Skip to main content

Main menu

  • Home
  • Content
    • Current
    • Archive
    • Special Volumes and Special Issue
  • Subscriptions
    • Subscribers
    • FAQ
    • Terms & Conditions for use of AJS Online
  • Instructions to Authors
    • Focus and paper options
    • Submit your manuscript
  • Site Features
    • Alerts
    • Feedback
    • Usage Statistics
    • RSS
  • About Us
    • Editorial Board
    • The Journal

User menu

  • Register
  • Subscribe
  • My alerts
  • Log in
  • Log out
  • My Cart

Search

  • Advanced search
American Journal of Science
  • Register
  • Subscribe
  • My alerts
  • Log in
  • Log out
  • My Cart
American Journal of Science

Advanced Search

  • Home
  • Content
    • Current
    • Archive
    • Special Volumes and Special Issue
  • Subscriptions
    • Subscribers
    • FAQ
    • Terms & Conditions for use of AJS Online
  • Instructions to Authors
    • Focus and paper options
    • Submit your manuscript
  • Site Features
    • Alerts
    • Feedback
    • Usage Statistics
    • RSS
  • About Us
    • Editorial Board
    • The Journal
  • Follow ajs on Twitter
  • Visit ajs on Facebook
  • Follow ajs on Instagram
Research ArticleArticles

Helium diffusion in natural zircon: Radiation damage, anisotropy, and the interpretation of zircon (U-Th)/He thermochronology

William R. Guenthner, Peter W. Reiners, Richard A. Ketcham, Lutz Nasdala and Gerald Giester
American Journal of Science March 2013, 313 (3) 145-198; DOI: https://doi.org/10.2475/03.2013.01
William R. Guenthner
* Department of Geosciences, University of Arizona, 1040 E. 4th Street, Tucson, Arizona 85721 U.S.A.
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
  • For correspondence: wrg@email.arizona.edu
Peter W. Reiners
* Department of Geosciences, University of Arizona, 1040 E. 4th Street, Tucson, Arizona 85721 U.S.A.
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Richard A. Ketcham
** Jackson School of Geosciences, University of Texas, Austin, Texas 78712 U.S.A.
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Lutz Nasdala
*** Institute for Mineralogy and Crystallography, University of Vienna, Althanstrasse 14, 1090, Vienna, Austria
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Gerald Giester
*** Institute for Mineralogy and Crystallography, University of Vienna, Althanstrasse 14, 1090, Vienna, Austria
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
  • Article
  • Figures & Data
  • References
  • Info & Metrics
  • PDF
Loading

Abstract

Accurate thermochronologic interpretation of zircon (U-Th)/He dates requires a realistic and practically useful understanding of He diffusion kinetics in natural zircon, ideally across the range of variation that characterize typically dated specimens. Here we present a series of date and diffusion measurements that document the importance of alpha dose, which we interpret to be correlated with accumulated radiation damage, on He diffusivity. This effect is manifest in both date-effective uranium (eU) correlations among zircon grains from single hand samples and in diffusion experiments on pairs of crystallographically oriented slabs of zircon with alpha doses ranging from ∼1016 to 1019 α/g. We interpret these results as due to two contrasting effects of radiation damage in zircon, both of which have much larger effects on He diffusivity and thermal sensitivity of the zircon (U-Th)/He system than crystallographic anisotropy. Between 1.2 × 1016 α/g and 1.4 × 1018 α/g, the frequency factor, D0, measured in the c-axis parallel direction decreases by roughly four orders of magnitude, causing He diffusivity to decrease dramatically (for example by three orders of magnitude at temperatures between 140 and 220 °C). Above ∼2 × 1018 α/g, however, activation energy decreases by a factor of roughly two, and diffusivity increases by about nine orders of magnitude by 8.2 × 1018 α/g. We interpret these two trends with a model that describes the increasing tortuosity of diffusion pathways with progressive damage accumulation, which in turn causes decreases in He diffusivity at low damage. At high damage, increasing diffusivity results from damage zone interconnection and consequential shrinking of the effective diffusion domain size. Our model predicts that the bulk zircon (U-Th)/He closure temperature (Tc) increases from about 140 to 220 °C between alpha doses of 1016 to 1018 α/g, followed by a dramatic decrease in Tc above this dose. Linking this parameterization to one describing damage annealing as a function of time and temperature, we can model the coevolution of damage, He diffusivity, and (U-Th)/He date of zircon. This model generates positive or negative date-eU correlations depending on the extent of damage in each grain and the date-eU sample's time-temperature history.

