Sei sulla pagina 1di 12

American Mineralogist, Volume 86, pages 473–484, 2001

Cathodoluminescence study of apatite crystals

JOCELYN BARBARAND1,* AND MAURICE PAGEL2


1
CRPG-CNRS, B.P. 20, 54501 Vandoeuvre-lès-Nancy Cedex, France
2
Département des Sciences de la Terre, UMR Orsayterre, Université Paris-XI, 91405 Orsay Cedex, France

ABSTRACT
Cathodoluminescence (CL) spectrometry represents a promising technique for the analysis of
trace-element concentrations and distributions in minerals. However, a higher precision and a stan-
dardization of the recording conditions are required to use CL spectral data quantitatively. A signifi-
cant step towards a more quantitative treatment of CL spectra is presented in this study.
A procedure to correct the spectra for the various efficiencies as a function of the wavelength of
the CL detector is proposed using low-pressure mercury-vapor and quartz-iodine lamps. CL spectra
presented in this study are thus corrected for the system response. Apatite CL spectra, which are
commonly composed of two broad bands centered at 3.5 and 2.2 eV, are deconvoluted to isolate
component bands and determine their areas. The crystallographic control by prismatic or basal sec-
tions of apatite on spectral intensities is significant and only prismatic sections should be used.
Signal decrease associated with electron bombardment (electron beam aging) is exponential and
appears drastic in the first hundred seconds but continues even after 15 minutes of beam bombard-
ment.
All observed CL bands could be correlated with a specific activator [rare earth elements (REE)
or manganese]. The 3.5 eV band is composed of three bands at 3.59 eV, 3.29 eV, and 2.87 eV. Ion
microprobe results and comparison between CL and photoluminescence data support Ce3+ activa-
tion for the origin of these bands.
The relationship between CL band intensity and REE concentration measured by ion microprobe
analysis demonstrates that CL also can provide semi-quantitative data for Gd3+, Ce3+, Dy3+, and Sm3+
when recording conditions are strictly controlled.

INTRODUCTION numerous overlapping bands. Less-intense bands are not ob-


Cathodoluminescence spectroscopy has been used for a long served while the true intensity of obvious bands is only esti-
time in petrographic studies to determine the provenance of mated. (3) Variation in luminescence can be linked with
mineral grains or to distinguish generations of growth within crystallographic controls or electron beam aging (Marshall
minerals (Marshall 1988 and references therein; Machel et al. 1988; Murray and Oreskes 1997).
1991). Recently, CL detectors have been attached to scanning The purpose of this study is to provide methodological and
electron microprobes allowing a better control of the electron analytical data leading to an improvement in the precision of
beam flux and general instrument stability, and permitting the acquisition of CL spectra, and to a standardization of the
higher magnification and spectrometric measurements in the mechanism of recording. A procedure using reference lamps is
UV and visible ranges. presented to correct for transmission variations with wave-
However, CL spectrometry is still not used as a microprobe lengths of the detector. CL spectra are deconvoluted to indi-
to reveal the presence of activators in the minerals or their con- vidualize most of the bands present and to obtain quantitative
centrations because of several limitations. (1) Spectra are data on these bands (intensity and area). Importance of elec-
strongly dependent of the recording system used. This depen- tron beam aging and the variation of the luminescence con-
dence is related to the variation of efficiency as a function of trolled by crystallographic orientations are investigated for some
wavelength of the diffraction gratings and photomultipliers. apatite samples. To test the use of CL spectrometry as a micro-
This system response implies modifications of the spectra in probe, data acquired from CL spectra on various REE activa-
intensity and shape of the bands between facilities (Townsend tors are compared with compositions determined by ion
and Rowlands 2000). (2) As an example, the luminescence of microprobe analysis. Finally, the results of a CL study of apa-
apatite under electron bombardment is caused by the presence tite crystals are presented to reveal chemical composition and
of Mn and rare earth elements (REE) that substitute for Ca associated growing conditions.
(Mariano 1988; Blanc et al. 1995). Spectra are composed of
INSTRUMENTATION

Methodology
* Current address: Department of Geological Sciences, Uni-
versity College London, Gower Street, London WC1E 6BT, CL images and spectra were acquired with an Oxford In-
U.K. E-mail: j.barbarand@ucl.ac.uk strument MonoCL2 system attached to a Philips XL30 scan-

