Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 40

Teledetección II– Practice

National University of San Marcos


Professional School of Geological Engineering
Undergraduate Unit

TERRASPEC
GROUP 3

COURSE: Remote Sensing II


TEACHER: MSc. Diana Pajuelo Aparicio
SEMESTER: 2021-I

MEMBERS:
• Cáceres Coaquira, Renzo
• Cubas Rivera, Ángel Cristofher
• Osorio Alcántara, Juan Piere
• Ramirez Sanchez, Jhonatan Alen

2021
Teledetección II– Practice

► CONTENT
1. Introduction
2. Theoretical framework
2.1. Reflectance spectroscopy
2.2. Spectrometers
3. Background
3.1. Before Terraspec
3.2. Previous studies
4. Terraspec
4.1. ASD TerraSpec Halo
4.2. ASD TerraSpec 4
4.3. Use
4.4. ASD Configuration
4.5. Collecting Data
4.6. Spectral libraries
4.7. Spectral Analysis
4.8. Complementary methods
4.9. Importance
4.10. Future projection
5. Applications
6. Conclusions
7. Bibliography
Teledetección II– Practice

1. Introduction

The development of Earth remote sensing has largely depended on the creation of new tools for
data acquisition and analysis. Portable ASD spectrometers are important tools and are widely
used for spectral measurements in many scientific fields and for the validation of satellite and
airborne spectral images, for more accurate image analysis and interpretation. What is notable
about imaging spectrometry and remote sensing is that details that are not visible to the human
eye can be detected.

2. Theoretical framework.
2.1. Reflectance spectroscopy.

It is the study of the interaction between energy as electromagnetic radiation (EMR) and matter.
Radiation can be absorbed , emitted or scattered by matter.

Reflectance spectroscopy can be defined as the technique that uses energy in the visible (0.4-
0.7), near infrared (0.7-1.0), and shortwave infrared (1.0) wavelength regions. 0-2.5 µm) of the
electromagnetic spectrum ( Fig. 1 ) to analyze minerals.

Fig.1 electromagnetic spectrum

The science and techniques of reflectance spectroscopy are based on the spectral properties of
materials. Certain atoms and molecules absorb energy based on their atomic structures.

The different physical and chemical properties of minerals give rise to their unique spectral
reflectance characteristics and many are ''infrared active'' due to the presence of molecules that
vibrate when they absorb light.

In the visible region , this is due to electronic transitions such as crystal field effects (atomic
energy level transitions), charge transfer (electronic transitions between elements), conduction
band transitions (electron transfer over spatially close energy levels) and the color center
phenomenon (energy levels induced by lattice defects). Much of this simply involves the release
of energy when an electron changes energy levels in an atom.
Teledetección II– Practice

Fig. 2 VNIR and SWIR Region

The absorption characteristics in the SWIR region are a function of the mineral composition.
They are a manifestation of the absorption of energy within the crystalline lattice from the
transitions of vibrational states ( Fig. 3). Since these vibrational states correspond to different
energy levels, the absorption features occur at well-defined wavelength positions. The energy
levels that define these wavelengths are a function of the size of the ionic radii of the cations
bonded to different molecules. Bonds will vibrate at different wavelengths depending on the
length of the bond. Because the bond lengths between a specific atom and molecule will be
consistent, it is possible to predict the compositions and compositional changes in the minerals
being analyzed by the wavelengths and shifts thereof. (Hunt, 1977).

Symmetrical Antisymmetric
Scissors
Stretch Stretch

Link Vibration
Modes:
Swinging Wiggling Torsion

Fig. 3 Modes of vibration of the molecules of a mineral.

Transitions between energy levels and compositional differences are manifested by absorption
characteristics at defined wavelengths ( Fig. 4). The common positions of absorption traits are
listed in Table I.
Teledetección II– Practice

Fig. 4. Types of energy level changes associated with different parts of the EM spectrum

Each mineral has a distinctive spectral signature, composed of various absorption characteristics,
which is a function of composition , crystallinity , concentration , water content and
environmental considerations. For mineral identification, that signature is compared to
characterized references for a validated identification. This can be done manually using
wavelength tables and overlaying references to the unknown or it can be done using computer
algorithms that operate against a database of references (i.e., a spectral library). For best results,
these libraries should be configured for specific bucket types and not allow the algorithm to
make the wrong choice.

Although computer algorithms are useful for data management and organization, care must be
taken when using them for phase identification and quantification. There are many variables
that are not taken into account and identifications not chosen by the user are not always
accurate. Additionally, important information about crystallinity, water, organic elements and
iron, among other things, is lost.

2.2. Spectrometers.

They are instruments that allow collecting and presenting spectral data . Although the main
objective of this operation is the identification of the minerals found in the different alteration
systems, there is other information that can be obtained from the data and that is of great
importance in the interpretation of the spectral information. These include chemical
substitution , crystallinity , water effects , paragenesis and temperature .

Modern portable spectrometers, such as those of the ASD Terraspec line, are equipped with
spectral libraries that facilitate the task of mineralogical identification.

3. Background.
3.1. Before Terraspec

Reflectance spectroscopy of minerals began in the early 20th century. XX, with Coblentz (1906)
on organic surfaces and minerals. With Lyon and Burns (1963) he focused on glass mineralogy,
as well as quartz and cristobalite as analogues of glass structure.
Teledetección II– Practice

In the early 1970s, Hunt and Salisbury (1970) initiated a systematic study of reflectance
spectroscopy of reference rocks and minerals, measuring the behavior in the visible (VIS: 350-
700 nm; ~28,500-14,300 cm-1 ), in the near infrared (NIR: 700-1300 nm; ~14,300-7700 cm-1)
and in the short infrared (SWIR: 1300 -2500 nm; ~7700-4000 cm-1) to assist in remote sensing
studies .

Over the last decade, the field of exploration remote sensing has undergone a fundamental
transformation, moving from image processing to the extraction of spectroscopic mineralogical
information, giving rise to the broader field of Spectral Geology and Remote Sensing (SGRS). ,
which encompasses technologies that contribute to the definition, confirmation and
characterization of mineral deposits.

Therefore, for certain types of mineral deposits, the recognition and detection of hydrothermal
alteration minerals ( e.g. Thompson and Thompson, 1996) are of utmost importance for
exploration, as they form a much larger target than the mineral itself. mineralization.

During the 1990s, the development of hand-held instruments allowed the direct collection of
rock and mineral spectra in the field, to verify remote sensing data (e.g. Landsat Thematic
Mapper). During this period, several instruments based on the VNIR-SWIR were developed, such
as the GER-IRIS (Geophysical Environmental Research, Inc., CSIRO, Australia), the ASD-FieldSpec
(Analytical Spectral Devices, Inc., now Malvern Panalytical, Longmont, Colorado, USA, hereinafter
referred to as ASD-Fieldspec), and the PIMA (Portable Infrared Mineral Analyzer, Integrated
Spectronics Pty. Ltd, Castle Hill, New South Wales, Australia, hereinafter referred to as PIMA).
Some of these instruments relied on solar illumination (e.g., GER-IRIS) while others used internal
light sources (e.g., PIMA) (Thompson et al., 1999). The ASD-Fieldspec included a contact probe
with internal light, but could also operate in sunlight. With the advent of these instruments and
simultaneous progress in software development, mineral reference databases were created
and/or expanded to aid interpretation. Some databases were available with the purchase of an
instrument.

3.2. Previous studies

Early field and laboratory work focused on the application of the SWIR region in the
identification of minerals of interest in alteration zones (Hunt and Ashley, 1979; Thompson et al
., 1999 ). The current work makes use of the full VIS-NIR-SWIR regions (350-2500 nm) to
characterize common and uncommon minerals in a variety of formation environments.

