Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

International Journal of Applied Earth Observation

and Geoinformation 5 (2004) 55–68

Analysis of spectral absorption features in hyperspectral imagery


Freek van der Meer a,b,∗
aDepartment of Earth Systems Analysis, International Institute for Earth Observation and Geoinformation ITC,
Hengelosestraat 99, P.O. Box 6, 7500 AA Enschede, The Netherlands
b Department of Technical Earth Sciences, Faculty of Civil Engineering and Geosciences, Delft University of Technology,
Mijnbouwstraat 120, 2628 RX Delft, The Netherlands
Received 8 May 2003; accepted 26 September 2003

Abstract
Spectral reflectance in the visible and near-infrared wavelengths provides a rapid and inexpensive means for determining
the mineralogy of samples and obtaining information on chemical composition. Absorption-band parameters such as the
position, depth, width, and asymmetry of the feature have been used to quantitatively estimate composition of samples
from hyperspectral field and laboratory reflectance data. The parameters have also been used to develop mapping methods
for the analysis of hyperspectral image data. This has resulted in techniques providing surface mineralogical information
(e.g., classification) using absorption-band depth and position. However, no attempt has been made to prepare images of the
absorption-band parameters. In this paper, a simple linear interpolation technique is proposed in order to derive absorption-band
position, depth and asymmetry from hyperspectral image data. AVIRIS data acquired in 1995 over the Cuprite mining area
(Nevada, USA) are used to demonstrate the technique and to interpret the data in terms of the known alteration phases
characterizing the area. A sensitivity analysis of the methods proposed shows that good results can be obtained for estimating
the absorption wavelength position, however the estimated absorption-band-depth is sensitive to the input parameters chosen.
The resulting parameter images (depth, position, asymmetry of the absorption) when carefully examined and interpreted by
an experienced remote sensing geologist provide key information on surface mineralogy. The estimates of depth and position
can be related to the chemistry of the samples and thus allow to bridge the gap between field geochemistry and remote sensing.
© 2003 Elsevier B.V. All rights reserved.
Keywords: Hyperspectral remote sensing; Spectroscopy; Absorption feature analysis; Hydrothermal alteration

1. Introduction and charge transfer processes (e.g., changes in en-


ergy states of electrons bound to atoms or molecules)
When light interacts with a mineral or rock, light associated with transition metal ions such as Fe, Ti,
of certain wavelengths is preferentially absorbed Cr, etc., determine largely the position of diagnostic
while at other wavelengths it is transmitted in the absorption features in the visible- and near-infrared
substance. Reflectance, is defined as the ratio of the wavelength region of the spectra of minerals (Burns,
intensity of light reflected from a sample to the inten- 1993; Adams, 1974, 1975). In addition, vibrational
sity of the light incident on it. Electronic transition processes in H2 O and OH− (e.g., small displacements
of the atoms about their resting positions) produce
∗ Fax: +31-53-487-4336/15-278-1328.
fundamental overtone absorptions in the mid- to
E-mail addresses: vdmeer@itc.nl, f.d.vandermeer@citg.tudelft.nl shortwave infrared part of the spectrum (Hunt, 1977).
(F. van der Meer). The position, shape, depth, and width of these absorp-

1569-8432/$ – see front matter © 2003 Elsevier B.V. All rights reserved.
doi:10.1016/j.jag.2003.09.001
56 F. van der Meer / International Journal of Applied Earth Observation and Geoinformation 5 (2004) 55–68