0003-004X/01/0004–473$05.00 473
474 BARBARAND AND PAGEL: CL STUDY OF APATITE CRYSTALS

ning electron microscope (SEM). The MonoCL optical con- the microscope. Analytical conditions were 15–20 kV with an
figuration has been described by Kearsley and Wright (1988). electron beam current between 200 and 400 mA. Photomicro-
The recording system consists of the following: a fully retract- graphs were taken with 1600 ISO films.
able parabolic aluminum mirror acting as light collector, placed Apatite crystals were arranged in rows and columns under
immediately above the sample surface; a silica window that a binocular microscope allowing easy location of grains. The
allows the passage of UV emissions; a scanning grating mono- samples were then mounted in epoxy resin, ground using 400-
chromator, with 1200 grooves per millimeter; a PA3 amplifier; and 600-grade silicon carbide papers, polished using 3 and 1
and a photomultiplier tube (PMT) as detector. The grating µm diamond pastes, and coated with carbon. Apatite crystals
monochromator is characterized by high wavelength resolu- sampled from different localities and geological environments
tion and linear scattering resulting, in turn, in an important loss were chosen for this CL study. The samples are described in
of intensity (Rémond et al. 1992). The PA3 amplifier provides Appendix 1.
a complete set of support electronics including power supplies Major-element analyses of apatite were obtained using a
and signal amplification. Photon pulses (photon counting mode) Cameca SX 50 microprobe and the PAP data-reduction pack-
are counted by the preamplifier. This arrangement provides a age. Operating conditions were 15 kV, a beam current of 10
better signal to noise ratio than the measure of average total nA, and a beam diameter of 5 µm. Four spectrometers are at-
current (linear mode). tached to the microprobe: three are equipped with PET, TAP,
The mirror system is retractable to simplify the use of the and LIF crystals, respectively, for the analysis of Cl-K-Ca-Ba,
SEM for secondary electron (SE) and back-scattered electron Na-Mg-Si-P-Sr, and Mn-Fe, whereas the fourth has a PC1 crys-
(BSE) imaging. The mirror is mounted symmetrically about tal for the analysis of F. Natural or synthetic standards were
the electron optic axis allowing collection of up to 80% of used to determine the element abundance: albite (Na), Mg ox-
emitted photons. The minimum magnification for the system ide (Mg), orthoclase (Si, K), hematite (Fe), Sr oxide (Sr), Mn
is constrained by the aperture size of the mirror and is approxi- oxide (Mn), synthetic chlorapatite (P, Ca, Cl), and topaz (F).
mately 100×. Counting time was 30 seconds for F, 20 for Mg, Si, Cl, Sr, Mn,
The spectral coverage is 200–840 nanometers (nm) and the and Fe, and 10 for Na, P, K, and Ca. Three replicate analyses
spectral dispersion is 2.7 nm per millimeter of slit width. The were averaged for each grain analyzed. Analyses in which the
width of the entrance and exit slits is usually 1 mm and, there- sum of the different oxides was lower than 97% were rejected.
fore, the spectral dispersion is 2.7 nm for the majority of the Trace-element abundances were determined with a Cameca
spectra presented. Reducing the slits allows a better resolution IMS 3f ion microprobe on gold-coated sections. The samples
but produces an important decrease in the measured intensity. were sputtered using a 10 mm diameter primary O– beam rang-
CL spectra (X-axis) can be expressed in wavelength (nm) ing in intensity from 2 to 5 nA. Positive secondary ions of REE
or in energy (electron volt, eV) by the relation: were analyzed at low mass resolution (M/DM = 400) and with
an energy filtering of –60 ± 10V. The ionic intensities, mea-
hc
E(eV) = or E(eV) = 1239.85 sured for each isotope, were normalized to that of 43Ca. Oxide
λ(nm) λ(nm)
interferences remaining after energy filtering were eliminated
where h is the Planck constant and c the speed of light. The using deconvolution techniques applied to the measurements
intensity (Y-axis) reports individual photon pulses. of 31 atomic masses in the range 138 to 180 (Fahey et al. 1987).
For this study, CL spectra were acquired at an accelerating The absolute atomic ratios normalized to Ca were then obtained
voltage of 25 kV, a 1–5 nA beam current, and entrance and exit from the deconvoluted ratios and the ion yields of each REE.
slit widths of 1 mm. The analyzed zone corresponds to a rect- These ion yields were determined from the analysis of the
angle of 40 × 30 µm (beam rastered at 3000×). Counting times of Durango standard apatite (Young et al. 1969; Reed 1986).
500 to 1000 milliseconds (ms) were chosen to avoid noisy spec- Photoluminescence (PL) spectra were measured on some
tra. The analytical procedure is a sequential acquisition and there- apatite crystals to study their UV luminescence and compare
fore the recording time varies between 7 and 15 minutes depending these results with those of CL. Luminescence spectra were in-
on the counting time and the analytical step. All spectra were re- vestigated with a UV laser (337 nm) that delivered pulses of
corded from the lower to the higher wavelengths. 10 ns duration and 0.1 cm–1 spectral width. The pulse energy
The photomultiplier voltage can be monitored by the soft- was maintained at about 10 mJ. The luminescence was observed
ware interface. The response curve of the PMT has been deter- with an incident angle of 90°, analyzed by INSTASPEC equip-
mined for a low-pressure mercury vapor lamp. The linear ment enabling time-resolving spectra acquisition and detected
response domain represents a small interval between 79 and by an intensified CCD matrix. Delay times and strobe pulse
84% of high voltage (HV) of the PMT. The same analysis was duration are between 20 ns and 9 ms; spectral detection ranges
carried out for an apatite crystal (Balazuc drill-core, Ardèche, between 350–900 nm. The system is combined with an optical
France) and the linear domain was found to be between 82 and microscope, enabling spectra acquisition at the microscopic
87% HV PMT. All the spectra presented in this study were there- scale. Band assignment was done from the data of Tarashchan
fore acquired within that range. (1978).
Panchromatic images were also obtained with an optical
CL Technosyn Mark II microscope. The components include a Correction of the CL spectra
small vacuum chamber with windows, an X-Y stage movement, Figure 1 illustrates the necessity of correcting the spectra
and a cold cathode electron gun centered on the optical axis of for the system response by comparing spectra of the same
BARBARAND AND PAGEL: CL STUDY OF APATITE CRYSTALS 475

