Вы находитесь на странице: 1из 27

This article was downloaded by:[Gersman, Ronen]

[University of Southern California]


On: 16 June 2008
Access Details: [subscription number 788774280]
Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954
Registered office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

International Journal of Remote


Sensing
Publication details, including instructions for authors and subscription information:
http://www.informaworld.com/smpp/title~content=t713722504

Mapping of hydrothermally altered rocks by the EO-1


Hyperion sensor, Northern Danakil Depression, Eritrea
Ronen Gersman ab; Eyal Ben-Dor b; Michael Beyth c; Dov Avigad a; Michael
Abraha d; Alem Kibreab d
a
Institute of Earth Sciences, The Hebrew University, Jerusalem, Israel
b
Department of Geography, Tel Aviv University, Israel
c
Geological Survey of Israel, Jerusalem
d
Eritrean Department of Mines, Asmara, Eritrea
Online Publication Date: 01 January 2008
To cite this Article: Gersman, Ronen, Ben-Dor, Eyal, Beyth, Michael, Avigad,
Dov, Abraha, Michael and Kibreab, Alem (2008) 'Mapping of hydrothermally altered rocks by the EO-1 Hyperion
sensor, Northern Danakil Depression, Eritrea', International Journal of Remote Sensing, 29:13, 3911 3936
To link to this article: DOI: 10.1080/01431160701874587
URL: http://dx.doi.org/10.1080/01431160701874587

PLEASE SCROLL DOWN FOR ARTICLE


Full terms and conditions of use: http://www.informaworld.com/terms-and-conditions-of-access.pdf
This article maybe used for research, teaching and private study purposes. Any substantial or systematic reproduction,
re-distribution, re-selling, loan or sub-licensing, systematic supply or distribution in any form to anyone is expressly
forbidden.
The publisher does not give any warranty express or implied or make any representation that the contents will be
complete or accurate or up to date. The accuracy of any instructions, formulae and drug doses should be
independently verified with primary sources. The publisher shall not be liable for any loss, actions, claims, proceedings,
demand or costs or damages whatsoever or howsoever caused arising directly or indirectly in connection with or
arising out of the use of this material.

Downloaded By: [Gersman, Ronen] At: 21:27 16 June 2008

International Journal of Remote Sensing


Vol. 29, No. 13, 10 July 2008, 39113936

Mapping of hydrothermally altered rocks by the EO-1 Hyperion sensor,


Northern Danakil Depression, Eritrea
RONEN GERSMAN*{{, EYAL BEN-DOR{, MICHAEL BEYTH, DOV
AVIGAD{, MICHAEL ABRAHA" and ALEM KIBREAB"
{Institute of Earth Sciences, The Hebrew University, Givat Ram, Jerusalem, 91904,
Israel
{Department of Geography, Tel Aviv University, Ramat Aviv, 69978, Israel
Geological Survey of Israel, 30 Malkhei Israel St., Jerusalem, 95501
"Eritrean Department of Mines, Liberty Avenue, 213, PO Box 272, Asmara, Eritrea
(Received 15 March 2007; in final form 1 June 2007 )
An EO-1 Hyperion scene was used to identify and map hydrothermally altered
rocks and a Precambrian metamorphic sequence at and around the Alid volcanic
dome, at the northern Danakil Depression, Eritrea. Mapping was coupled with
laboratory analyses, including reflectance measurements, X-ray diffraction, and
petrographic examination of selected rock samples. Thematic maps were
compiled from the dataset, which was carefully pre-processed to evaluate and
to correct interferences in the data. Despite the difficulties, lithological mapping
using narrow spectral bands proved possible. A spectral signature attributed to
ammonium was detected in the laboratory measurements of hydrothermally
altered rocks from Alid. This was expressed as spectral absorption clues in the
atmospherically corrected cube, at the known hydrothermally altered areas. The
existence of ammonium in hydrothermally altered rocks within the Alid dome
has been confirmed by previous studies. Spectral information of endmembers
mineralogy found in the area (e.g. dolomite) enables a surface mineral map to be
produced that stands in good agreement with the known geology along the
overpass. These maps are the first hyperspectral overview of the surface
mineralogy in this arid terrain and may be used as a base for future studies of
remote areas such as the Danakil.

1.

Introduction

The Hyperion imaging spectrometer, mounted on the NASA Earth Observing 1


(EO-1) platform, is the first high spectral resolution imaging spectrometer that
routinely acquires data from orbit (Pearlman et al. 2003, Ungar et al. 2003). During
the time passed since its launch in November 2000, major work has been dedicated
to checking its performances and capabilities (e.g. Cudahy et al. 2002, Green et al.
2003, Pearlman et al. 2003, Thome et al. 2003). Researchers have evaluated its
capabilities for studies of vegetation (Asner and Heidebrecht 2003, Datt et al. 2003,
Goodenough et al. 2003), marine sciences (Kruse 2003) and geological purposes
(Cudahy et al. 2001, Hubbard et al. 2003, Kruse et al. 2003). Studies that compared
*Corresponding author. Present address: University of Southern California,
Department of Earth Sciences, 3651 Trousdale Pkwy, Los-Angeles, CA 90089-0740,
USA. Email: gersman@usc.edu
International Journal of Remote Sensing
ISSN 0143-1161 print/ISSN 1366-5901 online # 2008 Taylor & Francis
http://www.tandf.co.uk/journals
DOI: 10.1080/01431160701874587

Downloaded By: [Gersman, Ronen] At: 21:27 16 June 2008

3912

R. Gersman et al.

the performances of the Hyperion and spaceborne multispectral sensors and


between the Hyperion and the airborne AVIRIS (Airborne Visible/Infrared Imaging
Spectrometer) imaging spectrometer concluded that the Hyperion has a low signalto-noise ratio (SNR), relative to other airborne sensors such as AVIRIS or HyMAP
(Datt et al. 2003, Hubbard et al. 2003, Kruse et al. 2003). The low SNR is probably
the major significant disadvantage of the sensor performance. Furthermore, it was
shown that the Hyperion sensor has severe problems arising from the line-curvature
effect, which eventually prevent adequate atmospheric rectification and accordingly
precise thematic evaluation (Goodenough et al. 2003, Neville et al. 2003).
Nevertheless, additional geoscience applications of the Hyperion data are needed
for a better understanding of the advances and limitations of imaging spectroscopy
as a routine research tool.
In this study we used a single Hyperion scene acquired over east Eritrea to identify
and map hydrothermally altered rocks and the Neoproterozoic basement within and
around the Alid volcanic dome, located at the northern termination of the Danakil
Depression. Geo-referenced ASTER scenes (Abrams et al. 2002) were used to
examine the area on a broader scale and as a reference for rectifying the Hyperion
scene. Our goal was to study the feasibility of the Hyperion data, calibrated and
coupled with a comprehensive field study and laboratory studies of an arid and wellexposed area, as a tool for use in locating natural resources.
The northern Danakil Depression and the Alid volcanic dome are of scientific and
economic interest (Souriot and Brun, 1992, Drury et al. 1994, Beyth 1996, Ghebreab
1998, Lowenstern et al. 1997, 1998). This area has never been studied by any
hyperspectral sensor and only rarely studied on the ground. Therefore the
opportunity of examining orbital hyperspectral data is significantly important in
order to demonstrate that hyperspectral data is capable of precisely mapping remote
areas.
Important ground targets in this work were hydrothermally altered rocks within
the Alid dome for their significant economical potential. Previous studies of Alids
hydrothermal hot springs and fumaroles (Beyth 1996, Duffield et al. 1997) assessed
its geothermal and epithermal gold potential. Both studies reported the existence of
ammonium sulphate minerals. Ammonium-bearing minerals are associated with
hydrothermally altered rocks, hydrocarbons in oil shale, and coal-bearing
formations (Krohn et al. 1993), which are found to occur in several economically
important geological environments (Krohn and Altaner 1987, Baugh et al. 1998).
Yet, the role and importance of geologic nitrogen is conspicuously absent from most
reviews of global nitrogen dynamics (Holloway and Dahlgren 2002). Knowledge of
the origin and mineralogical relations of ammonium minerals in known
hydrothermal systems is critical for the proper interpretation of remote sensing
data and for testing possible links to mineralization (Krohn et al. 1993).
2.