In Tasmania, studies with a PIMA-II spectrometer showed that white micas are classified from
phengite (peripheral zones) to sodium muscovite (proximal zones) in the stratiform Zn+Pb
deposit of Rosebery (Herrmann et al., 2001), which was supported by Yang et al. (2011), who
studied changes in white mica in the Hellyer VMS polymetallic deposit (Tasmania). VNIR-SWIR
spectroscopy was applied by Laakso et al. (2015, 2016) to differentiate phyllosilicates based on
their Al-OH and Fe-OH absorption characteristics in a study of the Izok Lake VMS reservoir,
Nunavut, Canada. These authors determined that proximal phyllosilicates had an average Al-OH
absorption feature (white mica) at 2203 nm versus a distal Al-OH absorption feature at 2201 nm,
and Fe-OH absorption features (biotite/ chlorite) at 2254 nm (proximal) versus 2251 nm for the
distal phyllosilicates.

In another type of deposit, epithermal gold deposits, Kerr et al. (2011) applied spectral analysis
in the Bobbys Pond and Western Avalon districts. At Bobbys Pond, a barren area was examined
and pyrophyllite, alunite and topaz were easily identified. It was determined that the detection
Teledetección II– Practice

of the presence of topaz, difficult to recognize in hand specimens or petrographically. In the


Western Avalon studies, in addition to identifying pyrophyllite and alunite, dickite was detected.
This mineral is an indicator of Au mineralization in high sulfidation epithermal systems. Other
studies, such as the Cerro La Mina epithermal prospect in Patagonia, Argentina (Ducart et al.,
2006), have also demonstrated the application of SWIR in the identification of advanced argillic
alteration in these systems using hand-held instruments and hyperspectral remote sensing.

For sediment-hosted lead zinc deposits (SEDEX), it was tested by Peter et al. (2015) in a pilot
study in the Selwyn Basin, Yukon, Canada, in reservoirs in the Howard's Pass and MacMillan Pass
districts. Mineralization (and the immediately adjacent host shale) at Howard's Pass is
characterized by phengite (52214 nm), while muscovite (42209 nm) is prevalent everywhere else.
At MacMillan Pass, siderite, muscovite, phengite and montmorillonite are identified spectrally
immediately below and within the mineralized zone. However, the mineralogy of ores (pyrite,
galena and sphalerite, with siderite, white mica (muscovite, phengite), etc.) and their lack of well-
developed hydrothermal alteration zones adjacent to the mineralization zone, is a challenge that
still exists. must be overcome, this is due to its low reflectance and weak absorption.

In a study of the Pebble Cu-Au-Mo porphyry deposit in Iliamna, Alaska, Harraden et al. (2013)
demonstrated that SWIR spectroscopy is suitable for distinguishing illite, sericite, kaolinite, and
pyrophyllite in drill core. These authors assigned alteration types to different ranges of Al-OH
absorption characteristics between 2160 and 2220 nm and produced a detailed map. These
authors observed that Au and Cu mineralization is closely associated with the lower wavelengths
of Al-OH absorption features associated with pyrophyllite and sericite and concluded that SWIR
spectroscopy is a useful tool for exploration, since clay minerals were genetically linked to
mineralization.

4. Terraspec.

It is a cost-effective technique for the identification of minerals in the field, that is, it allows the
acquisition of chemical analyzes in situ, non-destructively, quickly, and generates qualitative,
semiquantitative and quantitative results, due to external factors such as calibration,
interferences (atmospheric or mineralogical-lithological).

The ASD TerraSpec line of mineral analyzers provides portable, rapid, non-destructive analysis of
minerals in the laboratory or in the field. It features the portability of the ASD TerraSpec Mineral
Analyzer and its ability to quickly identify key minerals. When used as a mineral analyzer, results
are presented in seconds to speed mineral processing decisions.

It has two main types of spectrometers, which are part of this line, distinguishable according to
their specifications and application ( Table 1 and 2 )

4.1. ASD TerraSpec Halo

It is a full-range VIS-NIR-SWIR spectrometer that measures the visible, near-infrared and


shortwave infrared (350-2500 nm) regions. The instrument includes internal references for easy
operation and data management, and also features state-of-the-art mineral identification
software that allows data to be quickly captured in the field and easily managed in the
laboratory or core room. .
Teledetección II– Practice

4.2. ASD TerraSpec 4

Recognized as the de facto technology for mineralogical analysis. The ASD TerraSpec 4 Hi-Res
mineral analyzer brings new levels of efficiency to mineral exploration technology. This next-
generation mineral spectrometer offers improved performance in SWIR 1 and 2 regions and
6nm resolution

Spectrometer Specifications
 Spectral range (nm): 350-2500
 Spectral resolution: 3nm @ 700nm, 9.8nm @ 1400nm,
ASD TerraSpec Halo
8.1nm @ 2100nm

 Wavelength range: 350-2500nm


ASD TerraSpec 4  Resolution: 3nm at 700nm and 6nm at 1400/2100nm

Table 1 . Spectrometer specifications are described

Spectrometer Advantages Benefits Applications


ASD TerraSpec  Oil exploration
Halo  Allows you to  Laptop.  Mining production and
quickly capture  Fast and precise. mineral processing
data in the field  Without cables.  Mining exploration
and easily  Simple capture and  Remote sensing and
manage it in management of mineral geology
Teledetección II– Practice

the laboratory data.  Mining/energy


or witness  Immediate multi- exploration
room mineral and scalar  Alteration map
results.  Exploring green/brown
 Extensive spectral fields
library.  Mine mapping
 Easy handling features  Analysis of clay species
 No third party hardware.  Confirmation Sampling
 Spectrum management
Analysis of oil sands

 Allows accurate  Improved opticaland oil shales.
evaluation of configuration provides
 Oil exploration.
low more accurate spectral
 Mining production and
concentration, results. mineral processing.
low reflectance Industry-leading  Mining exploration.
ASD TerraSpec
minerals performance in SWIR  Mapping of deposits.
4
1/SWIR 2 regions.  Determination of
geochemical gradients.
 Delimitation of clay
species.
 Faster vectoring to the
site.
Table 2 . Comparison between portable spectrometers, depending on the study to be carried
out.

4.3. Use

In this case, more detailed information will be required for the TerraSpec HALO , which is ready
to use; however…

 It is recommended that you charge the batteries first

 It is recommended that you configure the Halo Manager software before using it in the
field

It comes with one default project and two default locations, however…

 You can immediately go to the field and collect data


 You can add new locations in addition to the default ones
 Comes with customizable location settings

HALO identifies up to 7 minerals and shows the 4 best results, and up to 7 scalar values.

As HALO boots, it runs a full set of diagnostic tests and automatically takes a white reference
scan from the external Spectralon drive.

 After the initial external white reference, HALO will continue to take white reference
scans internally. This takes about 20 seconds. (For the first half hour, it will stop to scan
every 2 minutes. After the first half hour, it will stop every 10 minutes).
 The main menu displayed ( Fig. 5 ), you CANNOT collect spectra while HALO is doing
an internal reference scan.
Teledetección II– Practice

 After HALO has restarted, you will see the menu screen.

Fig. 5 Terraspec menu screen

4.4. Halo Configuration (Fig. 6)


 GPS – On/Off
 Units – Feet, Meters
 Language - English, Spanish and Chinese
 Time zone: any of the universal time zones
 User Type - Technician, Geologist, where the Geologist settings will display all mineral
identifications, including those with a confidence of one star, the lowest confidence
level. Technician Settings will show only two and three star results.

