lake Ontario assembled from
1971; Rogers, 1968, 1971; Stoermer, 1968). nearly cloud-free scenes during the spring warming season. The thermal bar
progression and collapse is readily apparent in the thermal band and water
While this entire process is of interest, we are particularly quality gradients coincide with the location of the warm water in the visible
interested in the early nearshore formation of the thermal color composite.
bar. During this early period, both stream discharges and
nonpoint runoff carry the spring flush of pollutants into the
lake to be trapped in a very localized area where concen- and nonpoint source contributions on the entire regional
trations dramatically increase. Traditional sensing and mon- resource. Remote sensing offers the synoptic perspective to
itoring methods have not been able to study this very study an entire lake and potentially all the lakes as a system.
localized (nearshore) yet simultaneously lake-wide (i.e., By combining different remote sensing resources, both
the entire shore) process except in isolated cases. Yet it is small-scale, site-specific phenomena and lake-wide phenom-
clear from Fig. 1 that remote sensing offers a potential to ena can be studied. The daily coverage of many sensors
observe, at a minimum, the surface turbidity effects at both allows us to generate cloud-free composites at temporal
localized and whole lake scales. intervals to monitor dynamic processes. With the current
generation of sensors [AVHRR, SeaWiFS, and Landsat TM/
1.3. The use of remote sensing for water quality studies Enhanced Thematic Mapper+ (ETM+)], sediments (sus-
pended solids), algae (chlorophyll), and gelbstoffe (dissolved
To improve monitoring and mapping of regional water organic compounds) have successfully been monitored in the
quality parameters and processes in the Great Lakes, there are open ocean, and to some extent, in large lakes (Bukata,
a number of requirements: the need to be able to extend from Jerome, Kondratyev, & Pozdnyakov, 1991; Hudson, Moore,
a spatially localized (usually nearshore) study to the entire Bale, Dyer, & Aiken, 1994; Piech, Schott, & Stewart, 1978;
lake and from a temporally localized (days to weeks) study to Strong, 1978). With the next generation of sensors (AVIRIS,
long-term predictions of temporal processes and conditions; MODIS, NEMO, WARFIGHTER, etc.), the potential to
the need to relate what is learned from one localized study to detect and measure critical water quality parameters should
a similar, but different site, tens or hundreds of miles down improve (Murphy, Moore, Wilson, Youngs, & Morris, 1996;
shore; and the need to look at the impact of aggregate sites Richardson & Ambrosia, 1996).
110 J.R. Schott et al. / Remote Sensing of Environment 78 (2001) 108–117