The calibration of wavelength and intensity was achieved


with reference lamps whose emission spectra are known. These
lamps work well with this type of SEM because the vacuum
chamber door can be opened wide, allowing the passage of
light. The grating monochromator is calibrated with a low-pres-
sure mercury vapor lamp with well-defined emission lines at
546.073, 576.960, and 579.066 nm. The monochromator can
be driven mechanically to the appropriate positions, which en-
ables precise acquisition of data. The calibration of the appara-
tus can thus be easily controlled.
To cover the entire range of the spectrum, two lamps whose
spectra are broad bands with no line are needed. For the UV
range (200–400 nm), a deuterium lamp whose spectrum corre-
sponds to a broad band between 160 and 380 nm was used
(Fig. 2a). For the visible and near IR portions of the spectrum
(400–750 nm), a quartz-iodine lamp was selected (Fig. 2b).
For both lamps, the system response (Fig. 2c and 2d) corre-
sponds to the intensity of the reference spectrum divided by
the intensity of the measured spectrum. Correction of a CL spec-
FIGURE 1. CL spectra of Durango apatite acquired by sev- trum requires multiplication of the raw spectrum by the system
eral authors with different recording systems: (a) Roeder et al. response. All spectra presented in this study are corrected for
1987; (b) Murray and Oreskes 1997; (c) and (d) this study. the system response.
Note that the Y-axis scale is not the same for the different spec-
tra. Observation with a system efficient in the UV range (c) Deconvolution of the CL spectra
shows that the maximum of intensity is reached for Ce3+ and Using calcite, El Ali (1989) has demonstrated that CL spec-
not for Eu2+ as indicated by (a). Analytical conditions: (a) Nu- tra can be deconvoluted into gaussian curves (Kopp et al. 1995;
clide Luminoscope—spectra were recorded with a NM-3 mono- Habermann et al. 1996; Townsend and Rowlands 2000). For
chromator system including an SC-1 scanning control, a D-43 deconvolution of broad emission spectra into component bands,
extended spectral range (250–930 nm) side-looking photomul- luminescence spectra were plotted as a function of energy. This
tiplier detector, and model DR-1 digital radiometer; (b) Patco conversion needs the transposition of the Y-axis by a factor λ2/hc
ELM3R Luminoscope—spectra were recorded with a (Townsend and Rowlands 2000). The intensity is then expressed
SPISUARC spectrometer (Acton Research Corporation) in arbitrary units of photon counts.
mounted with a Princeton Instruments CCD camera and an Deconvolution was made with a least-squares fitting algo-
Olympus microscope (10 kV, 0.5 mA, 45 seconds exposure rithm that minimizes the difference between the experimental
time); (c) and (d) MonoCL2, 25 kV, 1.3 nA, magnification of spectrum and the sum of the gaussian curves. IGOR software
1000, slits width of 1 mm, 85% of high voltage of the photo- (Interactive Graphic Operation for Research) from Wavemetrics
multiplier tube (HV PMT), exposure time of 500 ms per nm. Inc. was used for deconvolution. Based on presumed band num-
bers and wavelengths, iterative calculations give the band po-
sitions that correspond to the best fit between the spectrum and
the sum of calculated bands. The usual procedure is to start
with a one or two bands and to increase the band numbers until
sample acquired with different CL devices. Differences among the deconvolution does not significantly improve. Band posi-
the three spectra presented in Figure 1 are associated with the tions can be constrained or assumed for calculation.
CL apparatus efficiency as a function of wavelength. For scan- Deconvolution of the CL spectra can reveal different acti-
ning gratings, the maximum efficiency is controlled by the vators and establish their relative importance, particularly where
“blaze” applied during manufacture of the grating (Townsend some activators are overlapped by others with broader or more
and Rowlands 2000). In the same way, PMT are strongly wave- intense emission bands. It permits precise measurement of the
length dependent. Spectral response of the PMT used in this band parameters, such as the area or the width at half height, a
study displays maximum efficiency at 260 nm and is reduced first step to quantification. Examples of deconvolution of CL
by a factor of 10 at 700 nm. Consequently, luminescence of a spectra are presented in Figure 3.
given CL apparatus represents biased information about the Deconvolution must be carried out with care and the fol-
original luminescence of the sample. The recorded signal dif- lowing limitations must be taken into account. (1) Define the
fers greatly from that of the original emission. Determination level at which no additional band needs to be added. (2)
of the system response of the CL detector is essential to correct Deconvolution of 10 bands in one spectrum represents the limit
spectra for discrepancies due to the method of acquisition so beyond which the results given by the algorithm are not valid.
that they may be compared quantitatively among laboratories. (3) Deconvolution is difficult in the 2.5–4 eV range of the spec-
Parts of the spectrum from a given CL apparatus where record- trum where most of the activators are unknown. (4) For a simi-
ing efficiency is weak could not, however, be compared. lar correlation between the sum of the gaussian and the
476 BARBARAND AND PAGEL: CL STUDY OF APATITE CRYSTALS

FIGURE 2. CL spectra of ref-


erence lamps and correction
curve for the MonoCL2 installa-
tion at the University of Nancy I.
(a) reference spectrum supplied
by the manufacturer (1), and
spectrum acquired with the
MonoCL2 (2) for the deuterium
lamp (CL conditions = 83% HV
PMT, slit widths of 1 mm); (b)
reference spectrum (Mazères
1997) (3) and spectrum acquired
with the MonoCL2 (4) for the
quartz-iodine lamp (CL condi-
tions = 83% HV PMT, slit widths
of 1 mm); (c) system response of
the MonoCL2 obtained with the
two previous lamps; and (d) de-
tail of system response for the
range 300–700 nm.

spectrum, two bands may lead to an ambiguous interpretation.


The addition of a third band refines the two possibilities (Fig.
4). (5) Finally, it appears difficult to superimpose two bands or
to discriminate bands whose energies are very close. An ex-
ample is the Sm3+ band at 2.19 eV, that cannot be isolated by
deconvolution (other than by inference with other Sm3+ bands
that are present at 2.07 and 1.89 eV) because of the presence of
the Mn2+ band at 2.18 eV. These limitations illustrate that not
all bands present in a CL spectrum may be assigned to a spe-
cific activator by deconvolution into gaussian curves.

CONTROLS ON CL EMISSION

Importance of the recording conditions


Instrumental conditions are important in the recording of
CL emission. Variation of beam intensity strongly modify the
CL spectra, the relative intensity, and the ratio between the dif-
ferent CL bands. Figure 5 shows spectra from an apatite grain
from the 1245 m Balazuc sandstone determined using differ-
ent beam currents. For a current <1 nA, only the band at 3.97
eV can be seen, whereas for a current >3 nA, the broad band
centered at 3.29 eV appears. A smaller band around 2 eV can
also be discriminated (2.07 and 1.91 eV). Operating conditions
must therefore be specified precisely for any CL spectra.