Fig. 6 Terraspec menu screen for the corresponding settings.

4.5. Collecting data


 HALO identifies up to 7 mineral matches from its mineral database and reports the top
4. (Halo Manager shows all 7 matches).
 HALO also calculates the corresponding mineral scalars and reports the first 7.
Teledetección II– Practice

 The star rating next to each mineral indicates confidence.

90% Confidence
60-90% Confidence
Less than 60% Confidence

Fig. 7 Kaolinite + Montomorillonite + Goethite

Fig . 8 Goethite + Montmorillonite + Kaolinite

4.5.1. Recommendations for collecting samples


Teledetección II– Practice

As much as possible, samples should be:

 Dry . Moisture masks diagnostic features of many minerals.

The sample surface must be dry. If you suspect moisture on the sample surface, hold the
sample against the HALO measurement window for 30 seconds before pressing the trigger.
The HALO light source will often dry the sample surface enough to get a good reading.

 Clean – free of organic material; free of oxidation surfaces

This includes mineral layers and excessive moisture. Organic residue such as soil, algae,
lichens and other plant materials that obscure underlying mineralogy, and while it is
possible to take data on weathered surfaces, this complicates interpretation as weathering
minerals show up on those identified by the HALO.

 Hold the sample steady and stable as much as possible against the HALO measurement
window while the system is scanning.

4.6. Spectral Libraries

The mineral libraries available with the purchase of a spectrometer are usually proprietary. They
include the most common VNIR-SWIR active minerals, especially those with strong SWIR
patterns. Libraries may be available on board the instrument, or through third-party software
such as TSG (CSIRO) or Specmin (Spectral International).

TerraSpec Halo Library (Analytical Spectral Devices, Inc., now Malvern Panalytical, Longmont,
Colorado, USA, hereinafter referred to as ASD-Halo), contains >500 spectra of >150 minerals
that are spectrally active in the regions VIS, NIR and SWIR, which represents about 50 sets of
spectral reflection data (processed as absolute reflectance) from the Kodama clay collection, that
is, ASD, Inc. evaluated the spectra collected from the various specimens and, based on their
quality, representativeness or uniqueness, incorporated them into the on-board library of the
ASD-Halo instrument ( Fig. 9 and 10 )

Following the success of the spectral library of the Kodama clay collection, another spectral
library is being built for all (spectrally active) minerals associated with critical metal deposits, i.e.
minerals carrying REE, U, Th and Nb found in carbonatites and alkaline intrusive rocks.

Kodama Clay Collection

The SGC National Mineral Reference Collection (NMC) contains many pure specimens that have
been characterized by X-ray diffraction (XRD) and/or electron probe microanalysis (EPMA). The
Kodama Clay Collection was the initial test (; Percival et al., 2016). This spectral library contains
information about the analyzed specimen, including the date of collection, the locality, and a
brief description of the sample; a text file of the data (relative reflection) so that it can be
manipulated by other programs for plotting; and an image of the layout generated in MS-
Excel1 .

For the Kodama clay collection, each spectrum was exported as an ASCII type file. Each set of
spectra was entered into MS-Excel1 and averaged. The average spectrum was used to create the
graphs and exported as a .pdf file. Data averaged with wavelength information were processed
to create a .txt file for ease of accessibility (Percival et al., 2016a).
Teledetección II– Practice

Fig . 9 Stacked VNIR-SWIR spectra of important alteration minerals.

Fig. 10. Impact of spectral and spatial resolution on a spectral signature.

4.7. Spectral Analysis

The process uses the mineral combination algorithm

 HALO uses a proprietary algorithm to match minerals.


 The unknown specter fits the onboard mineral library. Once the best-fit mineral
spectrum is matched, it is subtracted from the unknown spectrum.
 Using the remainder of the unknown spectrum, the process is repeated to generate up
to 7 mineral matches

Once you've collected your wraiths, sync them to HALO using Halo Manager.

 All data and new locations created in HALO are synchronized with Halo Manager and
stored in the correct project and locations.
 The project with the selected locations and their configuration in the Halo Manager sync
queue are uploaded to HALO.

4.7.1. Halo Manager


Teledetección II– Practice

 Halo Manager is the software that can download spectra and scalars on a Windows
computer.

You can create projects and locations.

Configure settings for locations.

View data, listen to audio notes, add written notes.

Export to third-party software.

 Important! You cannot move specters/data between projects or locations in Halo Manager!

You can rename locations in Halo Manager.

You can export spectra/data and rearrange.

Fig. 11 Halo manager interface with collection data


Teledetección II– Practice

Fig . 12 Verification of spectral signatures according to the station where it was taken.

File export

There are many options for exporting data in Halo Manager. The most direct way is to export all
samples in one location.

1. To select all samples collected under a location mark, check the "select all" box.
2. Click on the "Sample Actions" drop-down menu.
3. Choose "Export Selected"

Fig . 13 The 'select all' option is marked


Teledetección II– Practice

Fig . 14 The 'Sample Actions' option is marked and finally 'Export Selected'

Once the samples have been exported, you can navigate to them using Windows Explorer. They
are always stored in C:\Users\Public\Halo\Exports

Fig . 15 The data exported from the spectrometer is observed in different formats.

 Spectral data are exported in .ASD or .txt format as appropriate. ( Fig. 15)
 Mineral IDs, star ratings and scalar information can be found in the text file at the end.
Teledetección II– Practice

Fig. 16. Final data table.

You can upload or import spectral data to any third party software such as TSG easily by
navigating to the export folder.

Fig. 17. Export process to third-party users of final data.

Scalars

 Scalars describe properties of crystallinity or mineral composition.

 These tell the geologist about geological or geothermal conditions that can lead to
potential mineralization.

 When a mineral is identified, the database searches for the related scalar. The
appropriate scalar will be activated according to the minerals identified in the sample.
For example, Phengite, which contains Al bonds, activates the AL-OH scalar.

 They can be reported as a position in nanometers of a feature in the wavelength region


(Al-OH 2200) OR as the reflectance value of a ratio (eg Kx 0.9)
Teledetección II– Practice

Reported scalars
Reported when the HALO detects a mineral that has an Al-OH bond in the
Al-OH structure; useful for establishing geochemical gradients since this
characteristic changes with composition
Reported when HALO detects the mineral kaolinite; values > 1 indicative of
Kx kaolinite produced by weathering, values <1 associated with those produced
by alteration.
Reported when the HALO detects a mineral from the Illitas group; increasing
ISM
ISM value indicative of increasing thermal maturity or metamorphic grade.
Reported when the HALO detects a mineral from the Chlorite group;
CSM Increasing CSM value is indicative of increasing thermal maturity or
metamorphic grade.
It is always reported that the HALO detects a mineral that has an Mg-OH bond
Mg-OH
in the structure; useful for establishing geochemical gradients.
Reported when the HALO detects a mineral that has an Fe-OH bond in the
Fe-OH
structure; useful for establishing geochemical gradients.
Reported every time HALO detects an Fe3+ mineral; useful for establishing
Fe3t
Fe3+ mineralogy.
Ferrous Mineral Intensity. It is reported when the HALO detects an Fe3+
Fe3i
mineral; track the intensity of the Fe3+ absorption function.
Reported whenever the HALO detects a mineral that has an Al-OH, Fe-OH or
Al-Fe-Mg
Mg-OH bond in the structure.