Remote sensing, however, also has limitations. Optical was designed to produce high-resolution simulations for
remote sensing records only the surface of the world. In the node-to-node matching with aircraft and satellite imagery.
case of water there is some penetration of the surface, but The ALGE model has recently been adapted for use with
deeper water is obscured, limiting studies of vertical strati- Lake Ontario (Fig. 2). Modifications have enabled modeling
fication of the thermal bar and transport of materials. A vertical stratification, formation, and development of the
critical issue identified with the thermal bar is restricted thermal bar, inclusion of the Niagara River (80% of the total
mixing; the bar prevents nearshore waters from mixing influx to Lake Ontario), the St. Lawrence River outflow, and
evenly with core lake water as it normally would, trapping general lake circulation (including coriolis effects). The
nutrient- and pollutant-rich runoff nearshore. Adding a model incorporates lake bathymetry, inflow and outf low
hydrodynamic model to predict what cannot be seen would (temperature, volume and extent), hourly air temperature,
increase the ability to follow subsurface progressions of wind speed, wind direction, insolation (including cloud
thermal bars and sediments and would add the capability to cover effects), relative humidity, and radiosonde inputs for
further understand the physics of the processes, including upper air effects. These meteorological data are interpolated
inputs, outputs, and driving factors. to a lake-wide mean using up to seven meteorological sites
around the lake. The input parameters to ALGE are the
1.3.1. Hydrodynamic modeling initial lake temperature and the date of the spring turnover
To augment imaging studies, Rochester Institute of when the lake is assumed to be well mixed. These param-
Technology (RIT) is utilizing a hydrodynamic model ini- eters can be determined using remotely sensed images of
tially developed by the Department of Energy to predict the surface temperature. The starting date and initial start
movement and dissipation of thermal plumes discharged temperature have a dramatic effect on the ALGE’s predic-
into cooling lakes, rivers, and estuaries (Garrett & Hayes, tion of a thermal bar’s initial formation and progression.
1997; Garrett et al., in press). The four-dimensional (x,y,z,t) Remotely sensed data can both improve inputs to the
finite difference hydrodynamic model, termed ALGE, is hydrodynamic model and empirically calibrate the model.
capable of predicting temperatures, flow vectors, and mater- With the spatial resolution of Landsat, the images can be
ial transport. Cooling lake simulations include recirculation, used both as feedback and as verification to the model using
buoyancy-driven flow, and sediment deposition. ALGE also both lake-wide and localized (e.g., stream discharge) phe-
simulates wind-driven circulation and can combine wind nomena. Presently, AVHRR-derived temperature maps are
stress effects with tidal and buoyancy forces. Atmospheric used as input for a start date and initiation temperature for
energy exchange is modeled through turbulent sensible and ALGE. Images are observed until the surface of the lake
latent heat transfer, including the effects of clouds. ALGE appears to be ice-free and isothermal. The surface temper-

Fig. 2. Example outputs of the ALGE 3-D hydrodynamic model with two validation images. The surface temperature map images show the formation and the
two-phase propagation of the thermal bar (water temperature of 3.8 – 4.2 C) in Lake Ontario. Images (a) – (d) were taken from a time sequence of modeled
temperatures for spring warming in 1998 before the Niagara inflow was included. Images (e) – (h) are for the same conditions after the Niagara inflow and St.
Lawrence outflow were added. Images (k) and (l) are east – west cross-sections of the lake corresponding to the surface images (g) and (h). Images (i) and (j) are
AVHRR-derived temperature maps using a different color code and illustrate the need for incorporating the Niagara inflow.
J.R. Schott et al. / Remote Sensing of Environment 78 (2001) 108–117 111

