Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

Journal of Geochemical Exploration 172 (2017) 71–88

Contents lists available at ScienceDirect

Journal of Geochemical Exploration

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

Alteration mapping on drill cores using a HySpex SWIR-320m


hyperspectral camera: Application to the exploration of an
unconformity-related uranium deposit (Saskatchewan, Canada)☆
Magali Mathieu a,c, Régis Roy a,⁎, Patrick Launeau b, Michel Cathelineau c, David Quirt a
a
AREVA Resources Canada Inc., P.O. Box 9204, 810 - 45th Street West, Saskatoon, SK S7K 3X5, Canada
b
Laboratoire de Planétologie & Géodynamique/UMR-CNRS 6112, Université de Nantes, BP 92209, 44322 Nantes Cedex 3, France
c
Université de Lorraine, CNRS, CREGU GeoRessources Laboratory, BP 70239, F-54506 Vandœuvre-lès-Nancy, France

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

Article history: In mineral exploration, the search for geophysical, geochemical, and mineralogical pathfinders is of critical im-
Received 22 February 2016 portance, and many techniques have been used to detect the presence of mineralization and related host-rock
Revised 29 August 2016 alteration. In the case of unconformity-type uranium deposits that are spatially-linked to unconformities be-
Accepted 26 September 2016
tween sedimentary basin and underlying basement rocks, hydrothermal fluid-rock interactions produced ex-
Available online 28 September 2016
tended alteration envelopes which are used to target mineralization and are therefore important guides for
Keywords:
uranium exploration.
Hyperspectral imaging The purpose of this study is to evaluate the utility of hyperspectral mapping of alteration minerals in drill core
Hyperspectral mapping samples using the HySpex SWIR-320m hyperspectral camera that covers the 1300–2500 nm (SWIR) spectral
Drill core range. A series of representative samples, including pegmatite, pelitic gneiss, and sandstone, from the area
Alteration minerals around the Cigar Lake uranium deposit (Athabasca Basin, Saskatchewan, Canada) were analyzed by SWIR reflec-
Unconformity-related uranium deposit tance spectroscopy and optical microscopy. An algorithm was developed and implemented in the IDL language to
Athabasca Basin determine both the positions and depth of diagnostic absorption bands of various minerals, and to calculate semi-
Canada
quantitative estimates of mineral content. Subsequent petrographic analysis using optical microscopy validated
the hyperspectral mapping methodology.
Selected alteration minerals present in the core samples, such as clay minerals and carbonates, were mapped. The
“pseudo-modal” mineralogical composition for each sample was determined. The mineral maps highlight the
mineralogy and the main petrographic textures such as foliation, bedding, veins, and the geometry of pervasive
alteration.
This study offers new perspectives for the use of this method in mineralogical analysis of drill core samples and
the characterization of hydrothermal alteration. The proposed mineralogical semi-quantification methodology
may be used in other geological contexts and should improve the understanding of the mineralogical alteration
distribution around hydrothermal metal deposits.
© 2016 Elsevier B.V. All rights reserved.

1. Introduction of the electromagnetic spectrum (VNIR) between 350 and 1350 nm,
whereas vibrational processes occur in the SWIR portion of the electro-
Field-based short-wavelength infrared (SWIR) reflectance spec- magnetic spectrum (1350 to 2500 nm). Although the field of view
trometry has become a conventional tool for alteration mineral map- (FOV) analyzed by a field spectrometer is about a few square centime-
ping in the mineral industry since the mid-1990s. This technique ters, the penetration depth of this technique is generally between 30
allows the acquisition of reflectance spectra in the 350–2500 nm spec- and 100 μm, depending on the radiation wavelength (Buckingham
tral range and it is used to characterize alteration minerals in samples. and Sommer, 1983).
As summarized by Clark (1999), spectral absorption bands are caused Hyperspectral imaging is a technique that combines both spectral
by different phenomena depending on the analyzed spectral range. and spatial imaging methods. It is based on the representation of objects
Electronic processes occur mainly in the Visible Near-InfraRed portion in hundreds of narrow and contiguous spectral bands, with a spectral
resolution of 10 nm or less (Goetz et al., 1985). The acquired images cor-
☆ To be submitted to the Journal of Geochemical Exploration.
respond to a “hyperspectral cube” (Vane and Goetz, 1988) endowed
⁎ Corresponding author. with two spatial dimensions (i.e. X and Y) and an electromagnetic spec-
E-mail address: regis.roy@areva.com (R. Roy). tral dimension (i.e. Z). Hyperspectral imaging systems have been used

http://dx.doi.org/10.1016/j.gexplo.2016.09.008
0375-6742/© 2016 Elsevier B.V. All rights reserved.
72 M. Mathieu et al. / Journal of Geochemical Exploration 172 (2017) 71–88

in mineral exploration to map minerals of economic interest at different Unconformity-type uranium deposits are formed at the redox inter-
scales: from space-borne (Kruse et al., 2003; Kruse et al., 2012) and air- face between oxidized uranium-bearing fluids and a reducing environ-
borne scales (Kruse et al., 2012; Roy et al., 2009) that cover large areas, ment (Hoeve and Quirt, 1987; Jefferson et al., 2007). The mineralization
to more local scales, such as outcrops on the ground (Kurz et al., 2013) is intimately associated with alteration minerals that can be used in
or laboratory-scale on collected samples (Baissa et al., 2011). The first mineral exploration as proxies for uranium ore (Hoeve and Quirt,
synthetic hyperspectral image was created in the mid-1990s (Kruse, 1984; Quirt, 2010, 2013). To characterize such mineralogical alterations
1996) by obtaining spectral readings with a PIMA II (Portable Infrared in the field, hand-held infrared spectrometers are used, with spectral
Mineral Analyzer) spectrometer every 1 cm along a drill core. With data being acquired along drill cores generally using a contact probe
the advent of laboratory hyperspectral scanning systems in the late- (i.e. analyzed surface b 2 cm2) at discrete sampling intervals (e.g. every
2000s, it became possible to scan and to analyze drill core samples at 2 to 3 m).
a much higher resolution and to detail spectral variations over the entire To date, only a few studies on core sample material from uranium
sample. For example, white micas were mapped in Chile to characterize deposits using hyperspectral imagery have been performed (Zhang
alterations in porphyry copper pebbles (Agus, 2011), carbonate et al., 2013; Sun et al., 2015), however, this type of study has not been
microfacies were determined in the Agadir Basin on Morocco from sam- conducted on unconformity-type uranium deposits. The aim of this
ple analyses (Baissa et al., 2011), and Roache et al. (2011) used this tech- study is therefore to evaluate the utility of mapping host-rock alter-
nique to define an index that acts as a proxy for composition of the ations on sampled drill core material using hyperspectral imaging.
epidote-clinozoisite solid solution in mafic rocks to target Archean This analytical tool will increase the data density, as well as data quality,
gold mineralization in Australia. Hyperspectral imagery was also used to better identify alteration minerals due to higher spatial and spectral
by Kruse et al. (2012) to map the presence of both kaolinite and jarosite resolutions. A series of samples were collected from drill holes on the
in cores to determine boundaries between reduced versus oxidized ma- Waterbury/Cigar exploration project managed by AREVA Resources
terials that correspond to elevated silver values (Kruse et al., 2012). Canada Inc. (ARC; Fig. 1) that encompasses the world-class Cigar Lake
More recently, hyperspectral imaging has been used to estimate quanti- uranium deposit. The samples were scanned (hyperspectral imaging),
ties of light REE (e.g. Nd, Sm) in drill cores from Nevada (Turner et al., analyzed, and interpreted for mineralogy, with calibration using petrog-
2014), to characterize bitumen oil sands in the Western Canadian Sed- raphy by optical microscopy. A computer algorithm was developed to
imentary Basin (Speta et al., 2013), and to map kimberlites in Canadian classify and discriminate minerals based on both the position and the
Shield crustal rocks (Tappert et al., 2015). depth of diagnostic absorption bands.

Fig. 1. Simplified geological map showing the location of the Waterbury/Cigar project property and the Cigar Lake deposit within the Athabasca Basin. Other uranium deposits are
highlighted by red dots. Projection: UTM Zone 13, datum NAD 1983.
M. Mathieu et al. / Journal of Geochemical Exploration 172 (2017) 71–88 73