As the mineralogy of the system changes, the Al-OH feature shifts from left to right (2189 to
2200 to 2224), deepening and widening. By observing how the values of the Al-OH function
change, a geologist could tell whether they are in the paragonitic or fengitic part of the system.
( Fig. 18)

Fig. 18. Scalar spectral signatures.

The scalars are reported on the screen.

 Up to 9 scalars are reported, while 7 can be reported on the screen. All scalars can be
viewed in the Halo Manager.
 Minerals and scalars are listed separately.
 The scalars present only correspond to the minerals found and are ordered according to
the frequency of use.
Teledetección II– Practice

Fig. 19. Report on the scalars terraspec screen.

Complementary Methods

The use of hyperspectral and multispectral images helps to better delimit the study area and
map minerals of hydrothermal alteration.

Importance.

Portable spectrometers provide in situ data that allows the geologist to immediately integrate
multiple data sets (geophysical, geochemical, and mineralogical) on the ground, at the outcrop,
where this information can be quickly applied to better understand what is physically observed –
instant gratification . This was not possible before its invention. Reflectance spectroscopy is
changing the nomenclature of exploration and allows geologists to observe mineral associations
that provide a deeper understanding not only of the rocks themselves, but of the processes that
have transformed them into the precious metal resources so important to the world.

Future projection

Over the next decade, hyperspectral recording will be more fully incorporated into project
workflows. This will be on the application side. Advanced data science, such as machine learning,
is one of the emerging tools to analyze and integrate high-dimensional, high-volume core
hyperspectral image data with other geoscience datasets. It will also transcend the sampling
process where it will be integrated with other analytical methods. This will allow the data set to
become the fundamental base layer used in witness registration.
Teledetección II– Practice

Applications

The applications of the use of the Terraspec spectrometer will be made in application to
Porphyry copper deposits, by the author Neal Luke, Jamie Wilkinson, J. Mason and Z. Chang.

Spectral characteristics of propylitic alteration minerals as a vectorization tool for


porphyry copper deposits

Shortwave infrared reflectance (SWIR) spectroscopy is a rapid and effective method for
detecting and characterizing hydrothermal alteration associated with mineral deposits, and can
identify not only mineral species but also changes in element composition. main minerals.
Porphyry deposits represent large accumulations of valuable metals in the Earth's crust and
exhibit extensive alteration signatures, making them an attractive target for exploration,
especially using remote sensing, which can cover large areas quickly. Reflectance spectroscopy
has been widely applied in the sericic (phyllic), argillic, and advanced argillic alteration domains
because it is particularly effective in discriminating bright clay minerals. However, the propylitic
domain has remained relatively unexplored because propylitic rocks are typically dark and
produce relatively indistinct spectra.

This study used a portable ASD TerraSpec 4 spectrometer to collect SWIR spectra of rocks
surrounding the Batu Hijau porphyry Cu-Au deposit in Indonesia, where previous work has
identified systematic spatial variations in chlorite chemistry, a common propylitic alteration
mineral. Spectra were collected from 90 samples and processed using The Spectral Geologist
(TSG) software, as well as the Halo mineral identifier, to characterize the mineralogy and extract
the positions and depths of spectral absorption features, which were then correlated with the
geochemistry of the main elements. Two absorption features of chlorite located around 2250
nm and 2340 nm correlate with the Mg# (Mg/[Mg+Fe]) of chlorite, both in terms of wavelength
position and depth. As Mg# increases, the wavelengths of both features increase from 2249 nm
to 2254 nm and from 2332 nm to 2343 nm respectively, and the absorption depths also increase
significantly. In the spatial dimension, these feature variations act as reasonably strong vectors
towards the reservoir, showing systematic increases along a transect away from the center of the
porphyry, reaching a maximum at distances of around 1.6 km, which coincides with the spatial
trend shown by Mg#, as well as with several trace element indicators in chlorite. The hull slope
in the spectra between 1400 nm and 1900 nm is also shown to increase with Mg#, and the
Teledetección II– Practice

position of an absorption feature at 1400 nm increases with the Al:Si ratio, a parameter that also
tends to increase with proximity to porphyry deposits.

Depth variations of features, in particular, appear to represent a new finding in chlorite


reflectance spectroscopy; However, the causes are not entirely clear and require further
investigation. However, the systematic behavior provides a potentially useful new tool for
exploration in zones of propylitic alteration.

Remote sensing is a valuable tool in mineral exploration, providing a rapid way to identify and
map hydrothermal alteration products over large areas. Much work has been done on the
characterization of hydrothermal alteration associated with mineral deposits in terms of spectral
signatures (Yang & Huntingdon 1996; Herrmann et al. 2001; Sun et al. 2001; Jones et al. 2005),
specifically including porphyry deposits (Cudahy et al. 2001; Chang et al. 2011; Dilles 2012;
Zadeh et al. 2014; Halley et al. 2015). However, these studies focus on the dominant clayey,
argillic, and sericitic alteration assemblages. Regarding a particular scanning application, Chang
et al. (2011) demonstrated that it is possible to use the spectral characteristics of alunite in the
zone of advanced argillic alteration as indicators of porphyry deposits.

Research on SWIR spectroscopy of chlorite and epidote, two of the most common propylitic
alteration minerals, is scarce. However, the sensitivity of SWIR spectra to chemical changes in
minerals that are recognizable both in the laboratory has been recognized (King & Clark 1989;
Liebscher 2004; Bishop et al. 2008) as in hyperspectral images (Cudahy et al. 2001; Roache et al.
2011). This indicates that the spectral features could act as a vector for mineral deposits in the
propylitic domain, as well as for more proximal alteration zones.

The large footprint of propylitic alteration should present an attractive target for remote sensing
mineral exploration. However, the rocks that typify zones of propylitic alteration are usually dark
(low reflectance), which makes their characterization extremely difficult. Fortunately, the
availability of increasingly sensitive portable field spectrometers has opened the possibility of
addressing this problem. These spectrometers are an ideal tool for detailed investigation of the
spectral characteristics of rocks; They can produce high-resolution spectra, free from the effects
of atmospheric scattering and absorption, and have clearly detectable absorption characteristics
even with reflectance values as low as 1%.

This study used an ASD TerraSpec 4 spectrometer to collect infrared spectra of rocks from the
Batu Hijau porphyry copper system in Indonesia. It builds on previous research carried out as
part of the AMIRA P765A research project (AMIRA International Ltd) which included studies of
spatial variation in chlorite geochemistry (Wilkinson et al. 2015) and epidote (Cooke at al.
2014b). The most important thing is that the work used the same samples analyzed by
Wilkinson et al. (2015), so extensive prior knowledge of whole-rock geochemistry and mineral
chemistry was available. The main objectives were to characterize the infrared spectra of
propylitic rocks along the Batu Hijau alteration footprint, and to identify any spectral features
that could reflect systematic geochemical variation in the spatial domain that could therefore
act as vectors towards economic mineralization. In the longer term, it is hoped that such
features could be located by hyperspectral airborne sensors or even by satellite, such as ASTER
Teledetección II– Practice

or WorldView-3, which would allow a relatively rapid search of large areas of propylitic alteration
for porphyry deposits.

geology

Batu Hijau is a gigantic Cu-Au porphyry deposit located in the southwestern part of the island of
Sumbawa, Indonesia. It is located in the Sunda-Banda Arc, which hosts numerous magmatic-
hydrothermal deposits (Garwin et al. 2005), related to calcareous-alkaline magmatism resulting
from the subduction of the Indo-Australian plate under the Eurasian plate. Before the start of
open pit mining, Batu Hijau contained an estimated 1,644 million tonnes of ore with copper and
gold grades of 0.44% and 0.35 g/t respectively (Cooke et al. 2005).