Activator assignment
Band assignment was based principally on data from syn-
FIGURE 3. Deconvolution treatment of two CL spectra. The thetic chlorapatite crystals doped with various REE, Mn2+, and
bold curve corresponds to the sum of the gaussian curves. (a) U4+ (Blanc et al. 1995). These data are in good agreement with
Arendal apatite spectrum characterized by REE and Mn bands; previous work (Mariano and Ring 1975; Roeder et al. 1987;
(b) Ødegårdens Verk apatite spectrum with a high Ce band but Mariano 1988). CL band energies associated with the different
lacking the Mn band. activators presented in this study are the mean values of the
BARBARAND AND PAGEL: CL STUDY OF APATITE CRYSTALS 477

bands from numerous analyses carried out on the various


samples. The error for these energies is the standard deviation
of the mean when more than five spectra were used.
Two broad bands are usually present in the apatite crystal
spectra—one centered at 3.5 eV and the other at 2.2 eV—in
addition to several narrower bands that correspond to REE ac-
tivators. Trivalent rare earth ions have the electron configura-
tion 4f(k)[5s25p6], with the exceptions of La, Ce, Gd, and Lu.
The 5s25p6 electrons form two completely filled shells, which
partially shield the 4f subshell. These electron configurations
give rise to characteristic narrow luminescence bands through
an f → f transition (Yang et al. 1995). The interaction between
the 4f electron and the crystal field is weak and, therefore, the
emission band wavelengths vary little among different miner-
als (Rémond et al. 1992). Luminescence emission can also take
place in the 5d electron shell that interacts strongly with the
crystal field. The transition f → d yields a broad band (La, Ce,
Lu). The bands of Gd3+ (3.97 ± 0.006 eV), Dy3+ (bands at 2.57
± 0.006, 2.15 ± 0.005, and 1.86 ± 0.012 eV), Er3+ (3.08 eV),
Eu2+ (2.75 eV), Sm3+ (2.19 ± 0.006, 2.07 ± 0.003, and 1.91 ±
0.003 eV), and Tb3+ (3.25, 2.98, 2.83 ± 0.005, 2.53, and 2.27 ±
0.006 eV) can be identified by deconvolution of the CL spectra.
An Mn2+ band (2.15 ± 0.02 eV) is present as a broad band
that overlaps with the Dy3+, Sm3+, and Tb3+ bands. Mn2+ is a
transition metal ion that has an electron configuration of 3d5
and interacts strongly with the crystal field (d → d transition).

FIGURE 4. Deconvolution steps of a CL spectra. The


deconvolution into one or two bands does not provide a good
fit between the experimental spectra and the sum of the gaussian
peaks. Introduction of two peaks gives two contrasting solu-
tions with differing degrees of confidence. The addition of a
third peak markedly improves the result and leads to a repro-
ducible unique solution.

▼ FIGURE 5. The same apatite crystal from the Balazuc


sample analyzed at different beam currents: (a) 0.3 nA, (b) 1.3
nA, and (c) 4.9 nA. The intensity of bands increases with the
current density, and the band ratios (for example, Gd/Er) evolve
for the different analytical conditions (CL conditions = 25 kV,
slit widths of 1 mm).
478 BARBARAND AND PAGEL: CL STUDY OF APATITE CRYSTALS

The center and width of the Mn2+ band are very sensitive to length of 337 nm) has its optimal output near 400 nm. Spectra
their nearest neighbor atom surroundings in the given crystals acquired with lasers having a lower excitation wavelength (266
(Mariano 1988; Yang et al. 1995). The difference between the nm) are characterized by Ce3+ bands at 365 and 382 nm
position of Mn2+ band in this study and that proposed by Blanc (Tarashchan 1978). These positions are similar to those found
et al. (1995), 2.19 eV, may be explained by the fact that the by CL. Combination of CL and PL data allows intrinsic lumi-
spectra presented here are expressed in eV as opposed to nm in nescence for the UV band activation to be eliminated because
Blanc et al. work. This shifts the maximum of broad band to photon energy is insufficient to cause near band gap radiation.
lower energy as spectra are multiplied by K/wavelength (where These results argue in favor of Ce3+ activation for UV emission
K = 1239.85). This shift is found only for broad bands because in CL spectra.
their maximum is defined by several points. Introduction of Ce3+ ions in the apatite structure is claimed
After deconvolution, the broad band centered at 3.5 eV is to produce vacancies or defects, and therefore to be associated
composed of three bands at 3.59 ± 0.08 eV, 3.29 ± 0.06 eV, and with intrinsic luminescence (Barbin and Schvoerer 1997). This
2.87 ± 0.09 eV, respectively. This band is often the most in- hypothesis is not retained here because the introduction of La3+
tense spectral band, and its maximum intensity could be as- and Pr3+ which occur in the same Ca site [Ca(2); Hughes et al.
sociated with either one of the bands at 3.59 or 3.29 eV. The widths 1991], should result in high UV emission as well (Blanc et al.
at half-height of these two bands range between 0.3 and 0.6 eV. 1995).
The origin of these three bands (3.59, 3.29, and 2.87 eV) is The intensities or area ratios between the two bands at 3.3
still controversial and two causes are usually proposed: intrin- and 3.6 eV are, however, variable among the samples, and this
sic luminescence or Ce3+ activation (Barbin and Schvoerer variation is not easily associated with a single activation.
1997). Although Tb3+ and Tm3+ have bands in this range, their
activation can be dismissed because their intensity and width Luminescence variations
characteristics do not match the observed spectra. Crystallographic control on the CL signal. To study the
The term “intrinsic luminescence” is used improperly crystallographic dependence of CL in apatite crystals, CL spec-
(Bresse et al. 1995), but refers in this study to emission in rela- tra were acquired for various samples (Appendix 1) on pris-
tionship to lattice defects such as charge recombination cen- matic and basal sections or on prismatic sections but with a
ters, defects originally present in the crystal, or defects induced 90° rotation of the mount before new acquisition. These spec-
by electron bombardment. This type of luminescence has been tra were obtained from different zones to avoid electron beam
proposed for both silicate and carbonate minerals. Iacconi and aging but from the same sector as revealed by spectroscopic
Caruba (1980) showed that in silicate minerals, intrinsic lumi- images.
nescence forms a broad band localized in the near UV or vio- A centimeter-long apatite crystal from Durango, Mexico,
let-blue; in carbonates, Barbin and Schvoerer (1997) considered was cut perpendicular and parallel to the c-axis. Numerous
that blue or near UV luminescence is associated, at least partly, spectra were acquired for each section using the same analyti-
with lattice point defects because no correlation had been es- cal conditions (Fig. 6a and 6b). The shapes of the spectra are
tablished between an activator (Pb2+, Eu2+, or Ce3+) and this the same for both crystallographic orientations. The major dif-
luminescence. ference in the spectra is their intensity, with the mean intensity
Ce3+ activation has three bands at 3.54, 3.26, and 2.71 eV for the basal section lower than that for the prismatic face.
(Blanc et al. 1995). Ce3+ is characterized by an f → f transition For sample Gun-F (Gunlock, Utah), spectra were acquired
and also by an f → d transition located in the near UV (Cesbron before and after a 90° rotation of the slide. Normal orientation
et al. 1993). This implies that the emission bands are broad and acquisitions do not display significant difference (Fig. 6c).
are, therefore, compatible with the 3.59 ± 0.08, 3.29 ± 0.06, However, another area from the same sector of the crystal ana-
and 2.87 ± 0.09 eV bands. lyzed parallel to the c-axis (prismatic section) yields a strongly
This study suggests that these three bands are related to Ce3+ different spectrum.
activation. The three Ce3+ bands have been revealed by spec- For a Lestrezes (Margeride, France) granite apatite, normal
tral deconvolution. The relationship between the Ce3+ concen- rotation of the crystal led to different results (Fig. 6d). Spectra
tration and the 3.6–3.3 eV band areas is discussed below (see are composed mainly of the Ce3+ and Mn2+ bands. The 90° ro-
“Semi-quantitative determination of some REE contents with tation results in an inversion of the relative intensities of the
CL spectrometry” and Figs. 8c and 8d). This relationship dis- two bands: the Ce3+ band is significantly higher for the zone
plays quite a good correlation indicating that UV activation is analyzed in a section perpendicular to the c-axis but the Mn2+
controlled by the presence and concentration of Ce3+ ions. band is significantly higher for the zone analyzed in a section
Samples with a very low Ce3+ concentration appear to be char- parallel to the c-axis (Fig. 6d).
acterized by a weak UV emission. These results show that in a perspective of quantitative
Comparison of CL and PL spectra shows that when UV measurement of the activator content, crystallographic orien-
emission is high in the CL spectra it is also high in the PL tation must be carefully controlled. Analysis should be done
spectra. PL spectra are time-resolved, allowing discrimination on prismatic sections for maximum efficiency. Differences
of the different activators. UV emission is associated mainly between prismatic and basal sections may be explained by the
with Ce3+ and also with Eu2+, each being characterized by short diffusion mechanism of elements along the c-axis, as already
decay times. The Ce3+ band wavelength is shifted to a higher demonstrated for electron microprobe analysis for F and Cl
value (395–400 nm) because the laser used (excitation wave- (Stormer et al. 1993). Diffusion rate appears higher along the
BARBARAND AND PAGEL: CL STUDY OF APATITE CRYSTALS 479