2. Geological setting: Cigar Lake deposit composed mainly of Unit 1 graphite-rich pelitic gneiss. The Unit 1
gneisses are also divided in two subunits: i) a cordierite augen-gneiss,
The Athabasca sedimentary basin is located in the northern parts of rich in graphite and pyrite (in the belt center, Fig. 2) and ii) a fine-
both the provinces of Saskatchewan and Alberta, Canada (Fig. 1). The grained pelitic gneiss, with biotite and cordierite, that surrounds the
unmetamorphosed basin is filled by the Athabasca Group sediments, augen gneisses (Smellie and Karlsson, 1996, Bruneton, 1993; Fig. 2).
dominantly quartz arenite (sandstone). The sandstone is dominantly Pegmatite lenses are present and are mainly concordant with the gneiss
composed of detrital quartz, with minor heavy minerals (zircon, tour- foliation, although some are discordant.
maline, Fe-Ti oxides), and trace detrital mica. The matrix material con- The overlying micro-conglomeratic sandstones of the lower
tains kaolin and illite (Quirt, 2001). Detrital kaolinite was mostly Manitou Falls Formation fill the Athabasca Basin in the Cigar Lake area
diagenetically transformed to dickite, with later illitization affecting (Fig. 2).
both kaolin polymorphs.
The crystalline basement adjacent to, and underlying, the eastern
part of the Athabasca Basin in northern Saskatchewan consists of com- 2.2. Alteration
plexly deformed and strongly metamorphosed igneous and
supracrustal rocks of the western Canadian Shield of North America Three types of rock alteration are observed in the Cigar Lake deposit
(Jefferson et al., 2007), composed of Archean cratonic provinces, sepa- (Bruneton, 1987, 1993; Pacquet and Weber, 1993), two of which pre-
rated by Paleoproterozoic orogens and major crustal shear zones date the uranium mineralization. First, a retrograde metamorphic alter-
(Hoffman, 1988), including the Trans-Hudson Orogeny (THO). These ation has affected the basement during the late stages of the THO. It
crystalline basement rocks are unconformably overlain by the essential- consists of chloritization of cordierite, biotite, garnet, and amphibole,
ly undeformed Paleo- to Mesoproterozoic Athabasca Group that con- and sericitization of feldspar (Bruneton, 1993; Annesley et al., 2005).
sists mainly of orthoquartzitic sandstone with minor siltstone and During erosion, the basement has been altered at surface by meteoric
conglomerate. The rocks in the vicinity of the sub-Athabasca unconfor- water, which has led to the formation of a (paleo)regolith, in which ka-
mity, both sandstone and basement, host numerous high-grade olinite, illite, and hematite were the main alteration phases (Macdonald,
unconformity-type uranium deposits in locations of sandstone- 1985). Finally, a diagenetic-hydrothermal alteration related to the min-
basement fluid interaction that are typically constrained to basement eralizing event circa 1.5 Ga (Fayek et al., 2002) has been superimposed
structural corridors containing brittle reactivated faults associated on the previous alterations (Hoeve and Quirt, 1984).
with Paleoproterozoic Wollaston Group graphitic pelitic gneiss (Hoeve The hydrothermal host-rock alteration affects both the Athabasca
et al., 1980; Hoeve and Quirt, 1984; Jefferson et al., 2007; among others). sandstones and the basement rocks, in particular close to reactivated
Directly below the sub-Athabasca unconformity is a paleoregolith on fault damage zones (Fig. 2). In the sandstones, the alteration forms
the basement rocks (Macdonald, 1985), with an uppermost clay- (kao- haloes that are geometrically arranged around the orebody, extending
linite) and hematite-rich red zone. up to 300 m from the main mineralization. Six alteration zones have
The Athabasca Basin is known to be a very rich uranium region by its been described outwards from the deposit (Table 1).
abundance of large deposits (e.g. McArthur River, Cigar Lake, Key Lake, In the basement, two hydrothermal alteration zones are evident. The
Cluff Lake, Eagle Point, McClean Lake, and Rabbit Lake) and very high first zone, located near the unconformity, is completely altered to clay
uranium grades, with the McArthur River and Cigar Lake deposits con- minerals, with near-complete dissolution of original quartz. It is com-
tain ore at average grades of 14.9% and 17.8% U3O8 respectively, posed of illite and chlorite (mainly sudoite which replaces retrograde
representing 234.9 million lbs (106,549 t) of U3O8 for 592,000 t of ore chlorites), replacing the original rock-forming silicate minerals (biotite,
(Proven and Probable mineral reserves) at Cigar Lake deposit garnet, amphibole, feldspar; Halter, 1988), and local dravite (alkali-de-
(Cameco, 2014). ficient dravitic tourmaline: magnesiofoitite sub-group). This zone is
also frequently depleted in graphite and sulphides. Below this zone,
2.1. Geology the rock is clay-altered, but the original textures and quartz are pre-
served, as well as graphite and sulphides, and illite predominates over
The Cigar Lake deposit is located in the eastern part of the Athabasca sudoite.
Basin (Fig. 1), at the unconformity between the Manitou Falls Forma-
tion, a basal unit of the circa 1.7 Ga Athabasca Group (Ramaekers
et al., 2007) and the Wollaston Group, part of the Archean-Proterozoic 2.3. Mineralization
basement complex. It was discovered in 1981 by Cogema Canada Ltd.
(Bruneton, 1987, 1993). The main orebody forms a lens lying at and above the unconformity
The entire orebody forms a continuous lens approximately 2 km within the Athabasca Group sandstones; however, a minor part of the
long by 25 to 100 m width, lying at a depth of approximately 450 m uranium mineralization occurs in the sub-Athabasca basement rocks
below the surface. The high-grade core of the deposit, currently being (Fig. 2). Three stages of hydrothermal mineralization have been identi-
mined (Phase 1), is approximately 600 m in length and 100 m in fied (Bruneton, 1987): i) a main stage of uranium deposition in uranium
width. The host rocks of the deposit are strongly altered to clay min- oxide (pitchblende/uraninite: UO2) of two habits (radiating and
erals, dominantly Al-Mg di-trioctahedral chlorite (sudoite) and illite. euhedral) dated at 1400–1500 Ma (U-Pb, Fayek et al., 2002), ii) a Ni-
The mineralization is generally massive and high-grade (Bruneton, Co ± Fe arsenide/sulpharsenide stage with minor uranium remobiliza-
1993). The deposit contains Proven and Probable mineral reserves of tion, iii) late uranium remobilization and new pitchblende deposition,
592,000 t of ore at an average grade of 17.84% U3O8, containing 234.9 in particular around 450 to 200 Ma. The alteration envelopes and urani-
million lbs (106,549 t) of U3O8 (Cameco, 2014). um precipitation were generally synchronous, although cross-cutting
The basement rocks in the Cigar Lake area are dominantly relations are present (Pacquet and Weber, 1993; Alekandre et al.,
metasedimentary and belong to the Wollaston Group. They were meta- 2005). The high-grade mineralization typically contains sudoite and
morphosed and retromorphosed during the THO, with peak metamor- lesser illite, and is surrounded by a halo dominated by illite and sudoite.
phism around 1840 Ma (Annesley et al., 2005). At Cigar Lake, the Illite is predominant within the basin above the deposit within the
basement rocks are mostly comprised of two units (Bruneton, 1993; sandstones.
Fig. 2): i) graphitic biotite-cordierite pelitic gneiss (Unit 1), and ii) Some perched uranium mineralization occurs in the sandstone, up
calc-silicate gneiss, with amphibole and pyroxene (Unit 2). These to 300 m above the unconformity, in intensely fractured zones (Fig. 2,
rocks strike east-west, are steeply south-dipping (50–80°), and are Bruneton, 1993).
74 M. Mathieu et al. / Journal of Geochemical Exploration 172 (2017) 71–88

Fig. 2. Schematic cross-section showing the geology, geometry, and the alteration haloes of the Cigar Lake deposit (modified from Bruneton, 1987; Cramer, 1986; Fouques et al., 1986; and
Pacquet and Weber, 1993).

3. Sampling and methodology samples were scanned, a rock slab was cut from the core for preparation
of a thin section to be used in optical petrography for validation of the
3.1. Sampling and sample preparation spectral analyses.

The selected samples are of diamond drill core material from the 3.2. Analytical methods
Waterbury/Cigar project area. They consist of 16 barren drill core sec-
tions (10 to 30 cm long), taken from drill holes within the deposit and 3.2.1. Hyperspectral data acquisition at high spatial resolution
from 1 to 8 km away from the deposit. They represent different litholo- The linear HySpex SWIR-320m hyperspectral camera is equipped
gies and alteration types and degrees: from retrograde to hydrothermal with a mercury-cadmium-telluride (MCT or HgCdTe) detector array. It
alteration, and from weak to strong hydrothermal alteration degree. simultaneously acquires a full SWIR spectrum, with a spectral sampling
Two sets of samples were analyzed (Table 2): i) a set of basement interval of 5 nm between 1300 and 2500 nm, in 256 bands (lines: z-
rocks, composed of pegmatites and gneisses with different alteration direction of the hypercube) for 320 columns of the MCT that forms
degrees (drill holes WC449, WC426, WC422, WC372, and U382) and only one line (x-direction) of the hyperspectral image (Table 3). The
ii) a set of sandstone samples (drill holes WC458 and WC457). Each y-direction of the hyperspectral image is acquired in the laboratory
sample was sawn longitudinally to provide a half-core planar surface set-up by moving the platform holding the sample in front of the cam-
for analysis by scanning with the hyperspectral camera. Once the era at a constant speed, forming square pixels (Fig. 3a). A distance of
M. Mathieu et al. / Journal of Geochemical Exploration 172 (2017) 71–88 75

Table 1 Table 3
Alteration zones in the sandstone and the basement around Cigar Lake deposit (synthesis HySpex SWIR-320m technical specifications.
from Bruneton, 1987, 1993; Halter, 1988; partially in Pacquet and Weber, 1993; Smellie
and Karlsson, 1996). Manufacturer Norsk Elektro Optikk (NEO)