  • Zircon
  • He diffusion
  • thermochronology
  • radiation damage

INTRODUCTION

Over the last decade, numerous studies have used zircon (U-Th)/He (zircon He) thermochronology to interpret thermal histories and geologic processes. Accurate and realistic interpretations using this thermochronometer, as well as constraints on convenient indices like closure temperature (Tc) and the partial retention zone (PRZ), require quantitative understanding of the kinetics of He diffusion in zircon, including the effects of temperature, crystallographic orientation, and radiation damage. Most zircon He dating studies thus far have assumed that the kinetics measured on a few zircons from a limited number of locations (Reiners and others, 2002; Reiners and others, 2004; Wolfe and Stockli, 2010) apply to all zircons found in a wide range of geologic settings. This assumption may be appropriate, at least to first order, in some cases, as demonstrated by geologically consistent results from settings such as deep drill cores (Wolfe and Stockli, 2010). But several aspects of He diffusion in zircon are likely to be more complicated, which may lead to more complex results in some applications.

Anisotropic He diffusion in zircon is one such complication. Molecular dynamic simulations (Reich and others, 2007; Saadoune and others, 2009; Bengston and others, 2012) and laboratory measurements (Farley, 2007; Cherniak and others, 2009) have demonstrated that He diffusion is faster in the c-axis parallel direction than the c-axis orthogonal direction, at least in specimens with little or no radiation damage or other type of defects. These studies might suggest that grain aspect ratios may influence diffusion kinetics. Watson and others (2010) introduced analytical and numerical methods that allow consideration of the degree to which anisotropy affects both the calculation of bulk He loss from a zircon and step-heating results. However, results from our study suggest that anisotropy is a relatively minor problem for interpreting zircon He dates compared with the effects of radiation damage, which are less well understood.

Damage results from self-irradiation, primarily by recoils of heavy daughter nuclei upon emission of an alpha particle, but also by spontaneous fission events and the alpha particles themselves. Because radiation damage can be annealed at elevated temperatures (for example, Zhang and others, 2000), its extent in a grain can be predicted only roughly by calculating time-integrated self-irradiation doses. These doses can be calculated from the concentration of effective uranium (eU) as scaled for relative alpha production rate (eU = U + 0.235 × Th), and an estimate of the time since the sample was cooled below the threshold temperature for long-term damage annealing. Nasdala and others (2004a) showed that alpha doses calculated using a zircon's U-Pb date may overestimate the radiation damage present, as their Sri Lankan samples experienced annealing post-dating each zircon's U-Pb date. Instead, they calculated “effective alpha doses” by applying a correction factor that accounts for the partial long-term annealing. This correction factor was in part calibrated with Raman spectroscopy, a more direct technique than estimating alpha dose for quantifying damage. Unfortunately, this calibration was specific to Nasdala and other's (2004a) particular suite of zircons, which makes it difficult to broadly use in constraining the threshold temperature of long-term annealing. Furthermore, as we will discuss in greater detail in later sections, there is no consensus on how to model the kinetics of alpha recoil damage on either laboratory or geologic timescales. For samples with simple thermal histories, it may be possible to roughly estimate the duration over which radiation damage has accumulated from the density of spontaneous fission tracks. To the best of our knowledge, the kinetics of fission-track annealing are the only ones available that describe damage annealing of any type in zircon on geologic timescales (for example, Rahn and others, 2004; Tagami, 2005; Yamada and others, 2007), and if the apatite system serves as a comparison (for example, Shuster and Farley, 2009), then it is reasonable to expect that they correlate with the kinetics of alpha recoil damage annealing. Current estimates of the ZFT partial annealing zone (that is, threshold temperature for annealing) are 262 to 330 °C at an isothermal hold-time of 10 my (Yamada and others, 2007), although Garver and others (2005) have shown that annealing temperatures can be as low as 180 °C in heavily damaged zircons. Again, we leave a more detailed discussion of the limitations of comparing these two types of damage annealing to a later section, but for now assume that ZFT kinetics provide an estimate of the degree of structural annealing in zircon. Thus, as long as zircon He dates are either similar to ZFT dates, or if we can assume from geologic constraints that He dates record a pulse of rapid cooling from temperatures consistent with the ZFT partial annealing zone, then He dates can be used to estimate the duration over which radiation damage has accumulated, and this, combined with effective U concentration, provides an estimate of the “effective alpha dose.” In this paper, we report alpha doses following this assumption, unless stated otherwise.