FIGURE 6. The effect of crystal orientation on the CL spectra. Prismatic (a) and basal (b) sections of a Durango apatite crystal
(CL conditions = 25 kV, 1.3 nA, 50 × 50 µm analyzing zone, 85% HV PMT); (c) normal acquisition of three spectra from a Gun-
F apatite crystal. Spectrum 2 was acquired after a 90º rotation of the slide. Spectrum 3 represents another zone in the same sector
as 2; and (d) normal acquisition of a Lestrezes apatite crystal. Spectrum 2 was acquired after a 90º rotation of the slide. (CL
conditions = 25 kV, 4,9 nA, slits width of 1mm, 83% HV PMT, and acquisition time of 1000 ms).

c-axis, and could imply an important element loss recorded by were done with a voltage of 25 kV, a beam current of 4.9 nA.
lower intensity for spectra obtained on basal sections (Fig. 6a Sample location and SEM focusing were done in different parts
and b). Acquisition following a 90° rotation gives various results, of the mount with a beam current of 0.1 nA to avoid exposing
with significant changes in the spectra in some cases (Fig. 6d). the studied zone to the beam. Acquisition times of 500 to 1200
Electron beam aging. CL emission displays time depen- seconds were selected to examine signal variations for the rou-
dent changes in luminescent color and intensity. The decrease tine analysis (7–15 minutes).
of luminescence in some minerals such as quartz is referred as Luminescence fades rapidly during electron beam bombard-
electron beam aging (Marshall 1988). Possible mechanisms for ment (Fig. 7). Fading is usually composed of three parts: (1) a
the intensity decrease in synthetic phosphors include precipi- rapid decrease in the first 100 seconds, corresponding to an
tation of metal from the host and destruction of luminescence intensity reduction of 40 to 80% depending on the activator;
centers either by heating or oxygen loss (Marshall 1988). This (2) a much smaller loss intensity follows (100–400 seconds),
phenomenon of fading is the consequence of electronic bom- and (3) a period with little change (400–1200). Luminescence
bardment (rise in temperature, creation of an electric field). decrease seems to reach a plateau after 800–1000 seconds. Af-
Ramseyer et al. (1989), however, demonstrated for quartz, that ter 500 seconds, intensity is reduced to 50–85%. Consequently,
thermal effects are not responsible for alteration of the lumi- the first 100 seconds are critical for luminescence decrease.
nescence. The evolution of intensity with time is not the same for the
To examine this phenomenon in apatite, time intensity plots different activators. Mn2+ band intensity displays a drastic fall
were determined for different wavelengths of three activators and then a very small decrease. Dy3+ and Ce3+ bands intensities
Ce3+ (3.59 and 3.29 eV), Dy3+ (2.57 eV) and Mn2+ (2.15 eV), in decrease more gradually.
the Gun-F, Fish Canyon, and Arendal apatites. Measurements Rapid decrease of luminescence has been demonstrated in
480 BARBARAND AND PAGEL: CL STUDY OF APATITE CRYSTALS

FIGURE 7. Time vs. intensity plots for different activators. (a) Ce3+ band (at 3.59 eV) in a Gun F apatite crystal; (b) Ce3+ band
(at 3.59 eV) in a Fish Canyon apatite crystal; (c) Dy3+ band (at 2.57 eV) in an Arendal apatite crystal; (d) Mn2+ band (at 2.15 eV)
in a Gun F apatite crystal. The 2.15 eV band represents exactly the Mn2+ luminescence because the Arendal grain displays very
small bands for Dy3+ (2.15 eV) and Sm3+ (2.19–2.07 eV). For the three bands, CL signal fades rapidly during electron bombard-
ment (CL conditions = 25 kV, 4.9 nA, 83% HV PMT, and slit width of 1 mm).