Dimensions (L × W × H; cm)a 36 × 14 × 15.2


Alteration Description
Weight (kg) 7
zone
Optical device Lens
Sandstone Bleaching Loss of diagenetic hematite (30 cm × 100 cm × 300 cm)
Grey Beginning of quartz cement dissolution; contains Field Of View (FOV; degrees) 14
alteration increased amount of clay minerals and iron sulphides Type of detector SWIR/HgCdTe
zone1 (both marcasite and pyrite) Spatial pixels (line) 320
Quartz zone Quartz precipitation; fractured, with euhedral quartz Spectral range (nm) 1300–2500
filling most of the fractures Re-sampling interval (nm) User-defined
Grey Presence of strong clay alteration (illite), quartz Integration time (ms) User-defined
alteration dissolution; strong fracturing Number of bands 239
zone2 Pixel size (mm × mm) 0.5 × 0.5
Clay Increased clay content due to increased quartz a
Camera support not taken into account.
alteration dissolution; dominant clays are illite and sudoite;
zone local dravite accumulations
Appearance of massive irregular clay layers resulting
Spectral correction was done in two steps: i) a radiometric correction
from intense quartz dissolution
Massive clay Caps the ore zone; near-total quartz dissolution of photon counts in radiances, including factory calibration between
zone composed of clay minerals (sudoite and illite); local columns of the MCT detector, and ii) a reflectance conversion which is
significant quantities of hematite and siderite given in the scan direction (column by column) by dividing each pixel
Basement Totally Reduced zone; clay minerals are illite, sudoite; local by the mean of the Spectralon® digitized spectrum present in the
argillized dravite
head of each image and multiplied by the Spectralon® factory spectra.
zone Neither graphite nor pyrite are present
Altered zone Illitization and chloritization of biotite, garnet, A second spectral correction, done column by column, eliminated the
amphibole, and feldspar; graphite and pyrite are remaining defects such as lighting inhomogeneity and damaged pixels.
present; presence of secondary alteration minerals Then, all the images were gathered in a mosaic forming a single image
(dravite, APS, apatite)
that was used in further processing (Fig. 4). Electronic noise remained
after all corrections, therefore a Minimum Noise Fraction (MNF) trans-
formation (Green et al., 1988; Boardman and Kruse, 1994) was applied
30 cm between sample and camera produces an image with a spatial to the mosaic image to remove the noise based on 36 components. The
resolution of 0.5 mm/pixel providing picture elements that are smaller Environment for Visualizing Images (ENVI 5.2) software was used for all
than many of the mineral grains and veins of interest. The high spatial hyperspectral image processing.
resolution commonly allows resolution of the spectral signature of pri-
mary minerals and alteration minerals separately, instead of having a 3.2.2. Mineralogical mapping methodology
mixture of numerous components in one pixel, as with spectral readings In infrared reflectance spectroscopy, grain sizes and surface rough-
obtained using a field spectrometer centimeter-scale probe. Moreover, ness affect the intensity of the light reflection and the mineral chemical
the small neighborhood of pixels around one analysis in an image compositions drive light absorption intensities, with diminishment of
allows the extraction of statistics about the reliability of the light reflection in various wavelength intervals. In the case of clay min-
measurements. erals, micas, and tourmaline, those absorptions are mainly due to the
For the SWIR reflectance spectroscopy, a halogen lamp giving energy OH and/or H2O bond vibrations within the crystal lattices that produce
up to 2500 nm (Fig. 3a) is used. A white standard panel made of narrow absorption bands in the 1300 nm to 2500 nm SWIR domain
Spectralon®, considered as a perfect Lambertian surface reflecting 99% (Hunt and Salisbury, 1970). However, in this domain, other minerals
of the incident light up to 2500 nm, was used as a reference material like quartz and feldspar display no absorption features and thus cannot
necessary to convert the obtained spectral data into reflectance data. It be detected by SWIR reflectance spectroscopy (Hunt and Salisbury,
was placed ahead each set of samples to be scanned together (Fig. 3b). 1970). They only play the role of a diffusive medium in or on which

Table 2
Description of lithologies and alterations of samples. Grain size: F = fine, M = medium, C = coarse; Rock type: P = present, Crd = cordierite; Intensity of alteration: W = weak, S =
strong, X = fresh rock, Und. = Undefined.

Sample no. Drill hole Depth (m) Rock type Alteration

Graphitic metapelitic gneiss Pegmatite Sandstone Intensity Type

Crd augen gneiss Grain size

1 WC449 402.95–403.10 P S Yellow-pale green


2 434.50–434.60 P Und. Yellow-green
3 443.60–443.70 P X ±retrograde chlorite alteration
4 453.45–453.70 F X
5 469.50–469.80 M W
6 485.00–485.30 F Und. Green-yellowish
7 488.50–488.80 F X
8 512.00–512.10 P S Green
9 556.00–556.30 P Und. Yellow-green (feldspars)
10 WC426 437.60–437.90 M X
11 WC422 469.90–470.10 P C W
12 WC422 472.80–472.90 P C W
13 WC372 445.60–445.90 P Und. Light green
14 U382 78.50–78.90 P Und. Yellowish-green pale (cordieritic gneiss)
21 WC458 306.50–306.80 Dickite
20 WC457 323.00–323.30 Illite
76 M. Mathieu et al. / Journal of Geochemical Exploration 172 (2017) 71–88

Fig. 3. a) HySpex SWIR-320m camera: the device is mounted on the scanning system while samples are placed on a motorized platform (the red arrow shows the translational movement
of the platform during data acquisition). b) Core sample set on the motorized platform.

coating and or inclusion of other minerals can be detected. The aim of sandstone clay mineralogy using SWIR spectral data and to give a
this study is to work with this effect to detect in gneiss, pegmatite, semi-quantitative estimation of their proportions. Six minerals, includ-
and sandstone, the presence and abundance of key minerals for the ing illite, several kaolin polymorphs, chlorite, and tourmaline, can be
study of uranium ore deposits, particularly clay minerals and identified and semi-quantified from each reflectance spectrum acquired
tourmaline. with a field portable spectrometer (e.g. PIMA II, TerraSpec), using the
Once the hyperspectral images have been calibrated into reflectance, MINSPEC software developed by Earle. The MINSPEC algorithm is
minerals can be identified in each pixel of the entire hyperspectral based on the detection of characteristic absorption peaks (i.e. reflec-
image due to a spectral resolution high enough to resolve the relatively tance minima), as well as the shapes, shoulders, and slopes/inflections
narrow OH absorption features. A simple false color composite was pro- of the peaks and intervening spectrum.
duced by setting red, green, and blue channels of a display based on the Similarly, the algorithm developed in this study aims to classify a
wavelength of the absorption features, their depths, and their shoulder series of minerals by searching the position (in wavelength) of the
styles. Uniform material shows grey variations, whereas patches of ma- maximum and minimum absorption features. Short spectral ranges
terial containing narrow absorption features display various colors de- containing the potential shift of the absorption bands due to cation
pending on the choice of wavelength and the absorption depth. substitutions are defined for each mineral (or solid solution of
Preparation of the color composite images is a prospecting stage in the endmembers). They were practically chosen after comparison between
mineralogical analysis by facilitating the selection of wavelength char- spectral signatures present in the samples and reference spectral librar-
acteristics of a mineral. The final mineral identification is provided by ies (e.g. United States Geological Survey, Clark et al., 1993, 2007).
the determination of a relatively unique sequence of absorption fea- Clay minerals and micas are identified by the absorption band relat-
tures, depths, and shoulder styles that unambiguously characterizes its ed to the Al-OH bond, centered at 2200 nm (e.g. Post and Noble, 1993)
presence. and the absorption feature at 1900 nm linked to the H2O molecule. Sim-
It is also possible to identify minerals by using image classification. ilarly, illite spectra are characterized by a deep absorption band at
There are two types of methods: i) unsupervised, which are automatic, 1910 nm and an absorption band, due to the Al-OH bond, located be-
based on iterative processes with no operator intervention and ii) tween 2197 and 2207 nm. As seen in Fig. 5a, because of its quartz-
supervised, which are manual, with operator intervention. Because dominant (sandstone) background, the illite in the sandstone (spec-
most unsupervised methods do not demonstrate proper mathematical trum 1) displays stronger reflectance values than the illite found in feld-
sophistication, supervised methods have generally been used on spar of pegmatites (spectra 2a and 2b) or gneisses (spectrum 2c). The
hyperspectral data, such as the Spectral Angle Mapper method (SAM, Al-OH absorption band position also varies between sample spectra,
Kruse et al., 1993) and the Mixture-Tuned Matched Filtering method shifting from 2197 nm (spectra 1, 2a) to 2202 (spectrum 2b) to
(MTMF, Williams and Hunt, 2002). Recently, Sun et al. (2015) used 2207 nm (spectrum 2c). In Fig. 5b, the muscovite in the gneiss (spec-
the MTMF method to map alteration minerals (illite, chlorite, and cal- trum 1) and in the pegmatite (spectrum 2) displays shallower
cite) on one core sample from a uranium deposit in China. However, 1910 nm water and 2207 nm Al-OH absorption bands than the trace
these methods use the entire spectrum and do not focus on the narrow amounts of phengitic mica (spectrum 3) identified in pegmatite, also
spectral feature(s) of the key minerals being searched for. So, the results characterized by an Al-OH absorption band at 2217 nm.
from one sample cannot be applied to another. Thus, small variations Fig. 5c shows the typical chlorite spectral absorption bands related to
along the full spectrum may be overall greater than one absorption fea- Fe-OH and Mg-OH bonds. Mg-chlorite absorption bands are at 2251 nm
ture that is consequently missed. and 2333 nm, whereas Fe-chlorite absorption bands are shifted to
The accurate and precise measurement of OH and CO absorption higher wavelengths at 2256 and 2343 nm. Sudoite (Al-Mg di-
band depth positions within the wavelength spectrum is therefore rec- trioctahedral chlorite) which is known to be present in the Cigar Lake al-
ommended for mineral identification. However, they are sensitive to teration haloes could not be differentiated here because of its spectral
cation substitutions and may need to be adjusted for each sample. similarity to white micas/clays and other chlorites. In Fig. 5d, a kaolin
This adjustment was carried out through use of a programming method, polymorph is identified by two pairs of narrow absorption bands near
based on work by Baissa et al. (2011) and Agus (2011) that was adapted 1400 nm (OH) and 2200 nm (Al-OH). This mineral is present in both
for this study of Cigar Lake deposit alteration minerals. Earlier studies of the sandstone samples (no. 20 and 21). The OH (1384 and
were conducted by Earle (1996, 1997) to determine Athabasca 1413 nm) and Al-OH (2178 and 2207 nm) absorption band positions
M. Mathieu et al. / Journal of Geochemical Exploration 172 (2017) 71–88 77