Figure1 . Batu Hijau, a gigantic Cu-Au porphyry deposit located in the southwestern part of Sumbawa Island,
Indonesia

The Batu Hijau district is largely formed by an early to middle Miocene volcaniclastic sequence
comprising volcanic lithic breccia and volcanic sandstones with local layers of limestone. This
sequence is cut by several intrusive phases dating to the early to middle Miocene and including
andesites and andesite porphyries, quartz diorites, and porphyritic tonalites (Garwin 2002).
Mineralization is strongly associated with tonalitic porphyries in four main centers: Batu Hijau,
Sekongkang, Arung Ara and Katala. Batu Hijau is significantly larger and of higher grade than
the rest (Garwin 2002). At the Batu Hijau tonalite porphyry complex, at the core of the system, at
least three intrusive episodes occurred, and the intensity of mineralization and alteration
decreased with each subsequent intrusion (Clode et al. 1999).

Potassium alteration, at the core of the Batu Hijau system, is characterized by the replacement
of the hornblende mass in the host tonalites by biotite and magnetite. The surrounding
propylitic alteration can be divided into three subzones that, in order of increasing distance
outward, comprise (1) actinolite ± epidote ± chlorite; (2) epidote + chlorite; and (3) chlorite
(replacing mafic minerals) ± calcite ± albite (Garwin 2002; Clode et al. 1999). A transitional
chlorite-sericite (intermediate argillic zone) developed after the propylitic and potassic alteration
stages, overprinting the biotite (potassic) and actinolite zones (Idrus et al. 2009). Structurally
Teledetección II– Practice

controlled argillic and advanced argillic alteration is widespread and overprints earlier alteration
(Fig. 2; Garwin 2002; Clode et al. 1999; Idrus et al. 2009).

Figure2 Structurally controlled argillic and advanced argillic alteration is widespread and overprints earlier
alteration (Garwin 2002; Clode et al. 1999; Idrus et al. 2009).
Teledetección II– Practice

Methodology
Teledetección II– Practice

4. Results

4.1. Spectrally determined mineralogy

Spectral Assistant identified chlorite and epidote as the most abundant SWIR-active mineral
groups in the samples, with white mica and carbonates also making significant contributions.
(Figs. 4A, 5A, 5B). Chlorite, when identified, was almost always the primary mineral. Epidote was
common as the second most abundant SWIR-active mineral in the chlorite-dominated samples,
and in some as the primary mineral. The Halo results (Figs. 4B, 5C) were generally similar, but
identified clays (mainly smectites and vermiculites) as dominant in the biotite and actinolite
zones, with chlorite commonly attributed as a secondary mineral. Although the identification of
these phases must be treated with caution, their presence would be consistent with the later
argillic overprinting that has been well mapped in the proximal parts of the alteration system
(Fig. 2). The Halo also classified many spectra that the TSA could not, with most of them being
zeolite dominant, particularly in the very distal samples. The minerals contained in each group
are shown in Table 4.

Of the 90 samples, 46 produced spectra with the primary mineral identified by TSA as chlorite or
epidote (Fig. 5 D). Of this subset, 30 were attributed to contain primarily chlorite, 10 to contain
primarily epidote, and 6 to be unclear (probably containing approximately equal amounts of
each). Separate spectra collected in different areas of the same sample were very consistent (Fig.
6) indicating that the SWIR response is reproducible and varies more between samples (as a
function of spatial position and mineral abundance) than within samples.

Figure 3 Profiles of the spectra corrected by the hull ratio; This correction accentuates
the absorption features by eliminating the wide background variation

Figure 4 Halo results


Teledetección II– Practice

Figure 5 maps showing the special distribution of mineral occurrences (A,B,C)

Figure 6 Spectral Wizard identified chlorite and epidote as the most abundant SWIR-active mineral groups in the samples, with white mica and
carbonates also making significant contributions (5 A, B and C), of the 90 samples, 46 produced spectra with the primary mineral identified by TSA
as chlorite or epidote (5D)
Teledetección II– Practice

4.2. Chlorite

The chlorite chemistry of all Batu Hijau propylitic zones shows significant variation in terms of Fe
and Mg which have a strong inverse correlation (Fig. 7A) related to the well-established solid
solution between the iron and magnesium end members (Deer et al. 2009). The mass ratio
Mg/(Mg+Fe) (Mg#) ranges between 0.30 and 0.62, which translates into Mg/Fe molar ratios
between 0.78 and 0.51 with an average of 0.68± 0.01 (1σ). Comparison of the Mg# of chlorite
with the Mg# of the whole rock shows no correlation (Fig. 7B) suggesting a lack of protolith
control over chlorite composition. The Si content shows an inverse correlation with Al (Fig. 7C),
probably as a result of the Tschermak substitution: 𝐴𝑙3 ! + 𝐴𝑙3! = 𝑆𝑖4! + 𝑀2! (Deer et al. 2009). Ca
is also strongly correlated with Si and inversely with Al (Fig. 7D).

The chlorite-dominated spectra are characterized by two key absorption features centered
around 2250 nm and 2340 nm (Fig. 8), caused by the stretching of the Fe-OH and Mg-OH bond
Teledetección II– Practice

respectively (Herrmann et al. 2001). A third feature occurs around 2000 nm but is masked by a
large feature at 1910 nm, present in all samples, which is attributable to the presence of
molecular water. A feature around 1400 nm is also consistently present, caused by OH- (Clark et
al. 1990; Bishop et al. 2008), but it is not diagnostic of chlorite. In many samples, an absorption
feature around 2195 nm is probably the result of the presence of clay minerals containing Al-OH
groups (Clark et al. 1990; Herrmann et al. 2001). There is also a notable slope (hull) that
descends from ~1900 nm to ~1400 nm with variable gradient.

Within the spectra classified as dominant by chlorite, the most notable variation occurs in the
position and depth of the absorption features centered around 2250 nm and 2340 nm, which
are strongly coupled (Fig. 9). For absorption centered at 2250 nm, the exact wavelength position
of the feature minimum varies between 2255 nm and 2248 nm. For the absorption centered at
2340 nm, this varies between 2343 nm and 2328 nm. The maximum depth of both features
varies between 0.03 and 0.55 (fraction of the reflectance range).

The positions of these two features show a relatively strong inverse correlation with the Mg# of
chlorite in the samples (Fig. 10). Samples containing more Mg-rich chlorite correspond to
spectra in which the 2250 nm and 2340 nm features are shifted to lower wavelengths. Spectra
with absorption feature minima at lower wavelengths appear to be anomalous, with wavelength
values more in line with samples not classified as chlorite-dominated.

The depth of both features also shows a negative correlation with Mg#, which is considerably
stronger at depths below ~0.34 (Fig. 11). Data points with higher depth values appear to be
more in the range of depths shown by the epidote-dominated samples.

Changes in the hull slope are also observed between 1900 nm and 1400 nm, with the samples
richer in Fe having a steeper slope (Fig. 12). The mentioned features are illustrated in selected
spectra (Fig. 13).

The position of the OH absorption feature around 1400 nm also shows a systematic variation
that correlates quite well with the Al:Si ratio of the chlorite in the samples, with the Al-rich
chlorite producing spectra with the feature shifted to higher wavelengths (Fig. 14).

4.3. Epidote

The geochemistry of epidote shows the greatest variation with respect to Fe and Al (Table 3)
representing the solid solution between epidote and clinozoisite. Fe:Al molar ratios range
between 0.50 and 0.24, meaning that all samples (except one) fall into the epidote classification
(Franz & Liebscher 2004). Other major elements, including Ca, show remarkably little variation.