tion features are controlled by the particular crystal absence of water and hydroxyl, carbonate and sulfate
structure in which the absorbing species is contained determine the absorption features in the SWIR region.
and by the chemical structure of the material. Thus, The hydroxyl is generally bound to Mg or Al. The wa-
variables characterizing absorption features can be ter molecule (H2 O) gives rise to overtones as seen in
directly related to the chemistry and structure of the reflectance spectra of H2 O-bearing minerals. The first
sample. The absorption depth is an indicator for the overtones of the OH stretches occur at about 1.4 ␮m
amount of the material causing the absorption present and the combinations of the H–O–H bend with the
in a sample. Furthermore, the absorption-band depth OH stretches are found near 1.9 ␮m. OH-groups com-
is related to the grain or particle-size as the amount of monly occur in multiple crystallographic sites of a spe-
light scattered and absorbed by a grain is dependent cific mineral and are typically attached to metal ions.
on grain size. A larger grain has a greater internal Thus, there may be more than one OH feature. The
path where photons may be absorbed according to combination of the metal–OH bend plus OH stretch
Beers Law. In smaller grains there are proportionally occurs near 2.2–2.3 ␮m and is diagnostic of miner-
more surface reflections compared to internal photon alogy. Carbonates also show diagnostic vibrational
path lengths, if multiple scattering dominates, the absorption-bands due to the CO3 2− ion at 2.50–2.55
reflectance decreases with increasing grain size. and 2.30–2.35 ␮m, and three weaker bands occur near
Field and laboratory spectra have been used to re- 2.12–2.16, 1.97–2.00, and 1.85–1.87 ␮m.
late absorption features to the chemical composition Quantitative estimates of mineralogical composi-
of samples in the areas of both soil science and miner- tion and chemical analysis on the basis of spectro-
alogy as well as in the area of vegetation science. For scopic data have been demonstrated by many authors
the analysis of hyperspectral image data there are sev- (e.g., Adams, 1975; Swayze and Clark, 1990). Adams
eral techniques available to derive surface composi- (1975) and Cloutis et al. (1986) showed that the iron
tion (e.g., surface mineralogy) from a combination of absorption-bands near 1 and 2 ␮m shift as a function
absorption-band position and depth. However, no such of the Fe:(Fe + Mg) ratio. Similarly, King and Ridley
technique provides spatial information on the varia- (1987) showed this effect for olivines. Mustard (1992)
tion of absorption-band depth, position and shape de- showed that the Fe:(Fe + Mg) ratio can be estimated
spite the fact that these parameters are of vital use in from reflectance spectra. Duke (1994) found subtle
quantitative surface compositional mapping. shifts of the Al–OH absorption-band in muscovites
In this paper, first a review is given of the use of with aluminum composition. As Al is substituted
reflectance spectroscopy to estimate soil/rock geo- by Mg, the crystal becomes distorted causing slight
chemistry and foliar biochemistry using field and labo- changes in Al–OH bond lengths and thus shifts in
ratory spectroscopy. Then, we evaluate spectral feature the absorption-band position of the 2.2 ␮m absorp-
analysis techniques available to map absorption fea- tion appear (Duke, 1994). Hence Duke (1994) found
tures in hyperspectral image data. Thereafter, a simple a way of quantifying the Al content of muscovites
linear interpolation method is introduced to estimate from the Al–OH absorption-band position. Van der
absorption-band parameters from hyperspectral image Meer (1994, 1995) demonstrated that reflectance
data. Finally, in a case study the use of this technique spectroscopy provides a measure for the ratio of cal-
in the analysis of airborne hyperspectral image data cite:dolomite in limestones. In soil science, reflectance
for surface compositional mapping is demonstrated. spectroscopy has been used to quantify soil parame-
ters such as organic matter, total iron, exchangeable
Ca and Mg, and Ph (Ben-Dor et al., 1997; Shepherd
2. Spectral absorption feature analysis on and Walsh, 2002). Recently, Kariuki et al. (2003)
hyperspectral reflectance spectral data show that cation exchange capacity (CEC) can be
estimated from reflectance spectra. In the vegetation
Reflectance spectra of minerals are dominated in sciences, reflectance spectra have been used for years
the visible to near-infrared wavelength range by the to estimate foliar biochemistry. Since the pioneer-
presence or absence of transition metal ions (e.g., Fe, ing work of Curran (1989) on leaf biochemistry and
Cr, Co, Ni; Hunt, 1977; Burns, 1993). The presence or Wessmann et al. (1988) on remote sensing of canopy
F. van der Meer / International Journal of Applied Earth Observation and Geoinformation 5 (2004) 55–68 57

biochemistry many authors have demonstrated that review). Techniques that specifically use absorption-
leaf constituents (notably leaf pigments) can be quanti- band position and depth include (1) the relative
tatively assessed from spectral data (e.g., Curran et al., absorption-band-depth (RBD) approach of Crowley
1992; Fourty et al., 1996). Examples of biochemi- et al. (1989), (2) the spectral feature fitting (SFF)
cals that are estimated include leaf water (Bowman, technique of Clark et al. (1990a) and (3) the Tricorder
1989), chlorophyll (Lichtenthaler et al., 1996) and ni- (Crowley and Swayze, 1995) and TETRACORDER
trogen (Yoder and Pettigrew-Crosby, 1995). Recently, (Clark et al., 2003) algorithms developed at the
Kokaly and Clark (1999) presented a model for de- USGS spectral laboratory. These techniques work
termining the concentration of foliar components on so-called continuum removed reflectance spectra
from reflectance spectra using absorption-band-depth (Fig. 1), thus acknowledging that the absorption in a
analysis and multiple linear regression analysis to spectrum has two components: a continuum and indi-
estimate a relation between reflectance and the con- vidual features. The continuum or background is the
centration of several components (nitrogen and cellu- overall albedo of the reflectance curve. Removing this
lose). Curran et al. (2001) show another example of effectively scales the spectra to 100% when the spec-
biochemical analysis of leaf reflectance spectra using tral curve approaches the continuum. Mathematically
absorption-band position and depth analysis. this can be done as follows (Clark et al., 2003)
L(w) O(w)
Lc (w) = and Oc (w) = (1)
3. Analysis of spectral absorption features in Cl (w) Co (w)
hyperspectral image data where L(w) is the library spectrum as a function of
wavelength, w, O the observed spectrum, Cl the con-
There are various techniques to process hyper- tinuum for the library spectrum, Co the continuum for
spectral imagery in order to obtain surface compo- the observed spectrum, Lc the continuum-removed li-
sitional information on a pixel-by-pixel basis for the brary spectrum, and Oc is the continuum-removed ob-
entire image (see Van der Meer et al., 2001 for a served spectrum.