Previous work on He diffusion in zircon focused mostly on differences between specimens from the same rock sample with alpha doses greater or less than ∼2 × 1018 α/g (Hurley, 1952; Holland, 1954; Hurley and others, 1956; Reiners, 2005). At doses higher than ∼2 × 1018 α/g, zircon He dates in these studies become systematically younger with increasing damage. Nasdala and others (2004a) proposed this was likely due to the extensive inter-connection of boundaries between crystalline and amorphous domains at moderate degrees of radiation damage, at damage levels beyond the first percolation point as proposed by Salje and others (1999), which opens up a three-dimensional network of pathways for He migration. Subsequent modeling by Ketcham and others (2013) indicated that the important percolating phase may be damage from spontaneous fission, as alpha recoil damage percolation occurs at two orders of magnitude lower alpha dose. Reiners (2005) attempted to match the progressively younger dates with the trend line of decreasing fraction of remaining crystallinity as determined by another percolation-based model that accounts for the double-overlapping of damage cascades (Weber and others, 1994). This approach had only limited success, however, as only one dataset conformed to this model. This suggests a more sophisticated understanding of the damage-diffusivity relationship at high amounts of damage is required.

In contrast to the relatively well documented behavior at high extents of damage in zircon, little attention has been paid to potential damage-diffusivity relationships at low degrees of radiation damage, where the effects on He diffusion may be quite different. In apatite, for example, He diffusivity decreases with increasing damage (Shuster and others, 2006; Flowers and others, 2009; Gautheron and others, 2009; Shuster and Farley, 2009). This has been interpreted as a result of preferential partitioning (that is, “trapping”) of He in damage zones, impeding diffusion. This is manifest as positive correlations between apatite He date and (eU), which, among specimens from a sample that experienced a common time-temperature (t-T) history, is a proxy for relative extents of radiation damage.

In this study, we found that zircon He dates from some geologic settings also display positive date-eU correlations, which we interpret to be a result of damage at low alpha doses. We also observe negative date-eU correlations and, in some instances, both types of correlations may be present in a single sample. Throughout this paper, we use the span in eU concentrations for zircons from the same sample as a first-order proxy for each zircon's degree of radiation damage. Although more direct measurements of radiation damage (for example, Raman spectroscopy) would be ideal, for cases where such data are lacking and where zircons share a common thermal history (that is, all zircons experienced any annealing at the same time), date-eU correlations manifest the effect that damage has on He diffusivity. A quantitative explanation of these correlations requires a new, damage-based model for He diffusion in zircon. To develop this model, we conducted a series of step-heating diffusion experiments on zircons with self-irradiation doses spanning nearly three orders of magnitude (∼1016 to 1019 α/g). With the kinetics from these experiments, we parameterize a relationship between alpha dose and radiation damage, and He diffusivity, that accounts for decreases in diffusivity at low damage and increases in diffusivity at high damage. Finally, similar to the apatite radiation damage accumulation and annealing model (RDAAM, Flowers and others, 2009), we combine this new damage-diffusivity parameterization with a damage annealing model to use various date-eU correlations to constrain candidate t-T paths. Our new zircon damage and annealing model both explains zircon He datasets that do not conform to the canonical kinetics of Reiners and others (2004), and allows geologists to use zircon He date-eU correlations to place additional constraints on a sample's t-T history.

METHODS

(U-Th)/He Dating

To demonstrate correlations between zircon He date and eU that we interpret as resulting from radiation damage effects on He diffusivity, we show results from samples from both newly reported and previously published zircon He datasets. Zircon He dates from the Sri Lankan dataset were reported previously by Nasdala and others (2004a) while new data include samples from the Miocene Marnoso-Arenacea Formation in the Italian Apennines, sedimentary and basement units associated with the US Cordillera in Utah and Wyoming, and meta-sedimentary rocks from the Cooma Metamorphic Complex in Australia. Our Utah sample resides in the hanging wall of a major Sevier belt thrust sheet (Absaroka). Balanced cross-sections from this location suggest that this sample has undergone kilometer-scale tectonic and sedimentary burial (DeCelles, 1994). The other western US sample is a well sample from La Barge, Wyoming, that was collected in the footwall of the Hogsback thrust at a depth of ∼4 km.