α-quartz (Ramseyer et al. 1988), and the decrease along an concentrations, also determined by ion microprobe analysis,
exponential decay curve has been reported for apatite crystals have been investigated further for apatite grains recovered from
(Murray and Oreskes 1997). Under their operating conditions sandstones from a drill hole in the South of the Paris Basin and
(10 kV, 0.5 mA), aging was observed for more than 45 minutes for apatite crystals from several different settings (Appendix
of exposure. Time of exposure to the electron beam is also a 1). CL spectra were acquired first, followed by ion microprobe
critical factor during electron microprobe analysis (Stormer et analyses on the same areas of each apatite crystal. CL spectra
al. 1993). were obtained for Paris Basin samples only in the range 2.5–4
These experiments indicate that measured CL spectra may eV to study Gd3+, Ce3+, and Dy3+; spectra from the other samples
correspond to compromised information. CL bands may dis- cover the full range. Band area was calculated from the gaussian
play a reduced intensity after a specific period of time by elec- curve.
tron bombardment, rather than the real intensity. Great care is For Gd3+, the correlation between CL and concentration is
thus required when comparing different spectra. quite good if grains with more than 150 ppm Gd3+ are excluded
(Fig. 8a). Gd3+ determination by CL spectrometry is, however,
SEMI-QUANTITATIVE DETERMINATION OF SOME REE difficult because the broader Ce3+ band tends to mask the Gd3+
CONTENTS WITH CL SPECTROMETRY band when its intensity is great. The detection limit for Gd in
Correlation between the areas beneath Gd3+ and Dy3+ CL samples with such a high Ce3+ band is, therefore, poorer.
bands and concentration, as determined by ion microprobe The correlation for Dy3+ is more problematic because the
analysis, has been reported for apatite crystals recovered from 2.15 eV and 2.57 eV bands overlap respectively the Mn2+ band
a North Sea basin sandstone (Barbarand et al. 1996). In the and the Er3+ and Tb3+ bands. Deconvolution needs absolutely
present study, the relationships between CL band areas and REE to be carried on for the 2–2.7 eV range of the spectrum to char-
BARBARAND AND PAGEL: CL STUDY OF APATITE CRYSTALS 481

FIGURE 8. Relationships between element concentration, as determined by ion microprobe analysis, and band area for vari-
ous activators in apatite crystals of different settings (solid circles) and detrital apatite grains from Paris Basin sandstones (open
circles). (a) Gd3+ at 3.97 eV (correlation coefficent, r = 0.76); (b) Ce3+ at 3.59 eV (r = 0.75); (c) Ce3+ at 3.29 eV (r = 0.39); (d) Dy3+
at 2.57 eV (r = 0.74); (e) Sm3+ at 2.07 eV (r= 0.82), and (f) Sm3+ at 1.91 eV (r = 0.88). (CL conditions = 25 kV, 4,9 nA, slit widths
of 1 mm).

acterize the distinctive area of each band (Fig. 8b). ferent analytical conditions (such as variation in the vacuum or
The Ce3+ content varies widely among the analyzed grains in the sample’s position respect to the mirror). This difference
so that a better comparison can be made between concentra- is recorded also in the minimum amount of Ce3+ necessary to
tion and CL band area. For the 3.59 eV band, the relationship reveal the presence of the Ce3+ band: 400 ppm for the Paris
appears linear up to 1000 ppm but is less clear for higher con- Basin samples, but only a few parts per million for the others.
centrations (Fig. 8c). For the 3.29 eV band, the relationship is For Sm3+, the relation appears essentially linear for the two
linear for the whole range of concentration if the Paris Basin bands (2.07 and 1.91 eV) (Fig. 8e and 8f). The 1.91 band is
apatites, or those from the different settings are considered sepa- outside the Mn2+ band but the 2.07 band is inside its influence
rately (Fig. 8d). This difference could be associated with dif- and the fact that the same result was obtained for both bands
482 BARBARAND AND PAGEL: CL STUDY OF APATITE CRYSTALS

proves the validity of the deconvolution procedures. located on the southeast border of the French Massif Central
The relationships between the CL band area and activator near the Uzer fault (Pagel et al. 1996). Apatite grains were sepa-
content appear very poor for grains that have higher REE con- rated from Upper Triassic sandstone at 1245 meters depth,
tents and, especially higher amounts of the HREE (which rep- mounted in resin, and then polished. When viewed with an
resent the majority of CL activators). HREE may act as optical microscope, the apatite grains appear homogeneous with
quenchers for other REE when a certain concentration thresh- well-developed euhedral faces. These features are surprising
old is crossed. All energy is not transmitted integrally to radia- because, for most sandstones, the majority of detrital grains
tive energy, and the CL band intensity appears lower than might should be rounded. Secondary electron and backscattered elec-
be expected from the activator content. Such an hypothesis has tron images do not show any crystal zoning, although they do
been already proposed by Townsend and Rowlands (2000). reveal an array of small cavities (probably fluid inclusions)
CL spectrometry detection limits are therefore very low with (Figs. 9b and 9c). Panchromatic and monochromatic CL im-
the analytical conditions used, less than 50 ppm for Gd3+, Dy3+, ages, on the other hand, do show that apatite grains are com-
Tb3+, or Sm3+, and roughly 400 ppm for Ce3+. These values are posed of detrital rounded cores and euhedral overgrowths (Figs.
much lower than electron microprobe or BSE analyses. This 9a and 9d). Some grains are cut by fractures filled with bright
limit is, however, dependent on the chemical environment: ac- pink diagenetic apatite. Our CL study shows that these two
curacy is higher for Tb3+ at 2.27 eV or Dy3+ at 2.15 eV when portions of composite grains have different compositions. The
the Mn2+ band (2.15 eV) is absent from CL spectra. yellow cores are dominated by the Mn2+ band with only very
small bands for Gd3+, Ce3+, Dy3+, and Sm3+ (Fig. 10). Spectra of
USE OF CL SPECTROSCOPY AND SPECTROMETRY FOR the bright pink overgrowths are free of the Mn2+ band and have
THE STUDY OF APATITE CRYSTALS higher concentrations of Gd3+, Ce3+, Eu2+, and Sm3+. Color
The association of CL spectroscopy and spectrometry is a changes in the grains are thus linked to activation by Mn2+ (yel-
powerful petrographic tool to differentiate multiple fluid events low at 2.15 eV) in the cores and by Eu2+ (pink at 2.75 eV) at the
as well as to characterize elemental chemistry. An example is rims.
for apatite grains recovered from the Balazuc BA1 drill hole These results suggest that the diagenetic fluid that formed