1.0 Position
Apporent Reflectance (or Normalized Reflectance)

Area(left)
Continuum
Removed
0.8
Spectrum
Area(right)

Depth
0.6
Continuum

0.4
Kaolinite Spectrum

0.5 1.0 1.5 2.0 2.5


Wavelength (Micrometers)

Fig. 1. Definition of the continuum and continuum removal and subsequent definition of absorption feature characteristics.
58 F. van der Meer / International Journal of Applied Earth Observation and Geoinformation 5 (2004) 55–68

Crowley et al. (1989) developed a method of min- tional information is derived and results are vali-
eral mapping from imaging spectrometer data using dated.
RBD images generated directly from radiance data. These mapping methods described all produce
In essence, RBD images provide a local continuum validated surface compositional information (mostly
correction removing any small channel to channel ra- mineralogical maps), but they do not provide infor-
diometric offsets, as well as variable atmospheric ab- mation on the absorption-band position, depth and
sorption and solar irradiance drop off for each pixel asymmetry on a pixel by pixel basis although these
in the data set. To produce a RBD image, several parameters are used in the matching performed. Since
data channels from both absorption-band shoulders are absorption-band parameters are of importance to
summed and then divided by the sum of several chan- quantitative reflectance spectroscopy, here we pro-
nels from the absorption-band minimum. The result- pose a simple linear estimation method to derive such
ing absorption-band-depth image gives the depth of parameters from hyperspectral image data.
an absorption feature relative to the local continuum
which can be used to identify pixels having stronger
absorption-bands indicating that these may represent 4. Absorption-band position, depth and
a certain mineral. asymmetry mapping from hyperspectral image
Spectral feature fitting (embedded in the ENVI soft- data
ware, Clark et al., 1990a) uses continuum removed
pixel spectra, which are compared to continuum ref- As was discussed in previous sections, field or
erence spectra of known mineralogy. A least-squares laboratory reflectance spectra have been used to
fit is calculated band by band between each reference derive compositional information on samples. Usu-
end-member and the unknown (continuum removed) ally this involves a multiple-linear regression of
pixel spectra. A “scale” image is produced for each absorption-band parameters and chemical composi-
endmember selected for analysis by first subtracting tion. The following absorption-band parameters cal-
the continuum-removed spectra from one, thus invert- culated from continuum removed spectra (Fig. 1) are
ing them and making the continuum zero. A large scale often used: (1) the absorption-band position, (2) the
factor is equivalent to a deep spectral feature, while absorption-band depth and (3) the absorption-band
a small scaling factor indicates a weak spectral fea- asymmetry. The relative depth, D, of the absorption
ture. The total root-mean-square (RMS) error is used feature is defined as the reflectance value at the shoul-
to form an RMS error image for each endmember. The ders minus the reflectance value at the absorption-band
ratio of the scale image and the RMS image provides minimum. The depth of an absorption-band, D, can
a “fit” image that is a measure of how well the un- be defined relative to the continuum, Rc , as
known spectrum matches the reference spectrum on a Rb
pixel-by-pixel basis. D=1− (2)
Rc
The (Tricorder and its successor) Tetracorder
(Clark et al., 2003) uses spectral matching algo- where Rb is the reflectance at the band bottom and Rc
rithms carried out in a two step process. First, the is the reflectance of the continuum at the same wave-
local spectral slope (the “continuum”) is estimated length as Rb (Green and Graig, 1985). These Hull quo-
and removed both from reference and observed spec- tient spectra (also referred to as continuum removed
tra. Next, the identification of materials from their spectra) are used to characterize absorption features,
spectra is constrained by (1) the goodness of fit of a known to be attributed to a certain mineral of interest,
spectral feature to a reference, (2) reflectance level, in terms of their position, depth, width, and asymme-
(3) continuum slope, and (4) presence or absence try. The absorption-band position, λ, is defined as the
of key ancillary spectral features. The Tetracorder band having the minimum reflectance value over the
uses these reference continuum-removed-spectral wavelength range of the absorption feature. The asym-
features to compute a weighted fit between un- metry factor, S, of the absorption feature is defined as
known spectra and known library spectra. By means Aleft
S= (3)
of an expert system approach, surface composi- Aright
F. van der Meer / International Journal of Applied Earth Observation and Geoinformation 5 (2004) 55–68 59

where Aleft is the area of the absorption from starting Laboratory spectra as derived from field-based spec-
point to maximum point and the Aright is the area of the trometers such as the ASD Fieldspec and the GER
absorption from maximum absorption point to the end 3700, etc. can be considered continues. Imaging de-
point (shoulder) of the absorption. Values for S range vices discretize the spectrum into a number of broader
from 0 to infinity where S equals 1 for a symmetric (10–20 nm wide) bands and sample the spectrum every
absorption feature. Features in which the area on the 10–20 nm. Thus, Eqs. (2) and (3) need to be approxi-
left-hand side is greater than the area on the right-hand mated for the case of imaging spectrometer data. Here,
side (i.e., absorption features skewed to longer wave- I propose a simple linear approximation method of the
length) will result in asymmetry values greater than absorption feature parameters. I performed all calcula-
1. Features in which the area on the left-hand side is tions on image data using a script that can be run in the
smaller than the area on the right-hand side (i.e., ab- band math operator of the ENVI software. The band
sorption features skewed to shorter wavelength) will center wavelength position is usually used for further
result in asymmetry values between 0 and 1. Okada calculations, thus providing spectral measurements on
and Iwashita (1992) produced waveform characteris- a discrete number of wavelength positions character-
tics on hyperspectral imagery, however the algorithms izing an absorption feature. To accommodate this, a
were not very well documented. simple linear method is proposed to calculate the ab-
The absorption-band parameter definition as dis- sorption feature parameters from image data. Fig. 2
cussed above assumes nearly continuous (contiguous) explains graphically the procedure followed.
spectral data, whereas imaging spectrometers acquire First, for the absorption feature of interest the image
data in a large number of discrete spectral bands. bands are determined that would serve as the shoulders