We report single-grain zircon He dates, which were analyzed by a number of researchers over the last decade at both Yale University and the University of Arizona. Analytical methods were similar for all samples. Mineral separation followed standard crushing, sieving, and magnetic and density separation procedures. Analysts used methods described by Reiners (2005) that included Nd:YAG and CO2 laser heating, cryogenic purification, and quadrupole mass spectrometry for 4He analysis, and isotope high-resolution inductively coupled plasma spectrometry for U and Th analysis. Results are reported using the method of Hourigan and others (2005) for the alpha ejection correction.

4He Diffusion Experiments

Sample selection.

For new diffusion experiments we selected samples based on three criteria: 1) the sample was large enough to allow preparation of crystallographically oriented slabs with high aspect ratios (∼10), 2) U and Th concentrations were uniform and slabs came from zircon interiors so as to avoid alpha ejection loss or diffusive rounding of He, and 3) the extent of structural radiation damage was characterized by Raman analyses. In addition to characterizing the degree of damage, we also calculate the alpha dose of each sample.

Because we desire samples with both a rapid high temperature to low temperature cooling history, and samples whose radiation damage can be measured by some direct method (Raman, IR spectroscopy, TEM, et cetera), we have selected several zircon specimens (RB140, BR231, M127, G3, and N17) from the Sri Lankan dataset of Nasdala and others (2004a). The geo- and thermochronologic characteristics of these zircons, as well as their structural damage, have been well characterized with a number of different techniques. Results suggest that they have all experienced a similar, probably geologically rapid, cooling event at about 420 to 440 Ma. As a check for uniform U and Th distribution in our Sri Lankan zircons, we rely on the backscattered electron and cathodoluminescence images reported by Nasdala and others (2004a), which demonstrated that all of the zircons from the Sri Lankan suite possess little or no U and Th zonation (images for M127 were not detailed by these authors, but this zircon was subjected to the same analyses and yielded similar results). We also note that the Raman spectra for the Sri Lankan zircons were uniform, further evidence against zonation. We calculated alpha doses using the previously reported zircon He dates for each sample (Nasdala and others, 2004a), except in the case of N17, which—due to the fact that it loses significant amounts of He at room temperature—was calculated using a date consistent with the other Sri Lankan samples (430 Ma). Our calculated doses range from 4.7 × 1017 to 8.2 × 1018 α/g (table 1).

View this table:
  • View inline
  • View popup
  • Download powerpoint
Table 1

Zircon slab data

We also measured diffusion properties on a zircon specimen from the Mud Tank carbonatite in Australia, to provide an example with relatively low amounts of damage. We calculate an “effective alpha dose” using an age of 300 Ma, which corresponds to the timing for regional exhumation associated with the Alice Springs orogeny as determined from Rb-Sr dates on biotite and apatite fission track dates (Green and others, 2006). Mud Tank has been used previously in He and Pb diffusion studies (for example, Cherniak and Watson, 2003; Cherniak and others, 2009) and adequately satisfies our other two selection criteria. Uniform U and Th concentration was checked in part through the use of Raman spectroscopy, detailed below.

Sample preparation, Raman spectroscopy.

As a direct quantification of radiation damage, we use the full width at half-maximum (FWHM) of the v3(SiO4) Raman band near the 1000 cm−1 Raman shift (for example, Nasdala and others, 2001). This FWHM broadens from initially <2 cm−1 for well crystallized to >30 cm−1 for severely radiation-damaged zircon (Nasdala and others, 1995). Nasdala and others (2004a) obtained FWHM numbers for the Sri Lankan samples used in our study, and we include these values in table 1.

FWHM numbers for the Mud Tank zircon have not been previously reported and we therefore measured new Raman spectra for this zircon. As a check for possible damage annealing caused by the step-heating experiments, we also measured Raman spectra on pieces of M127 after step-heating and after the final degassing by laser heating. We obtained several Raman spectra at room temperature with a dispersive Horiba Jobin Yvon LabRAM HR 800 spectrometer. This system was equipped with an Olympus BX41 optical microscope, an Olympus 100× objective (n.a. = 0.90), a diffraction grating with 1,800 grooves/mm, and a Si-based, Peltier-cooled charge-coupled device (CCD) detector. Spectra were excited with the He–Ne 632.8 nm emission (3 mW at the sample). We calibrated the spectrometer using the Rayleigh line and neon lamp emissions. The wavenumber accuracy was better than 0.5 cm−1, and the spectral resolution was determined at ∼0.8 cm−1. Band fitting was done after appropriate background correction, assuming Lorentzian-Gaussian band shapes. We corrected our measured FWHMs for the experimental band broadening (that is, apparatus function), and real FWHMs were calculated according to the simplified procedure of Dijkman and van der Maas (1976). Total uncertainties of corrected FWHMs are assessed to vary between ±0.4 cm−1 (FWHM values smaller than 6 cm−1) and ±1.2 cm−1 (FWHM values of ∼20 cm−1).