Geological setting

The Danakil Depression is the terminus of an embryonic spreading axis the


northern part of the Afar triangle, which formed between the African plate on the
west and the Danakil microplate on the east (figure 1). According to palaeomagnetic
data, the Danakil microplate rotated 10u anti-clockwise during the past 4 million
years (Souriot and Brun 1992, Ghebreab 1998). The Alid volcanic centre is located
along the axis of the NNWSSE trending Danakil depression, informally known as
the Alid graben (Duffield et al., 1997). The graben is topographically and

3913

Downloaded By: [Gersman, Ronen] At: 21:27 16 June 2008

Hyperion mapping in the Danakil Depression, Eritrea

Figure 1. The working area. (a) The Alid-area map is based on an ASTER scene and after
Abbate et al. (2004), Beyth (1996), Duffield et al. (1997) and Gersman et al. (2006). (b)
Regional tectonic setting, modified after Lowenstern et al. (1997). Dark grey is mostly
Precambrian basement, pale grey is mostly Cenozoic volcanics and sediments. Straight lines
represent major spreading axes, in which the Danakil depression extends northwest from the
Afar depression. The small frame represents the location of (a).

structurally limited on its western side by east-dipping normal faults which form the
Red Sea escarpment (Drury et al. 1994, Duffield et al. 1997, Beyth et al. 2003).
Neoproterozoic basement rocks, related to the East African Orogen (Drury et al.
1994, Ghebreab, 1999a,b; Beyth et al. 2003, Ghebreab et al. 2005) on the western
side of the graben rise up to 2500 m. The eastern topographic and structural
boundary is marked by a steep and abrupt 300 m high west-dipping normal fault
escarpment. Basement rocks are locally exposed at the base of this scarp and are
unconformably overlain by a post-Miocene sequence of intercalated sedimentary
and volcanic deposits (Duffield et al. 1997, Lowenstern et al. 1997).
The Alid Dome, which rises about 900 m above sea level was interpreted by
Duffield et al. (1997), Lowenstern et al. (1997) and Lowenstern et al. (1998) to be a
structural dome rather than a classical volcano. They suggested that a shallow
felsicgranitic magma body which lies about 25 km beneath the top of the
mountain is responsible for a 1 km uplift of the overburden rocks. The volcanic
sequences in the Alid include basalts, pumiceous rhyolites (some with xenoliths of
basement rocks and of older rhyolite and basalt) and a sedimentary sequence with
pillow basalt, submarine debris flow and shallow marine sediments (see Alid series
in figure 1). Calculated 40Argon/39Argon ages of the volcanic succession near Alid
range from 1.2 million years to 14 500 years (Duffield et al. 1997). Younger basalt

Downloaded By: [Gersman, Ronen] At: 21:27 16 June 2008

3914

R. Gersman et al.

flows are estimated to be Holocene in age. Outcrops of Neoproterozoic pelitic


biotitekyaniteschist of the metamorphic Bizen Domain (Beyth et al. 2003,
Ghebreab et al. 2005) are exposed on the eastern side of the dome beneath the
volcanicsedimentary succession, as a result of the structural updoming (Beyth
1996, Duffield et al. 1997, Lowenstern et al. 1997, 1998, Gersman et al. 2006). Much
of the pyroclastics and lava flows within the Alid dome have been strongly altered
by hydrothermal activity (Beyth 1996, Duffield et al. 1997, Lowenstern et al. 1997,
1998, Gersman et al. 2006). Beyth (1996) and Duffield et al. (1997) reported
kaolinite, gypsum, anhydrite, montmorillonite, illite and ammonium sulphate
minerals (tschermigite NH4Al(SO4)2 12H2O, mascagnite (NH4)2(SO4), kokatite
(NH4)2Ca(SO4)2 H2O) around hydrothermal fumaroles and hot springs within the
Alid dome. The continuing extension of crustal spreading across the Alid graben is
creating north/northwest-trending faults and fissures (Duffield et al. 1997). A set of
eastwest fractures was documented by the authors in the recent basaltic cover north
of the Alid dome.
3.

Methods

The study combined analysis of hyperspectral data as acquired from orbit


(Hyperion) followed by a comprehensive field study, and a laboratory spectral
and mineralogical study of rock samples.
3.1

Field work

A field reconnaissance was carried out between 5 and 15 February 2005.


Geographical locations were measured by a Magellan GPS (SportTrack colour)
with an average accuracy ,7 m. Samples for laboratory studies included fresh and
surface-weathered sides of representative rocks. Samples of hydrothermally altered
rocks were taken from two sites within the Alid caldera: fumaroles in Darere
(14.877uN, 39.915uE) and hot springs in Illegedi (14.882uN, 39.926uE). Samples of
homogeneous alluvial fans and playas for atmospheric calibration (Ben-Dor et al.
1994, Clark et al. 2002, Crouvi 2002, Rowan and Mars 2003) were collected from
measured homogeneous areas of 60 m660 m, representing a square of 4 pixels in an
image of 30 m pixel size.
3.2

Laboratory work

3.2.1 Mineralogy. The mineralogy of fine grained samples was studied using the
X-ray diffraction (XRD) technique for bulk mineralogy (e.g. Warren 1990). The
samples included hydrothermally altered rocks and alluvial/aeolian deposits for the
atmospheric calibration. The XRD analyses were carried out on bulk powder using
a Philips diffractometer PW 1830/3710/3020 of the Geological Survey of Israel.
3.2.2 Petrography. The petrography of high grade pelitic schists and orthogneiss,
basaltic/andesitic dykes, low-grade metavolcanic phyllite, mylonite, graphite schist
and amphibolite gneiss were studied under an optical polarized microscope.
3.2.3 Reflectance spectral analysis. Spectral measurements were carried out using
the Analytical Spectral Devices Inc. (ASD) field spectrometer (FieldSpec pro FR)
with a spectral range of 3502500 nm. The spectral resolution is 3 nm at 700 nm and
10 nm at 1400/2500 nm. Sampling interval is 1.4 nm at 3501050 nm and 2 nm at
10002500 nm (http://www.asdi.com). The measurements were performed at the

Downloaded By: [Gersman, Ronen] At: 21:27 16 June 2008

Hyperion mapping in the Danakil Depression, Eritrea

3915

Remote Sensing laboratory in Tel-Aviv University using a contact probe and a


built-in illumination source (A 4.5 W HalogenKrypton lamp with AC to DC
current converter).
The reflectance spectra were calibrated against a Halon plate (Weidner and Hsia
1981). Black current calibration and white reference measurements were performed
approximately every 30 min and the stability of the spectrometer was examined
against a standard sample (dolomite).
Each rock sample was measured in duplicate on both its fresh and weathered
sides. Each calibration sample composed of unconsolidated sand/pebbles was mixed
between every duplicate recording. The spectra of each calibration sample were
averaged to one representative spectrum. The collected spectra were run through the
continuum removal (CR) technique (Clark and Roush 1984) in order to recognize
and allocate diagnostic spectral features.
3.3

Hyperion data processing

Very close to the field trip we ordered the acquisition of a single Hyperion scene that
took place on 2 February 2005. Processing the scene was carried out using the
ENvironment for Visualizing Images (ENVI 2003) software packages, version 4.0
and other accessory codes to re-process the data.
The Hyperion sensor provides continuous spectral coverage of 242 bands, in a
10 nm (average) sampling interval over the reflected spectrum from 400 to 2500 nm.
The instrument consists of two detectors. The VNIR (VIS + NIR) detector covers a
spectral range of 4001000 nm in 70 channels and the SWIR detector covers the
range of 9002500 nm in 172 channels (Ungar et al. 2003). There are 21 channels in
the overlapping spectral range. Each scene is collected as a narrow strip, covering a
ground area of approximately 7.7 km in the across-track direction, and 75 km (in
our case) in the along-track direction in a pushbroom configuration, from an
altitude of 705 km (Beck 2003, Ungar et al. 2003). The ground sample distance (pixel
size) is 30 m for all bands.
3.3.1 General. The processing of the Hyperion data is schematically summarized
in figure 2. Pre-processing steps included the removal of overlapping and inactive
bands (17, 5878 and 225242, after Barry 2001), and of the first and last samples
of the scene. The radiometric calibration to convert raw DN to W m22 mm21 sr21
units was done after Beck (2003). In the end this pre-processing step enables
determining cube dimensions of 196 spectral bands, 254 columns and 2500 lines.
Several processes were performed removing the atmosphere to filter out some data
noises, and estimate the quality of the data as a spectral database.
3.3.2 SNR estimation. SNR was estimated from the corrected radiometric image.
Three homogeneous targets were selected for this purpose (figure 3) and the SNR
was estimated according to Green et al. (2003) and Kruse et al. (2003). The SNR
values are 90 : 1 in the VIS range, 60 : 1 at SWIR-I (,1000 to ,1600 nm (excluding
the water vapour signal at ,1400 nm)) and about 35 : 1 in SWIR-II (,2000
2400 nm). Kruse et al. (2003) reported SNR values of about 25 : 1 for the SWIR-II
range, which is in good agreement with our data.
3.3.3 Atmospheric correction. A two-stage atmospheric correction was applied to
the scene, separated by a removal of vertical lines (de-striping) and residual
noise stages. The two stages were radiative-modelled atmospheric correction and a

R. Gersman et al.

Downloaded By: [Gersman, Ronen] At: 21:27 16 June 2008

3916

Figure 2.

