You are on page 1of 11

Chemical Geology 207 (2004) 135 – 145

www.elsevier.com/locate/chemgeo

Apatite fission-track chronometry using laser ablation ICP-MS


Noriko Hasebe a,*, Jocelyn Barbarand b, Kym Jarvis c,
Andrew Carter a, Anthony J. Hurford a
a
Research School of Earth Sciences, University College and Birkbeck College, London WC1E 6BT, UK
b
Department des Sciences de la Terre, Université Paris Sud, France
c
NERC ICP-MS Facility, Kingston University, Surrey KT1 2EE, UK
Received 27 August 2003; accepted 13 January 2004

Abstract

Comparison is made of apatite fission-track (FT) ages measured by using conventional thermal neutron irradiation and Laser
Ablation-Inductively Coupled Plasma-Mass Spectrometry (LA-ICP-MS) to determine sample uranium content. Spontaneous
track densities were determined on specified apatite grains from the Durango, Fish Canyon Tuff, and Mount Dromedary
Complex age standards. Neutron-induced track densities and ICP-MS uranium values were determined for the same specified
grains. Ages determined using the two approaches show reasonable consistency, indicating the possibility of routine fission-
track analysis using LA-ICP-MS. Uncertainties and assumptions in the LA-ICP-MS approach include the laser ablation of
chemically etched mineral surfaces, the suitability of reference materials for calibration, and the derivation of FT ages using the
alternatives of selection of a kf value or a revised zeta comparative approach. The use of LA-ICP-MS in FT chronometry would
dramatically increase the speed of analysis and hence sample throughput, as well as avoiding the need for neutron irradiation
and the production, transport and handling of radioactive samples. With the acknowledged dependency of FT annealing in
apatite on composition, such grain-by-grain, multi-element analysis could provide a more robust definition of the specific track
annealing characteristics of each apatite crystal.
D 2004 Elsevier B.V. All rights reserved.

Keywords: Apatite; Fission-track chronometry; Laser ablation

1. Introduction variety of geological applications (see, e.g. Green et


al., 1989; Gallagher et al., 1994; Fitzgerald et al.,
1.1. Uranium measurement in conventional FT 1995; Carter, 1999; Kohn and Green, 2002). Conven-
method tionally in the FT method, uranium concentration is
determined using a neutron activation technique to
Fission-track (FT) chronometry is widely used to induce fission in a proportion of 235U atoms, record-
establish both age and thermal history in a wide ing the fission events in an external detector to give
essentially a map of uranium distribution (e.g.
Gleadow, 1981; Hurford and Green, 1982). A major
* Corresponding author. Present address: Institute of Nature and strength of this approach is that data are collected
Environment Technology, Kakuma-machi, Kanazawa, Ishikawa 920 from identical areas on individual grains and their
1192, Japan. mirror-images in the detector, particularly important

0009-2541/$ - see front matter D 2004 Elsevier B.V. All rights reserved.
doi:10.1016/j.chemgeo.2004.01.007
136 N. Hasebe et al. / Chemical Geology 207 (2004) 135–145

when analysing sedimentary samples where variable amounts of solid samples in many fields (e.g. envi-
provenance and/or mineral chemistry (including ura- ronmental, geological, and archeological sciences).
nium content) may result in grains having different The methodology does not require special sample
ages. However, this methodology contains a number preparation such as sample dissolution and carbon
of uncertainties, including the absolute value of the film coating, and is therefore time efficient. In the
238
U spontaneous fission decay constant (kf) (Bigazzi, earth sciences a wide range of elements including rare-
1981), the efficiency of irradiation in terms of induc- earths are routinely measured on a single-grain basis
ing fission of 235U (Green and Hurford, 1984), and the (e.g. Jarvis and Williams, 1993; Odegard, 1999) to
geometry factor which relates to registration of in- study, for example, mineral growth process through
duced fission tracks on an external detector (Gleadow metamorphism (e.g. Rubatto, 2002) and the fractional
and Lovering, 1977). To overcome these issues, the coefficient of target element (e.g. Ertel et al., 2001).
FT community adopted a system calibration based on Initially, the use of ICP-MS methods for U – Pb geo-
the comparison of sample measurement with the chronology required zircon with uranium contents
analysis of standards whose ages are known—the zeta generally >50 Ag/g (e.g. Hirata, 1997; Horn et al.,
calibration (Fleischer and Hart, 1972; Hurford, 1990; 2000). The method has now matured such that it offers
Hurford and Green, 1983). Using the zeta approach spatial resolutions down to f 10 Am, and for U – Pb
and common age standards has resulted in a unified applications, permits rapid analyses ( f 30 ages/h)
system of calibration which has been highly effective with precision equivalent to that of the ion probe,
over the last 12 years and has produced a wealth of provided a f 40– 60 Am diameter ablated pit is used.
useful data in diverse geosciences applications. Importantly, high precision in situ isotopic analysis is
Whilst zeta circumvents many of the technical now possible with laser-ablation multi-collector ICP-
uncertainties, the quality of sample irradiation remains MS for U – Th isotopic compositions (e.g. Stirling et
a fundamental limitation on the quality of data. The al., 2000; Horn et al., 2000). Whereas zircon has been
number of research reactors able to deliver suitable the main focus of LA-ICP-MS geochronological stud-
levels of neutron thermalisation for FT analysis (Hur- ies, apatite has seen comparatively little research (e.g.
ford, 1990; Tagami and Nishimura, 1992) is limited Kimura et al., 2000; Gleadow et al., 2002). One area
and recently environmental, safety and political pres- that has been briefly mentioned is the viability of using
sures have led to a number of suitable reactor facilities LA-ICP-MS methods for uranium measurement for FT
closing. A consequence is that FT samples are having analysis (e.g. conference abstracts by Cox et al., 2000;
to be sent across the globe for irradiation with the Svojtka and Kosler, 2002). The aim of this study is to
ensuing real problems of loss, damage, disturbance build on these preliminary proposals and provide a
through customs inspection, longer turn-round times, more complete assessment of the suitability of LA-
transport of modestly active samples and lack of direct ICP-MS methods for uranium mapping in apatite, to
collaboration with reactor staff. With the surviving determine whether it is a viable alternative to the
research reactors getting older and a low probability neutron irradiation methods currently used in FT
of replacement (especially with positions that give analysis.
suitable thermal/epithermal flux ratios), alternative
approaches to uranium determination merit serious
exploration. 2. Experimental details