ature on this date is used as the initial water temperature for


Lake Ontario in the model. Landsat’s calibration and res-
olution should provide more accurate temperature maps,
providing the model more accurate inputs. Once modeling
has predicted the location of the thermal bar during its
propagation, improved Landsat temperature maps can be
used to verify that ALGE is modeling the surface hydro-
dynamics correctly. In particular, interactions of local dis-
charges with the thermal bar can be observed with Landsat.
Since each discharge has a separate set of inputs, they can
be independently adjusted until the ALGE predictions
match the detailed thermal patterns observed in that region
by Landsat.
Additionally, Landsat’s spatial resolution allows a closer
look at local effects which are not presently in ALGE. At
this point, the model of Lake Ontario only includes one Fig. 3. Variation in the emissivity of water over the 8 – 14-mm spectral range
inflow, the Niagara River (Fig. 2). This is the primary source (from Melchor, 1941).
of water for the lake, but smaller discharges may provide the
bulk of sediments and pollutants. With the higher spectral
and spatial resolution, effects of smaller discharges can be the long-wave infrared region at  0.985 (Fig. 3) (Melchor,
observed and significant effluents that should be incorpo- 1941), even under the varying conditions typically experi-
rated into ALGE can be identified. enced on a large body of water (variable wind speed, Singh,
In summary, a hydrodynamic model has been developed 1994; and differing material concentrations, Masuda, Takish-
and implemented that models, on a whole-lake basis, the ima, & Takayama, 1988). Large lakes incorporate more than a
formation and development of the thermal bar. It also few pixels. Wind across large lakes produces a nearly
includes the major inflow and outflow and can be expanded constant mixing effect, tending to bring the lake to a uniform
to include local discharges and nonpoint pollutants. This temperature and making the temperature extremely stable.
model is also capable of mass transport calculations The need for data over a large lake to perform vicarious
throughout the lake. However, the model is driven by and calibration drives the water quality research. Already, its
must be calibrated and validated by accurate local and lake- apparent that any one of the Great Lakes would make an
wide temperature measurements. The circulation in large ideal target for both calibration and water quality study.
lakes is driven largely by buoyancy effects caused by However, the fact that the Great Lakes also develop thermal
thermal variations in the water. One primary way to evaluate bars makes them an even more desirable target. The thermal
and control the model’s performance is to compare meas- bar provides dramatic thermal contrast for the calibration
ured and observed temperatures at the surface, the one place study, as well as driving the deterioration of nearshore water
lake-wide measurements are possible. To accomplish this, quality in the spring.
we need a well-calibrated, high-spatial resolution imaging
sensor capable of mapping the detail associated with near-
shore phenomena (early thermal bar formation and storm 2. Technical approach and results
discharges) as well as whole-lake processes.
This is just one example of why we need well-calibrated, Over the two vicarious calibration collection seasons,
high-resolution thermal sensors on-board instruments like summer 1999 and summer 2000, Landsat 7 Band 6 has
Landsat. The rest of this paper focuses on the thermal proven to be incredibly stable. The offset has remained
calibration of the Landsat ETM+ instrument which is critical constant since launch and the internal gain, if it is changing,
to the future success of studies like the one discussed here. is changing at a rate of 0.04% per year (Markham, 2000).
The focus of this study was to verify the absolute radiometry
1.4. Thermal calibration of the internal gain and offset. A combination of airborne
instrument measurements and ground samples was collected
Validation of the thermal calibration of Landsat ETM+ is a and the corresponding Landsat scenes were acquired to
fundamental goal of this study. To minimize error in this compare the image predicted sensor-reaching radiance to
vicarious method of calibration, the appropriate targets must the ground truth predicted sensor-reaching radiance.
be selected. The qualities of a good thermal calibration target
are: (1) a well-known emissivity, (2) large (enough to cover at 2.1. On-board thermal calibration
least a few pixels), (3) homogeneous, and (4) thermally
stable. In the right proportions, water satisfies all of these The ETM+ on-board thermal calibration uses essentially
requirements. The emissivity of water is nearly flat through the same two-step approach that was used to calibrate
112 J.R. Schott et al. / Remote Sensing of Environment 78 (2001) 108–117