The methodology of processing and interpreting the Hyperion data.

fine-tuning atmospheric correction using the empirical line (EL) technique. Noise
reduction steps were performed between the two atmospheric correction stages.
3.3.4 Radiative-modelled atmospheric correction. The principal atmospheric correction was carried out by using the Atmospheric CORrection Now (ACORN)
software that is based on licensed MODTRAN technology to determine the
atmospheric parameters (Kruse et al. 2003; ACORNTM 5.0, 2004). A few correction
sessions were carried out with different input values of estimated water vapour,
atmospheric visibility, average altitude and artefact suppression type (ACORNTM
5.0, 2004). The Thompson Ramo Woolridge (TRW) curvature function of the
Hyperion (Pearlman et al. 2003) was integrated in the ACORN correction.
3.3.5 Noise reduction. Line curvature estimation and correction: line curvature
(smile effect), which significantly exists in pushbroom systems, refers to an acrosstrack shift from a centre wavelength, which is due to the change of dispersion angle
with field position (Goodenough et al. 2003, Neville et al. 2003). For the VNIR
bands, the shifts range between 2.6 and 3.5 nm whereas for the SWIR bands the
shifts are less than 1 nm (Goodenough et al. 2003). The VNIR variations are about
30% with respect to the 10 nm full width at half maximum (FWHM) of the Hyperion

3917

Downloaded By: [Gersman, Ronen] At: 21:27 16 June 2008

Hyperion mapping in the Danakil Depression, Eritrea

Figure 3. Locations of regions for SNR estimation (a) and their SNR values (b): fan 1 (1),
fan 2 (2), and stream (3).

instrument (ACORNTM 5.0, 2004) and cannot be ignored as they may alter the pixel
spectral response and reduce the total classification accuracies.
Datt et al. (2003) and Goodenough et al. (2003) reported that the smile effect has
a visual expression, as a vertical brightness gradient in the first eigenband of the
radiometric corrected Hyperion cube. To estimate the line curvature effect, we
applied the minimum noise fraction (MNF) transformation (Green et al. 1988,
Jensen 2005) to the raw data (DN) image and to the radiometric corrected image. In
both of the images the brightness gradient was clearly evident in the VNIR range
and was less significant in the SWIR range, where the across-track shifts are
reported to be rather small (figure 4). The line curvature effect was partly corrected
by the ACORN 5.0 radiative-model atmospheric correction using the TRW
parameters measured prior the launch (ACORNTM 5.0, 2004). However, as the
smile effect is a dynamic phenomenon and might change from one scene to another
(K. Staenze, personal communication), the spectral calibration data may have
changed from the provided TRW parameters. Goodenough et al. (2003) and
Pearlman et al. (2003) reported that changes of about 1 nm are possible since the
sensor launch. This leads to residual curvature effects remaining and calls for
specific correction (see below).
Vertical stripe removal (de-striping): a de-striping procedure was applied to the
data in order to overcome a spatial problem which further reduced the remaining
curvature effect. Vertical stripes are often seen in data acquired using pushbroom

Downloaded By: [Gersman, Ronen] At: 21:27 16 June 2008

3918

R. Gersman et al.

Figure 4. The removal of the line-array effect as expressed by the first MNF band of the
VNIR range at three processing levels. Note the visual improvement between the raw data (a),
the ACORN 5.0 smile corrected data (b) and the de-striped data (c).

(line-array) technology and may be caused by factors such as detector nonlinearities,


movement of the slit with respect to the focal plane, and temperature effects (Kruse
et al. 2003). The correction of the vertical stripes is based on adjusting the brightness
of each image column (in all bands) based on a calculated offset relative to the
average detector response of the scene (Kruse 1988, Kruse et al. 2003). In this study
the de-striping process followed the first atmospheric correction to minimize
changes in the raw data. The procedure was done according to Kruse (1988) and
Kruse et al. (2003). Asner and Heidebrecht (2003), Cudahy et al. (2002), Datt et al.
(2003) and Goodenough et al. (2003) have also reported successful vertical destriping results using similar methods. The de-striping process reduced the residual
line-array effect to a negligible level (figures 4 and 5) and slightly improved the SNR,
especially in the NIRSWIR range (figure 6).
Residual noise removal: a MNF procedure (ENVI 2003, Green et al. 1988, Jensen
2005) was used to reduce the residual noise after the de-striping process. According
to Datt et al. (2003), it is best to handle the VNIR and SWIR data separately and
combine them after the MNF filtering process, rather than use the entire spectral
range. In this study the MNF filtering process (forward MNF, removing the noisy
channels, backward MNF) yielded similar results when all the bands were treated as
a whole, and when the VNIR and the SWIR bands were treated separately. The
transition between the VNIR and the SWIR sensors is reflected in places as a
spectral jump between bands 57 (932.41 nm) and 79 (932.64 nm) (figure 7).
Misalignment of the SWIR and VNIR arrays, which was reported by
Goodenough et al. (2003) exists in the MNF of the separate detectors but was
not visible in the MNF result of all the 196 bands (whole range). Based on the lower
jump between bands 57 and 79 (figure 7) of our data presented, we chose to work

3919

Downloaded By: [Gersman, Ronen] At: 21:27 16 June 2008

Hyperion mapping in the Danakil Depression, Eritrea

Figure 5.

Results of vertical stripes removal: band 216 (a) before and (b) after de-striping.

with the MNF transformation of the entire spectral range (figure 8). The inverse
MNF procedure included the first nine eigenbands based on the asymptotic curve
behaviour obtained in eigenvalues higher than 9.

Figure 6.

The effect of de-striping on the SNR.

Downloaded By: [Gersman, Ronen] At: 21:27 16 June 2008

3920

R. Gersman et al.

Figure 7. Comparison between results of residual noise reduction. Reflectance of the MNF
treatment of the entire spectral range (196 bands) and separate MNF of VNIR and SWIR
ranges (VNIR + SWIR). Gaps are due to removal of water vapour absorption bands (119
132, 165190) and VNIRSWIR overlapping (bands 58, 79).

Figure 8. Comparison between MNF eigenvalues of the radiomertically corrected, atmospherically corrected and de-striped data. The first 35 MNF bands of the total 196 are
included.

3921

Downloaded By: [Gersman, Ronen] At: 21:27 16 June 2008

Hyperion mapping in the Danakil Depression, Eritrea

Figure 9. Locations of possible sites for empirical line (EL) correction: (1) basalt (sample
err-86); (2) aeolian deposit (sample err-84); (3) rhyolite pebbles (sample err-83, this sample
was collected about 1500 m NNW of the ROI location); (4) fine sand (sample err-51c). See
table 1 for the EL combinations and table 2 for compositions of samples err-84 and err-51c.