1.2. Alternative method for uranium measurement 2.1. Conventional FT method

This study considers the suitability of Laser Abla- Aliquots of the apatite FT age standards from
tion-Inductively Coupled Plasma-Mass Spectrometry Durango, Fish Canyon Tuff and Mount Dromedary
(LA-ICP-MS) as a method for direct measurement of Complex (e.g. Hurford, 1990; Green, 1985) were
uranium concentrations for determination of single- mounted in resin, polished and the spontaneous tracks
grain fission-track ages. LA-ICP-MS has been used etched using standard techniques (Hurford et al.,
widely to measure the chemical composition of small 1999), before irradiation in close contact with low-
N. Hasebe et al. / Chemical Geology 207 (2004) 135–145 137

uranium mica detectors in thermal neutron facility Table 2


4VGR of the reactor PLUTO, Harwell, UK (Cd ratio ICP-MS and laser operating conditions
for Au > 500). Tracks from the induced fission of 235U ICP-MS: instrument operating conditions
Instrument Thermo Elemental
were recorded in the external mica detectors. Glass
PlasmaQuad PQ2+ STE
dosimeters Corning CN5 ( f 12.2 Ag/g natural urani- Forward power 1.5 kW
um, Bellmans et al., 1995) or NIST 612 (37 Ag/g Reflected power <5 W
depleted uranium, Pearce et al., 1997) together with Plasma gas argon
detectors were included to monitor each irradiation. Coolant gas flow rate 14 l min 1
Carrier gas flow rate 0.975 l min 1
After irradiation, the mica external detectors from age
Auxiliary gas flow rate 1.5 l min 1
standards and dosimeters were etched in 48% HF acid
at 20 jC for 30 min and the track densities counted ICP-MS: data acquisition parameters
using a Zeiss Axioplan optical microscope with a Dwell time 100 As
100  dry objective and 1250x total magnification. Channels 20
Sweeps 100
Ten to fifteen grains were analysed for each apatite
Data acquisition time 5s
standard, the grain location and specific area counted Data acquisition mode scanning
being recorded on photomicrographs. Table 1 provides Isotopes measured 44
Ca, 232Th and 238
U
a summary of the FT analytical data.
Laser: operating conditions
Laser type CETAC LSX100 Nd:YAG
2.2. LA-ICP-MS
Wavelength 266 nm
Laser mode Q switched
Uranium measurements were carried out using Laser output energy 0.422 mJ
LA-ICP-MS on exactly the same grains of apatite Shot repetition rate 10 Hz
standards used in the FT analyses to permit direct Sampling scheme Zig-zag raster area f 50  50 Am
comparison of results. Analysis was made at the NERC
ICP-MS facility, Kingston University, using a Cetac
LSX-100 (wavelength 266 nm) laser with an energy of FT and LA-ICP-MS data from equivalent areas. ICP-
0.422 mJ measured as energy output from the laser MS and laser ablation operating parameters are given in
head. Scan speed was 10 Am/s with firing repetition rate Table 2. Ablated material was introduced into a VG
of 10 Hz and an ˜ 10A beam diameter. The analysis was Elemental PlasmaQuad 2+ STE, and 238U concentra-
carried out using the raster mode, i.e. ablation took tions were calibrated against NIST 610 and NIST 612
place in a zig-zag pattern. The space between lines of standard glasses (Pearce et al., 1997). Standards were
ablation was set to be 25 Am to avoid overlap of two measured every f 30 min using the same procedure as
adjacent trenches. For each grain the total ablated area for apatite to monitor the stability of the signal, together
was broadly equivalent to that counted for fission with the gas blanks. In total six datasets were provided
tracks (e.g. 50  50 Am), allowing direct comparison for a whole measurements. One dataset contains three