A B

1.00 1.00

Depth1
C1
Shoulder 2
(S2) Shoulder 1
Absorption-banddepth (%R relative to continuum)

Depth2 (S1)
C2

Absorption depth Absorption point 1


(A1)
0.90 0.90
Absorption point 2
(A2)

Continuum removed spectrum

0.80 0.80

Absorption wavelength

2.10 2.15 2.20 2.25 2.30


Wavelength(micrometers)

Fig. 2. Definition of the parameters used in the linear interpolation of absorption feature characteristics.
60 F. van der Meer / International Journal of Applied Earth Observation and Geoinformation 5 (2004) 55–68

of the absorption feature by visual inspection using wavelength, a positive value if the absorption feature
expert knowledge on spectroscopy. By definition, there is skewed toward the longer wavelength. Given that
is a short wavelength shoulder (shoulder 2, denoted S2 calculation of areas (as in Eq. (3)) leads to erroneous
in Fig. 2) and a long wavelength shoulder (shoulder estimates in coarse sampled spectral data sets, the
1, denoted S1 in Fig. 2). Next, the data are continuum asymmetry parameter of Eq. (10) is a good approxi-
removed using the mentioned shoulders as starting and mation as will be demonstrated in the next section.
ending points. Subsequently, two bands are selected
as the absorption points which will be used in the
interpolation (points A1 and A2 in Fig. 2). Then, the 5. Case study
coefficients C1 and C2 are calculated as
 To demonstrate the use of absorption feature map-
C1 = (depth1 )2 + (S1 − A1 )2 (4) ping as described we will apply the technique to
and hyperspectral image data on an area of hydrother-
 mal alteration. Here data from NASA’s Airborne
C2 = (depth2 )2 + (S2 − A2 )2 (5) Visible/Infrared Imaging Spectrometer (AVIRIS) ac-
From this, the interpolated wavelength position (e.g., quired in 1995 from the Cuprite mining area situated
the wavelength of maximum absorption) can be found some 30 km. south of the town of Goldfield in west-
by interpolating between the shoulders and absorption ern Nevada are used (Fig. 3). The AVIRIS data were
points in the spectrum as converted to reflectance using the empirical line cali-
bration method described by Roberts et al. (1985).
absorption wavelength The Cuprite mining site (Albers and Stewart, 1972;
 
C1 Abrams et al., 1977) is an area of extensive hydrother-
=− × (A1 − A2 ) + A1 (6) mal alteration related to early Miocene volcanism dur-
C1 + C 2
ing which dacite and andesite flows were extruded and
or
hot, acidic brines began circulating through the series
absorption wavelength of rhyolitic basalts, rhyolitic welded ash flows and air
  fall tuffs.
C2
= × (A1 − A2 ) + A2 (7) Silicified rocks form a large irregular patch extend-
C2 + C 1
ing from the middle to the south end of the area. The
The associated absorption-band depth is derived as silicified core represents the most intensely altered
rocks at Cuprite containing quartz, calcite and minor
absorption depth
  alunite and kaolinite. Opalized rocks contain abun-
S1 − absorption wavelength dant opal and as much as 30% alunite and kaolinite.
= × depth1 (8)
S1 − A 1 Locally, an interval of soft, poorly exposed material
or mapped as argillized rock separates fresh rock from
opalized rock. In the argillized rocks, plagioclase is
absorption depth altered to kaolinite, and glass is altered to opal and
 
absorption wavelength − S2 varying amounts of montmorillonite and kaolinite. The
= × depth2 (9)
A2 − S 2 distribution of these alteration assemblages is charac-
teristic for a fossilized hot-spring deposit, which are
The asymmetry factor of the absorption feature is cal-
often mined for gold. The circular distribution of the
culated as
mineral zones, demonstrates that alteration occurred
asymmetry = A−B=(absorption wavelength−S2 ) along a central vent with the lateral mineral zoning
controlled by a decrease of acidity and temperature.
− (S1 − absorption wavelength) (10)
As input to the absorption-band mapping ap-
This operation returns 0 for a perfect symmetric ab- proach, continuum removed spectra were used on a
sorption feature, a negative value for a absorption fea- selected portion of the wavelength spectrum where
ture that is asymmetric and skewed toward the short the absorption-band of interest is found. In this first
F. van der Meer / International Journal of Applied Earth Observation and Geoinformation 5 (2004) 55–68 61

Fig. 3. Location of the study area, hydrothermal alteration map (after Albers and Stewart, 1972) and (inset) AVIRIS image with the mapped
area outlined.