3.3.6 Fine tuning atmospheric correction using the EL method. After the first
treatment process, in which dead pixels were gently removed, curvature alignment
was corrected, and the atmospheric attenuation was removed by ACORN, a fine
tuning stage was applied to the data. This was done by generating meticulous
calibration alignment using ground targets and the EL procedure (Ben-Dor et al.
1994, Kruse 1998; Clark et al. 2002; ENVI 2003). For that purpose we examined
three combinations of areas with different albedos: dark, semi-dark, semi-bright and
bright (figure 9, table 1). The selected areas included recent basalt as the dark target
(1 in figure 9); rhyolite pebbles (semi-dark target, 3 in figure 9); fine sand from a
Table 1. Components used for the empirical line (EL) corrections. Combination A yielded the
best result.
Components for EL
EL correction
A
B
C

Fine Sand

Rhyolite Pebbles

Aeolian Deposit

Recent Basalt

x
x
x

x
x
x

Downloaded By: [Gersman, Ronen] At: 21:27 16 June 2008

3922

R. Gersman et al.

fluvial plain (semi-bright target, 4 in figure 9); and an aeolian deposit (bright target,
2 in figure 9). Selection of the best EL result was done after comparison between the
Hyperion-corrected reflectance curves of eight known ground targets which served
as validation sites and their corresponding laboratory-measured spectra. The
comparison included visual examination and scoring of four spectral parameters
(VNIR absorptions, SWIR absorptions, curve shape and total reflectance) and
spectral correlation using the CORREL function of the EXCEL program. The
correlations were applied to the VNIR range and the SWIR range separately, as well
as to the entire spectral range, excluding the strong water absorption regions at 1400
and 1900 nm.
3.3.7 Classification. After setting up the atmospheric correction and receiving a
confident ground reflectance cube, we moved forward for the thematic part. In this
stage, we used two methods for end-member selection to classify the area. The first
method followed the approach suggested by Kruse (1998) and Kruse et al. (2003),
namely, the Pixel Purity Index (PPI). The basic concept is to reduce the data to
manageable levels by finding those pixels in the image that can explain every other
pixel in terms of a mixing model without a priori knowledge of the ground surface
(Boardman 1993, 1998, Boardman et al. 1995, Kruse 1998). The spectrally purest
(extreme) pixels in the image are allocated by repeatedly projecting n-dimensional
scatter-plots onto a random unit vector. Given a user-threshold, those pixels that
were recorded as extreme most of the times are defined as pure pixels (Kruse et al.
1997). The PPI and the n-dimensional visualizer tools enabled selecting the major
end-members in the scene. In the second method, diagnostic absorption bands of
defined ground targets were identified and mapped based on a priori knowledge of
the composition and the laboratory-measured reflectance signature of these targets.
The Spectral Angle Mapper (SAM) (Kruse et al. 1993, Jensen 2005) is an excellent
tool for that purpose and thus was used for classifying the images pixels. The
strategy was to perform a general mapping using the PPI, and then to conduct an
intimate mapping procedure for specific targets using the previous knowledge.
3.3.8 Geometric correction. The Hyperion image was geo-referenced relative
to a rectified ASTER scene, using a first-degree polynomial of selected
ground control points (GCP). Geo-referencing was carried out using the
Environmental Systems Research Institute (ESRI) ArcView software, version 9.0
(http://www.esri.com/).
4.
4.1

Results
Mineralogical data

4.1.1 XRD analysis. The modal compositions of the hydrothermally altered rock
samples from Darere and Illegedi (figure 1) are similar and include quartz,
potassium-feldspar, plagioclase, kaolinite and sulphates (table 2). The samples used
for the EL atmospheric removal process included mainly quartz, potassium-feldspar
and plagioclase, minor mica/illite and traces of chlorite and amphibole.
4.1.2 Laboratory reflectance measurements. The reflectance of samples from
Darere (all err-41 and dr-2 samples in figure 10 and table 2) exhibited strong
absorption features of kaolinite and ferric iron and minor signatures of alunite.
Other reflectance curves exhibited similarities to spectra of potassium-feldspar and
opal (Clark et al. 1993, figure 10(b)). Two samples from Illegedi (samples err-42 and

Downloaded By: [Gersman, Ronen] At: 21:27 16 June 2008

Hyperion mapping in the Danakil Depression, Eritrea

3923

Table 2. Mineralogical compositions of selected samples of hydrothermally altered rocks and


of alluvial plains, as detected by the X-ray diffraction method. Samples err-37, err-39 and err51 consists of fine arkosic sand from fluvial plains. Sample err-84 is from an aeolian deposit.
Sample No.
Mineral
Quartz
K-feldspar
Plagioclase
Kaolinite
Mica/Illite
Chlorite
Alunite
Jarosite
Gypsum*
Amphibole
Tridymite
Cristobalite
Calcite

Hydrothermally altered, Alid

Calibration samples

dr-2 Err-41 Err-41b Err-42 Err-43 Err-69b Err-37 Err-39 Err-51c Err-84

xxx
tr

xx
xx
xx
x

xx
xx
xx

xxx

?
xx
x

tr

?
tr
tr
x
x

?
xx
xxx

xx
x
xx

xx
xx
xx

xx
xx
xx

xxx
?
x

x
tr

x
tr

x
tr

x
x

tr

tr

tr

xx
xx
x

xx
?
x

*Including anhydrite. xxx: dominant; xx: major; x: minor; tr: traces.

Figure 10. (a) Laboratory reflectance spectra of samples from Darere, Alid. Alunite
signatures are strongly masked by kaolinite. The shoulder at ,1770 nm (sample dr-2) implies
its existence. (b) Comparison with orthoclase and opal from Clark et al. (1993). Sample names
(e.g. err-41, dr-2) indicate separate samples from the same location. See table 2 for
mineralogical compositions of selected samples.

Downloaded By: [Gersman, Ronen] At: 21:27 16 June 2008

3924

R. Gersman et al.

err-43 in figure 11 and table 2) were analysed, showing in the first sample strong
features of jarosite and gypsum, and opal features in the second sample, both
confirmed by the XRD analysis. Continuum removal (CR) of the reflectance curves
from 1500 to 1600 nm revealed an absorption feature at 1558 nm in four samples of
hydrothermally altered rocks from Alid (figure 12): err-41c, err-41b-c (from Darere),
err-42a, and err-42b (from Illegedi). This feature is attributed to the existence of
ammonium in silicate minerals (Krohn and Altaner 1987, Baugh et al. 1998, Bishop
et al. 2002). Absorption at 2254 nm (possibly due to iron-hydroxyl vibration) was
found to be the most diagnostic spectral feature of the pelitic schist. More
characteristic features are the absorptions of the combination vibration modes of

Figure 11. Laboratory reflectance spectra of samples from Illegedi, Alid. See figure 10 and
table 2 for information of sample names and composition. The small absorption at 430 nm is
diagnostic to ferric iron and jarosite (Hunt and Ashley 1979, Cloutis et al. 2006). Spectra of
sample err-42 are similar to opal (Clark et al. 1993) in shape (also features at ,1130 and
2250 nm).

3925

Downloaded By: [Gersman, Ronen] At: 21:27 16 June 2008

Hyperion mapping in the Danakil Depression, Eritrea

Figure 12. (a) Reflectance curves of ammonium-bearing minerals and of samples of


hydrothermally altered rocks which might contain ammonium. (b) The continuum removal
(CR) of the absorption feature at ,1560 nm. The minerals are from Clark et al. (1993) and
Grove et al. (1992). Rock samples from Alid are marked as err-##.

magnesium-hydroxyl (Mg-OH) at 2324 nm10 nm and of the aluminium hydroxyl


(Al-OH) at ,2200 nm in clay minerals.
4.2

Processing of the hyperspectral data

4.2.1 Atmospheric corrections. The most successful radiative-based correction was


obtained using the 1.5pb mode (ACORNTM 5.0, 2004), which also partially
corrected the line array effect (figure 4). However, residual atmospheric absorptions
of oxygen and carbon dioxide were still apparent. To fine tune the ACORN results,
we applied the EL approach. The best end-member combination for the EL
correction comprised the fresh basalt and an aeolian deposit (table 1), which is
dominated by quartz (sample err-84 in table 2). Nevertheless, the existence of calcite
in sample err-84 (table 2) may have introduced a minor false absorption feature at
,2300 nm throughout the entire image.
4.2.2 Classification. Unsupervised classification: end-members for classification
were chosen using a Pixel Purity Index (PPI, Boardman et al. 1995) procedure
with 15 000 iterations with a threshold value of 1.5. A total of 18 411 pixels were
chosen in this process. Seventeen regions of interest (ROI) were finally chosen, in
sizes that range from 19 to 82 pixels (17 10073 800 m2). Two classification
sessions were carried out: the first used the entire spectral range of the Hyperion
(426.822395.5 nm) excluding the water vapour regions (at ,1400 and ,1900 nm)
and the second used the SWIR range only (2062.552395.5 nm). Both classifications
successfully recognized the hydrothermally altered rocks and schist within the Alid