Table 1
Fission-track apatite ages calculated by conventional neutron irradiation and LA-ICP-MS methods
Apatite No. of qs, Ns qi, Ni Dosimeter qd, Nd P(v2), tconv, Ma tlaser, Ma U, ppm
sample grains 106 cm 2 106 cm 2 glass 106 cm 2 % ( F 1r) ( F 1r)
DUR 11 0.18 323 1.79 3153 CN5 1.65 1144 80 26.8 F 1.9 29.7 F 1.9 14.9
FCT 14 0.24 145 1.05 626 SRM612 0.943 1045 53 33.7 F 3.5 30.6 F 2.7 19.7
MDC 12 0.97 839 3.47 2990 SRM612 2.45 1356 0.02 106 F 6 94.6 F 4.5 28.4
DUR = Durango; FCT = Fish Canyon Tuff; MDC = Mount Dromedary Complex.
Ns,i,d = numbers of tracks counted for spontaneous, induced and dosimeter measurements; qs,i,d = densities of spontaneous, induced and
dosimeter tracks cm 2, tconv = fission track pooled  age calculated from pooled Ns and Ni for all grains counted using zeta values shown in
Table 2; P(v2) = probability of v2 for (n  1) degrees of freedom (Galbraith, 1981); U = average uranium content of all grains measured by LA-
ICP-MS; tlaser = weighted mean of grain ages calculated from individual Ns and 238U content measured by LA-ICP-MS.
138 N. Hasebe et al. / Chemical Geology 207 (2004) 135–145

measurements for each standard glass. Data reduction fission produces an etchable, observable track; and (b)
was carried out in Microsoft Excel. 44Ca was used as an the distance travelled by the fission fragment (Laslett et
internal standard using chemical data obtained previ- al., 1982). Assuming no preferred orientation in the
ously (e.g. Jarvis and Williams, 1993; Barbarand et al., fission process and no variation in distance travelled for
2003). Blank correction was performed prior to internal different spontaneous fission events, the probability
standard correction. All samples and standards were may be calculated from (Wagner and Van den haute,
corrected for change in 44Ca throughout the run. 1992):
However, the gas blank was not internally standardised
in any way since it does not contain calcium. Corrected the area of a part of the semi  sphere
integrals for NIST 610 and NIST 612 were used to
representing track orientation in which
construct calibration lines which were forced through
the origin. a track intersect an observed surface
Px ¼ ð2Þ
the area of the semi  sphere
representing all posible track orientation
3. Measurement of uranium content Lx
¼
L
3.1. Induced FT method

The uranium concentration of each apatite grain where Px = probability of a FT intersecting the ob-
was determined by comparing its induced-track served surface for a nucleus which is x Am away from
density in an external detector with that of a the observed surface, and L = distance travelled by a
uranium dosimeter glass (also recorded in an ex- fission fragment (Am).
ternal detector) using the equation: Assuming the number of fission events (Dx) at x
Am away from the observed surface is homoge-
qi A235dos g  REDdos neous, the general registration efficiency R is calcu-
U¼  Udos   ð1Þ
qd A235 g  RED lated as

Z L
where U = total uranium content of apatite (Ag/g),
qi = induced track density for an apatite grain mea- PxDxdx Z L
0
sured in an external detector (cm 2), qd =induced R¼2 Z L ¼2 Pxdr ¼ L ð3Þ
0
track density for the dosimeter glass measured in an Dxdx
external detector (cm 2), Udos = the uranium content 0

of the dosimeter glass (Ag/g), A235dos = 235U abun-


dance in weight for the dosimeter glass (g/g), Therefore, RED can be estimated as half the in-
A235 = 235U abundance in weight for the apatite (g/g), duced FT length in apatite ( f 7.9 Am after chemical
g = geometry factor, necessary because tracks etching, Iwano and Danhara, 1998, see also Gleadow
recorded in the external detector have a 2k geometry et al., 1986) plus an unetchable track range (and
(originating from only one side of the observed therefore invisible in apatite) which is visible if the
surface), whilst tracks in apatite have a 4k geometry range is in the mica ( f 1.0 Am) (Iwano and Danhara,
(coming from both sides of the viewed surface); equal 1998). However, measurement of track lengths in
to 0.5 for the c-axis parallel apatite sections analysed glass and the registration efficiency in an external
in this study. REDdos = a registration efficiency factor detector are less well studied. Accordingly we use a
for 235U induced tracks in the dosimeter glass (cm), factor of 1 for REDdos/RED as a first-order estimation of
RED = a registration efficiency factor for 235U induced U concentration by the induced FT method. Consid-
tracks in the apatite (cm). ering that the material exerts a major control on the
The probability of a FT intersecting the observed registration efficiency (e.g. the difference between
surface is a function of two distances: (a) the distance apatite and zircon is more than 30%), the real uranium
between the observed surface and a nucleus whose concentration might be much more (or less) than the
N. Hasebe et al. / Chemical Geology 207 (2004) 135–145 139

estimated concentration derived using the ratio RED-


dos/RED. However, the relative relationship between 1
2
grains would be preserved. 3
4

Relative signal Intensity


Uncertainty of the measured uranium concentration 5
6
was calculated using the formation of Green (1981); all

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
s   2
1 2 1
Uncertainty ¼ U þ ð4Þ
Ni Nd

where Ni = number of counted induced tracks recorded


in the apatite external detector, and Nd = number of
counted induced tracks recorded in the dosimeter
glass external detector.
0 100 200 300 400 500