FIGURE 9. Different images of a single apatite grain from a Balazuc Triassic sandstone. The images represent (a) Technosyn,
(b) SE, (c) BSE, and (d) CL-SEM at 350 nm.
BARBARAND AND PAGEL: CL STUDY OF APATITE CRYSTALS 483

FIGURE 10. CL spectra in a detrital core and an overgrowth in apatite grains from a Balazuc BA1 sandstone (CL conditions
= 25 kV, 1.3 nA, analyzed zone of 20 × 20 µm).

the apatite rims was enriched in REE compared with the origi- d’apatite et histoire thermique de la bordure sud-est du Massif Central, 492 p.
Ph.D. dissertation, Université Henri Poincaré, Nancy.
nal fluid from which the cores crystallized and also had a lower Barbarand, J., Pagel, M., Blanc, Ph., Braun, J.-J., Chaussidon, M., Vetil, C., and
concentration of Mn2+. The mobility of REE during diagenesis Walgenwitz, F. (1996) Combined cathodoluminescence spectra, fission-track
in the Balazuc Triassic sandstone has also been shown by the age and chemistry of detrital apatite grains: towards a better understanding of
fission-track annealing. International conference on cathodoluminescence and
presence of REE in diagenetic anhydrite crystals (Blanc et al. related techniques in geosciences and geomaterials, September 2–4, Nancy,
1994). The low Mn2+ content indicates a reducing environment, France. Book of Abstracts, 19–20.
Blanc, Ph., Arbey, F., Cros, P., Cesbron, F., and Ohnenstetter, D. (1994) Applica-
an interpretation supported by the absence of U in the apatite tions de la microscopie électronique à balayage et de la cathodoluminescence à
overgrowth as evidenced by fission track mapping. des matériaux géologiques (sulfates, carbonates, silicates). Bulletin de la Société
Correction for the system response of the detector, Géologique de France, 165, 4, 341–352.
Blanc, Ph., Baumer, A., Cesbron, F., and Ohnenstetter, D. (1995) Les activateurs de
deconvolution of the spectra into component bands, assignment cathodoluminescence dans des chlorapatites préparées par synthèse
of all activators, and control of the recording conditions (i.e., hydrothermale. Compte Rendu de l’Académie des Sciences, Paris, 321, IIa,
analysis on prismatic sections with the shortest exposure to the 1119–1126.
Bresse, J.F., Remond, G., and Akamatsu, B. (1995) Cathodoluminescence micros-
beam and using the same voltage and intensity), allow CL spec- copy and spectroscopy of semiconductors and wide bandgap insulating materi-
trometry to obtain semi-quantitative data on some elements and als. 4th European workshop on modern developments and applications in mi-
crobeam analysis, May 14-19, St Malo, France, Book of Abstracts, 213–254.
their distributions. CL spectra associated with CL spectroscopy Cesbron, F., Ohnenstetter, D., Blanc, Ph., Rouer, O., and Sichere, M.-C. (1993)
can be used to recognize and characterize heterogeneities in Incorporation de terres rares dans des zircons de synthèse: étude par
apatite crystals. This approach is a highly useful way to test cathodoluminescence. Compte Rendu de l’Académie des Sciences, Paris, 316,
IIa, 1231–1238.
the sample for homogeneity and to reveal different generations El Ali, A. (1989) Etude des roches carbonatées de réservoirs d’hydrocarbures par
of materials, which is particularly valuable in deciding on apa- résonance paramagnétique électronique et cathodoluminescence, 190 p. Ph.D.
tite standards for use in chemical or geochemical purposes. This dissertation, Université de Paris VII, Paris.
Fahey, A.J., Goswami, J.N., McKeegan, K.D., and Zinner, E. (1987) 26Al, 244Pu,
fast and inexpensive technique should be used routinely in geo- 50
Ti, REE, and trace elements abundances in hibonite grains from CM and CV
chronological (e.g., U/Pb for example) or thermochronometric meteorites. Geochimica and Cosmochimica Acta, 51, 329–350.
Habermann, D., Neuser, R.D., and Richter, D.K. (1996) REE-activated
(fission-track analysis) studies to properly characterize the cathodoluminescence of calcite and dolomite: high-resolution spectrometric
material. analysis of Cl emission (HRS-CL). Sedimentary Geology, 101, 1–7.
Hughes, J.M., Cameron, M., and Mariano, A.N. (1991) Rare-earth-element order-
ACKNOWLEDGMENTS ing and structural variations in natural rare-earth-bearing apatites. American
Mineralogist, 76, 1165–1173.
This work was undertaken with the support of Elf and the GPF program. Iacconi, P. and Caruba, R. (1980) Trapping and emission centres in X-irradiated
J.B. was supported by a grant from the Ministère de l’Enseignement et de la zircon (III): influence of trivalent rare earth impurities. Physical State Solide a,
Recherche and by an Elf contract. The authors thank Philippe Blanc (Université 62, 589–596.
Paris VI) for the introduction to CL, Marc Chaussidon (CRPG-CNRS) for the Kearsley, A. and Wright, P. (1988) Geological applications of scanning
ion microprobe analysis, Gérard Panczer (LCPML, Université Lyon I) for the cathodoluminescence imagery. Microscopy and Analysis, September, 49–51.
photoluminescence spectra and Jeanne Paquette and Darryl Henry for construc- Kopp, O.C., Fuller, E.L., and Owen, M.R. (1995) Interpretation of cathodo-lumi-
tive comments on the manuscript and Walter E. Trzcienski for his editorial work. nescence spectra obtained from dolomite and calcite gangue minerals, and
dolostone breccias in the central Tennessee Zinc District (U.S.A.). Scanning
REFERENCES CITED Microscopy Supplement 9, 211–223.
Machel, H.G., Mason, R.A., Mariano, A.N., and Mucci, A. (1991) Causes and emis-
Barbin, V. and Schvoerer, M. (1997) Cathodoluminescence et géosciences. Compte sion of luminescence in calcite and dolomite. In C.E. Barker and O.C. Kopp,
Rendu de l’Académie des Sciences, Paris, 325, IIa, 157–169. Eds., Luminescence microscopy: quantitative and qualitative aspects, SEPM,
Barbarand, J. (1999) Cinétique de cicatrisation des traces de fission dans les cristaux 9–25.
484 BARBARAND AND PAGEL: CL STUDY OF APATITE CRYSTALS