Downloaded By: [Gersman, Ronen] At: 21:27 16 June 2008

3926

R. Gersman et al.

dome, and identified kaolinite-rich areas in the southern part of the scene, which has
a spectral signature similar to that of the altered rocks of the Alid dome (figures 13
and 14). The kaolinite-rich rocks were not visited in the field and are suspected to be
hydrothermally altered.
Supervised classification: This classification was applied to the CR image limited
to the spectral range between 1507 nm (band 136) and 1598 nm (band 145). The goal
of this session was to detect the slight absorption at 1558 nm, which is attributed to
the ammonium-related absorption, as seen in the laboratory spectral reflectance of
the hydrothermally altered samples (figure 12). Band rationing of the CR Hyperion
bands 140 and 141 (1548 nm/1558 nm) was performed prior to the classification.
These bands represent the wavelength location of the ammonium-related absorption
which was discovered in the laboratory-measured reflectance curves, as appears in
the Hyperion spectral resolution (figure 12). The ratioing emphasized hydrothermally altered rocks within Alid as well as three different rock units: black slate,
schist rocks and kaolinite-rich rocks. Four different shapes of CR absorption
features were chosen as end-members for the classification (figure 15). The endmembers included schist rocks, kaolinite-rich rocks (neither was reached in the field
trip), black-slate outcrops and hydrothermally altered rocks in the Alid dome. The
classification result of the entire scene shows that the Alid-type end-member
appears both in the kaolinite-rich and in the black-slate units (figures 14 and 15);
each of these units is characterized by a different reflectance signature. The fourth
end-member is the schist rocks, which differs in shape and depth from the other endmembers.
The Alid-type end-member is spatially associated with hydrothermally altered
rocks within the Alid area (Duffield et al. 1997, Gersman et al. 2006, figure 14).
Furthermore, rocks in the eastern escarpment of the Alid graben are classified both
as hydrothermally altered and as ammonium-bearing (figure 14).
5.

Classification accuracy assessment

A quantitative assessment of the mapping results was not conducted due to two
main reasons: (1) No accurate, detailed spatial and spectral surveys were performed
in the field (see Crouvi 2002, Levin 2002). This is also the reason why a wide-used
error matrix for accuracy assessment was not established. (2) The mineralogical
results of the XRD analysis estimated the bulk mineralogy of the rock samples and
were not quantitative. On the basis of these estimations, the mineralogical
distribution maps could be ranked on a scale of absent to dominant only.
A visual assessment of the classification accuracy of the entire scene can be made
from the map, integrating the major at-surface geological/lithological elements and
the ground observations. This map consists of ground targets that were identified by
the PPI procedure, some of which were visited in the field. Ground targets that were
visited in the field but were not recognized by the PPI procedure are also included.
The map in figure 13 classifies the end-members generated by the PPI procedure.
The main end-members comprise the Neoproterozoic basement, Tertiary and
Quaternary volcanic rocks, fluvial and aeolian deposits and hydrothermally
altered rocks. Mapping with these end-members was carried out using the SAM
procedure.
Schist in the southwest corner of the map is classified as alluvial fan on the map,
probably because of highly weathered surface. Dolomite patches intercalated
between schistose areas are well mapped. Yet, the northern border of the dolomite

3927

Downloaded By: [Gersman, Ronen] At: 21:27 16 June 2008

Hyperion mapping in the Danakil Depression, Eritrea

Figure 13. Integrated thematic map using the Hyperion scene. The map units were chosen
either from the targets identified by the Pixel Purity Index (PPI) procedure, or according to
the laboratory-measured spectrum of samples collected during the field trip. The map is
overlaid on a base image that used an RGB combination of bands 292011 (true colours).

unit merges with an alluvial fan, which is dominated by dolomitic debris. The black
slate is carefully classified towards south, but is also found in the western slope of
Alid which most probably is an erroneous classification. Patches within the black
slate domain are shaded areas, mistakenly classified as basalt. The identification of a

Downloaded By: [Gersman, Ronen] At: 21:27 16 June 2008

3928

R. Gersman et al.

basaltic cover around Alid, however, is correct. The topography has a major effect
on the apparent reflectance (Richter 1998, Levin 2002, Feng et al. 2003, Riano et al.
2003), hence the Alid slopes, which represent the areas rough terrain, are poorly
classified. Yet, a basic distinction between acidic volcanic rocks (a mixture of
rhyolite, ignimbrite and obsidian) and basaltic flows was obtained, as well as a
distinction between the aeolian and alluvial covers. The classification procedure of
the Hyperion did recognize hydrothermal sites within the Alid dome (figure 14): the
sites of Illegedi, Ghinda and Humbebet, described by Duffield et al. (1997), were
identified, whereas Abakri and Asaela (figure 14), both strongly shaded, were
missed. The site of Darere, which was visited in the field trip, was not spectrally
recognized, probably due to relatively dense vegetation that was present at the time
of the overpass.
A group of pixels northeast of Alid were classified as hydrothermally altered
rocks (figure 14). The classified pixels are concentrated in a group and are not
aligned on a residual vertical stripe, nor limited to shaded areas. This pattern
suggests that the classification of those pixels is probably correct.
The classification procedure identified outcrops of schist within Alid (figures 13
and 14). The distribution of the schist outcrops on the map does not necessarily
represent their exact location in the field, since this area is subject to strong

Figure 14. (a) Map of hydrothermally altered areas within the Alid dome and on the eastern
fault escarpment. The classification of the surrounding volcanic rocks is not included; (b)
mapping of the ammonium signal in the entire scene. There are distinct legends for (a) and (b).

Downloaded By: [Gersman, Ronen] At: 21:27 16 June 2008

Hyperion mapping in the Danakil Depression, Eritrea

3929

topographic and shading effects. Yet, the existence of kyanite schist at Illegedi is
known from the field trip (Gersman et al. 2006) and was reported by Duffield et al.
(1997). The identification of rhyolites along the western border of the map, from
Alid towards northeast, is most probably erroneous. In addition, the SWIR-based
classification of alluvial and of aeolian plains yielded poor results. This area has
remains of the vertical stripes a strong residual noise, especially in the SWIR range
of the Hyperion sensor, which could be the cause of this poor classification result.
6.

Mapping of the hydrothermally altered rocks of Alid

The mineralogical assemblage detected around fumaroles and hot springs resembles
the advanced argillic alteration type (Hunt and Ashley 1979, Heald et al. 1987). This
type of alteration contains kaolinite or pyrophyllite and alunite, together with
possible quartz, diaspore, opal and potassium-mica (Hunt and Ashley 1979). The
existence of ammonium sulphate minerals in Alid (tschermigite NH4Al(SO4)2
12H2O, mascagnite (NH4)2(SO4), kokatite (NH4)2Ca(SO4)2 H2O) was reported
by Beyth (1996) and Duffield et al. (1997). Tschermigite and alum (KAl(SO4)2) are
salts that are formed where ammonium- and sulphate-bearing waters react with
crustal rocks at sub-boiling temperatures (up to 66uC in the thermal pools of Alid;
Duffield et al. 1997) and oxidizing atmospheric conditions. Since the primary
dissolved anion within the acidic thermal pools in Alid is sulphate (up to 1767 ppm
in Illegedi) and the main dissolved cations are ammonium, calcium and potassium
(Duffield et al. 1997, Lowenstern et al. 1997), salts containing these constituents are
likely to precipitate as the solutions evaporate (Duffield et al. 1997). These
conditions may be favourable for the incorporation of ammonium ions into alunite
(Altaner et al. 1988, Krohn et al. 1993), aluminous clay minerals (Krohn and
Altaner 1987, Sucha et al. 1998, Bishop et al. 2002, Holloway and Dahlgren 2002) or
potassium-feldspar (Krohn and Altaner 1987, Krohn et al. 1993, Baugh et al. 1998,
Holloway and Dahlgren 2002), depending on the redox conditions of the fluids
(Krohn and Altaner 1987). Incorporation of ammonium into minerals occurs when
an ammonium ions replace alkali cations (usually potassium, which is close in
radius) in the crystal structure. Potassic and sericitic alteration, and advanced
argillic alteration (with alunite) will create potassium substitution sites. Conversely,
kaolinitization, where potassium-bearing minerals are removed (which is the case in
the kaolinite-rich rocks of Alid), and silification will tend to prevent high
ammonium values (Ridgway et al. 1990, Baugh et al., 1998).
Mapping of ammonium in previous studies utilized different absorption features
of the nitrogen-hydrogen (NH) bond representing the vibrational combination
n1 + n3 where n1 and n3 are fundamental vibration modes of the ammonium ion
(NH4 + ) at 3040 cm21 and 3140 cm21, respectively (Krohn and Altaner 1987). Baugh
et al. (1998) used the absorption features at 2120 nm to map buddingtonite, whereas
Bishop et al. (2002) pointed out 1558 nm as a significant ammonium feature. The
fact that the important ammonium-related feature at 2120 nm is absent from our
laboratory spectra should make us very cautious in calling upon ammonium to
explain the 1558 nm feature. Nevertheless, we could not find another mechanism to
explain the observed absorption (Krohn et al. 1993; Baugh et al. 1998, Bishop et al.
2002). Trace amounts of ammonium in minerals might not be detected by the XRD
method: the XRD pattern of buddingtonite (NH4AlSi3O8), for example, can be
overlapped by the potassium-feldspar pattern (Krohn et al. 1993, Baugh et al. 1998)
and that may be the case in sample err-41 (major potassium-feldspar, table 2).