case study, focus was on clay mineral absorption fea- centered at 2.3898 ␮m (shoulder 1) were used. The
tures in the (Shortwave infrared wavelength region) two absorption points used for the interpolation were
SWIR around 2.20 ␮m and carbonate absorption fea- AVIRIS band 201 centered at 2.2805 ␮m (absorption
tures in the SWIR around 2.30 ␮m. Two bands on point 2) and AVIRIS band 210 centered at 2.3700 ␮m
the shoulders of the feature were selected to perform (absorption point 1) for the carbonate feature.
the continuum removal. For the Al–OH clay min- The resulting images are shown in Fig. 4. These
eral absorption features, AVIRIS band 178 centered images can be used to (1) relate absorption feature pa-
at 2.0509 ␮m (shoulder 2) and AVIRIS band 199 rameters to measures of mineral chemistry and (2) to
centered at 2.2606 ␮m (shoulder 1) were used. The describe the hydrothermal alteration system in terms of
two absorption points used for the interpolation were surface composition (mineralogy) and alteration inten-
AVIRIS band 184 centered at 2.1110 ␮m (absorption sity. In particular, the mineralogy as expressed in the
point 2) and AVIRIS band 196 centered at 2.2307 ␮m Al–OH absorption parameter estimates may be of use.
(absorption point 1) for the Al–OH feature. Table 1 summarizes the key mineral groups for each of
For the carbonate feature, the AVIRIS band 197 cen- the alteration phases theoretically to be found in this
tered at 2.2407 ␮m (shoulder 2) and AVIRIS band 212 high sulfidation alteration system (see also Fig. 5).
62 F. van der Meer / International Journal of Applied Earth Observation and Geoinformation 5 (2004) 55–68

Fig. 4. Derived absorption feature parameters: (A) Al–OH position of maximum absorption wavelength, (B) carbonate absorption position
of maximum absorption wavelength, (C) Al–OH depth of the absorption feature (in percentage reflectance relative to the continuum), (D)
carbonate absorption-depth (in percentage reflectance relative to the continuum), (E) asymmetry of the absorption feature ((++) strongly
skewed to longer wavelength, (+/0) weakly skewed to longer wavelength, (0/−) weakly skewed to shorter wavelength, (−−) strongly
skewed to shorter wavelength), and (F) carbonate asymmetry of the absorption feature ((++) strongly skewed to longer wavelength, (+/0)
weakly skewed to longer wavelength, (0/−) weakly skewed to shorter wavelength, (−−) strongly skewed to shorter wavelength).
F. van der Meer / International Journal of Applied Earth Observation and Geoinformation 5 (2004) 55–68 63

Table 1
Hydrothermal alteration as they occur at Cuprite and associated main mineral phases (key minerals indicative for the phase are indicated
in italics)
Phase 1 Phase 2 Phase 3 Phase 4 Phase 5 Phase 6

Propylitic Potassic Argillic Sericitic Advanced argillic Low-temperature advanced


K-feldspar K-feldspar Kaolinite ordered Kaolinite Pyrophyllite Buddingtonite
Chlorite Sericite Illite/smectite Dickite Alunite Kaolinite
Carbonate Montmorillonite Gypsum Sericite Alunite
Montmorillonite

A plot of the Al–OH absorption features (depth and rite. Alunite has a rather symmetric Al–OH absorp-
asymmetry versus wavelength of maximum absorp- tion feature (asymmetry value of 0), whereas kaoli-
tion) from spectral library data is shown in Fig. 6 along nite, jarosite and in particular gypsum would have a
with the interpretation of the surface composition in positive asymmetry (because their Al–OH absorption
terms of the main alteration phases. features are skewed towards longer wavelengths).
The wavelength position of the Al–OH (Fig. 4) The carbonate absorption features show a wave-
can be interpreted in terms of alteration phases. length ranging from around 2.30 –2.35 ␮m, thus from
Short wavelength Al–OH absorption is indicative for the pure dolomite (theoretical wavelength of absorp-
the presence of buddingtonite and alunite indicating tion is 2.30 ␮m) to the pure calcite (theoretical wave-
relative high alteration (low temperature advanced length of absorption is 2.35 ␮m). In addition, chlorite
and advanced argillic alteration phases). Interme- (at 2.32 ␮m), epidote (at 2.33 ␮m), talc (at 2.30 ␮m)
diate wavelength Al–OH absorption is confined to and pyrophyllite (at 2.32 ␮m) show absorption fea-
the argillic alteration phases (kaolinite, gypsum), tures in the 2.30–2.35 ␮m region (Clark et al., 1990b).
whereas the potassic alteration phase (dominated by Most of these minerals are indicative for propylitic
montmorillonite) and the propylitic alteration phase to potassic alteration phases. Pyrophyllite, however,
is dominated by chlorite thus giving relatively long would are interpreted as advanced argillic alteration.
wavelength Al–OH features. The absorption-band Hence, I interpret the carbonate absorption feature
depth (Fig. 6) of the Al–OH feature increases with maps (Fig. 4) as indicative of both calcium-carbonate
intensity of the hydrothermal alteration both because as well as Mg–OH absorption mostly corresponding
of the increase in volume of alteration minerals as to the propylitic and potassic alteration phases.
well as because of the increase in depth of mineral In order to assess the sensitivity of the method to
related absorption-bands. The advanced argillic (and the selected input parameters, we performed a sensi-
low temperature advanced alteration) phase, dom- tivity analysis. A set of 20 pixels known from field
inated by buddingtonite and alunite, show intense survey to contain predominantly alunite were iden-
(deep) absorption features, whereas the argillic and tified on the image. The absorption-band-depth and
the potassic and propylitic alteration phase mineral the absorption-band position were estimated using
generally show less pronounced absorption features. Eqs. (6) and (9) for two sets of absorption param-
The use of the asymmetry feature for mapping and eters varying the values of the parameters A1 and
analysis of the Cuprite data is less straightforward (see A2 . By plotting the pixel spectra we measured, by
Fig. 6A). In principle, by the definition of the asym- hand, the true absorption-band-depth and position and
metry, the minerals with positive asymmetry values calculated the difference between the true and esti-
have Al–OH features skewed to longer wavelengths mated absorption-band-depth and position for the two
and those with a negative asymmetry value corre- parameter-sets. Results are plotted in Fig. 7. These
spond to absorption features skewed to shorter wave- show that the difference between the true and esti-
lengths. Typical minerals that would exhibit a small mated absorption wavelength position lies between
negative (or near zero to small positive) asymmetry the two extremes of plus and minus 15 nm. difference.
value are montmorillonite, buddingtonite and chlo- This seems throughout acceptable given the spec-
64 F. van der Meer / International Journal of Applied Earth Observation and Geoinformation 5 (2004) 55–68