Sample preparation, slab orientation.

To control for crystallographic direction, we oriented each millimeter-scale sample with single-crystal X-ray diffraction analysis. Samples were attached individually to a glass fiber and placed in a Nonius Kappa CCD diffractometer. Ten frames with a step width of 2° were taken with Mo–Kα radiation. We registered several hundred Bragg reflections, which was more than sufficient to determine the sample's crystallographic orientation. A small glass capillary was oriented parallel to the sample's crystallographic [001] direction and then glued onto the specimen. For the grinding and polishing process, we attached our samples to a glass slide, with the glass capillary oriented either parallel or perpendicular to the slide. The attachment was done with an acetone-soluble glue that hardens, and can be dissolved, at room temperature (UHU hart). After the top polished side was finished, we detached our samples from the glass slide, turned them over, and attached them again, to produce plane-parallel, doubly polished slabs. Temperatures never rose above ∼40 °C throughout the entire preparation process. The slab thicknesses (in the range of 40-110 μm) were chosen, depending on slab sizes, to get aspect ratios of 10:1 or higher. This process produced two oriented slabs per sample, one in the c-axis parallel direction (PAR_C), the other c-axis orthogonal (ORT_C).

Step-heating experiments.

We conducted our diffusion experiments on a He extraction/measurement line at the University of Arizona and used the cycled, step-heating procedure and projector-bulb furnace setup of Farley and others (1999). Slabs were held isothermally for durations between 10 and 1590 minutes, and the gas released by each step was cryogenically purified and analyzed for 4He with a quadrupole mass spectrometer. In general, we maintained a similar time-temperature schedule for all slabs: an initial low temperature step at 150 °C, followed by a prograde series of steps in 10 degree increments to 500 °C, followed by a retrograde series of steps to 265 °C, and a final prograde cycle back up to 500 °C. Due to differences in slab size, and in order to release more than just a few percent of gas, some deviations in the length of certain temperature steps were necessary. Time-steps on the initial retrograde cycle often varied and some slabs required additional cycling between 400 and 500 °C. The schedules for samples G3 and N17 involved lower maximum temperatures (383 °C and 270 °C, respectively) and several short time steps (10-30 minutes) because of their high diffusivities. After the step-heating extractions, we completely degassed each sample by laser heating to measure the remaining fraction. A significant fraction of gas was accidentally pumped away and lost during the final degassing of one sample, M127_PAR_C. As such, we calculate the total amount of gas for this sample using the measured U and Th concentration, the zircon He date, and our measured slab dimensions. The same calculation from nearly all of our other samples agrees with observed releases within a few percent. We include this sample in all subsequent tables and figures.

RESULTS

Zircon He Dates: Positive and Negative Date-eU Correlations

We report zircon He dates for all previously unpublished samples in table 2. These data are plotted in figures 1 and 2 and show positive, negative, and sometimes both types of correlations between date and eU in the same sample between date and eU. Each correlation (except for Cooma) represents a collection of single grain dates from a single igneous or sedimentary sample. Importantly, all of the grains in a given correlation have experienced the same t-T history for igneous samples, and the same post-depositional t-T history for sedimentary samples.

View this table:
  • View inline
  • View popup
Table 2

Zircon (U-Th)/He data

Fig. 1.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 1.

Positive date-eU correlations. Individual points in each dataset represent single grain dates (2 sigma error).

Fig. 2.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 2.

Negative date-eU correlations. Individual points in each dataset represent single grain ages (2 sigma error).