3.2. LA-ICP-MS calibration and U content U238 content (µg/g)

Fig. 1. Relative LA-ICP-MS signal intensity of 238U for NIST610


Calibration of the LA-ICP-MS analyses was made and 612 standard glasses. Six sets of measurement were made
by measurement of 6 datasets on NIST 610 and 612 during the analyses with a calibration line obtained for each set
glasses interspaced with the apatite sample determi- based on triplicate measurements for each standard glass. Signal
nations. The weighted mean of 238U relative intensity intensity is relative figures after ablated volume correction using Ca-
was calculated from triplicate analysis for each stan- 44 intensity.
dard glass set after ablated volume correction using
Ca-44 intensity and CaO content, and compared
lysed by the FT method. Some apatite grains of the
against 238U concentration based on their published
Mount Dromedary standard did not give a suffi-
uranium concentrations (Pearce et al., 1997) and
ciently strong signal and therefore failed the accep-
isotopic ratio (235U: 0.24% in number of atoms;
tance criteria defined as being a sample signal
Carpenter and Reimer, 1974). Calibration lines (rela-
greater than three standard deviations of the blank.
tive signal intensity vs. 238U concentration) remained
The average uranium content for each sample is
stable within error throughout the measurements (Fig.
listed in Table 1.
1) and this enabled adoption of a mean calibration line
to calculate 238U concentration to within f 10% of
3.3. Comparison of uranium values
analytical error. Total uranium concentration of apa-
tites (Ag/g) was calculated from measured 238U con-
Fig. 2 plots uranium concentrations of apatite
tent (Ag/g) and natural isotope abundance ratio of
grains estimated from neutron-induced tracks and
uranium (235U: 0.72%/238U: 99.28% in number of
from LA-ICP-MS. The data for Durango apatite form
atoms, and 0.71% and 99.29% in weight), making the
a cluster because the analysed apatites are fragments
reasonable assumption of a natural 235U/238U ratio in
of a single large megacryst which has been crushed. In
all apatite samples. Therefore
the absence of zoning and between-grain variation,
238
U broadly similar uranium values would be expected for
U¼ 238 A
ð5Þ all fragments. The Fish Canyon Tuff data show a
broadly linear relationship between the two estimates
where U = sample total uranium content (Ag/g), of uranium. The data points are below the one-to-one
238
U = uranium-238 content (Ag/g) obtained from correlation line, uranium concentrations determined
LA-ICP-MS, and A238 = abundance of 238U in weight by induced-track density being underestimated rela-
(0.9929). tive to the LA-ICP-MS data. This could result from
Uranium concentrations were successfully mea- the approximation of REDdos in the induced FT ap-
sured by LA-ICP-MS on most apatite grains ana- proach, implying that REDdos should be larger when
140 N. Hasebe et al. / Chemical Geology 207 (2004) 135–145

broad cluster with no significant relationship between


the results from the two different approaches used to
determine uranium concentration.

4. Calculation of FT ages

Comparison of ages calculated for the same apatite


grains by both the conventional FT and LA-ICP-MS
approaches represents for chronometry a more useful
assessment of the utility of the new approach.

4.1. Conventional FT ages

Conventional FT ages were determined for each


apatite grain using the zeta approach (Hurford, 1990;
Hurford and Green, 1983) with data presented in
Table 1. The zeta calibration approach (Fleischer
Fig. 2. Uranium contents in Ag/g of individual apatite grains
and Hart, 1972; Hurford and Green, 1983) is a
determined by LA-ICP-MS and neutron-induced fission-track consensus approach adopted by the FT community
density. Error on x-axis is F 10% assessed from reproducibility of to avoid absolute evaluation of neutron fluence and
measurements on standard glasses. Error on y-axis is F 1 standard selection of a value 238U spontaneous fission decay
error derived from track counts. Dur; Durango, FCT; Fish Canyon constant kf (Hurford, 1990). By definition, the cali-
Tuff, MDC; Mount Dromedary Complex.
bration factor zeta must be determined for a dosimeter
glass by analysis of age standards. The zeta values of
compared to RED, which would be consistent with a 319 F 9 for CN5 and 311 F 10 for SRM 612 used in
fission fragment travelling further in glass than in this study were derived from analysis of the apatite
apatite. The Mount Dromedary Complex data show a age standards from Durango, Fish Canyon Tuff and

Table 3
Fission-track data for zeta calibration against apatite age standards
Apatite standard No. of grains qs, 106 cm 2 Ns qi, 106 cm 2 Ni qd, 106 cm 2 Nd fFr P(v2), %
CN5 dosimeter glass
DUR-1 11 0.20 299 1.89 2884 1.6508 1144 368 F 26 67
DUR-2 22 0.14 420 0.86 2632 1.3104 1453 301 F 18 15
FCT-1 20 0.19 252 1.37 1826 1.2996 1441 310 F 22 94
FCT-2 20 0.22 313 1.26 1809 1.1751 1303 274 F 18 99
MDC 21 0.86 1219 1.74 2452 1.1896 1319 336 F 15 0.5
Weighted mean f CN5 319 F 9

SRM 612 dosimeter glass


DUR-1 11 0.14 337 1.57 3847 2.4621 1365 292 F 19 14
DUR-2 11 0.12 254 1.57 3274 2.4621 1365 330 F 24 74
FCT-1 14 0.20 121 1.13 669 0.9425 1045 327 F 38 22
FCT-2 20 0.21 262 1.45 1824 1.2617 1399 307 F 22 27
MDC 10 0.92 375 3.50 1433 2.4459 1356 311 F 20 73
Weighted mean f SRM612 311 F 10
DUR = Durango; FCT = Fish Canyon Tuff; MDC = Mount Dromedary Complex.
Ns,i,d = numbers of tracks counted for spontaneous, induced and dosimeter measurements; qs,i,d = densities of spontaneous, induced and
dosimeter tracks cm 2; P(v2) probability of v2 for (n  1) degrees of freedom (Galbraith, 1981).
N. Hasebe et al. / Chemical Geology 207 (2004) 135–145 141