Alteration Zone
Stable Alteration Mineral
Mineral (absorption feature)
Suite
Propylitic Potassic Argillic Sericitic Advanced Low-temp
(phyllic) argillic advanced

Opa (Si-O-Si, Si-O-H)


Quartz Quartz (Si-O-Si)l

K-feldspar (Si-O-Al)
Adularia (Si-O-Al)
Felsdspars Albite (Si-O-Al)
Buddingtonite (NH-O-H)

Dickite (Al-O-H)
Kaolinite ordered (Al-O-H)
Kaolins Kaolinite disordered (Al-O-H)
Halloysite (Al-O-H)

Montmorillonite (Al-O-H)
Micas lllite/Smectite (Al-O-H)
Smectities Sericite (Al-O-H)
(2:1) Pyrophyllite (Al-O-H)
Chlorite (Mg, Fe-OH)

Alunite (Al-O-H)
Anhydrite (none)
Sulfates Jarosite (Fe-O-H)
Gypsum (H-O-H)

Carbonates Calcite (CO3 )

3+
Iron Oxides Goethite (Fe )

Tourmaline (B-O-H)
Diaspore (O-H)
Other Epidote (Mg-O-H)
Minerals Topaz (F-O-H)
Zeolites (H-O-H)

Hydrogen-Ion Activity

Fig. 5. Major alteration zones characterizing the high sulfidation hydrothermal alteration system found at Cuprite and the mineral assemblages
that are commonly found in these zones (modified after Thompson and Thompson, 1996).

tral resolution of the AVIRIS system (approximately change of the parameter set has an effect on individual
10 nm) and the signal to noise ratio (estimated as 50:1 (single pixel) estimates of the absorption wavelength
for a bright 80% reflective target and 10:1 for a dark position, but has relatively little effect on the total set
20% reflective target in the SWIR at 2200 nm). The of 20 estimates. The absorption-band-depth, however,
F. van der Meer / International Journal of Applied Earth Observation and Geoinformation 5 (2004) 55–68 65

Gypsum
6

Phase 3
5
Asymmetry

Phase 5
2
Kaolinite
Phase 6 Albite Phase 1 Jarosite
Kaolinite
Alunite
1 Illite Chlorite
Quartz
Buddingtonite Phase 2
Montmorillonite
0
2.1 2.12 2.14 2.16 2.18 2.2 2.22 2.24 2.26 2.28 2.3
(A) Wavelength (micrometer)
30

Alunite
Absorption-band depth (%R)

25
Buddingtonite

Phase 5
20

Phase 6 Kaolinite
15 Kaolinite

Gypsum
10
Montmorillonite
Chlorite
Illite
Jarosite
Albite
Phase 1
5
Phase 3
Phase 2
Quartz
0
2.1 2.12 2.14 2.16 2.18 2.2 2.22 2.24 2.26 2.28 2.3

(B) Wavelength (micrometer)


Fig. 6. Plot of (A) asymmetry versus absorption wavelength and (B) depth vs. absorption wavelength for key minerals in the alteration
system based on laboratory library spectra (phase 1: propylitic, phase 2: potassic, phase 3: argillic, phase 4: sericitic, phase 5: advanced
argillic, phase 6: low-temperature advanced).
66 F. van der Meer / International Journal of Applied Earth Observation and Geoinformation 5 (2004) 55–68

30
S1 = 2.0509 nm.
S2 = 2.2606 nm.
25 A1 = 2.2307 nm.
True-Estimated Absorption band-depth (%R)

A2 = 2.1110 nm.
N=20
20

15

10

0
-20 -15 -10 -5 0 5 10 15 20

-5
(A)

25
S1 = 2.0509 nm.
S2 = 2.2606 nm.
A1 = 2.2208 nm. 20
A2 = 2.1210 nm.
True-Estimated Absorption band-depth (%R)

N=20 15

10

0
-20 -15 -10 -5 0 5 10 15 20
-5

-10

-15
(B) True-Estimated Absorption wavelength (nm.)