Fig. 4. Mosaic gathering all examined samples. Pegmatite: samples no. 1, 2, 3, 8, 9, 13; gneisses: no. 4, 5, 6, 7, 10, 11, 12, 14; sandstones: no. 20 and 21. Color composite: 1983 nm (red),
1681 nm (green), 1394 nm (blue).

suggest the presence of dickite. Kaolinite, with different band positions, Based on all of these absorption band positions, an algorithm was
was not identified. developed to classify minerals in two steps (see flowchart in Fig. 6).
Carbonate spectra are characterized by an absorption band between The first step (Fig. 6, box 1) looks for the maximum and minimum ab-
2300 nm and 2360 nm due to the CO3 group. Some basement samples sorption feature position in a selection of characteristic ranges deter-
show a dolomite signature, with a characteristic CaMg(CO3)2 absorption mined for each mineral type listed in Table 4. The second step (Fig. 6,
feature at 2324 nm (Fig. 5e) resulting from the combination of CaCO3 box 2) uses the combination of minimum and maximum found in step
(calcite) and MgCO3 (magnesite) absorption bands. Water absorption one to classify each pixel in a class number. Some of the minerals (e.g.
bands at 1400 and 1934 nm are due to residual water in carbonate min- carbonates, tourmaline, and kaolins) have sufficiently contrasting signa-
eral microporosity trapped during crystal growth (Baissa et al., 2011). tures to be classified using only with the position of their maximum ab-
In Fig. 5f, the dravite spectrum is characterized by absorption bands sorption features (Fig. 6, box 2a). Other minerals having common
at 1408 and 1442 nm for OH, at 2207 nm for Al-OH, at 2241 nm for Fe- absorption bands (e.g. micas and kaolins at 2207 nm) require additional
OH, and at 2368 nm for Mg-OH. This signature is observed in one peg- slope or index calculations to differentiate them (Table 5 and Fig. 6, box
matite sample (no. 13). 2b).
78 M. Mathieu et al. / Journal of Geochemical Exploration 172 (2017) 71–88

Fig. 5. HySpex SWIR-320m spectra of minerals characterizing Waterbury/Cigar Lake project samples: a) illite, b) muscovite and phengite, c) chlorite, d) dickite, e) dolomite, and f) dravite.
Continuum removal has been applied to enhance the differences in shape between spectra. The short vertical lines indicate wavelength location of important absorption features that are
discussed in the text.

In summary, class membership needs three conditions: i) absorption linked to the water (around 1900 nm, Eq. (1)), ii) mica index value de-
feature positions have to be defined in a nanometer wide range, ii) a fined by Eq. (2), and iii) spectral slope characteristics of some mineral
maximum is necessary to identify the spectral shape, and iii) for some mixtures.
minerals (e.g. micas, clays and chlorites) additional constraints are
necessary. ref 1808 þ ref 2080
DepthH2 O ¼ 1000  ð1Þ
Since illite and muscovite are similar in chemistry, their differentia- 2  ref min
tion requires comparisons in multiple ranges of the spectrum. This is
done using other parameters, such as: i) depth of the absorption feature with: ref1808 the reflectance value at 1808 nm
M. Mathieu et al. / Journal of Geochemical Exploration 172 (2017) 71–88 79

Fig. 6. Processing flowchart of the algorithm developed and implemented in IDL to classify minerals from a hyperspectral image.

Table 4
Spectral ranges (in nm) defined to select maximum and minimum absorption feature positions to classify minerals.

Class Carbonates Kaolins Tourmaline Micas Chlorites


(dravite)

Reflectance minimum (research interval 1) 2106–2402 2153–2192 2324–2397 2139–2294 2187–2275


Reflectance maximum (research interval 1) 2331–2417 2178–2207 – 2207–2338 2251–2343
Reflectance minimum (research interval 2) – 2192–2241 – 2294–2377 2275–2387
Reflectance maximum (research interval 2) – 2207–2260 – – –

ref2080 the reflectance value at 2080 nm position in wavelength of the first minimum of reflectance (refmin, in
refmin the minimum reflectance value in the range 1808–2080 nm the range 2187–2275 nm). If a positive slope is found in this range,
  the spectrum is considered as a chlorite and illite/white mica mixture,
2  ref max with dominant chlorite.
Indexmica ¼ 1000  −1 ð2Þ
ref min þ ref 2 min The “chlorite index” (same expression as the “mica index” Eq. (2),
with different spectral ranges) is used to classify the various chlorites,
with: refmin the first minimum reflectance value in the range 2139– without species discrimination.
2294 nm The “pseudo-modal”1 composition is the spectral proportion of iden-
refmax the maximum reflectance value in the range 2207–2338 nm tifiable minerals present in each sample and is defined by Eq. (3).
ref2min the second minimum reflectance value in the range 2294–
X
2377 nm Pixelmineral
An example is presented in Fig. 7 to illustrate the parameters defined Proportionmineral ð%Þ ¼  100 ð3Þ
Np
by the equations above. Mixtures of spectrally-distinct white mica (e.g.
muscovite, phengite, paragonite), or illite, with chlorites can be ob- with: Proportionmineral the proportion of a mineral (class) in the cho-
served in the spectra of most of the samples. They are characterized sen sample, in %
by additional absorption features due to the relative weights of illite/ ∑Pixelmineral the sum of the pixels belonging to the mineral (class)
white micas 2200 nm Al-OH absorption band and the chlorite chosen in the sample selected
2250 nm Fe-OH absorption band. When the reflectance value is lower Np the total number of pixels in the selected sample
for the Al-OH feature (i.e. white mica/illite characteristic absorption
band) than for the Fe-OH feature (i.e. chlorite characteristic absorption 3.2.3. Petrographic analysis
band), the slope of the reflectance in the 2190–2300 nm range (the po- Validation of the mineral identification results for alteration min-
sition in wavelength of refmin and refmax, defined in the Eq. (2)) is calcu- erals obtained using hyperspectral imagery was carried out through op-
lated. If a negative slope is found in this range, the spectrum is tical microscopic petrographic analysis on thin sections. For this
considered as an illite/white mica and chlorite mixture, with dominant purpose, thin sections were made from the drill core samples and
illite/white mica. When the reflectance value is lower for the Fe-OH fea-
ture (i.e. chlorite characteristic absorption band) than for the Al-OH fea- 1
"Pseudo-modal" composition because it takes into account only the minerals having a
ture (i.e. white mica/illite characteristic absorption band), the slope of spectral signature in the SWIR range (i.e. excludes mainly quartz, unaltered feldspars, and
the reflectance is calculated in the range defined by 2187 nm and the sulfides).
80 M. Mathieu et al. / Journal of Geochemical Exploration 172 (2017) 71–88

Table 5
Spectral ranges (in nm) defined to classify minerals based on additional spectral features.

Mineral Membership conditions for the mineral class

Minimum reflectance position in Maximum reflectance in Other necessary conditions


the the
range (nm) range (nm)

Magnesite 2300–2310 2331–2402 –


Dolomite 2310–2330 2331–2402 –
Calcite 2330–2335 –
Rhodocrosite 2335–2336 2331–2402 –
Kaolinite 2160–2170 2207–2260 –
Dickite 2170–2180 –
Mixed Illite-Dickite – 2207–2338 positive slope between 1379 and 1398 nm during mica research
Paragonite 2190–2195 –
Illite (basement) 2192–2210 DepthH2O (1900 nm) N 1100
Illite (sandstone) 2192–2210 IndexMica N 750
Muscovite 2206–2210 DepthH2O (1900 nm) b 1200 IndexMica b 400
Phengite 2217–2222 –
Micas + Chlorites – negative slope between wavelength refmin and wavelength refmax during mica
research
Chlorites b2260 IndexChlorite N 16
Chlorite + Micas or – negative slope between 2187 nm and wavelength refmin during chlorite research
Sudoite
Tourmaline (Dravite) 2367–2369 – –