Landsat 4 and 5 TM thermal bands (Barker, 1985). A pair a small correction for the effect of the atmosphere above the
of on-board thermal reference levels is introduced by the aircraft. This can be estimated using MODTRAN according
calibration wand during the scan mirror reversal at the end to (Eq. (1)):
of each line (Fig. 4). This provides two known radiance
Ls ðsÞ ¼ tðs  hÞLa ðhÞ þ Lu ðs  hÞ ð1Þ
levels (one from the monitored wand temperature and one
from a monitored blackbody whose radiance is reflected where Ls(s) and La(h) are the spectral radiance values
onto the detectors by a mirror on the calibration wand). predicted to reach the spacecraft and observed by the aircraft
Thus, the digital count (DC)-to-radiance calibration at the or measured on the ground, respectively, s and h indicate the
wand has a full two-point update for every mirror oscil- sensor’s location in space (s) or at some elevation (h),
lation. This would complete the calibration process if the t(s  h) is the transmission from altitude h to space and
calibration wand were ahead of all of the ETM+ optics. Lu(s  h) is the path radiance due to the air column between
However, as shown in Fig. 4, the scan mirror, telescope, and the aircraft and the spacecraft. For airborne acquisitions,
scan line corrector optics are all ahead of the calibration since most of the atmospheric effects are in the lower
wand. These optical elements and their support structure atmosphere, the correction to space is small and any errors
will attenuate the radiance reaching the scan mirror and due to lack of knowledge of the intervening atmosphere
emit additional radiance into the optical path within the should also be quite small (Table 1). For some dates where
instrument. Since viewing the calibration wand does not calibrated aircraft data were not available, surface temper-
include these effects, the contribution of these elements was atures were propagated to space using MODTRAN.
estimated using a fore-optics radiometry model developed
prior to launch. The radiometry model is empirically 2.2. Calibration campaign 1999
calibrated based on a number of thermal sensors that
monitor the temperature of various elements in the optical The summer of 1999 was the first flight season for RIT’s
assembly ahead of the calibration wand. This calibration Modular Imaging Spectrometer Instrument (MISI), flown in
generates a correction to the per scan calibration that should a Piper Aztec aircraft. The MISI instrument has four thermal
account for the fore-optics contribution. Thus, by monitor- channels, as well as a 64-channel visible – near-infrared
ing the temperature of the calibration wand, on-board imaging spectrometer. Our emphasis here will be on the
blackbody, and optical components, the image data can thermal channels, two of which are close matches to the
theoretically be calibrated. Landsat ETM+ Band 6 [one is shifted slightly (  0.5 mm
This study provides an independent check of the cal- lower) and the other uses a spare Landsat spectral filter for a
ibration by comparing targets of known radiance to image- very close match]. The thermal sensors have a 2 mrad field
derived Landsat radiances. Using simultaneous airborne of view, a system noise level better than 0.1 K and MISI was
image and ground truth acquisitions, the surface-leaving flown at altitudes up to 5000 ft yielding a 10 = ft (3 m)
radiance was predicted and extrapolated to space-reaching spatial resolution. The details of the MISI instrument can be
radiance using the MODTRAN radiation propagation model found in Feng, Schott, and Gallagher (1994) and Schott,
(Berk, Bernstein, & Robertson, 1989). Schott, Gallagher, Gallagher, Nordgren, Sanders, and Barsi (1999).
and Barsi (1997) describe the details behind this approach. MISI’s on-board calibration system is similar to that of
For large, uniform, thermally stable targets, the predicted Landsat, a two-point calibration system updating every scan
and observed airborne image radiances should differ only by line. The essential difference is that MISI’s two full aperture

Fig. 4. The Landsat 7 optical path. The calibration wand moves into the optical path once every scan line, providing both visible and thermal calibration targets.
J.R. Schott et al. / Remote Sensing of Environment 78 (2001) 108–117 113

Table 1
Error propagation of the effects of atmospheric uncertainty on sensor-reaching radiance (apparent temperature) using MODTRAN
Lu Lu error Error in sensor-reaching
T (K) t t error (W/m2 sr mm) (W/m2 sr mm) radiance (app. temp.)
MODTRAN (1.5 km to space) 285 0.820 0.00500 1.850 0.10000 0.0498
305 0.820 0.00500 1.850 0.10000 0.0446
MODTRAN (3 km to space) 285 0.915 0.00043 0.622 0.01226 0.0446
305 0.915 0.00043 0.622 0.01226 0.0400
MODTRAN (7 km to space) 285 0.973 0.00003 0.078 0.00049 0.0016
305 0.973 0.00003 0.078 0.00049 0.0018
Targets at two different temperatures are propagated from various altitudes to space and the error in radiance-reaching space is predicted based on uncertainties
in estimating the atmospheric profile.