Downloaded By: [Gersman, Ronen] At: 21:27 16 June 2008

3930

R. Gersman et al.

Sample err-42 (quartz dominant, table 2), contains trace amounts of mica/illite.
Drits et al. (1997) pointed out that XRD analyses of ammonium-bearing illite did
not always indicate the presence of structural ammonium in the lattice. In this case
reflectance spectroscopy seems to be more sensitive to the existence of ammonium
than XRD (Krohn et al. 1993, Drits et al. 1997, Baugh et al. 1998). Krohn et al.
(1993) claimed that minimum concentration for spectral detection of ammonium
from whole-rock samples is approximately 0.06% of (NH4)2O by weight, depending
upon the samples lithology. It is suggested, therefore, that the absorption at
1558 nm is due to the existence of ammonium-bearing phyllosilicates, which were
not detected by the XRD analysis.
The detection of the feature at 1558 nm in the laboratory spectra has led us to seek
an equivalent feature in the Hyperion dataset. As the SNR around 1558 nm is
relatively high and no overlapping with other chromophors exists, this band was
further studied and emphasized four major lithologies (figure 14(a), figure 15):
hydrothermally altered areas in Alid (Alid type), kaolinite-rich areas, black slate/
phyllite, and possibly pelitic schist, all maybe containing ammonium-bearing
minerals (Krohn et al. 1993, Baugh et al. 1998, Holloway and Dahlgren 2002). The
spatial association between the hydrothermally altered areas within Alid and the
areas where the feature attributed to ammonium was detected (figure 14(b)), and the
identification of the 1558 nm feature in a probably hydrothermally altered zone
(kaolinite-rich type, figure 14(b)) support the possibility that the spectral absorption
feature at 1558 nm may represent ammonium at the surface. The 1558 nm signal
could have been distinguished from areas which did not display this feature
(figure 16). It was not aligned on residual vertical stripes, nor restricted to shaded
areas. It also did not originate from the reflectance curves used for the EL correction
and it is not a carbon dioxide or an oxygen residual absorption. The laboratory
spectral measurements and studies made by Beyth (1996) and Duffield et al. (1997)
on the same area led us to suggest that the Hyperion was able to detect the 1558 nm

Figure 15. (a) Continuum removal (CR) of the apparent reflectance image, in the range of
15001600 nm (shaded areas). A laboratory-measured feature from a hydrothermally altered
rock (dashed line) is added for comparison. (b) Examples of apparent reflectance curves of
end-members from (a).

3931

Downloaded By: [Gersman, Ronen] At: 21:27 16 June 2008

Hyperion mapping in the Danakil Depression, Eritrea

Figure 16. Continuum removal (CR) of the reflectance spectra used for the empirical line
correction (err-84, err-86) and of a feature at 1558 nm.

spectral feature around this specific Alid area. This finding together with other
mapping units shows that despite the Hyperion limitations, it is useful for mapping
mineralogical in remote areas. Furthermore, it stresses the need for hyperspectral
data and technology in general.
7.

Summary and conclusions

A careful processing of the Hyperion data followed well-known techniques and


mapped the area well according to previous knowledge. A ground data reconnaissance was performed along with careful laboratory analyses of the samples. The
unsupervised classification approach separated between different types of rock
groups, whereas the supervised classification pinpointed the unique spectral feature at
around 1558 nm, arguably ammonium-related. This feature has a strong spatial
association with kaolinite-rich, hydrothermally altered rocks. Absorption features in
the SWIR-1 are significant because they are positioned in a high SNR spectral range
that enables ammonium-bearing minerals to be detected quickly and nondestructively by a variety of airborne and ground-based sensors (Krohn et al. 1988,
1993) and now, presumably, also by the spaceborne Hyperion sensor. The ability of
the Hyperion to detect ammonium spectral signatures has not been previously
reported. While the mapping of the ammonium-related feature is arguable, the
suggested mineralogical mapping results demonstrate that the extraction of
information from hyperspectral data is yet to be completed, and points out the
significance of the smaller, indicative absorption features for this purpose. The need
and use of hyperspectral technology for mapping remote areas and the potential
success of such mapping are well demonstrated here.

Downloaded By: [Gersman, Ronen] At: 21:27 16 June 2008

3932

R. Gersman et al.

Acknowledgements
Without the support of Mr H. Goder, Israel ambassador to Eritrea, this project
could not have been accomplished. We thank A. Mushkin, S. Peeri and Y. Golani
for helping with the ASTER and Hyperion images, T. Medhin and staff geologists
of the Department of Mines, Asmara, for sharing with us their profound knowledge
during the field trip, A. Sandler for the XRD analysis and A. Kultunov for helping
with the processing of the Hyperion data. Careful and thoughtful comments by Ed
Cloutis and an anonymous reviewer helped to improve this manuscript and are
greatly appreciated. This work was funded by the US Agency for International
Development, Bureau for Economic Growth, Agriculture and Trade, Project
No. C23-001 and Award No. Ta-MOU-03-C23-001.
References
ABBATE, E., WOLDEHAIMANOT, B., BRUNI, P., FALORNI, P., PAPINI, M., SAGRI, M.,
GIRMAY, S. and TECLE, T.M., 2004, Geology of the Homo-bearing Pleistocene
Dandiero Basin (Buya region, Eritrean Danakil Depression). Rivista Italiana di
Paleontologia e Stratigrafia, 10, pp. 534.
ABRAMS, M., HOOK, S. and RAMACHANDRAN, B., 2002, ASTER User Handbook. Available
online at: http://asterweb.jpl.nasa.gov/documents.asp (accessed 21 January 2008).
ACORNTM 5.0, 2004, Tutorial, ImSpec LLC, Advanced Imaging and Spectroscopy, ImSpec
LLC.
ALTANER, S.P., FITZPATRICK, J.J., KROHN, M.D., BETHKE, P.M., HAYBA, D.O., GOSS, J.A.
and BROWN, Z.A., 1988, Ammonium in alunites. American Mineraogist, 73, pp.
145152.
ASNER, G.P. and HEIDEBRECHT, K.B., 2003, Imaging spectroscopy for desertification studies:
comparing AVIRIS and EO-1 Hyperion in Argentina drylands. IEEE Transactions on
Geoscience and Remote Sensing, 41, pp. 12831296.
BARRY, P., 2001, EO-1/Hyperion science data users guide, level 1_B. TRW Space, Defense &
Information Systems, No. HYP.TO.01.077, Rev. Public release L1_B.
BAUGH, W.M., KRUSE, F.A. and ATKINSON, W.W., JR., 1998, Quantitative geochemical
mapping of ammonium minerals in the Southern Cedar Mountains, Nevada, using
the Airborne Visible/Infrared Imaging Spectrometer (AVIRIS). Remote Sensing of
Environment, 65, pp. 292308.
BECK, R., 2003, EO-1 User guide, V.2.3. Available online at http://eo1.usgs.gov and http://
eo1.usgs.gov/documents.php (accessed 21 January 2008). University of Cincinnati,
Ohio.
BEN-DOR, E., KRUSE, F.A., LEFKOFF, A.B. and BANIN, A., 1994, Comparison of three
calibration techniques for utilization of GER 63-channel aircraft scanner data of
Makhtesh Ramon, Negev, Israel. Photogrammetric Engineering and Remote Sensing,
60, pp. 13391354.
BEYTH, M., 1996, Preliminary assessment of the Alid geothermal field, Eritrea. Israel
Geological Survey Current Research, 10, pp. 124128.
BEYTH, M., AVIGAD, D., WETZEL, H.-U., MATHEWS, A. and BERHE, S.M., 2003, Crustal
exhumation and indications for Snowball Earth in the East African orogen: north
Ethiopia and east Eritrea. Precambrian Research, 123, pp. 187201.
BISHOP, J.L., BANIN, A., MANCINELLI, R.L. and KLOVSTAD, M.R., 2002, Detection of soluble
and fixed NH4 + in clay minerals by DTA and IR reflectance spectroscopy: a potential
tool for planetary surface exploration. Planetary and Space Science, 50, pp. 1119.
BOARDMAN, J.W., 1993, Automated spectral unmixing of AVIRIS data using convex
geometry concepts. In Summaries, Fourth JPL Airborne Geoscience Workshop, JPL
Publication 93-26, vol. 1 (Pasadena, CA: Jet Propulsion Laboratory), pp. 1114.
BOARDMAN, J.W., 1998, Leveraging the high dimensionality of AVIRIS data for improved
sub-pixel target unmixing and rejection of false positives: mixture tuned matched