were examined with an Olympus BX 51 optical microscope for selected to emphasize the spectral variations of phyllosilicates. These
performing classical optical identification of the mineral phases in variations are revealed with a red, green, and blue color composite de-
each sample. The microscope was equipped with an AxioCam ICc 1 cam- fined using bands at 2348 nm, 2270 nm and 2197 nm, respectively. Be-
era to take photographs of the thin sections. cause illite and muscovite spectra are similar, pixels with a white mica/
There were two objectives of this petrographic analysis: i) the iden- phyllosilicate composition appear in green. For gneiss samples, these
tification of the dominant mineral phases in typical sectors of the minerals delineate the foliation (mica and clay-altered feldspar: sam-
hyperspectral maps and cross-checking of the mineral identifications ples no. 4, 5, 6, and 7), whereas in the pegmatites they are found in un-
obtained by the two methods, and ii) mineral mapping for each thin deformed feldspar crystals (samples no. 3 and 9). Dolomite appears in
section. magenta. However, dickite (only identified in sandstone samples (i.e.
The optical microscopic mineral mapping was carried out by visual samples no. 20 and 21) is also indicated by green on Fig. 8b (normalized
interpretation of images digitized with a VHX 2000 numerical micro- reflectance), as is illite. This situation shows that minerals presenting a
scope (Keyence) and processed using the Keyence image processing similar spectral behavior can display the same colors in one 3-band
software. The classical mineral identification was done using cross- color composite image, so it is then necessary to use a series of color
polarized transmitted light. These interpretative mineral maps were composites to differentiate all of these minerals. However, this is not
superimposed on the hyperspectral maps. Mineral proportions (relative the scope of these images which are only quick illustrations of the
%) of each mineral present in the thin sections were calculated using the data explorations that are used in preparing for the determination of
image processing software. the mineral classification flow chart used for the hyperspectral mineral
mapping.
4. Results
4.2. Mineral mapping
4.1. Color composite images
The mineral maps, or spectral-class images, resulting from the flow
Color composite images were prepared in order to quickly highlight chart classification described in Table 5 are shown in Fig. 9. These
the minerals that characterize the samples without significant amounts maps allow fast identification of the main spectrally-active mineral spe-
of data processing. Fig. 8 shows an example of a color composite image cies, especially those resulting from alteration events. The mineral

Fig. 7. Spectral ranges and bands used to calculate the water absorption depth near 1900 nm and the micas index. a) Spectral ranges used for the parameter research. 1: minimum of
reflectance research window for the H2O feature; 2: research window for the first minimum of reflectance for Al-OH mica feature (refmin); 3: research window for the maximum of
reflectance for the Al-OH mica feature (refmax); 4: research window for the second minimum of reflectance for the Al-OH mica feature (ref2min). b) representation of the refmin, refmax,
and ref2min parameters to differentiate illite from muscovite.
M. Mathieu et al. / Journal of Geochemical Exploration 172 (2017) 71–88 81

Fig. 8. HySpex SWIR-320m color composites defined with bands sensitive to the spectral properties of illite (red indicates 2348 nm; green indicates 2270 nm; blue indicates 2197 nm):
a) composite image in reflectance, b) composite image with normalized reflectance (Roy et al., 2009).

species that can be mapped with the described algorithm are: related to the presence of feldspars, and a ‘chlorite-mica mixture or
i) carbonates: magnesite, dolomite, calcite, rhodocrosite, ii) the kaolin sudoite’ (14%), with unclassified pixels representing 10% of the sample.
polymorphs: kaolinite, dickite, dickite-illite mixture, iii) the Sample no. 2 is chlorite dominant (50%) and the “chlorite-mica mixture
dioctahedral micas: paragonite, illite, muscovite, phengite, iv) chlorite or sudoite” class represents 17%, while illite is found in low amount
and/or mica-chlorite mixtures, and v) dravite. The discrimination be- (6%). Samples no. 3, 8, and 9 have signatures similar to sample no. 1,
tween the inherited retrograde metamorphic minerals (muscovite and but with double to quadruple the number of unclassified pixels. Sam-
tri-octahedral chlorite) and the hydrothermal minerals is often rather ples no. 8 and 9 are the only samples in which the minor presence of
difficult. However, these minerals only have a small abundance in al- carbonates (1 to 11%), as veins or impregnations, was detected. Sample
tered rocks around the deposits relative to the hydrothermal minerals. no. 13 is the only sample in which dravite was identified (8%), accompa-
The “unclassified” class represents pixels with no particular signature nied by illite (19%), however, the majority of the sample is mainly com-
in the SWIR domain (e.g. quartz, unaltered feldspar, or sulphides) or posed of “unclassified” pixels (63%) that correspond to unaltered
that exhibit too noisy a response to be reliably identified. Some of the (fresh) crystals of feldspar.
minerals listed above are not encountered in the Cigar Lake area (e.g. The gneiss samples present various degrees of alteration and the
magnesite, rhodocrosite) or only are present in trace amounts (e.g. spectral classification mapping highlights well their metamorphic tex-
paragonite, phengite). These minerals are included in the classification tures, in particular foliation. Chloritized biotite and elongated quartz
algorithm to provide a quality control on the mapping. Subpixel-size and feldspar delineate the foliation of samples no. 4 to 7. In detail, this
mineral mixtures are classified in separate classes to differentiate foliation is characterized by chlorite (sample no. 4), illite (sample no.
them from pure minerals. Note that sudoite currently cannot be reliably 6), or a mixture of these two minerals (samples no. 5 and 7), present
classified because the sudoite spectrum is very similar to a combination as an alteration product of feldspars. The proportion of “unclassified”
of a mica and a chlorite signature in the SWIR range. Pure “illite from pixels varies from 25 to 93% resulting from the presence of spectrally-
sandstone” and “illite from gneiss”, mixed with spectrally undetectable translucent quartz embedded in altered feldspar and unaltered feld-
feldspar, classes have distinguishable mica index values (Table 5) due to spars. Dominant minerals are the “chlorite-mica mixture or sudoite”
the contrasting translucent properties of the quartz and quartz + feld- for samples no. 4, 5, and 7. The mica-chlorite mixture (32%) and illite
spar backgrounds. Calculated mineralogical proportions are presented from feldspars (24%) are prominent in sample no. 6. In sample no. 10,
in Table 6 for each sample. composed of two lithologies (gneiss and pegmatite), “unclassified”
Altered pegmatite samples are dominated by illite, present as the pixels, again mainly quartz and unaltered feldspars, dominate (84%), es-
major alteration product of feldspars, and chlorite. Sample no. 1 is main- pecially in the gneissic part of the sample. Alteration minerals in sample
ly composed of a mica-chlorite mixture (37%) and illite (37%) that is 10, mostly present in the pegmatitic zone, are the “chlorite-mica
82 M. Mathieu et al. / Journal of Geochemical Exploration 172 (2017) 71–88

Fig. 9. HySpex SWIR-320m spectral-class image of the Waterbury-Cigar Lake project samples. Pegmatites: samples no. 1, 2, 3, 8, 9 and 13; gneisses: samples no. 4, 5, 6, 7, 10, 11, 12 and 14;
sandstones: samples no. 20 and 21 (see Table 2). See color key for image color-class assignment. Note that unclassified (black) pixels correspond to quartz, feldspar, and sulphide minerals,
and also to small phyllosilicate grains that display a noisy spectral signature.

mixture or sudoite” and illite. Samples no. 11, 12, and 14 are similar to featureless quartz (i.e. “unclassified” pixels) in both samples is low
the gneissic part of the sample no. 10 in texture, structure, and compo- (17–18%) relative to the other.
sition, with foliation being visible in the three samples and the mapped
mineralogy being dominated by chloritized biotite in an illitic matrix 4.3. Hyperspectral validation through petrographic study
that is classified as “chlorite-mica mixture or sudoite” (~20%).
The mineral map of sample no. 20, a sandstone with micro- Most samples were also petrographically-examined, and four thin
conglomerate bed, highlights the sedimentary bedding emphasized by sections were studied quantitatively in great detail: three pegmatite
alternating beds of contrasting grain sizes that also show contrasting samples (samples no. 1 and 8 from drill hole WC449, and sample no.
diagenetic matrix mineralogy. The fine-grained sandstone beds contain 13 from drill hole WC372) and one gneiss sample (sample no. 6, drill
dominant illite and coarse-grained micro-conglomerate beds contain hole WC449). Fig. 10 presents the comparison between the mineral
dickite. The narrow transition between these beds is characterized by maps obtained from petrographic identification and hyperspectral
a dickite-illite mixture. In sample no. 21, dickite is the spectrally- image processing (columns 1 and 2, with the hyperspectral images (col-
dominant matrix clay mineral (83%). The proportion of spectrally- umn 3)).
M. Mathieu et al. / Journal of Geochemical Exploration 172 (2017) 71–88 83

Table 6
Proportion of minerals obtained from the SWIR spectral classification. Pegmatites: samples no. 1, 2, 3, 8, 9, and 14; gneisses: samples no. 4, 5, 6, 7, 10, 11, 12, and 14; sandstones: samples
no. 20 and 21.