the Mount Dromedary Complex (Table 3). However, Because we observe the number of fission decay
these original and multiple zeta calibration experi- events as a spontaneous track density, the equation is
ments were made on separate aliquots and were changed to:
completely independent of the analyses presented in
kf
this study. qs ¼ N238 ðekD t  1ÞRsp k ð7Þ
kD
4.2. LA-ICP-MS FT ages where qs = spontaneous FT density at the observed
surface (cm 2), Rsp = a registration factor by which
238
For the LA-ICP-MS approach, the FT age of each U in a unit volume would leave spontaneous
grain was calculated using an absolute approach tracks on an observed surface (cm), and k = a con-
which required selection of a kf value. Although there stant, which would be variable depending on exper-
is a diversity of values for kf, those direct determi- imental factors such as etching and observation
nations which are independent of absolute neutron conditions.
fluence measurement centre around 8.5  10  17 Here N238 (cm 3) is calculated using the content of
year 1. Of these analyses, we have chosen to use in uranium-238:
this study the precise value of (8.46 F 0.06)  10 17
year 1 (Spadavecchia and Hahn, 1967; Bigazzi, 238
U  106  d
N238 ¼ NA  ð8Þ
1981). M
The general age equation is
where NA = Avogadro’s number, 238U = uranium-238
kf content measured by LA-ICP-MS ( Ag/g), d = apatite
Dt ¼ N238 ðekD t  1Þ ð6Þ density (taken as 3.19 g/cm3), and M = the mass of
kD 238
U. Therefore,
where Dt = number of decay events of 238U during !
time t in a unit volume (cm 3), kf = 238U sponta- 1 qs kD M
t¼ ln 1 þ ð9Þ
neous fission decay constant (8.46  10 17 year 1 kD kf NA238 U106 dRsp k
Spadavecchia and Hahn, 1967), kD = 238U total
decay constant (1.55125  10 10 year 1 Jaffey et In this study we adopted as Rsp a half of the mean
al., 1971), N238 = current number of uranium-238 etchable spontaneous fission-track length in apatite
atoms in a unit volume (cm 3), and t = FT age from geologically rapidly cooled samples without
(years). subsequent thermal disturbance (i.e. unannealed): 7.5

Fig. 3. Single-grain FT apatite ages calculated by using 238U content determined by LA-ICP-MS plotted against FT ages determined by
conventional zeta age calibration. Continuous lines represent reference age of each apatite standard (e.g. Hurford, 1998). Error bars are F 1
standard error.
142 N. Hasebe et al. / Chemical Geology 207 (2004) 135–145

Am = 7.5  10 5 cm (Gleadow et al., 1986), and we portion of the grain that has been polished away, and
have set k as 1. from up to half the distance of a fission-track length
The error of a grain age is given by below the polished surface, in total a maximum depth
of f 10 Am. Since we cannot determine uranium
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
concentrations for the missing portion of the apatite,
1
t þ d2 ð10Þ we have to rely on data collected from the polished
Ns2
grain. For the conventional FT method, neutron irra-
diation is used to induce fission forming new tracks
where Ns = the number of counted spontaneous tracks, that register in an external detector. The tracks
and d = uncertainties in uranium concentration (10% = recorded in the detector are produced by uranium up
0.1). to 10 Am (broadly half a fission-track length) below
The results are shown in Fig. 3 with the weighted the grain surface, an approach that assumes no sig-
mean age for each sample listed in Table 1. nificant variation in uranium between the missing and
preserved portions of the apatite, i.e. zoning is unim-
portant over a 20 Am range.
5. Discussion A similar assumption must be made for LA-ICP-
MS analysis, with ablation to greater than half a FT
These first results demonstrate that a LA-ICP-MS length sampling uranium in area not equivalent to that
approach to calculating a FT age is viable and offers used for FT measurements.
an alternative to reactor-based uranium determinations
and mapping. Although the age data produced in this 5.2. Etched FT density and laser ablation
study are in reasonable agreement with conventional
FT results, several issues need to be resolved in order LA-ICP-MS measurements were made on apatites
to enhance the LA-ICP-MS approach. that had been etched to reveal the spontaneous track
densities. If etching removes U or Ca (which is
5.1. Depth drilled by laser ablation measured as the internal standard) selectively, the
calculated 238U content would be biased. Currently it
Firstly, the depth of the ablated trench, estimated in is unknown whether any such losses occur and the
this study to be 25 F 5 Am by microscopic observation potential significance of any such removal. Further-
for most grains although occasionally deeper, needs more, the levels of etching may change the ablation
better control. In some cases the laser drilled com- characteristics of apatite, whilst differences in the
pletely through the smaller grains. The Durango numbers of etched tracks on an apatite surface may
apatite tends to have shallower pits than other two affect the volume of material ablated and the particle
samples. The wide scatter of U contents and FT ages distribution introduced into the ICP potentially affect-
for the MDC apatite (Figs. 2 and 3) could be ing the ionising efficiency, although previous studies
explained by the in-grain heterogeneity of U content, on fractionation (Horn and Gunther, 2002) suggest this
up to a factor of f 4 between, for example, rim and to be unlikely. It is interesting that the apatite which
core of one grain (Kohn, unpublished data). The shows greatest variation in age is that from the Mount
variation in ablated depth between samples could be Dromedary Complex (Fig. 3) which has the highest
caused by chemical composition which may change spontaneous track density ( f 3.5  106 cm 2), sig-
the optical property and therefore alter the efficiency nificantly greater than in the other two samples,
of energy absorption. In addition, the spontaneous although the scatter is most probably caused by ura-
track density could also change the ablation character- nium heterogeneity within a grain.
istics of material. The aim of ablation is to determine
the uranium concentration of a volume equivalent to 5.3. Age equation
that used to measure spontaneous fission-track densi-
ties. The spontaneous tracks counted on the observed The equation used in this study to calculate the
surface result from fission of uranium both from a LA-ICP-MS ages gave reasonable results when
N. Hasebe et al. / Chemical Geology 207 (2004) 135–145 143