A comparison between figures 1 and 2 highlights several features of both types of correlations. In figure 1, the correlations are generally positive and show an increase in date with eU. In all samples the oldest dates in each one span a range from roughly 30 Ma to as great as 300 Ma, while eU concentrations are as low as ∼100 ppm, but no greater than ∼1500 ppm. In contrast, samples with negative date-eU correlations (fig. 2) tend to have older maximum dates and higher eU concentrations. These correlations include new results from the Archean basement exposed in the Bighorn Mountains, Wyoming and the Minnesota River valley, Minnesota, and placer zircons from Sri Lanka (Nasdala and others, 2004a). The oldest dates in each of these samples range from approximately 300 Ma to nearly 1.0 Ga and are almost all significantly older than any dates shown in figure 1. Although concentrations of eU overlap somewhat with samples shown in figure 1, the highest concentrations in the Minnesota, Sri Lankan and Big Horns samples are all greater than 2000 ppm, and a few Sri Lankan grains are greater than 4000 ppm. We also observe differences in the shape of the negative correlation for each sample: some are broadly continuously negative (for example Minnesota, Bighorns), whereas others appear to have a date plateau followed by a steep decline at high damage amounts (for example Sri Lanka). The rollover from reproducible dates to negative trends begins at different eU concentrations in each sample, which further suggests that each correlation has a unique form. Similar to the Sri Lankan grains, a composite sample, Cooma, shows a drop off in dates at a threshold eU; however, instead of a date plateau this appears to have a slight positive correlation at lower eU concentrations. Thus, we observe positive, negative, and sometimes both types of date-eU correlations in certain samples. In general, negative correlations occur in samples with old maximum dates (100-1000 Ma) and high eU concentrations (>2000 ppm), and positive correlations occur in samples with young maximum dates (10-100 Ma) and low eU concentrations (10-1500 ppm).

Raman Spectroscopy

The Mud Tank zircon yielded measured Raman FWHMs in the range of 2.0 to 2.3 cm−1. These FWHMs refer to the main SiO4 stretching band, which was observed at 1008.2 to 1008.4 cm−1. After mathematical correction for the artificial band broadening due to the spectrometer's limited spectral resolution, we transformed the measured FWHM values to real FWHMs of 1.7 to 2.0 cm−1. Both parameters are indistinguishable from Raman values of synthetic ZrSiO4 of 1008.3 cm−1 Raman shift and 1.8 cm−1 FWHM (Nasdala and others, 2002). Consequently, the Mud Tank material represents an extremely low degree of radiation damage, which is close to, or even below, the detection sensitivity of the Raman technique. Mud Tank's low damage is further supported by a low calculated alpha dose of 1.2 × 1016 α/g and a broad-band yellow cathodoluminescence (Nasdala and others, 2004b), which is only observed at extremely low defect densities (Nasdala and others, 2011).

After step-heating, the FWHM for M127 was 11.2 to 13.2 cm−1 observed at 1001.5 to 1002.5 cm−1. Compared to the published results for unannealed M127 (13.7-14.7 cm−1 at 999-1000 cm−1), these values represent a minor degree of annealing and suggest that our standard heating schedule is not substantially annealing the amount of damage in our samples. Following final laser heating, the FWHMs were lowered to ∼2.0 cm−1 at ∼1008 cm−1, which is close to the values for synthetic, undamaged zircon.

Diffusion Experiments

Results of step-heating experiments are shown in table 3 and as Arrhenius trends in figure 3. For the Arrhenius trends, we use the fractional gas loss equation for a plane sheet geometry to calculate D/a2 values at each temperature step (Fechtig and Kalbitzer, 1966). A striking feature in all of these plots is the non-linear behavior of diffusivities in the initial prograde temperature steps. Other studies have observed such behavior as well (Reiners and others, 2002; Reiners and others, 2004) and this non-linearity often manifests as a convex-up curve. In RB140, BR231, and M127, the curve is positioned above the linear Arrhenius trend, whereas in Mud Tank, G3, and N17 the curve is positioned below the linear Arrhenius trend. As temperature increased, however, the trend became linear after the first tenths to couple of percent of gas was released in nearly all samples (except for N17). This is apparent in a plot of ln(a/a0) as a function of cumulative gas released (fig. 4). Furthermore, as was observed in previous studies (Reiners and others, 2002; Reiners and others, 2004), this behavior seemed to disappear after the highest temperatures were reached in the initial prograde path and was almost completely absent in all subsequent steps (fig. 4, see table 3 for corresponding fraction degassed). Interestingly though, a subtle return to this type of non-linear behavior was apparent in the lowest retrograde temperature steps of some slabs.