Drill hole WC 449 WC 426 WC 422 WC 372 U 382 WC 457 WC 458

Sample number 1 2 3 4 5 6 7 8 9 10 11 12 13 14 20 21
Clay proportion (%) 89 82 66 78 50 66 7 54 57 16 41 40 29 29 82 83
Unclassified 11 17 34 22 48 33 93 44 29 84 59 60 63 70 18 17
Magnesite 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0
Dolomite 0 0 0 0 0 0 0 1 11 0 0 0 0 0 0 0
Calcite 0 1 0 0 0 0 0 0 3 0 0 0 0 0 0 0
Rhodochrosite 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0
Kaolinite 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0
Dickite 0 0 0 0 0 0 0 0 0 0 0 0 0 0 7 83
Dickite + Illite 0 0 0 0 0 0 0 0 0 0 0 0 0 0 18 0
Paragonite 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0
Illite (basement) 37 7 34 7 3 24 1 34 34 4 2 6 19 4 16 0
Illite (sandstone) 0 0 2 0 0 0 0 0 0 0 0 0 0 0 40 0
Muscovite 0 0 0 0 0 1 0 3 2 0 0 0 0 0 0 0
Phengite 0 2 0 0 0 0 0 1 0 0 0 0 0 0 0 0
Micas + Chlorite 37 6 21 28 9 32 1 10 14 4 3 5 4 4 2 0
Chlorite + Mica or Sudoite 14 17 6 27 25 8 4 4 4 6 23 18 5 20 0 0
Dravite 0 0 0 0 1 0 0 0 0 0 0 0 8 0 0 0
Chlorites 2 50 2 17 13 1 1 1 2 2 13 9 0 0 0 0

Thin sections of the sandstone samples showed that discrete clay cross-cuts the thin section in its lower part (Fig. 10, D3 and D4). The
mineral phases are very difficult to distinguish using optical petrogra- spectral-class map highlights these different zones when they are
phy. Coarse quartz grains are generally coated by, and the intergranular thick enough, however, the chloritized biotite zone, illite and biotite
matrix contains, flaky clay minerals that are too small to be identified at chloritized zone and carbonate zone are not hyperspectrally-mapped
the spatial resolution of the optical microscope and of the hyperspectral because they are too thin (Fig. 10, D3 and D4). A group of large
images. chloritized biotites (greater than the pixel size, i.e. 500 μm) situated
near the lower border of the thin section is well mapped.
4.3.1. Mineral identification
Quantitative petrographic mineral mapping (proportions in Table 7) 4.3.2. Mineral proportions
of sample no. 1 yielded an estimated 36% quartz and 45% phyllosilicate. The relative proportions of minerals observed in the thin sections
The phyllosilicates comprise approximately 8% of chloritized biotites were calculated using the image processing software included with
(tri-octahedral chlorite) with intercalated muscovite (Fig. 10, A1 and the Keyence digital microscope. To provide a level comparison,
A2), and chloritized (27%) and illitized (10%) feldspar. Some chloritized spectrally-inactive minerals (quartz, unaltered feldspars, sulphides,
biotites are associated with rutile. Feldspars are strongly altered and graphite), and areas of “unclassified” pixels were not taken into account
transformed into sudoite and hydrothermal illite (10% illite, but this is in the comparison between the two techniques. For the hyperspectral
likely an underestimate because of the dissemination of this mineral). imaging, the sudoite proportion was defined as the sum of two classes
Muscovite and relic phengites (i.e. high temperature micas) are visible. “mica-chlorite mixture or sudoite” and “mica-chlorite mixture”, which
Pixels spectrally classified as chlorite correspond to the position of the leads to an overestimation of this mineral. Illite and muscovite were
chloritized biotites observed in the thin section (Fig. 10, A3 and A4). gathered in a single class because they are most often spectrally-
Feldspars transformed into sudoite are classified as “chlorite-mica mix- inseparable. The proportion of the illite-dickite class was divided in
ture or sudoite”, or as a mica-chlorite mixture. two and distributed to both the illite class and the dickite class, solely
Sample no. 8 is composed of almost 50% unaltered to slightly-altered based on the hypothesis that the mixture is composed of equal propor-
feldspars. From petrography, the main alteration is of feldspar with clay tions of dickite and illite.
mineral phases, comprising 44%, with only 9% quartz present. The clay The bivariate comparison diagrams of Fig. 11 show linear best-fit
minerals are illite (28%) and sudoite (16%). A carbonate vein cross- lines and correlations obtained between hyperspectral imaging mineral
cuts the thin section laterally (Fig. 10, B1 and B2). It is well-mapped proportions and comparable data from the image processing of petro-
with the hyperspectral imaging, although its width is overestimated graphic thin sections. The mineral proportions are similar for pegmatite
(Fig. 10, B3 and B4). Illite is also well-mapped. samples (Fig. 11a, samples no. 1, 8 and 9b), with slope for the best-fit
From petrography, sample no. 13 is mainly composed of unaltered line being close to 1. However, for the gneiss sample, the mineral pro-
(spectrally-“unclassified”) feldspars (64%). A vein of dravite cross-cuts portions are not correlated (Fig. 11a, sample no. 6b) because the mineral
the thin section, representing 10% of the sample. This sample also con- grains are much smaller than in the pegmatite samples, resulting in a
tains minor illite/muscovite (Fig. 10, C1 and C2). The spectral-class less precise spectral classification using hyperspectral imagery data.
map also highlights the dravite veining (Fig. 10, C3 and C4). Some mus- Concerning mineral proportions, including all minerals, illite and
covite (Fig. 10, C2 and C3) is observed petrographically, but illite is only muscovite show the best correlation between the two techniques
classified from hyperspectral data and is not observed in thin section. (Fig. 11b). The chlorite proportions are poorly correlated, with the
This may be due to illitized feldspars being present in the sample at a proportion being overestimated or underestimated by hyperspectral
depth greater than the thin section thickness (see Discussion and imaging depending on the sample. Sudoite proportion is well corre-
conclusions). lated between the two techniques in pegmatite samples, but not for
Because of the small mineral grain size in gneiss sample no. 6 (Fig. the gneiss sample, probably due to fine-scale intermixing of the
10, D1), the petrographic interpretation was done by classifying each sudoite and illite.
part of the thin section by the dominant mineralogy (Fig. 10, D2). The hyperspectral carbonate proportion is two times more than
There are two quartz-rich zones, with an illitic matrix, situated at oppo- obtained from the microscopic data. Thus, the carbonate veins are
site corners of the thin section. Between these zones, there are layers of approximately twice as wide on the hyperspectral-class mineral
chloritized biotites and featureless illitized feldspar. A carbonate vein maps than on the optical petrographic mineral maps. This feature is
84 M. Mathieu et al. / Journal of Geochemical Exploration 172 (2017) 71–88

Fig. 10. Comparison between thin section images and spectral-class maps for the pegmatite samples no. 1 (A), no. 8 (B), and no. 13 (C), and for a gneiss sample (no. 6, D). 1) Thin section
images are in transmitted, polarized, and analyzed light; 2) mineral maps from thin section interpretation; 3) spectral-class maps, and 4) layering of the thin section mineral maps over the
spectral-class maps. Ilt: illite; Sud: sudoite; Chl: chloritized biotite; Ms.: muscovite; Cb: carbonate; Drv: dravite; Qz: quartz.

Table 7 likely due to the vein obliquity, which produces a broader signature
Proportion of minerals obtained from optical petrography. of the mineral in the vein wider than the thin section plan view (sur-
Drill hole WC 449 WC 372 face) indicates because the surrounding minerals are essentially
spectrally transparent (quartz, unaltered feldspars; see Discussion
Sample number 1 6b 8 13
Clay proportion (%) 45 88 44 25
and conclusions).
Rock type Pegmatite Gneiss Pegmatite Pegmatite
Quartz 36 11 9 9 5. Discussion and conclusions
Unaltered feldspar 11 0 45 64
Biotite + Graphite 8 1 0 0
The application of hyperspectral imagery to the mineralogical map-
Illite/Muscovite 10 47 28 8
Sudoite 27 0 16 0 ping of host-rock alteration halos has been developed successfully using
Dravite 0 0 0 10 the example of the samples from the area around Cigar Lake uranium
Tri-octahedral chlorite (biotite 8 41 0 9 deposit. The approach allowed the identification and differentiation of
replacement) a wide variety of minerals belonging to most alteration phases already
Carbonate 0 0 2 0
known for this area.
M. Mathieu et al. / Journal of Geochemical Exploration 172 (2017) 71–88 85

Fig. 11. Comparison of mineralogical proportions obtained by thin section analysis (X-axis) and by hyperspectral image classification (Y-axis) showing correlations between the methods:
a) Samples: no. 1 (pegmatite), 6b (gneiss), 8 (pegmatite), and 9b (pegmatite); b) Minerals: chlorite, sudoite, illite + muscovite, and carbonates.

5.1. Mineralogical interpretation illite), as well the different types of chlorites (tri-octahedral chlorite,
di-trioctahedral sudoite), and ii) the significant uncertainty on abso-
Illitization and/or chloritization of feldspars which represent the lute mineral abundances due to the impossibility of quantifying the
two main alteration phases in basement lithologies around uncon- quartz and fresh feldspar contents. However, the relative propor-
formity deposits was identified and mineral proportions were tions of chlorite and K-micas are easily obtained, as are their spatial
quantitavely estimated in hyperspectral mineral maps. Also, local oc- distributions at all scales from core pieces (decimeter scale) to thin
currences of hydrothermal minerals such as carbonate or dravite sections (sub-millimeter-scale).
veins were successfully identified and mapped. The main limitations Also because of the issue with spectrally-inactive minerals like
of the hyperspectral methods are: i) the difficulty in spectrally dis- quartz and unaltered feldspar, the alteration mineral proportions are
criminating the different di-octahedral K-phyllosilicates (muscovite, commonly overestimated, relative to thin section mineral mapping.
86 M. Mathieu et al. / Journal of Geochemical Exploration 172 (2017) 71–88

Fig. 12. Schematic sketch illustrating the explanation for local overestimation of clay mineral proportions in spectral-class maps. The A-B section is represented in depth by the scheme at
the bottom. The points 1 and 2 localized on quartz minerals have spectral signatures of clays (1: chlorite-mica mixture or sudoite; 2: illite) due to the investigation depth of the spectral
technique and the spectrally-transparent characteristic of quartz. Sud: sudoite; Ilt: illite; Chl: chloritized biotite; Ms: muscovite; Qz: quartz.