compared with independent age data. However, The factor e would be determined by analysis of age
some of the constants in the age equation may standard sample, as with the present zeta calibration:
be poorly evaluated. The registration efficiency
238
(Rsp: equivalent to a half of spontaneous track U
e¼ ðekD tstd  1Þ ð12Þ
length) could vary F 3% according to chemical qs
composition (Barbarand et al., 2003) and affect
calculated ages by a similar amount. The apatite where tstd represents a reference age of the age
density (d) also depends on its chemical compo- standard sample.
sition and cell parameters, and varies F 1.5% (Bar-
barand, unpublished data), resulting in an additio 5.4. Reference material
nal uncertainty in calculated age. However, such
factors are of minor importance when compared to The uncertainty in ICP-MS measurement of urani-
the present uncertainties of up to 10% in 238U um concentration was estimated in this study as up to
measurement. f 10%, based on repeated measurements of NIST
The lack of an internationally agreed value for kf standard glasses (Fig. 1). Consideration of the results
remains a problem (Hurford, 1998 and references from the Durango apatite suggest that the real uncer-
therein). Published kf values range from 5 to 12 tainty may be larger: despite being the crushed frag-
(10 17 year 1) (Bigazzi, 1981), the majority falling ments of a single large crystal, the dispersion of
within two groups which differ by f 15%; choice of Durango results exceeds the nominal 10% error bars
kf could cause this difference of up to 15% to for LA-ICP-MS shown in Fig. 2. The source of this
propagate into the age. Ideally, a consensus agree- additional apparent uncertainty in apatite measure-
ment should be sought for a single value of kf to be ment may arise from a matrix effect: differences in
used in FT chronometry; this might be an easier task the behaviours between apatite and standard glasses
than in the past where the complexity of neutron through ablation and ionisation processes. Reduced
fluence measurement was intimately associated in the errors may result from using a standard of similar
discussion. However, even if such a consensus was composition and ablation-transformation efficiency
reached, it would still be necessary to evaluate a (Odegard, 1999; Greenwood and Jarvis, 2002).
composite procedural value, the factor Q of Wagner
and Van den haute (1992), which encompasses 5.5. Multi-elemental analysis
differences in etching, counting, microscope, obser-
vation and other experimental variation. An alterna- In this study, we have measured only the 238U
tive comparative procedure, avoiding absolute concentration to obtain the largest gain from a small
evaluation of kf, would be analogous to the zeta amount of sample. However, an advantage of the LA-
approach in the conventional FT method (Hurford ICP-MS lies in multi-element analysis. The annealing
and Green, 1983). A calibration factor (let us call it behaviour of fission tracks in apatite is largely depen-
epsilon, e) would include experimental differences in dent on crystal structure and elements substituted in
etching and spontaneous track counting (k in the age that structure such as Cl and rare earth elements
Eq. (9) above) and registration factor (Rsp) and (Barbarand et al., 2003). The level of chlorine substi-
would be defined as: tution appears the most influential factor (Green et al.,
1985, 1986; Crowley et al., 1990; Donelick, 1993),
qs kD M and thus determination of chlorine content may have
e¼ 238 an important bearing on robust interpretation of apa-
kf NA U106 dRsp K
tite FT data. Although chlorine measurement by ICP-
MS is difficult mainly because of poor ionization
which, when substituted into the age equation, gives: efficiency (e.g. Tran et al., 2001), potentially it may
  be possible to measure both uranium, chlorine and
1 qs rare earths at the same time. This would represent a
t¼ ln 1 þ e  238 ð11Þ
kD U significant advance for the FT method.
144 N. Hasebe et al. / Chemical Geology 207 (2004) 135–145