blackbodies are ahead of all the optics (Fig. 5). Whereas the absence of the thermal bar, the underflights focused on
Landsat required a prelaunch radiometric model to offset the flight lines that included large bays, open lake, major river
propagation of radiance through the optical system, MISI’s discharges into the lake, and power plant thermal dis-
calibration includes the effect of the optical elements. The charges. These provide a gradient of temperatures for
MISI instrument successfully acquired data four times in calibration, but are less than ideal because a uniform large
1999 under Landsat 7. These data were acquired along the area is harder to obtain. Extensive ground truth campaigns
south shore of Lake Ontario and the north shore of Lake were conducted concurrent with the underflights. The
Erie. The data collection program involved flying the ground truth included deployment of reflectance panels,
aircraft at multiple altitudes over the same targets from field and water spectral reflectance measurements, water
several hundred feet to about 5000 ft. The timing of the sampling and laboratory analysis of coloring agents, and
Landsat launch, coupled with a very warm winter and early most importantly, for our purposes, surface water temper-
development of the thermal bar, resulted in only one ature measurements. Measurements were made from piers
collection of a weak thermal bar (low thermal contrast) in and several small boats deployed in the embayments and
Lake Ontario. In addition to being one of our prime coastal waters. Fig. 6 shows sample MISI thermal images
scientific interests, the thermal bar is also an ideal cal- and ground truth points. Measurements included Global
ibration target since it provides multiple large regions of Positioning System (GPS) location and surface temperature
uniform temperature water (i.e., many Landsat pixels). In readings from thermistors floated just at the surface on the
bottom of small styrofoam floats. Because surface waters in
these lakes are well mixed by the nearly constant wind and
long fetch, a surface gradient is not expected in these waters
and the thermistor measurements of temperature are used as
surface temperature values.
Regrettably, the MISI laboratory blackbody calibration
describing the temperature-to-DC relationship was discov-
ered to be invalid due to a difference in the in-flight and
laboratory readout electronics. By the time this was discov-
ered at the end of the season, changes had been made to the
readout circuitry so a proper recalibration using the flight
circuitry was impossible. To overcome this limitation, an
empirical calibration of the MISI was performed using
ground truth data from three collections spanning the
1999 collection season. For each collection, ground truth
data were converted to in-band sensor reaching spectral
radiance estimates using an equation of the form (Eq. (2)):
R
ðtðh; lÞ½eðlÞLTl þ ð1  eðlÞÞLdl  þ Lul ðh; lÞÞbðlÞdl
La ðhÞ ¼ R ð2Þ
bðlÞdl

where La(h) is the in-band spectral radiance [w/cm2 sr mm]


reaching the aircraft, t(h,l), L dl, and L ul(h) are
Fig. 5. The MISI line scanner design. The scan mirror rotates in the cross- MODTRAN-derived spectral estimates of transmission,
track direction while the aircraft moves forward. The calibration sources are downwelled spectral radiance and upwelled spectral
located in the backscan of the mirror and are imaged every scan line. A
pyramid mirror lies at the focal point of the Cassegrain telescope and
radiance, respectively, e(l) is the spectral emissivity of
currently redirects the incoming energy to two spectrometers, a broadband water, LTl is the Planckian spectral radiance from a target
visible focal plane and the cooled thermal focal plane. at temperature T, and b(l) is the spectral response function
114 J.R. Schott et al. / Remote Sensing of Environment 78 (2001) 108–117

Fig. 6. The Ginna Nuclear Power Plant discharge plume as imaged by MISI from 4000 ft. White dots represent areas where ground truth was collected. This
image has not been roll corrected.

of the MISI spectral channel. The MODTRAN runs used particular spectral channel. Temperatures, which were thus
radiosondes from the nearest airport (usually Buffalo, NY estimated for the blackbody, could be regressed against the
corrected for local surface temperature and water vapor voltage recorded for that blackbody to generate a voltage-to-
values; Schott 1997). temperature calibration. Fig. 7 shows a plot of temperature
The counts-to-radiance relationship for each flight line versus voltage derived in this manner using flight lines
could then be solved for using: where ground truth data were available. The plot indicated
that the temperature readout circuit was very stable (as
DCðiÞ ¼ mLa ðh; iÞ þ b ð3Þ
expected) over the flight season and provided the basis for
where DC(i) is the count for the ith target, La(h,i) is the analysis of all subsequent data. The error in blackbody
predicted sensor reaching spectral radiance for the ith target equivalent temperature values used to generate this curve is
and m and b are sensor gain and bias (which, for this study, approximately 1.2 K as opposed to a value of a few tenths
must be assumed constant over a flight line). Since DC Kelvin that could have been expected using a proper
values for each on-board blackbody and the voltage readout laboratory calibration.
from their temperature monitoring thermistors were re- This empirical calibration procedure for the MISI black-
corded, the thermistor readout was calibrated using the bodies was evaluated by inverting MISI aircraft radiance
following procedure. Blackbody DCs were converted to values to ground temperature values for ground truth sites.
spectral radiance and corresponding blackbody temperature This inversion used the multialtitude or profile calibration
using Eq. (3) and a Planckian lookup table built for the technique described by Schott (1979). This is a wholly in-