Downloaded By: [Gersman, Ronen] At: 21:27 16 June 2008

Hyperion mapping in the Danakil Depression, Eritrea

3933

filtering. In Summaries of the 7th Annual JPL Airborne Geoscience Workshop,


Pasadena, CA (Pasadena, CA: Jet Propulsion Laboratory), p. 51.
BOARDMAN, J.W., KRUSE, F.A. and GREEN, R.O., 1995, Mapping target signatures via partial
unmixing of AVIRIS data. In Summaries, Fifth JPL Airborne Earth Science
Workshop, JPL Publication 95-1, 1 (Pasadena, CA: Jet Propulsion Laboratory), pp.
2326.
CLARK, R.N. and ROUSH, T.L., 1984, Reflectance spectroscopy: quantitative analysis
techniques for remote sensing applications. Journal of Geophysical Research, 89, pp.
63296340.
CLARK, R.N., SWAYZE, G.A., GALLAGHER, A.J., KING, T.V.V. and CALVIN, W.M., 1993, The
US Geological Survey, Digital Spectral Library: Version 1: 0.2 to 3.0 microns. US
Geological Survey Open File Report 93-592.
CLARK, R.N., SWAYZE, G.A., LIVO, K.E., KOKALY, R.F., KING, T.V.V., DALTON, J.B.,
VANCE, J.S., ROCKWELL, B.W., HOEFEN, T. and MCDOUGAL, R.R., 2002, Surface
reflectance calibration of terrestrial imaging spectroscopy data: a tutorial using
AVIRIS. In Summaries of the 11th Annual JPL Airborne Geoscience Workshop,
Pasadena, CA (Pasadena, CA: Jet Propulsion Laboratory), pp. 3858.
CLOUTIS, E.A., HAWTHORNE, F.C., MERTZMAN, S.A., KRENN, K., CRAIG, M.A.,
MARCINO, D., METHOT, M., STRONG, J., MUSTARD, J.F., BLANEY, D.L., BELL, J.F.
III. and VILAS, F., 2006, Detection and discrimination of sulfate minerals using
reflectance spectroscopy. Icarus, 184, pp. 121157.
CROUVI, O., 2002, Geomorphic mapping using hyperspectral remote sensing: the Wadi
Raham alluvial fan as a case study. MSc Thesis, Institute of Earth Sciences, The
Hebrew University, Jerusalem.
CUDAHY, T.J., HEWSON, R.D., HUNTINGTON, J.F., QUIGLEY, M.A. and BARRY, P.S., 2001,
The performance of the satellite-borne Hyperion Hyperspectral VNIR-SWIR imaging
system for mineral mapping at Mount Fitton, South Australia. In Proceedings of the
IEEE 2001 International Conference on Geoscience and Remote Sensing, Sydney,
IEEE, New Jersey, pp. 913.
CUDAHY, T.J., RODGER, A.R., BARRY, P.S., MASON, P., QUIGLEY, M.A., FOLKMAN, M.A.
and PEARLMAN, J., 2002, Assessment of the stability of the Hyperion SWIR module
for hyperspectral mineral mapping using multi-date images from Mount Fitton,
Australia. In Proceedings of the IEEE 2002 International Conference on Geoscience and
Remote Sensing, Toronto, IEEE, New Jersey, pp. 2428.
DATT, B., MCVICAR, T.R., VAN NIEL, T.G., JUPP, D.L.B. and PEARLMAN, J.S., 2003,
Preprocessing EO-1 Hyperion hyperspectral data to support the application of
agricultural indexes. IEEE Transactions on Geoscience and Remote Sensing, 41, pp.
12461259.
DRITS, V.A., LINDGREEN, H. and SALYN, A.L., 1997, Determination of the content and
distribution of fixed ammonium in illite-smectite by x-ray diffraction: application to
North Sea illitesmectite. American Mineralogist, 82, pp. 7987.
DRURY, S.A., KELLEY, S.P., BERHE, S.M., COLLIER, R.E.L. and ABRAHA, M., 1994,
Structures related to Red Sea evolution in northern Eritrea. Tectonics, 13, pp.
13711380.
DUFFIELD, W.A., LEAKE, W., BULLEN, T.D., KAHSAI, G., CLYNNE, M.A., KIDANE, W.,
FOURNIER, R.O., THEODORES, T., JANIK, C.J., LANPHERE, M.A., LOWENSTERN, J. and
SMITH, J.G., 1997, Geothermal potential of the Alid Volcanic Center, Danakil
Depression, Eritrea. US Geological Survey, USA, Ministry of Energy, Mines and
Water Resources, Eritrea.
ENVI, 2003, ENVI Users Guide, Research Systems, Inc., Boulder, CO, USA.
FENG, J., RIVARD, B. and SANCHEZ-AZOFEIFA, A., 2003, The topographic normalization of
hyperspectral data: implications for the selection of spectral end members and
lithologic mapping. Remote Sensing of Environment, 85, pp. 221231.