6. Conclusions Bellmans, F., De Corte, F., Van den haute, P., 1995. Comparison of
SRM and CN U-doped glasses: significance for their use as
thermal neutron fluence monitors in fission track dating. Radiat.
This study has demonstrated that LA-ICP-MS is a Meas. 24, 153 – 160.
viable alternative method for measuring uranium Bigazzi, G., 1981. The problem of the decay constant kf of 238U.
content for apatite FT analysis. LA-ICP-MS analysis Nucl. Tracks 5, 35 – 44.
appears to offer a fast and efficient way of acquiring Carpenter, B.S., Reimer, G.M., 1974. Calibrated glass standards for
fission track use. NBS Spec. Pub., 260.
data with a level of precision comparable to the
Carter, A., 1999. Present status and future avenues of source region
current neutron-irradiation technique. However, be- discrimination and characterization using fission track analysis.
fore this method can be used routinely, several issues Sediment. Geol. 124, 31 – 45.
need to be resolved. Firstly, assessment must be made Crowley, K.D., Cameron, M., Schaefer, R.L., 1990. Annealing of
of any possible linkage between the level of etching etchable fission-track damage in F-, OH-, Cl- and Sr-apatite: 1.
and spontaneous track density to degradation of Systematics and preliminary interpretations. Nucl. Tracks Radiat.
Meas. 17, 409 – 410.
uranium ICP-MS measurement. This requires a better Cox, R., Kosler, J., Sylvester, P., Hodych, J., 2000. Apatite fission-
understanding of how laser ablation works on etched track (FT) dating by LAM-ICO-MS analysis. Abstract of Gold-
apatite surfaces. Secondly, consideration must be schmidt 2000. J. Conf. Abstr. 5, 322.
given to adopting a consensus value for kf. Or we Donelick, R.A., 1993. Apatite etching characteristics versus chem-
could avoid kf by developing a comparative calibra- ical composition. Nucl. Tracks Radiat. Meas. 21, 604.
Ertel, W., O’Neill, H.St.C., Sylester, D.B., Dingwell, D.B, Spettel,
tion approach analogous to zeta, linking analysis of B., 2001. The solubility of rhenium in silicate melts: Implica-
samples to that of standards. Thirdly, a more appro- tions for the geochemical properties of rhenium at high temper-
priate ICP-MS standard is required, of similar com- atures. Geochim. Cosmochim. Acta 65, 2161 – 2170.
position to apatite, to reduce uncertainties in analysis. Fitzgerald, P.G., Sorkhabi, R.B., Redfield, T.F., 1995. Uplift and
denudation of the central Alaska range: a case study in the use of
apatite fission track thermochronology to determine absolute
uplift parameters. J. Geophys. Res. 100, 20175 – 20191.
Acknowledgements Fleischer, R.L., Hart, H.R., 1972. Fission track dating: techniques
and problems. In: Bishop, W.W., Miller, J.A., Cole, S. (Eds.),
Noriko Hasebe’s stay at UCL was supported by Calibration of Hominoid Evolution, 135 Scottish Academic
the Japanese Society of Promotion of Science 2000. Press, Edinburgh, p. 170.
Galbraith, R.F., 1981. On statistical models for fission track counts.
Jocelyn Barbarand was supported at UCL by a
Math. Geol. 13, 471 – 488.
Marie Curie fellowship of the European Community Gallagher, K., Hawkesworth, C.J., Mantovani, M.S.M., 1994. The
programme Energy, Environment and Sustainable denudation history of the onshore continental margin of SE
Development under Contract No. ENK6-CT-1999- Brazil inferred from apatite fission-track data. J. Geophys.
50002. Analytically, the ICP-MS measurements at Res. 99, 18117 – 18145.
Gleadow, A.J.W., 1981. Fission-track dating methods: what are the
the NERC Kingston University facility have been
real alternatives? Nucl. Tracks 5, 3 – 14.
enabled by grant ICP/179/1200. Fission-track stud- Gleadow, A.J.W., Lovering, J.F., 1977. Geometry factor for external
ies at UCL are supported by NERC-JREI grant detectors in fission track dating. Nucl. Track Detect. 1, 99 – 106.
NER/F/S/2000/00/165 ‘‘Defining low-temperature Gleadow, A.J.W., Duddy, I.R., Green, P.F., Lovering, J.F., 1986.
thermal history in the Earth’s upper crust by U – Confined fission track lengths in apatite: a diagnostic tool
for thermal history analysis. Contrib. Mineral. Petrol. 94,
Th/He dating and fission-track analysis’’. We thank
405 – 415.
Robin Clayton and Barry Kohn for comments on an Gleadow, A.J.W., Belton, D.X., Kohn, B.P., Brown, R.W., 2002.
earlier draft. Jan Kosler and an anonymous reviewer Fission track dating of phosphate minerals and the thermochro-
gave us helpful comments. [PD] nology of apatite. In: Kohn, M., Rakovan, J., Hughes, J.M.
(Eds.), Phosphates Geochemical, Geobiological and Material
Importance, Reviews in Mineralogy and Geochemistry, 48 Min-
eralogical Society of America, USA, pp. 579 – 630.
References Green, P.F., 1981. A new look at statistics in fission-track dating.
Nucl. Tracks 5, 77 – 86.
Barbarand, J., Carter, A., Wood, I.G., Hurford, A.J., 2003. Compo- Green, P.F., 1985. Comparison of zeta calibration baselines for
sitional and structural control of fission-track annealing in apa- fission-track dating of apatite, zircon and sphene. Chem. Geol.,
tite. Chem. Geol. 198, 107 – 137. Isot. Geosci. Sect. 58, 1 – 22.
N. Hasebe et al. / Chemical Geology 207 (2004) 135–145 145