Fig. 7. Empirically derived blackbody temperature versus thermistor voltage for MISI calibration in 1999. Linearity indicates the system is stable over the
collection season.
J.R. Schott et al. / Remote Sensing of Environment 78 (2001) 108–117 115

Fig. 8. The results of the 1999 calibration validation campaign using empirically calibrated MISI data from three dates, 11 May 1999, 07 June 1999, and 03
September 1999.

scene technique based on the availability of radiometrically After one year, the conclusion was to agree that there was
calibrated data. The root mean square error between the an offset error, and rely on more data to verify the slope was
ground truth and profile-derived temperatures obtained for not an artifact.
18 samples from 3 days and four flight lines was 1.3 K
with most of the error due to lack of knowledge of the on- 2.3. Calibration campaign 2000
board calibration.
With MISI finally calibrated using ground truth, the The analysis to date for the summer of 2000 season is
aircraft-derived radiances were extrapolated to space using based primarily on surface temperature measurements.
a MODTRAN atmosphere. The preliminary results of the Ground truth surface temperatures were converted to sur-
comparison to Landsat from 1999 are shown in Fig. 8. face-leaving radiances and extrapolated to space using
While still exhibiting the linear behavior expected of a MODTRAN and Equation 5. Essentially, this is the same
thermal instrument, there is definitely an error in the offset procedure as used with the underflights, except for the
and appears to be a significant error in gain. The errors in introduction of a greater error in predicting the effects of
the MISI data are such that, while a bias of 2 –3 K is the lower atmosphere.
strongly suspected, the error in gain error could have Fig. 9 illustrates the results from three collection dates in
potentially been due to MISI (though later it will be 2000. While the offset error still appears, the slope of the
associated with Landsat). If the error does, in fact, include line is essentially unity, meaning the gain error is gone. This
both a bias and a gain effect, errors at about 300 K are only is in much closer agreement with the Palluconi data from
2 K but grow to 4 K for colder (280 K) targets, far too high 1999 and upon inquiries, with the Palluconi data from 2000
to be acceptable for any quantitative studies with the data. (Palluconi, 2000). The offset is still on the order of 2 K, but
These error estimates are consistent with initial results from the error is no longer temperature dependent.
Palluconi (1999) for a single target at about 292 K, who
estimated that the Landsat observed radiances are high by 2.4. Compiled calibration results
2.3 K. The estimate using the regression results in Fig. 8
would suggest a difference of 2.2 K for an apparent at Compiling both year’s data still generates an error in
satellite temperature of 292 K. gain. Since this gain error did not appear in the Palluconi

Fig. 9. The results of the 2000 calibration validation campaign using ground measurements extrapolated to space from three dates, 05 July 2000, 26 July 2000,
and 07 September 2000.
116 J.R. Schott et al. / Remote Sensing of Environment 78 (2001) 108–117