In the spectra dominated by epidote (Fig. 15) two absorptions are again observed around 2250
nm and 2340 nm, caused by Fe--OH bonds (Clark et al. 1990). However, the most diagnostic
feature occurs at ~1550 nm and the presence of epidote in a sample is most easily identified by
this. Once again, absorptions are observed at ~1400 nm and ~1910 nm.

The range of wavelength positions of the 2250 nm and 2340 nm absorption features (2251-2255
nm and 2336-2342 nm respectively) is significantly smaller than that in chlorite. The depth of
the 2340 nm feature is generally greater than in chlorite (mean = 0.3 vs. 0.2), but that of the
2250 nm absorption is approximately the same (Table 3). No significant variations in absorption
traits that correlate with chemical composition (e.g., Fe:Al ratio) were observed, however this
may be due, in part, to paucity of data.
Teledetección II– Practice

4.4. Spatial patterns in the spectral and chemical response of chlorite

A number of spectral and chemical characteristics of chlorite vary spatially, mainly related to the
distance from the Batu Hijau porphyry center and the Sekongkang prospect.

4.4.1. 2250 and 2340 absorption positions

The wavelength positions of the features at 2250 nm and 2340 nm are lowest near the center of
the Batu Hijau system, and show a systematic shift to longer wavelengths away from the center,
providing a strong vector to the reservoir (Fig . 16). Plotting the data as a function of radial
distance from the center reveals a general increase in wavelength positions from the center up
to ~1.2 km, beyond which values generally stabilize (Fig. 17), coinciding approximately with the
edge of the actinolite subzone. The 2340 nm feature also follows this trend in the northwest
with respect to the center of the Sekongkang porphyry (Fig. 16B).
Teledetección II– Practice

Figure 10 Mg# of chlorite plotted against wavelength position at absorption


Figure 9 Plots showing the coupling of absorption
minimum for: (A) 2250 nm absorption; and (B) the absorption at 2340 nm. Both
characteristics at 2250 nm and 2340 nm. (A) Wavelength
characteristics show a reasonably strong negative correlation. Outliers are
positions at absorption minima showing a positive
present at the lower wavelengths (shown in gray). Linear regression lines are
correlation. (B) Relative absorption depths showing a
shown (excluding outliers). Vertical error bars are 1σ standard error
positive correlation.
(propagated through ratio calculation). Horizontal error bars are ±0.5 nm,
representing the measurement accuracy of the spectrometer.

4.4.2. 2250 and 2340 absorption depth

The absorption depths of 2250 nm and 2340 nm show similar patterns, systematically
deepening away from the center of the porphyry (Fig. 18). The vector towards mineralization
remains strong, although less clear than in the case of wavelength positions. The Sekongkang
prospect is not distinguished by variation in feature depth, which is interesting as a potential
discriminator between well-mineralized and poorly mineralized systems. Unlike wavelength
positions, absorption depths continually deepen away from the center (Fig. 19), although the
values are unusually high around 1 km to 1.5 km.

4.4.3. Chemistry of the main elements

The Mg/Fe substitution of chlorite, reflected by Mg#, is probably the main control on absorption
position and possibly depth. This also shows changes
Teledetección II– Practice

Figure 11 Mg# of chlorite plotted against relative depth of


absorption features for: (A) 2250 nm absorption; and (B) the
absorption at 2340 nm. Both characteristics show a negative
correlation. Depths greater than ~0.34 are considered outliers
possibly due to epidote in the sample (shown in gray). Vertical
error bars are 1σ standard error (propagated through ratio
calculation). Horizontal error bars calculated from the
spectrometer signal-to-noise ratio are not shown because they
were determined to be negligible (±0.025%)

systematic, but slightly more complicated, from the center outwards (Fig. 20). Chlorite, within
about 1 km of the center, is enriched in Mg and then shows a rapid decrease in the Mg:Fe ratio
up to about 1.5 km; beyond this, the relative Mg content increases progressively up to the
sampling limit at about 4.5 km. Samples at distances less than 500 m from the center
(corresponding to tonalite and carbonate-hosted chlorite) are anomalous, as previously
identified in the trace element chemistry of chlorite (Wilkinson et al. 2015).

The Al:Si ratio in chlorite shows a fairly strong inverse correlation with distance, particularly at
distances greater than 3 km from the center (Fig. 21). However, this is not reflected in any
obvious spatial variation in the position of the 1400 nm feature (which is believed to be
correlated with Al/[Al+Si]). This may be due, in part, to the fact that no samples beyond 3 km
showed chlorite-dominated spectra.
Teledetección II– Practice

5.. Discussion

The results of this study support the effectiveness of reflectance spectroscopy in discriminating
alteration zones (e.g., Sun et al. 2001; Jones et al. 2005; Zadeh et al. 2014). In detail, the minerals
identified in the spectra are generally consistent with the alteration mineralogy documented
around Batu Hijau (Clode et al. 1999; Garwin 2002). The predominance of clay minerals in the
central biotite zone probably represents a sensitive response to traces of late-stage
sericite/paragonite and intermediate argillic alteration (Fig. 2; Garwin 2002). It is not surprising
that biotite was not identified in the spectra given its notable low reflectance (Cudahy et al.
2001) especially when found next to highly reflective smectites. However, vermiculite was
identified, which can form from the hydrothermal alteration of biotite. The presence of chlorite
in the biotite zone and its predominance (along with epidote) in the propylitic zone are
consistent with the expected propylitic alteration assemblages (e.g., Cooke et al. 2014a).
Actinolite, when present, generally occurs in proportions too low to be detected. The zeolite
minerals identified in the distal samples have been previously observed and attributed to a final
stage of low temperature alteration (Clode et al. 1999).

Chlorite geochemical variation has previously been recognized as an effective vector of


mineralization in the Batu Hijau system (Wilkinson et al. 2015). The major element geochemistry
derived from electron microprobe analysis coincides, as expected, with that previously reported
using LA-ICP-MS (Wilkinson et al. 2015) and shows that the ratios of Mg# and Al:Si in chlorite
vary spatially with respect to the reservoir. The lack of correlation between chlorite and whole-
rock Mg# suggests that rock composition is not a primary control of chlorite chemistry. An
increase in the Fe content of chlorite away from the porphyry center at a distance of ~1.5 km
has been proposed as a result of the advection and cooling of Fe-enriched hypersaline brines
(Wilkinson et al. 2015). Al:Si ratios in chlorite vary as a result of substitution at the tetrahedral
site (Deer et al. 2009) and are related to the temperature of the fluid (Cathelineau 1988). Ca
concentrations may be controlled by this reaction, and it is likely that other major 2+ ions, such
as Mg2+ and Fe2+, as well as trace elements such as Sr2+, are also involved.

These mineral chemical patterns are reflected in the SWIR spectral features, especially those
centered around 2250 nm and 2340 nm. It is well established that the Mg# of chlorite influences
the wavelength positions of these features, with more Fe-rich chlorites causing shifts to higher
wavelengths (Yang & Huntington 1996; Herrmann et al. 2001; Jones et al. 2005; Bishop et al.
2008), and this study confirms this relationship. The characteristics of this region are attributed
to the stretching and bending overtones of metal-OH bonds (Hunt 1977), which will be affected
by the mass and/or ionic radius of the cations involved.