Downloaded By: [Gersman, Ronen] At: 21:27 16 June 2008

3934

R. Gersman et al.

GERSMAN, R., BEN-DOR, E., AVIGAD, D., BEYTH, M., ABRAHA, M. and KIBREAB, A., 2006,
Hyperspectral remote sensing as a tool for geological exploration, Northern
Danakil Depression, Eritrea. The Geological Survey of Israel, Report GSI/08/06,
Jerusalem.
GHEBREAB, W., 1998, Tectonics of the Red Sea region reassessed. Earth Science Reviews, 45,
pp. 144.
GHEBREAB, W., 1999a, Pan-African and Red Sea tectonics of Eastern Eritrea. Faculty of
Science and Technology, Uppsala University, Uppsala.
GHEBREAB, W., 1999b, Tectono-metamorphic history of Neoproterozoic rocks in eastern
Eritrea. Precambrian Research, 98, pp. 83105.
GHEBREAB, W., TALLBOT, C.J. and PAGE, L., 2005, Time constraints on exhumation of the
East African Orogen from field observations and 40Ar/39Ar cooling ages of low-angle
mylonites in Eritrea, NE Africa. Precambrian Research, 139, pp. 2041.
GOODENOUGH, D.G., DYK, A., NIEMANN, K.O., PEARLMAN, J.S., CHEN,Hao, HAN, T.,
MURDOCH, M. and WEST, C., 2003, Processing Hyperion and ALI for forest
classification. IEEE Transactions on Geoscience and Remote Sensing, 41, pp.
13211331.
GREEN, A.A., BERMAN, M., SWITZER, P. and CRAIG, M.D., 1988, A transformation for
ordering multispectral data in terms of image quality with implications for noise
removal. IEEE Transactions on Geoscience and Remote Sensing, 26, pp. 6574.
GREEN, R.O., PAVRI, B.E. and CHRIEN, T.G., 2003, On-orbit radiometric and spectral
calibration characteristics of EO-1 Hyperion derived with an underflight of AVIRIS
and in situ measurements at Salar de Arizaro, Argentina. IEEE Transactions on
Geoscience and Remote Sensing, 41, pp. 11941203.
GROVE, C.I., HOOK, S.J. and PAYLOR, E.D., 1992, Laboratory reflectance spectra for 160
minerals 0.42.5 micrometers. JPL Publication 92-2, Jet Propulsion Laboratory,
Pasadena, CA.
HEALD, P., FOLEY, N. and HAYBA, D., 1987, Comparative anatomy of volcanic-hosted
epithermal deposits: acid-sulfate and adularia-sericite types. Economic Geology, 82,
pp. 126.
HOLLOWAY, J.M. and DAHLGREN, R.A., 2002, Nitrogen in rock: occurrence and
biogeochemical implications. Global Biogeochemical Cycles, 16, pp. 11181135.
HUBBARD, B.E., CROWLEY, J.K. and ZIMBELMAN, D.R., 2003, Comparative alteration
mineral mapping using visible to shortwave infrared (0.42.4 mm) Hyperion, ALI, and
ASTER imagery. IEEE Transactions on Geoscience and Remote Sensing, 41, pp.
14011410.
HUNT, G.R. and ASHLEY, R.P., 1979, Spectra of altered rocks in the visible and near infrared.
Economic Geology, 74, pp. 16131629.
JENSEN, J.R., 2005, Introductory Digital Image Processing. A Remote Sensing Perspective, 3rd
edn (Upper Saddle River, NJ: Prentice Hall).
KROHN, M.D. and ALTANER, S.P., 1987, Near-infrared detection of ammonium minerals.
Geophysics, 52, pp. 924930.
KROHN, M.D., ALTANER, S.P. and HAYBA, D.O., 1988, Distribution of ammonium minerals
at Hg/Au-bearing hot springs deposits. In Proceedings of the Bulk-Mineable Precious
Metal Deposits of the Western United States Symposium, Geological Society of
Nevada, R.W. Schafer, J.J. Cooper, and P.G. Vikre (Eds), Geological Society of
Nevaza, Reno, pp. 661680.
KROHN, M.D., KENDALL, C., EVANS, J.R. and FRIES, T.L., 1993, Relations of ammonium
minerals at several hydrothermal systems in the western U.S. Journal of Volcanology
and Geothermal Research, 56, pp. 401413.
KRUSE, F.A., 1988, Use of airborne imaging spectrometer data to map minerals associated
with hydrothermally altered rocks in the Northern Grapevine Mountains, Nevada,
and California. Remote Sensing of Environment, 24, pp. 3151.

Downloaded By: [Gersman, Ronen] At: 21:27 16 June 2008

Hyperion mapping in the Danakil Depression, Eritrea

3935

KRUSE, F.A., 1998, Advances in hyperspectral remote sensing for geologic mapping and
exploration. In Proceedings 9th Australasian Remote Sensing Conference, Sydney,
Australia, http://www.hgimaging.com/FAK_Pubs.htm
KRUSE, F.A., 2003, Preliminary results hyperspectral mapping of coral reef systems using
EO-1 Hyperion, Buck Island and U.S. Virgin Islands. In Proceedings of the 12th JPL
Airborne Geoscience Workshop, JPL Publication 04-6 (Pasadena, CA: Jet Propulsion
Laboratory), pp. 157173.
KRUSE, F.A., BOARDMAN, J.W. and HUNTIGTON, J.F., 2003, Comparison of airborne
hyperspectral data and EO-1 Hyperion for mineral mapping. IEEE Transactions on
Geoscience and Remote Sensing, 41, pp. 13881400.
KRUSE, F.A., LEFKOFF, A.B., BOARDMAN, J.B., HEIDEBRECHT, K.B., SHAPIRO, A.T.,
BARLOON, P.J. and GOETZ, A.F.H., 1993, The spectral image processing system
(SIPS) interactive visualization and analysis of imaging spectrometer data. Remote
Sensing of Environment, 44, pp. 145163.
KRUSE, F.A., RICHARDSON, L.L. and AMBROSIA, V.G., 1997, Techniques developed for
geologic analysis of hyperspectral data applied to near-shore hyperspectral ocean
data. In Proceedings, ERIM 4th International Conference, Remote Sensing for Marine
and Coastal Environments, vol. I (Ann Arbor, MI: Environmental Research Institute
of Michigan), pp. I-233I-246.
LEVIN, N., 2002, Quantitative mapping of the soil rubification process on the
coastal sand dunes of Israel using an Airborne CASI hyperspectral sensor. The
sand dunes f Ashdod as a case study. MA thesis, Department of Geography
and Human Environment, Tel-Aviv University, Tel-Aviv, Israel [in
Hebrew].
LOWENSTERN, J.B., JANIK, C.J., TESFAI, T. and FOURNIER, R.O., 1997, Geochemical study of
the Alid hydrothermal system, Danakil Depression, Eritrea. Stanford Geothermal
Program Report SGP-TR-150, Stanford University, pp. 3744.
LOWENSTERN, J.B., CLYNNE, M.A. and BULLEN, T.D., 1998, Comagmatic A-type granophyre
and rhyolite from the Alid volcanic center, Eritrea, northeast Africa. Journal of
Petrology, 38, pp. 17071721.
NEVILLE, R.A., SUN, L. and STAENZ, K., 2003, Detection of spectral line curvature in imaging
spectrometer data. In Algorithms and Technologies for Multispectral, Hyperspectral,
and Ultraspectral Imagery IX, S.S. Shen, and P.E. Lewis (Eds). Proceedings of the
SPIE, 5093, SPIE, Bellingham, WA, pp. 144154.
PEARLMAN, J.S., BARRY, P.S., SEGAL, C.C., SHEPANSKI, J., BEISO, D. and CARMAN, S.L.,
2003, Hyperion, a space-based imaging spectrometer. IEEE Transactions on
Geoscience and Remote Sensing, 41, pp. 11601173.
RIANO, D., CHUVIECO, E., SALAS, J. and AGUADO, I., 2003, Assessment of different
topographic corrections in Landsat-TM data for mapping vegetation types. IEEE
Transactions on Geoscience and Remote Sensing, 41, pp. 10561061.
RICHTER, R., 1998, Correction of satellite imagery over mountainous terrain. Applied Optics,
37, pp. 40044015.
RIDGWAY, J., APPLETON, J.D. and LEVINSON, A.A., 1990, Ammonium geochemistry in
mineral exploration a comparison of results from the American cordilleras and the
southwest Pacific. Applied Geochemistry, 5, pp. 475489.
ROWAN, L.C. and MARS, J.C., 2003, Lithologic mapping in the mountain pass, California
area using Advanced Spaceborne Thermal Emission and Reflection Radiometer
(ASTER) data. Remote Sensing of Environment, 84, pp. 350366.
SOURIOT, T. and BRUN, J.P., 1992, Faulting and block rotation in the Afar triangle, East
Africa: the Danakil crank-arm model. Geology, 20, pp. 911914.
SUCHA V., ELSASS, F., EBERL, D.D., KUCHTA, L., MADEJOVA, J., GATES, W.P. and
KOMADEL, P., 1998, Hydrothermal synthesis of ammonium illite. American
Mineralogist, 83, pp. 5867.

Downloaded By: [Gersman, Ronen] At: 21:27 16 June 2008

3936

Hyperion mapping in the Danakil Depression, Eritrea

THOME, K.J., BIGGAR, S.F. and WISNIEWSKI, W., 2003, Cross comparison of EO-1 sensors
and other Earth resources sensors to Landsat-7 ETM + using Railroad Valley Playa.
IEEE Transactions on Geoscience and Remote Sensing, 41, pp. 11801188.
UNGAR, S.G., PEARLMAN, J.S., MENDENHALL, J.A. and REUTER, D., 2003, Overview of the
Earth Observing One (EO-1) mission. IEEE Transactions on Geoscience and Remote
Sensing, 41, pp. 11491159.
WARREN, B.E., 1990, X-Ray Diffraction (New York: Dover).
WEIDNER, V.R. and HSIA, J.J., 1981, Reflection properties of pressed polytetrafluoroethylene powder. Journal of the Optical Society of America, 71, pp. 856861.

Вам также может понравиться