Fig. 7. Results of sensitivity analysis for the estimation of absorption-band parameters for 20 pixels of relatively pure alunite outcrops
using two different parameter sets (see text for discussion).

seems more sensitive to the choice of absorption pa- tion depth being smaller) estimates. This can be at-
rameters. The parameter set shown in Fig. 7B, where tributed to the shape of the absorption features, which
the spectral ‘distance’ between the shoulder and the tends to be convex curved rather than linear starting
parameters A1 and A2 is larger, shows overall slightly from the shoulders. Nevertheless, linears seem a good
better (difference between true and estimated absorp- first order approximation.
F. van der Meer / International Journal of Applied Earth Observation and Geoinformation 5 (2004) 55–68 67

6. Conclusions References

Reflectance spectroscopy has been used to derive Abrams, M.J., Ashley, R.P., Rowan, L.C., Goetz, A.F.H.,
estimates of soil and rock geochemistry and foliar Kahle, A.B., 1977. Mapping of hydrothermal alteration in
the Cuprite mining district, Nevada using aircraft scanner
biochemistry using, to date, mostly field and labora- images for the spectral region 0.46–2.36 ␮m. Geology 5, 713–
tory spectroscopic techniques. These techniques most 718.
often make use of absorption feature characteristics Adams, J.B., 1974. Visible and near-infrared diffuse reflectance:
(e.g., absorption-band wavelength position, depth and spectra of pyroxenes as applied to remote sensing of solid
objects in the solar system. J. Geophys. Res. 79, 4829–4836.
asymmetry), which are combined with geochemical Adams, J.B., 1975. Interpretation of visible and near-infrared
analysis in a multi-linear regression to find empirical diffuse reflectance spectra of pyroxenes and other rock forming
relationships with chemistry of a sample. Absorption minerals. In: Karr, C. (Ed.), Infrared and Raman Spectroscopy
features have also been used as input to spectral feature of Lunar and Terrestrial Materials. Academic Press, New York,
fitting techniques that allow mapping surface compo- pp. 91–116.
Albers, J.P., Stewart, J.H., 1972. Geology and mineral deposits
sition (mainly mineralogy) from hyperspectral image of Esmeralda County, Nevada. Nevada Bureau of Mines Geol.
data. However, by this process, the absorption feature Bull. 78, 1–80.
analysis is confined to input data in a feature fitting. In Ben-Dor, E., Inbar, Y., Chen, Y., 1997. The reflectance spectra
this paper, a simple linear interpolation method is in- of organic matter in the visible near-infrared and short
wave infrared region (400–2500 nm) during a controlled
troduced to estimate absorption-band parameters from decomposition process. Rem. Sens. Environ. 61, 1–15.
hyperspectral image data. By applying this to AVIRIS Bowman, W.D., 1989. The relationships between leaf water status,
data from Cuprite, it has been demonstrated that ab- gas exchange, and spectral reflectance in cotton leaves. Rem.
sorption feature maps correspond favorably with the Sens. Environ. 30, 249–255.
Burns, R.G., 1993. Origin of electronic spectra of minerals in
main alteration phases characterizing the hydrother-
the visible and near-infrared region. In: Pieters, C.M., Englert,
mal system studied. Thus, the derived feature maps P.A.J. (Eds.), Remote Geochemical Analysis: Elemental and
allow enhancing the analysis of airborne hyperspec- Mineralogical Composition. Cambridge University Press, New
tral image data for surface compositional mapping. York, pp. 3–30.
From this sensitivity analysis it can be concluded that Clark, R.N., Gallagher, A.J., Swayze, G.A., 1990a. Material
absorption-band depth mapping of imaging spectrometer
linear estimation techniques to model absorption fea- data using the complete band shape least-squares algorithm
ture parameters is a simple first-order approximation, simultaneously fit to multiple spectral features from multiple
however for accurate depth estimates and modeling materials. In: Proceedings of the Third Airborne Visible/Infrared
complex (doublet and triplet) absorption-bands, higher Imaging Spectrometer (AVIRIS) Workshop, JPL Publication
order polynomials are required. Hence, the next step 90-54, pp. 176–186.
Clark, R.N., King, T.V.V., Kleijwa, M., Swayze, G.A., Vergo,
is to expand in the direction of biochemical and geo- N., 1990b. High spectral resolution reflectance spectroscopy of
chemical mapping using polynomial fitting of spectral minerals. J. Geophys. Res. 95, 12653–12680.
absorption feature analysis from airborne and space- Clark, R.N., Swayze, G.A., Livo, K.E., Kokaly, R.F., Sutley,
borne hyperspectral data. Nevertheless, the linear S.J., Dalton, J.B., McDougal, R.R., Gent, C.A., 2003. Imaging
spectroscopy: earth and planetary remote sensing with the
technique presented in this paper gives a simple ap-
USGS Tetracorder and expert systems. J. Geophys. Res., in
proximation that can be easily implemented by users press.
(for example using the band math operator of ENVI® ) Cloutis, E.A., Gaffey, M.J., Jackowski, T.L., Reed, K.L., 1986.
that are not familiar with programming languages. Calibrations of phase abundance, composition, and particle size
distribution of olivine-orthopyroxene mixtures from reflectance
spectra. J. Geophys. Res. 91, 11641–11653.
Crowley, J.K., Swayze, G.A., 1995. Mapping minerals, amorphous
materials, environmental materials, vegetation, water, ice, and
Acknowledgements other materials: the USGS Tricorder Algorithm. In: Summaries
of the Fifth Annual JPL Airborne Earth Science Workshop,
I would like to thank Prof. Peter Atkinson (Univer- JPL Publication 95-1, pp. 39–40.
Crowley, J.K., Brickey, D.W., Rowan, L.C., 1989. Airborne
sity of Southampton) and an anonymous reviewer for imaging spectrometer data of the Ruby mountains, Montana:
the valuable comments and suggestions that helped in mineral discrimination using relative absorption-band-depth
improving the manuscript. images. Rem. Sens. Environ. 29, 121–134.
68 F. van der Meer / International Journal of Applied Earth Observation and Geoinformation 5 (2004) 55–68