proven to be extremely stable though both calibration


validation investigations have shown an error in the offset.
The 0.237 W/m2 sr mm bias observed here agrees with an
independent validation effort to within 0.051 W/m2 sr mm or
0.4 K. The average of the two studies, 0.261 W/m2 sr mm, is
the correction that the Landsat Project Science Office will
implement in its processing system. The newly calibrated
images will be available through the EDG mid-December
2000, but images already processed and purchased can
easily be corrected for the calibration error by subtracting
the offset from the image radiance.
A model has been created to map lake-wide thermally-
Fig. 10. The combined results from the two years of validation of Landsat 7.
driven hydrodynamic processes in Lake Ontario and the
Because 11 May 1999 and 07 June 1999 were during the initial check-out
period, they were removed from the analysis. This validation reveals that potential exists for using Landsat as both input and valida-
the Landsat calibration is predicting image radiances that are too high by tion to the model. Due to two warm winters, the thermal bar
0.237 W/m2 sr mm. was weaker than normal so a detailed mapping of the early
formation of a thermal bar using Landsat 7 has not been
data or the 2000 ground truth data, the possible source of observed yet. Long-term quantitative thermal data will be
error focused on the use of MISI. More recently, however, required to refine and calibrate the models that will allow us
the Landsat Project Science Office has found that because to understand and forecast hydrodynamic processes and
the Landsat instrument was functioning in a very different material transport within large lakes.
operating mode during the first three months after launch, In order to continue refining the hydrodynamic model
the initial check-out data may be significantly different from and for use in other studies, Landsat must continue to be a
data collected in normal operations. Because the imager was resource for quantitative studies. While the instrument
only on for limited amounts of time during that period, the appears to have remained stable since launch, there is as
temperatures of the instrument components were much yet no well-accepted source for the bias error documented
cooler than those once the instrument was in normal here. Thus, continued efforts to monitor and update the
operation mode. Since the only two scenes from this thermal calibration of the ETM+ sensor are recommended.
check-out period are in the 1999 data set (none of the
Palluconi data were), the check-out period calibration may
be different than normal operation calibration.
Given that the validation needs to apply to the normal References
operational period, the 5/11/99 and 6/7/99 data were
Aubert, E. J., & Richards, T. L. (1981). IFYGL — The International Field
removed from the analysis, in order to compile a final offset Year for the Great Lakes ( p. 410). Ann Arbor, MI: National Oceanic
value. Fig. 10 shows the compiled 1999 and 2000 validation and Atmospheric Administration.
data. The final offset value determined by this study Barker, J. L. (Ed.) (1985). Proceedings of Landsat-4 Science Character-
indicates that Landsat image radiances are too high by ization Early Results Symposium, 2/83, NASA Goddard, NASA Con-
0.237 W/m2 sr mm, about 2 K. ference Publication 2355, NASA/GSFC.
Berk, A., Bernstein, L. Sl., & Robertson, D. C. (1989). MODTRAN: a
The difference between this and the Palluconi study is moderate resolution model for LOWTRAN 7. GL-TR-89-0122, Spec-
0.051 W/m2 sr mm, about 0.4 K at 300 K. Compiling all the tral Sciences.
data, the final offset value is 0.261 W/m2 sr mm. The Bukata, R. P., Jerome, J. H., Kondratyev, K. Y., & Pozdnyakov, D. V.
Landsat Project Science Office has accepted these results (1991). Estimation of organic and inorganic matter in inland waters:
and begun the process of making the changes in the optical cross sections of Lakes Ontario and Ladoga. Journal of Great
Lakes Research, 17, 461 – 469.
appropriate software and calibration files. For scenes pur- Feng, X., Schott, J. R., & Gallagher, T. W. (1994). Modeling the perform-
chased from the Eros Data Gateway (EDG) before Decem- ance of a high speed scan mirror for an airborne line scanner. Optical
ber 2000, the user will have to manually subtract this offset Engineering, 33 (4), 1214 – 1222.
error from the radiance image product. Scenes purchased Garrett, A. J., & Hayes, D. W. (1997). Cooling lake simulations compared
from EDG after December 2000 will include this correction, to high resolution thermal imagery and dye tracer data. Journal of
Hydraulic Enginering, 123, 885 – 894.
so it will include the latest in calibration efforts. Garrett, A. J., Irvine, J., Evers, M., Smyre, T. K., King, J., Ford, A. D., &
Levine, C. (1998). An imagery-based hydrodynamic simulation of an
effluent stream entering the Clinch River. Photogrammetric Engineer-
3. Conclusions and recommendations ing and Remote Sensing.
Hudson, S. J., Moore, G. F., Bale, A. J., Dyer, K. R., & Aiken, J. (1994). An
operational approach to determining suspended sediment distributions
Most uses of Landsat data require calibrated data, includ- in the Humber estuary by airborne multi-spectral imagery. Proceedings
ing the validation and calibration of the ALGE hydrodyn- of the First International Airborne Remote Sensing Conference, 3,
amic model presented here. The thermal band itself has 10 – 20.
J.R. Schott et al. / Remote Sensing of Environment 78 (2001) 108–117 117