Green, P.F., Hurford, A.J., 1984. Thermal neutron dosimetry for Jaffey, A.H., Flynn, K.F., Glendenin, L.E., Bentley, W.C., Essling,
fission track dating. Nucl. Tracks 9, 231 – 241. A.M., 1971. Precision measurement of the half-lives and spe-
Green, P.F., Duddy, I.R., Gleadow, A.J.W., Tingate, P.R., Laslett, cific activities of 235U and 238U. Phys. Rev. 4, 1889 – 1906.
G.M., 1985. Fission-track annealing in apatite: track-length Jarvis, K.E., Williams, J.G., 1993. Laser ablation inductively cou-
measurements and the form of the arrhenius plot. Nucl. Tracks pled plasma mass spectrometry (LA-ICP-MS): a rapid technique
10, 323 – 328. for the direct, quantitative determination of major, trace and
Green, P.F., Duddy, I.R., Gleadow, A.J.W., Tingate, P.R., Laslett, rare-earth elements in geological samples. Chem. Geol. 106,
G.M., 1986. Thermal annealing of fission tracks in apatite: 1. A 251 – 262.
qualitative description. Chem. Geol., Isot. Geosci. Sect. 59, Kimura, J., Danhara, T., Iwano, H., 2000. A preliminary report
237 – 253. on trace element determinations in zircon and apatite crystals
Green, P.F., Duddy, I.R., Gleadow, A.J.W., Lovering, J.F., 1989. using excimer laser ablation-inductively coupled plasma mass
Apatite fission-track analysis as a paleotemperature indicator spectrometry (ExLA-ICPMS). Fission Track News Lett. 13,
for hydrocarbon exploration. In: Naeser, N.D., Mc Culloh, 11 – 20.
Th.H. (Eds.), Thermal History of Sedimentary Basins. Spring- Kohn, B.P., Green, P.F. (Eds.), 2002. Low Temperature Thermo-
er-Verlag, New York, pp. 181 – 195. chronology: from Tectonics to Landscape Evolution. Tectono-
Greenwood, J., Jarvis, K., 2002. NERC ICP-MS Facility: An over- physics, vol. 349. 368 pp.
view of capabilities and applications. Abstract of Young Scien- Laslett, G.M., Kendall, W.S., Gleadow, A.J.W., Duddy, I.R., 1982.
tist Meeting of International Association of Geoanalysists, UK. Bias in measurement of fission-track length distributions. Nucl.
Hirata, T., 1997. Soft ablation technique for laser ablation-induc- Tracks 6, 79 – 85.
tively coupled plasma mass spectrometry. J. Anal. At. Spectrom. Odegard, M., 1999. Preparation of synthetic calibration materials
12, 1337 – 1342. for use in the microanalysis of oxide minerals by direct fusion in
Horn, I., Gunther, D., 2002. Transport efficiency and particle size high-purity graphite electrodes: preliminary results for quartz
distribution of laser induced aerosols. Abstracts of the 12th and rutile. Geostand. Newsl. 23, 173 – 186.
annual V. M. Goldschmidt Conference, Davos, Switzerland, Pearce, N.J.G., Perkins, W.T., Westgate, J.A., Gorton, M.P, Jackson,
Special Supplement of Geochim. Cosmochim. Acta, 66, A341. S.E., Neal, C.R., Chenery, S.P., 1997. A compilation of new and
Horn, I., Rudnick, R.L., McDonough, W.F., 2000. Precise elemental published major and trace element data for NIST SRM 610 and
and isotope ratio determination by simultaneous solution nebu- NIST SRM 612 glass reference materials. Geostand. Newsl. 21,
lization and laser ablation-ICP-MS: application to U – Pb geo- 115 – 144.
chronology. Chem. Geol. 164, 281 – 301. Rubatto, D., 2002. Zircon trace element geochemistry: partitioning
Hurford, A.J., 1990. Standardization of fission track dating calibra- with garnet and the link between U – Pb ages and metamor-
tion: Recommendation by the Fission Track Working Group of phism. Chem. Geol. 184, 123 – 138.
the I. U. G. S. Subcommission on Geochronology. Chem. Geol., Spadavecchia, A., Hahn, B., 1967. Die Rotationskammer und
Isot. Geosci. Sect. 80, 171 – 178. einige Anwendungen. Helv. Phys. Acta 40, 1063 – 1079.
Hurford, A.J., 1998. Zeta: the ultimate solution to fission-track anal- Stirling, C.H., Lee, D.C., Christensen, J.N., Halliday, A.N., 2000.
ysis calibration or just an interim measure? In: Van den haute, P., High-precision in situ 238U – 234U – 230Th isotopic analysis using
De Corte, F. (Eds.), Advances in Fission-Track Geochronology. laser ablation multiple-collector ICPMS. Geochim. Cosmochim.
Kluwer Academic Publishing, Netherlands, pp. 19 – 32. Acta 64, 3737 – 3750.
Hurford, A.J., Green, P.F., 1982. A user’s guide to fission track Svojtka, M., Kosler, M., 2002. Fission-track dating of zircon by
dating calibration. Earth Planet. Sci. Lett. 59, 343 – 354. laser ablation ICPMS. Abstracts of the 12th annual V. M. Gold-
Hurford, A.J., Green, P.F., 1983. The zeta age calibration of fission- schmidt Conference, Davos, Switzerland. Special Supplement
track dating. Isot. Geosci. 1, 285 – 317. of Geochim. Cosmochim. Acta, vol. 66, p. A756.
Hurford, A.J., Platt, J.P., Carter, A., 1999. Fission-track analysis of Tagami, T., Nishimura, S., 1992. Neutron dosimetry and fission-
samples from the Alboran sea basement. In: Zahn, R., Comas, track age calibration: insights from intercalibration of uranium
M.C., Klaus, A. (Eds.), Proceedings of the Ocean Drilling Pro- and thorium glass dosimeters. Chem. Geol., Isot. Geosci. Sect.
gram, Scientific Results, 161, pp. 295 – 300. 102, 277 – 296.
Iwano, H., Danhara, T., 1998. A re-investigation of the geom- Tran, M., Sun, Q., Smith, B.W., Winefordne, J.D., 2001. Determi-
etry factors for fission-track dating of apatite, sphene and nation of F, Cl, and Br in Solid Organic Compounds by Laser-
zircon. In: Van den haute, P., De Corte, F. (Eds.), Advances Induced Plasma Spectroscopy. Appl. Spectrosc. 55, 739 – 744.
in Fission-Track Geochronology. Kluwer Academic Publish- Wagner, G.A., Van den haute, P., 1992. Fission Track Dating. Enke
ing, Netherlands, pp. 47 – 66. Verlag-Kluwer Academic, Netherlands. 285 pp.

You might also like