Two reasons can explain this phenomenon: i) the investigation depth of matrix will be a mixture between the two minerals. Therefore, it was
this technique is deeper (i.e. 30 to 100 μm) than the thin section thick- necessary to define mixture classes (e.g. mica + chlorite, dickite + illite)
ness (30 μm); ii) the transparency of the silicates in the SWIR range re- in the hyperspectral mineral classification algorithm.
veals alteration products that are present behind (below) the Gneiss samples showed lower reflectance values due to their dark-
spectrally-inactive minerals and thus are not visible optically. These lat- ness (mean reflectance of 0.2) relative to pegmatite (mean reflectance
ter alteration minerals can be derived from feldspar or from other min- of 0.4) and sandstone (mean reflectance of 0.6) samples. The gneiss
erals to form a coating on quartz and feldspar. spectra are consequently flat and noisy, and thus classification of alter-
ation minerals in these samples is more difficult. This phenomenon
5.2. Technical considerations could be minimized by additionally using a Spectralon® panel with a
lower reflectance (e.g. 20% or 50% reflectance), but minerals with a
To illustrate the issue of clay mineral proportion overestimation, an higher reflectance could then be oversaturated.
example is taken using sample no. 1. The signatures of a “chlorite-
mica mixture or sudoite” (spectrum 1 from point 1, Fig. 12) or the signa-
ture of illite (spectrum 2 from point 2, Fig. 12) can be induced by the in- 5.3. Conclusions
teraction between the incident light and alteration minerals present 50
to 100 μm beneath the quartz or feldspar crystals. The same phenome- This project on mineral alteration mapping on drill cores using a
non is responsible for dickite and illite mineral proportion overestima- HySpex SWIR-320m hyperspectral camera has highlighted the applica-
tion in sandstone samples (no. 20 and 21). tion of hyperspectral imaging on core samples coming from an
Mineral particle size also has an impact on the spectral classification. unconformity-type uranium deposit environment in Canada. Because
Some alteration minerals are intimately associated, because they occur of a fine spatial resolution (0.5 mm/pixel) and a fine spectral resolution
as pseudomorphs of the parent mineral (e.g. illite and chlorite replacing (5 nm), it has been possible to acquire a spectral signature at the scale of
feldspars), with the resulting matrix being composed of mixed mineral the mineral network (rock texture), thus minimizing the effects of mix-
grains of much b50 μm in size. However, even though the spatial reso- tures typical when using field spectrometer analyses. The computing
lution obtained from a hyperspectral camera (500 μm) is better than method developed to mineralogically classify the hyperspectral data,
that obtainable from field spectrometers, the spectral signature of this based primarily on the position of minima and maxima of reflectance,
M. Mathieu et al. / Journal of Geochemical Exploration 172 (2017) 71–88 87

allows for an automation of the mineral mapping, as long as the min- Clark, R.N., Swayze, G.A., Wise, R., Livo, K.E., Hoefen, T.M., Kokaly, R.F., Sutley, S.J., 2007.
The U. S. Geological Survey, Digital Spectral Library splib06a, U.S. Geological Survey.
erals display detectible spectral features in the SWIR spectral domain. Data Series 231.
The mineral maps obtained provide information about the sample Cramer, J.J., 1986. A natural analog for a fuel waste disposal vault. Proceedings, 2nd Inter-
structure, such as foliation, bedding, and veining that can be present national Conference on Radioactive Waste Management, Canadian Nuclear Society,
Winnipeg, Man, pp. 697–702.
in the core samples. Petrographic observations have validated the Earle, S., 1996. MINSPEC1 and CALSPEC1 interactive graphical programs for quantitative
hyperspectral mineralogical mapping. The pseudo-modal compositions analysis of clay minerals from relectance data and for assessment of PIMA II calibra-
obtained relate directly to the surface spectral proportions of these min- tion spectra. Internal Report, Grasswood Geosciences (13 p).
Earle, S., 1997. MINSPEC3 program for estimation of clay contents in Athabasca Group
erals. However, some overestimation of the alteration mineral propor- sandstones from reflectance spectral data. Internal Report, Grasswood Geosciences.
tions due to the technique has to be kept in mind. The information Fayek, M., Harrison, T.M., Ewing, R.C., Grove, M., Coath, C.D., 2002. O and Pb isotopic anal-
regarding the mineralogical composition also then enables a partial yses of uranium minerals by ion microprobe and U-Pb ages from the Cigar Lake de-
posit. Chem. Geol. 185 (3–4), 205–225. http://dx.doi.org/10.1016/S0009-
differentiation between retromorphic minerals and hydrothermal
2541(01)00401-6.
minerals. Fouques, J.P., Fowler, M., Knipping, H.D., Schimann, K., 1986. The Cigar Lake uranium de-
In the particular case of the Cigar Lake uranium deposit, illitization posit: discovery and general characteristics. In: Evans, E.L. (Ed.), Uranium Deposits of
and/or chloritization of feldspars, which represent the two main alter- Canada, Canadian Institute of Mining and Metallurgy. 33, pp. 218–229 Special Vol.
Goetz, A.F.H., Vane, G., Solomon, J.E., Rock, B.N., 1985. Imaging spectrometry for earth re-
ation phases in the basement rocks around unconformity-type deposits, mote sensing. Science 228 (4704), 1147–1153. http://dx.doi.org/10.1126/science.
were identified and quantitatively estimated in hyperspectral mineral 228.4704.1147.
maps. This work represents an important progression in the automated Green, A.A., Berman, M., Switzer, P., Craig, M.D., 1988. A transformation for ordering mul-
tispectral data in terms of image quality with implications for noise removal. IEEE
detailed mapping of mineral distribution on drill cores. From the geo- Trans. Geosci. Remote Sens. 26 (1), 65–74.
logical point of view, the mineral maps indicate that the host-rock Halter, G., 1988. Zonalité des altérations dans l'environnement des gisements d'uranium
alterations similar to those observed at/near the Cigar Lake uranium de- associés à la discordance du protérozoïque moyen (Saskatchewan, Canada). Univer-
sity of Nancy.
posit also occur up to 6 km from the deposit, and up to 150 m below the Hoeve, J., Quirt, D., 1984. Mineralization and host rock alteration in relation to clay min-
unconformity in pegmatite samples. This observation shows that eral diagenesis and evolution of the Middle-Proterozoic, Athabasca Basin, northern
hyperspectral mineral mapping of drill cores could in the future help Saskatchewan, Canada. Saskatchewan Research Council, SRC Technical Report. 187
(187 p).
in defining the pathfinders for mineral exploration. Hoeve, J., Quirt, D., 1987. A stationary redox front as a critical factor in the formation of
high-grade, unconformity-type uranium ores in the Athabasca basin, Saskatchewan,
Canada. Bull. Mineral. 110, 157–171.
Acknowledgements Hoeve, J., Sibbald, T.I.I., Ramaekers, P., Lewry, J.F., 1980. Athabasca Basin unconformity-
type uranium deposits: a special class of sandstone-type deposits. In: Ferguson, J.,
The authors would like to thank AREVA Mines (Paris, France) for fi- Goleby, A.B. (Eds.), Uranium in the Pine Creek Geosyncline: Vienna. International
Atomic Energy Agency, pp. 57–594.
nancial support of this study. The authors are also very grateful to Pat- Hoffman, P.E., 1988. United plates of America, the birth of a Craton: early Proterozoic as-
rick Ledru, Craig Cutts, and Amber Doney of AREVA Resources Canada sembly and growth of Laurentia. Annu. Rev. Earth Planet. Sci. 16, 543–603. http://dx.
for their geological and logistical support. The authors would like to doi.org/10.1146/annurev.ea.16.050188.002551.
Jefferson, C.W., Thomas, D., Quirt, D., Mwenifumbo, C.J., Brisbin, D., 2007. Empirical
thank Manuel Giraud, Erwan Le Menn, and Hervé Loyen of the LPG models for canadian unconformity-associated uranium deposits. In: Milkereit, B.
(Nantes, France) for the adaptation to the support for the drill core sam- (Ed.), Proceedings of Exploration 07: Fifth Decennial International Conference on
ples, and Lawal Aboubacar Moussa (Université de Lorraine, CREGU Mineral Exploration, pp. 741–769.
Kruse, F.A., 1996. Identification and mapping of minerals in drill core using hyperspectral
GeoRessources Laboratory) for his assistance in the image processing image analysis of infrared reflectance spectra. Int. J. Remote Sens. 17 (9), 1623–1632.
portion of the thin section analysis. Kruse, F.A., Bedell, R.L., Taranik, J.V., Peppin, W.A., Weatherbee, O., Calvin, W.M., 2012.
Mapping alteration minerals at prospect, outcrop and drill core scales using imaging
spectrometry. Int. J. Remote Sens. 33 (6).
References Kruse, F.A., Boardman, J.W., Huntington, J.F., 2003. Comparison of airborne hyperspectral
data and EO-1 Hyperion for mineral mapping. IEEE Trans. Geosci. Remote Sens. 41
Agus, A.J.L., 2011. Mapping White Mica in Milled Porphyry Copper Pebbles Using (6), 1388–1400. http://dx.doi.org/10.1109/TGRS.2003.812908.
Hyperspectral Imagery : An Exploratory Study. University of Twente (57p). Kruse, F.A., Lefkoff, A.B., Boardman, J.B., Heidebrecht, K.B., Shapiro, A.T., Barloon, P.J., Goetz,
Alekandre, P., Kyser, K., Polito, P., Thomas, D., 2005. Alteration mineralogy and stable iso- A.F.H., 1993. The spectral image processing system (SIPS) - interactive visualization
tope geochemistry of paleoproterozoic basement hosted unconformity-type uranium and analysis of imaging spectrometer data. Remote Sens. Environ. 44, 145–163.
deposits in the Athabasca Basin, Canada. Econ. Geol. 100 (8), 1547–1563. Kurz, T.H., Buckley, S.J., Howell, J.A., 2013. Close-range hyperspectral imaging for
Annesley, I.R., Madore, C., Portella, P., 2005. Geology and thermotectonic evolution of the geological field studies: workflow and methods. Int. J. Remote Sens. 34 (5),
western margin of the trans-Hudson Orogen: evidence from the eastern sub- 1798–1822.
Athabasca basement, Saskatchewan. Can. J. Earth Sci. 42 (4), 573–597. http://dx.doi. Macdonald, C., 1985. Mineralogy and geochemistry of the sub-Athabasca regolith near
org/10.1139/E05-034. Wollaston Lake. In: Sibbald, T.I.I., Petruk, W. (Eds.), Geology of Uranium Deposits. Ca-
Baissa, R., Labbassi, K., Launeau, P., Gaudin, A., Ouajhain, B., 2011. Using HySpex SWIR- nadian Institute of Mining and Metallurgy 32, pp. 155–158 Special Volume.
320m hyperspectral data for the identification and mapping of minerals in hand Pacquet, A., Weber, F., 1993. Petrography and mineralogy of alteration halos around the
specimens of carbonate rocks from the Ankloute formation (Agadir Basin, Western cigar Lake ore-body and their relation to the mineralization process. Can. J. Earth
Morocco). African Earth Sciences]–>J. Afr. Earth Sci. 61 (1), 1–9. http://dx.doi.org/ Sci. 30, 674–688. http://dx.doi.org/10.1139/e93-055.
10.1016/j.jafrearsci.2011.04.003. Post, J.L., Noble, P.N., 1993. The near-infrared combination band frequencies of
Boardman, J.W., Kruse, F.A., 1994. Automated spectral analysis: a geological example dioctahedral smectites, micas, and illites. Clay Clay Miner. 41 (6), 639–644.
using AVIRIS data, north Grapevine Mountains, Nevada. Proceedings, ERIM Tenth Quirt, D.H., 2010. Is Illite still a Pathfinder Mineral for the Geological Environment of Ath-
Thematic Conference on Geologic Remote Sensing, Environmental Research Institute abasca Unconformity-Type Uranium Deposits?? GeoCanada 2010, Calgary. May
of Michigan, Ann Arbor, MI, pp. I-407–I-418. 2010, 4 p http://www.cspg.org/cspg/documents/Conventions/Archives/Annual/
Bruneton, P., 1987. Geology of the Cigar Lake uranium deposit (Saskatchewan, Canada). 2010/0858_GC2010_Is_Illite_Still_a_Pathfinder_Mineral.pdf.
In: Gilboy, C.F., Vigrass, L.W. (Eds.), Economic Minerals of Saskatchewan, Quirt, D.H., 2001. Kaolinite and dickite in the Athabasca Sandstone, northern Saskatche-
pp. 99–119 Special publication no. 8. wan, Canada. Saskatchewan Research Council, Publication No. 10400-16D01 28 p.
Bruneton, P., 1993. Geological environment of the Cigar Lake uranium deposit. Can. Quirt, D.H., 2013. Clay and clay-size minerals as vectors to unconformity-type uranium-
J. Earth Sci. 30, 653–673. http://dx.doi.org/10.1139/e93-054. gold deposits. The 15th International Clay Conference, July 7–11 2013, Rio de Janeiro,
Buckingham, W.F., Sommer, S.E., 1983. Mineralogical characterization of rock surfaces Brazil http://15icc.org/abstracts/abstracts/669.
formed by hydrothermal alteration and weathering - application to remote sensing. Ramaekers, P., Jefferson, C.W., Yeo, G.M., Collier, B., Long, D.G., Catuneanu, O., Bernier, S.,
Econ. Geol. 78 (4), 664–674. http://dx.doi.org/10.2113/gsecongeo.78.4.664. Kupsch, B., Post, R., Drever, G., McHardy, S., Jircka, D., Cutts, C., Wheatley, K., 2007. Re-
Cameco, 2014. Cigar Lake. http://www.cameco.com/businesses/uranium-operations/ vised geological map and stratigraphy of the Athabasca group, Saskatchewan and Al-
canada/cigar-lake; (accessed December, 2015). berta. In: Jefferson, C.W., Delaney, G. (Eds.), EXTECH IV: Geology and Exploration
Clark, R. N., 1999. Manual of remote sensing, Volume 3, Remote Sensing for the Earth Sci- Technology of the Proterozoic Athabasca Basin, Saskatchewan and Alberta. Geological
ences, chap. Spectroscopy of Rocks and Minerals, and Principles of Spectroscopy, Survey of Canada Bulletin. 588, pp. 155–191.
pp. 3–58. John Wiley and Sons, New York. Roache, T.J., Walshe, J.L., Huntington, J.F., Quigley, M.A., Yang, K., Bil, B.W., Blake, K.L.,
Clark, R.N., Swayze, G.A., Gallagher, A.J., King, T.V.V., Calvin, W.M., 1993. The U. S. Geolog- Hyvarinen, T., 2011. Epidote-clinozoisite as a hyperspectral tool in exploration for Ar-
ical Survey, Digital Spectral Library: Version 1: 0.2 to 3.0 microns. U.S. Geological Sur- chean gold. Australian Journal of Earth Sciences]–>Aust. J. Earth Sci. 58 (7), 813–822.
vey Open File Report 93-592 (1340 pp). http://dx.doi.org/10.1080/08120099.2011.608170.
88 M. Mathieu et al. / Journal of Geochemical Exploration 172 (2017) 71–88