Markham, B. (2000). Personal communication. NASA/GSFC, Goddard, MD. precision temperature measurements. Proceedings — 14th Conference
Masuda, K., Takishima, T., & Takayama, Y. (1988). Emissivity of pure and on Great Lakes Research, 618 – 624.
sea waters for the model of sea surface in the infrared window regions. Schott, J. R. (1979). Temperature measurement of cooling water discharged
Remote Sensing of Environment, 24, 313 – 329. from power plants. Photogrammetric Engineering and Remote Sensing,
Melchor, C. V. (1941). The refractive index of liquid water in the near 45 (6), 753 – 761.
infrared spectrum. Journal of Optical Society of America, 31, 244 – 247. Schott, J. R. (1997). Remote sensing: the image chain approach. Oxford
Menon, A. S., Dutka, B. J., & Jurkovic, A. A. (1971). Preliminary bacter- University Press.
iological investigation of the Lake Ontario Thermal Bar. Proceedings — Schott, J. R., Gallagher, T. W., & Barsi, J. (1997). Calibration procedures
14th Conference on Great Lakes Research, 59 – 68. for evaluation of in-flight radiometry performance of thermal infrared
Murphy, R., Moore, G., Wilson, A., Youngs, K., & Morris, K. (1996). satellite sensors. Presented at EOS and SPIE Europto Conference,
Remote sensing of coastal estuaries. In: SeaWiFS Technical Report, Aerospace Remote Sensing, London.
Vol. 33, NASA Tech Memo 104566, Vol. 33. Schott, J. R., Gallagher, T. W., Nordgren, B., Sanders, L., & Barsi, J.
Palluconi, F. (1999). Personal communication. NASA/JPL, Pasadena, CA. (1999). Radiometric calibration procedures and performance for the
Palluconi, F. (2000). Personal communication. NASA/JPL, Pasadena, CA. Modular Imaging Spectrometer Instrument (MISI). Proceedings of the
Piech, K. R., Schott, J. R., & Stewart, K. M. (1978). The blue-to-green Earth International Airborne Remote Sensing Conference, ERIM.
reflectance ratio and lake water quality. Photogrammetric Engineering Singh, S. M. (1994). Effect of surface wind speed and sensor view zenith
and Remote Sensing, 44, 1303 – 1319. angle dependence on emissivity for SST retrieval from thermal infrared
Richardson, L., & Ambrosia, V. G. (1996). Algal accessory pigment detec- data: ATSR. International Journal on Remote Sensing, 15, 2615 – 2625.
tion using AVIRIS image-derived spectral radiance data. In: Proceed- Stoermer, E. F. (1968). Nearshore phytoplankton populations in the Grand
ings of the 6th Annual Airborne Earth Science Workshop. Jet Propulsion Haven, Michigan vicinity during thermal bar conditions. Proceedings —
Laboratory, NASA. 11th Conference on Great Lakes Research, 137 – 150.
Rogers, G. K. (1968). Heat advection within Lake Ontario in spring and Strong, A. E. (1978). Chemical whitings and chlorophyll distributions in
surface water transparency associated with the thermal bar. Proceed- the Great Lakes as viewed by Landsat. Remote Sensing of Environment,
ings — 11th Conference on Great Lakes Research, 480 – 486. 7, 61 – 72.
Rogers, G. K. (1971). Field investigation of the thermal bar in Lake Ontario:

View publication stats

You might also like