Figure 15 SWIR spectrum of a sample composed


predominantly of epidote. Diagnostic features include deep
Fe-OH absorption features at ~2250 and ~2340 nm, and OH
Teledetección II– Practice

The apparent increase in absorption depths with decreasing Mg# in chlorite is not documented
in any other study and may represent a new finding, but caution must be taken with this
interpretation. It is important to consider whether or not Mg# is linked to increases in
absorption depths, or whether both factors are independently controlled by another variable
that shows the same spatial pattern. If Mg# directly controls the absorption depth, it is
postulated to be the result of higher mass Fe cations having a stronger effect on the stretching
and vibrations of the metal-OH bond.

An alternative explanation is that an increase in the abundance of chlorite in a sample causes an


increase in the depth of its unique absorption features. This is because in any mixed
(polymineral) spectrum the prominence of a trait attributable to a mineral depends directly on

Figure 17 Position of the wavelength at the minimum for (A) the 2250 nm
Figure 18 Spatial variation of (A) the depths of the 2250 nm absorption
absorption features; and (B) 2340 nm absorption features in chlorite
features; and (B) 2340 nm in chlorite. Larger red circles correspond to
plotted against lateral distance to the nearest mineralized center.
shallower absorption features. A fairly systematic decrease is observed
Systematic outward increases are observed up to a distance of about 1.6
towards the center of Batu Hijau along the southwestern transect for
km, and wavelengths decrease slightly beyond that. Green dots indicate
both absorption features. There is no clear trend towards Sekongkang
samples with distances measured away from the center of Sekongkang
Teledetección II– Practice

its proportion in the sample (Thompson et al. 1999; Herrmann et al. 2001). However, this
explanation would be in contradiction to studies that have found that chlorite abundance
decreases away from porphyry centers (Norman et al. 1991) instead of increasing. In the Batu
Hijau samples, the proportion of chlorite has not been precisely determined, so this possibility
cannot currently be tested.

The effect of mixed mineral assemblages on the SWIR spectra of rocks is perhaps the biggest
problem encountered in this study. Although both spectral analysis techniques employ spectral
unmixing at their core, their effectiveness is limited, especially when the number of contributing
minerals is large and when the minerals have overlapping characteristics. For example, the
chlorite-epidote-calcite assemblage, which is common in porphyry systems, will have
overlapping features in the 2340 nm region (Dalton et al. 2004). At Batu Hijau, calcite is relatively
rare and therefore presumably has little effect, however epidote is abundant, especially as a
secondary mineral in the chlorite-dominated spectra. This is likely to affect the two important
characteristics of chlorite at 2250 nm and 2340 nm.

Figure 20 Mg# chlorite plotted against lateral distance to nearest


mineralized center. The data show a sharp decline in Mg content up to
a distance of about 1.6 km from the center, followed by a steady
increase to what could be background levels. No significant
relationship is observed in samples measured far from the center of
Sekongkang (green). Vertical error bars are 1σ standard error.

Figure 19 Relative depths of (A) absorption features at 2250 nm;


and (B) absorption features at 2340 nm in chlorite plotted against
lateral distance to the nearest mineralized center. A general
deepening is observed moving away from the center of Batu HIjau,
with slightly elevated values around 1.3 km to 1.6 km. Note that
the same trend is not observed in the Sekongkang survey, indicating
its possible usefulness as a discrimination tool.

In terms of wavelength positions, the narrower range of wavelength positions occupied by


Figure 21. Al/(Al+Si) content of chlorite as a function of lateral distance to
epidote-related absorptions (Table 3) could create a biasAinsystematic
the mineral. the chlorite spectra
decrease toward
is observed withmore
distance to Batu Hijau
that is not evident in the Sekongkang data. Vertical error bars are 1σ est
error
Teledetección II– Practice

central positions in such samples. However, the fact that chlorite Mg# is correlated with feature
positions (Fig. 10) with the same trend observed in previous studies on epidote-free samples
(Yang & Huntington 1996; Herrmann et al. 2001; Jones et al. 2005; Bishop et al. 2008) indicates
that any interference from epidote is limited. In terms of depth, the presence of epidote should
deepen the 2340 nm absorption feature, which could explain the anomalous absorption depth
values in the plots demonstrating correlations between absorption depth and chemistry (Fig.
11). Given this, the absorption feature at 2250 nm is likely to be a better indicator of chlorite
composition than that at 2340 nm. This point is also made in Herrmann et al. (2001) because the
secondary Al-OH feature in white mica also overlaps with the chlorite feature at 2340 nm.
However, use of the 2250 nm feature may be limited in some deposits because it has been
shown to disappear completely where there is widespread weathering (Suryantini 2003).

Two other spectral features were shown to vary with chlorite chemistry: the hull slope between
1400 nm and 1900 nm, and the position of the 1400 nm absorption feature. The increase in hull
slope in the more Fe-rich chlorite has been previously recognized (Thompson et al. 1999) and is
attributable to the strong and broad absorption caused by electronic effects (charge transfer) in
Fe2+ in the VNIR (Hunt 1977; Clark et al. 1990). However, this characteristic occurs in many Fe-
containing minerals, so it may simply reflect an abundance of these in the rock, rather than a
specific response to chlorite, thus explaining the lack of strong systematic spatial variation. The
position of the 1400 nm wavelength appears to be related to the Al:Si ratio in chlorite. Changes
in this absorption have been previously observed in chlorite (King and Clark 1989) and are
probably due to the modification of the OH vibration frequencies as a result of the Tschermak
substitution (Duke 1994). Unlike hull slope, this feature appears to offer considerable promise
for use in exploration using portable or core-based SWIR spectrometers. However, both
features are inapplicable to satellite remote sensing due to strong atmospheric scattering effects
at these wavelengths (Duke 1994).

This study has been unable to conclude anything about possible mineral chemical controls on
epidote spectral variations due to a clear lack of overlapping data: of the 16 samples that could
be classified as epidote dominated, only five had geochemical data of accompaniment. This is
unfortunate because epidote has shown promise as a vectoring tool in porphyry deposits
(Cooke et al. 2014b) and the spectral variation across the epidote-clinozoisite solid series is
certainly recognizable (Roache et al. 2011). This is an area where future study would be useful.

6. Possible application to satellite remote sensing

The results of this study indicate that porphyry deposits could be located using absorption
characteristics at 2250 nm and 2340 nm, which differ between the Mg-rich chlorite that may be
dominant in the interior propylitic zone (Wilkinson et al. 2015) and Fe-rich chlorite which may
predominate further away, or which may develop above buried porphyry systems (Halley et al.
2015). Precise wavelength positions can only be exploited, for the moment, using field
spectrometers or hyperspectral sensors capable of collecting high-resolution spectra. Therefore,
it should be possible to locate porphyry deposits using hyperspectral imaging, and previous
work has demonstrated discrimination of Fe chlorite from Mg chlorite in hyperspectral data
(Cudahy et al. 2001). However, the small wavelength shifts documented here are currently
impossible to localize using the wide bandwidths (0.02-0.2 µm) typical of satellite remote
sensing.

Fortunately, the relative change in the depth of absorption features is potentially recognizable
using broadband multispectral images collected by the ASTER (Advanced Thermal Emission
Radiometer) satellite sensor. ASTER Bands 7 and 8 measure surface reflectance in the ranges
Teledetección II– Practice

2235-2285 nm and 2295-2365 nm, which individually encompass the two absorption features of
interest.

Deeper absorption features should produce a darker response (lower reflectance) and ASTER
band ratio images could perhaps be used to enhance these features.