Roy, R., Launeau., P., Carrere, V., Pinet, P., Ceuleneer, G., Clénet, H., Daydou, Y., Girardeau, J., Tappert, M.C., Rivard, B., Fulop, A., Rogge, D., Tappert, R., Stalder, R., 2015. Characterizing
Amri, I., 2009. Geological mapping strategy using visible near-infrared–shortwave in- kimberlite dilution by crustal rocks at the snap Lake diamond mine (northwest terri-
frared hyperspectral remote sensing: application to the Oman ophiolite (Sumail Mas- tories, Canada) using SWIR (1.90–2.36 μ m) and LWIR (8.1–11.1 μm) hyperspectral
sif). Geochem. Geophys. Geosyst. 10. http://dx.doi.org/10.1029/2008GC002154 imagery collected from drill Core. Econ. Geol. 110 (6), 1375–1387. http://dx.doi.
Article N°Q02004. org/10.2113/econgeo.110.6.1375.
Hunt, G.R., Salisbury, J.W., 1970. Visible and near-infrared spectra of minerals and rocks: I. Turner, D., Rivard, B., Groat, L., 2014. Rare earth element ore grade estimation of mineral-
silicate minerals. II. Carbonates. Modern Geology. 1, pp. 283–300. ized drill core from hyperspectral imaging spectroscopy. Geoscience and Remote
Smellie, J., Karlsson, F., 1996. A reappraisal of some Cigar Lake issues of importance to per- Sensing Symposium (IGARSS). 2014 IEEE International, pp. 4612–4615 http://dx.
formance assessment. (Technical Report No. TR 96–08). SKB. doi.org/10.1109/IGARSS.2014.6947520 July 13–18 2014, Québec (Canada).
Speta, M., Rivard, B., Feng, J., Lipsett, M., Gingras, M., 2013. Hyperspectral imaging for the Vane, G., Goetz, A., 1988. Terrestrial imaging spectroscopy. Remote Sens. Environ. 24,
characterization of athabasca oil sands drill core. Geoscience and Remote Sensing 1–29.
Symposium (IGARSS). 2013 IEEE International, pp. 2184–2187 http://dx.doi.org/10. Williams, A.P., Hunt, E.R., 2002. Estimation of leafy spurge cover from hyperspectral imag-
1109/IGARSS.2013.6723248 July 21–26 2013, Melbourne (Australie). ery using mixture tuned matched filtering. Remote Sens. Environ. 82 (2–3), 446–456.
Sun, Y., Zhao, Y., Qin, K., Nie, J., Li, H., 2015. Geological application of HySpex ground Zhang, J.-L., Huang, Y.-J., Wang, J.-U., Zhou, M., Wu, D., Xuan, Y.-X., 2013. Hyperspectral
hyperspectral remote sensing in gold and uranium ore deposits. Proceedings of the drilling core logging and 3D mineral mapping technology for uranium exploration.
2015 Asia-Pacific Energy Equipment Engineering Research Conference (AP3ER Uranium Geology 29 (4), 249–255.
2015), pp. 392–395, June 13–14 2015, Zhuhai (China). AER-Advances in Engineering
Research 9, pp. 392–395.

You might also like