Curran, P.J., 1989. Remote sensing of foliar chemistry. Rem. Sens. Okada, K., Iwashita, A., 1992. Hyper-multispectral image analysis
Environ. 30, 271–278. based on waveform characteristics of spectral curve. Adv. Space
Curran, P.J., Dungan, J.L., Macler, B.A., Plummer, S.E., Peterson, Res. 12, 433–442.
D.L., 1992. Reflectance spectroscopy of fresh whole leaves for Roberts, D.A., Yamaguchi, Y., Lyon, R.J.P., 1985. Calibration
the estimation of chemical concentration. Rem. Sens. Environ. of airborne imaging spectrometer data to percent reflectance
39, 153–166. using field spectral measurements. In: Proceedings of the 19th
Curran, P.J., Dungan, J.L., Peterson, D.L., 2001. Estimating the International Symposium on Remote Sensing of Environment,
foliar biochemical concentration of leaves with reflectance Environmental Research Institute of Michigan (ERIM), Ann
spectrometry, testing the Kokaly and Clark methodologies. Rem. Arbor., USA, pp. 679–688.
Sens. Environ. 76, 349–359. Shepherd, K.D., Walsh, M.G., 2002. Development of reflectance
Duke, E.F., 1994. Near-infrared spectra of muscovite, Tschermak spectral libraries for characterization of soil properties. Soil
substitution, and metamorphic reaction progress: implications Sci. Soc. Am. J. 66, 988–998.
for remote sensing. Geology 22, 621–624. Swayze, G.A., Clark, R.N., 1990. Infrared spectra and crystal
Fourty, Th., Baret, F., Jacquemoud, S., Schmuck, G., Verdebout, chemistry of scapolites: implications for martian mineralogy. J.
J., 1996. Leaf optical properties with explicit description of its Geophys. Res. 95, 14481–14495.
biochemical composition: direct and inverse problems. Rem. Thompson, A.J.B., Thompson, J.F.H. (Eds.), 1996. Atlas of
Sens. Environ. 56, 104–117. alteration: a field and petrographic guide to hydrothermal
Green, A.A., Graig, M.D., 1985. Analysis of aircraft spectrometer alteration minerals. Geological Association of Canada, Mineral
data with logarithmic residuals. In: Vane, G., Goetz, A.F.H. Deposits Division, 119 pp.
(Eds.), Proceedings of the Airborne Imaging Spectrometer Data Van der Meer, F., 1994. Sequential indicator conditional simulation
Analysis Workshop, Pasadena, USA, NASA-JPL Publication and indicator kriging applied to discrimination of dolomitization
85-41, pp. 111–119. in GER 63-channel imaging spectrometer data. Nonrenewable
Hunt, G.R., 1977. Spectral signatures of particulate minerals in Resour. 3, 146–164.
the visible and near-infrared. Geophysics 42, 501–513. Van der Meer, F., 1995. Spectral reflectance of carbonate mineral
Kariuki, P., Van der Meer, F., Verhoef, P., 2003. Cation exchange mixtures and bidirectional reflectance theory. Rem. Sens. Rev.
capacity (CEC) determination from spectroscopy. Int. J. Rem. 13, 67–94.
Sens. 24, 161–167. Van der Meer, F., De Jong, S., Bakker, W., 2001. Imaging
King, T.V.V., Ridley, W.I., 1987. Relation of the spectroscopic spectrometry: basic analytical techniques. In: van der Meer,
reflectance of olivine to mineral chemistry and some remote F., de Jong, S. (Eds.), Imaging Spectrometry: Basic Principles
sensing implications. J. Geophys. Res. 92, 11457–11469. and Prospective Applications. Kluwer Academic Publishers,
Kokaly, R.F., Clark, R.N., 1999. Spectroscopic determination Dordrecht, The Netherlands, pp. 17–61.
of leaf biochemistry using band-depth analysis of absorption Wessmann, C.A., Aber, J.D., Peterson, D.L., Melillo, J., 1988.
features and stepwise multiple linear regression. Rem. Sens. Remote sensing of canopy chemistry and nitrogen cycling in
Environ. 67, 267–287. temperate forest ecosystems. Nature 335, 154–156.
Lichtenthaler, H.K., Gitelson, A., Lang, M., 1996. Non-destructive Yoder, B.J., Pettigrew-Crosby, R.E., 1995. Predicting nitrogen and
determination of chlorophyll content of leaves of a green and chlorophyll content and concentrations from reflectance spectra
an aurea mutant of tobacco by reflectance measurements. J. (400–2500 nm) at leaf and canopy scales. Rem. Sens. Environ.
Plant Physiol. 148, 483–493. 53, 199–211.
Mustard, J.F., 1992. Chemical composition of actinolite from
reflectance spectra. Am. Mineral. 77, 345–358.

You might also like