7.Conclusions

SWIR reflectance spectroscopy has proven to be a powerful tool for characterizing propylitic
alteration associated with porphyry reservoirs. The collection of high-resolution spectra using
portable spectrometers such as the TerraSpec 4, combined with comparison algorithms such as
those used by TSG and Halo, allows rapid identification of spectrally active minerals, even in the
case of low-grade propylitic "green rocks". reflectance. From the spectra, chlorite and epidote
were found to be the most abundant spectrally active minerals in the Batu Hijau rocks, and they
were most abundant in the nearest propylitic zone. Spectral signatures of the biotite zone
commonly indicate the presence of smectites and other phyllosilicates attributable to the later
stage, to the intermediate argillic and sericitic overprint.

Chlorite chemistry can be used as a powerful vector for the mineral (Wilkinson et al. 2015) and
this appears to translate into features in the SWIR spectra of the chlorite-dominated samples.
The Batu Hijau chlorite varies significantly in terms of Mg#, which is probably the main cause of
the changes in the wavelength position of the absorption features around 2250 nm and 2340
nm, and possibly the cause of the variation in the depth of these features. The positions and
depths of both features act as good vectors for mineralization: the wavelength positions shift
from 2254 to 2249 nm and from 2343 to 2332 nm, and the absorption depths decrease from
about 35% to 5% respectively. , moving away from ~1.6 km to 500 m from the site. This
correlates with a general increase in chlorite Mg# from approximately 1.6 km towards the
margins of mineralization. Variations in the depth of absorption features, in particular, appear to
be a new finding and one potentially more applicable to remote sensing than the well-
established changes in position, since changes in absorption depth can potentially be
recognized by the satellite remote sensing. The exact cause of the changes in absorption depth
is not entirely clear, but the fact that this can act as a vector of mineralization is significant,
especially if these observations can be recognized in other porphyry systems.

The extent to which these results apply to other porphyry deposits remains unclear.
Furthermore, there is currently little information to suggest that variations in the geochemistry
of alteration minerals are specific to fertile (well-mineralized) rather than weakly mineralized or
sterile systems. However, we note that the position of the 2250 nm peak and the depths of the
2340 nm and 2250 nm absorption features do not appear to highlight the weakly mineralized
Sekongkang prospect in the Batu Hijau district. Future work could include investigation of
chlorite spectral variation in other porphyry systems of variable mineralization, and more
detailed investigation of the effects of epidote. We conclude that SWIR spectroscopy holds
great promise as a practical exploration tool in the “green rock” environment.

Conclusions

 IR Spectroscopy is a powerful, non-destructive technique for identifying the presence of


minerals containing H 2 O, -OH and Fe-O in rocks.

 HALO comes with a world-class spectral library, which greatly improves certainty in
mineral identification.
Teledetección II– Practice

 This technique is particularly suitable for mapping the extent of alteration zones in
economic mineral systems.

 HALO comes with a world-class spectral library, which greatly improves certainty in
mineral identification.

 In the Batu Hijau porphyry Cu deposit, Indonesia, Neal et al. (2018) determined that
chlorite absorption features near 2250 and 2340 nm correlate with Mg# (Mg/[Mg+Fe])
with respect to the wavelength position and depth of the features. As Mg# increases,
the minimum absorption wavelengths at 2249 and 2332 nm were observed to increase
to 2254 and 2343 nm, respectively. Reflectance spectroscopy is useful for studying
chlorite minerals in zones of propylitic alteration according to Neal et al. (2018).
Teledetección II– Practice

Bibliography

 Agar B. and Coulter D. (2007) Remote sensing for mineral exploration – A decade
perspective 1997 2007. Pp 109 136 in: Exploration in the New Millennium, Proceedings
of the Fifth Decennial International Conference on Mineral Exploration (B. Milkereit,
editor).
 ASD Inc. (2013) ASD TerraSpec Mineral Analyzer opens new uranium exploration
potential in the Athabasca Basin. https://cdn2.hubspot.net/hub/45853/file-33386196-
pdf/docs/uravan_case_study.pdf (Accessed May 2, 2018).
 Caiza K. (2018). Estimation of hydrothermal alteration zones through interpretation of
ASTER-type satellite images and use of the terraspec equipment in the eastern area of
Cerro de Pasco, Peru. Central University of Ecuador.
 Clark, R.N. (2004) Spectroscopy of rocks and minerals, and
 principles of spectroscopy. pp. 17 55 in: Infrared Spectroscopy in Geochemistry,
Exploration Geochemistry, and Remote Sensing (PL King, M.S. Ramsey, and G.A.
Swayze, editors). Mineralogical Association of Canada, Short Course, 33.
 Hauff, P.L. (2005) Applied Reflectance Spectroscopy (users' manual; version 4.1).
Spectral International, Inc., Arvada, Colorado, USA.
 Hunt, G.R. and Ashley, R.P. (1979) Spectra of altered rocks in the visible and near
infrared. Economic Geology, 74, 1613-1629.
 Kerr, A., Rafuse, H., Sparkes, G., Hinchey, J., and Sandeman, H. (2011) Visible/infrared
spectroscopy (VIRS) as a research tool in economic geology: Background and pilot
studies from Newfoundland and Labrador. Current Research, Newfoundland and
Labrador Department of Natural Resources, Geological Survey Report 11-1, 145 166.
 Laakso, K., Peter, J.M., Rivard, B., and White, H.P. (2016) Short-wave infrared spectral
and geochemical characteristics of hydrothermal alteration at the Archean Izok Lake
Zn- Cu-Pb-Ag volcanogenic massive sulfide deposit, Nunavut, Canada: Application in
exploration target vectoring. Economic Geology, 111, 1223 1239.
 Lyon, RJP and Burns, E.A. (1963) Analysis of rocks and minerals by reflected infrared
radiation. Economic Geology, 58, 274-284.
 Neal, L.C., Wilkinson, J.J., Mason, P.J., and Chang, Z. (2018) Spectral characteristics of
propylitic alteration minerals as a vectoring tool for porphyry copper deposits. Journal
of Geochemical Exploration, 184, 179 198.
 Vargas G.; Ríos A.; Huamán D. (2010). Mapping of Hydrothermal Alteration Minerals
using the ASTER multispectral image in the Pallcamachay Project – Ancash, Peru. XV
Peruvian Congress of Geology.
 Percival, JB, Olejarz, AD, English, MLR, Belley, PM, Flynn, T., Laudadio, AB, and Stirling,
JAR (2016a) Spectral Library: The Kodama Clay Collection. Geological Survey of Canada
Open File 7923. http://dx.doi.org/ 10.4095/297564.
 Vargas P. Introduction to Reflectance Spectroscopy (VNIR –SWIR). Reflex Halo.
 Sparkes G. and Sandeman H. (2011) Visible/infrared spectroscopy (VIRS) as a research
tool in economic geology; background and pilot studies from New Foundland and
Labrador.
 Thompson, A.J.B., Hauff, P.L., and Robitaille, A.J. (1999) Alteration mapping in
exploration: application of short-wave infrared (SWIR) spectroscopy. Society of
Economic Geologists Newsletter, No. 39, 1, 16 27.
 Yang, K., Huntington, J.F., Gemmell, J.B., and Scott, K.M. (2011) Variations in
composition and abundance of white mica in the hydrothermal alteration system at
Teledetección II– Practice

Hellyer, Tasmania, as revealed by infrared reflectance spectroscopy. Journal of


Geochemical Exploration, 108, 143 156.
 Zhang, G., Wasyliuk, K., and Pan, Y. (2001) The characterization and quantitative
analysis of clay minerals in the Athabasca Basin, Saskatchewan: Application of
shortwave infrared reflectance spectroscopy. The Canadian Mineralogist, 39, 1347—
1363.

Alunite
Teledetección II– Practice

You might also like