Mariano, A.N. (1988) Some further geological applications of cathodoluminescence. geology. Scanning Microscopy, 9, 1, 43–62.
In D.J. Marshall, Eds., Cathodoluminescence of geological materials, p. 94– Young, E.J., Myers, A.T., Munson, E.L., and Conklin, N.M. (1969) Mineralogy and
123. Unwin Hyman, Boston. geochemistry of fluorapatite from Cerro de Mercado, Durango, Mexico. U.S.
Mariano, A.N. and Ring, P.J. (1975) Europium-activated cathodoluminescence in Geological Survey Professional Paper, 650-D, D84–D93.
minerals. Geochimica et Cosmochimica Acta, 39, 649–660.
Marshall, D.J. (1988) Cathodoluminescence of geological materials. Unwin Hyman,
Boston, 146 p. MANUSCRIPT RECEIVED DECEMBER 21, 1999
Mazères, S. (1997) Mise en oeuvre d’un microspectrofluorimètre pour l’étude de MANUSCRIPT ACCEPTED DECEMBER 23, 2000
microéchantillons en fluorescence stationnaire et résolue dans le temps, 200 p. MANUSCRIPT HANDLED BY WALTER E. TRZCIENSKI JR.
Ph.D. dissertation, Université Paul Sabatier, Toulouse.
Murray, J.R. and Oreskes, N. (1997) Uses and limitations of cathodoluminescence
in the study of apatite paragenesis. Economic Geology, 92, 368–376.
Pagel, M., Braun, J.-J., Disnar, J.R., Martinez, L., Renac, C., and Vasseur, G. (1996)
Thermal history constraints from combined organic matter, clay minerals, fluid
inclusions, apatite fission tracks and stable isotopes studies at the Ardeche
palaeomargin (BA1 drill hole, GPF program). Journal of Sedimentary Research, APPENDIX 1: BRIEF DESCRIPTIONS OF APATITE
67, 235–245.
Ramseyer, K., Baumann, J., Matter, A., and Mullis, J. (1988) Cathodoluminescence SAMPLES
colours of α-quartz. Mineralogical Magazine, 52, 669–677. 1. Durango apatite, Cerro de Mercado, Mexico (Young et
Ramseyer, K., Fischer, J., Matter, A., Eberhardt, P., and Geiss, J. (1989) A
cathodoluminescence microscope for low intensity luminescence. Journal of al. 1969), a fission-track age determination standard, is reported
Sedimentary Petrology, 59, 619–622. to be generated by hydrothermal reworking of a primary mag-
Reed, S.J.B. (1986) Ion microprobe determination of rare earth elements in acces-
sory minerals. Mineralogical Magazine, 50, 3–15.
netite-apatite igneous body.
Rémond, G., Cesbron, F., Chapoulie, R., Ohnenstetter, D., Roques-Carmes, C., and 2. Fish Canyon apatite, San Juan Mountains, Colorado
Schvoerer, M. (1992) Cathodoluminescence applied to the microcharacterization (Steven et al. 1967), is a fission-track age determination stan-
of mineral materials: a present status in experimentation and interpretation.
Scanning Microscopy, 6, 1, 23–68. dard from a welded tuff.
Roeder, P.L., MacArthur, D., Ma, X.P., Palmer, G.R., and Mariano, A.N. (1987) 3. Apatite crystals—Arendal, Bamble, Ødegårdens Verk,
Cathodoluminescence and microprobe study of rare earth elements in apatite. Skaaland Mine (Norway), Faraday (Ontario, Canada), and Gun-F
American Mineralogist, 72, 801–811.
Steven, T.A., Mehnert, H.H., and Obradovich, J.D. (1967) Age of volcanic activity (Utah)—of different environments and characterized by various
in the San Juan Mountains, Colorado. United State Geological Survey Profes- Cl/F ratio. A complete description is given in Barbarand (1999).
sional Papers 575-D, D47–D55.
Stormer, J.C., Pierson, M.L., and Tacker, R.C. (1993) Variation of F and Cl X-ray
Samples courtesy of A. Hurford and R. Siddall.
intensity due to anisotropic diffusion in apatite during electron microprobe analy- 4. Lestrezes apatites, recovered from a sample of the
sis. American Mineralogist, 78, 641–648. Margeride granite from the French Variscan orogen.
Tarashchan, A.N. (1978) Luminescence of minerals. Kiev, Naukova dumka.
Townsend, P.D. and Rowlands, A.P. (2000) Information encoded in cathodo-lumi- 5. Paris Basin apatites, separated from three Triassic
nescence emission spectra. In M. Pagel, V. Barbin, Ph. Blanc, and D. (Keuper) sandstones from the same borehole located in the
Ohnenstetter, Eds., Cathodoluminescence in Geoscience, p. 41–57. Springer, South of the Paris Basin (France), and sampled respectively at
Berlin.
Yang, C., Homman, N.P.-O., Malmqvist, K.G., Johansson, L., Halden, N.M., and 2104 (BP1), 2110 (BP2), and 2104m (BP3). Samples courtesy
Barbin, V. (1995) Ionoluminescence: a new tool for nuclear microprobes in of F. Walgenwitz (Elf).

Potrebbero piacerti anche