View this table:
  • View inline
  • View popup
Table 3

Step heating results

Fig. 3.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 3.

Arrhenius plots for various zircon slabs. In both plots, non-linear trends are observed in the initial temperature steps (white markers). Linear regression for obtaining kinetic parameters performed on steps following the high-temperature reached in the initial prograde path (gray markers).

Fig. 4.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 4.

Ln(a/a0) plotted as a function of cumulative fraction of He released in step-heating experiments. This term describes the deviation of D/a2 at any given time step from the D/a2 determined from linear regression. See Reiners and others (2004) for the derivation. The inset is a magnified version of the main plot. The rectangular box in the main plot indicates the span of this inset. Despite the initial high deviation in ln(a/a0) at low cumulative fraction released, almost all samples approach a value of 0 (no deviation) after roughly the first percent of gas is released.

To derive kinetic parameters for each slab we linearly regress all steps following our initial highest temperature step. Table 4 shows the activation energy (Ea) and frequency factor/diffusion dimension (D0/a2) parameters we obtain from these post-high temperature steps assuming a plane sheet geometry. We also calculate frequency factors (D0) using the half-width measurement for each slab and these are listed in table 4 as well. With the exception of one slab (Mud Tank PAR_C), values of Ea for all crystalline slabs show a relatively restricted range (155 to 172 kJ/mol), compared with the six order of magnitude range in frequency factors (5.03 × 10−3 to 146 cm2/s). This large span in D0 values is shown in figure 5. The ORT_C samples (oriented orthogonal to c-axis and dominated by c-axis parallel diffusion) in this study approximate a (log-log) linear relationship between D0 and alpha dose. Most previously published D0 values are also reasonably consistent with this trend except for one sample from Reiners and others (2002) (98PRGB4) and the two samples analyzed in Wolfe and Stockli (2010) (ZKTB4050 and ZKTB1516). Interestingly, D0 values for the PAR_C samples in this study remain constant over much of the same range in alpha dose, as figure 5 highlights.

View this table:
  • View inline
  • View popup
  • Download powerpoint
Table 4

Kinetic parameters from step-heating experiments

Fig. 5.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 5.

Pre-exponential factor (D0) versus radiation damage for samples analyzed in this study, and those previously published. Dashed lines have been added to highlight trends. Our ORT_C samples (all samples from this study with diffusion parallel to c-axis), as well as most of the published results, fall along a trend of decreasing D0 with increasing radiation damage. The PAR_C samples (all samples from this study with diffusion orthogonal to c-axis), however, are constant with increasing damage. Also note that these two sets of values become similar at damage levels equivalent to those of M127. The previously published data mostly falls along one of these trends. We represent the data that agrees with our ORT_C samples in light gray triangles (FCT and 98PRGB18 from Reiners and others, 2002; and 1CS15 and M146 from Reiners and others, 2004), and the data that agrees with our PAR_C samples in black triangles (98PRGB4 from Reiners and others, 2002; and ZKTB4050 from Wolfe and Stockli, 2010). The gray triangle without a black border represents a single point that does not fall along either trendline (ZKTB1516 from Wolfe and Stockli, 2010). Error bars are for 1 sigma error, if reported (errors for some samples are not available).

Figure 6 shows Arrhenius trends for each sample using the kinetics derived above. The difference in diffusivity between the oriented Mud Tank slabs is ∼1 order of magnitude at the same temperature, but this pair has diffusivities that are roughly 3 orders of magnitude greater than the BR231, RB140, and M127 slabs. The RB140 pair of slabs also has only ∼1 order of magnitude difference in diffusivities at the same temperature. Another pair of slabs, M127, shows almost no difference between the two directions (∼0.1 log units). The parallel oriented RB140 slab and the orthogonally oriented RB140 slab have a difference in their diffusivities that is ∼1 order of magnitude at the same temperature. In contrast, at higher alpha doses, the diffusivity of G3 (also orthogonally oriented) is roughly 6 orders of magnitude greater than that of BR231, and N17's diffusivity is approximately 10 orders of magnitude greater.

Fig. 6.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 6.

Comparison of Arrhenius trends for samples shown in figure 4. Kinetic parameters used in this comparison are from post-high temperatures steps (blue markers in fig. 4). We have converted from values of D/a2 to D using the half-width of each slab.

The relationship between alpha dose and diffusivity is more clearly demonstrated in figure 7, which shows diffusivity at a constant temperature (180 °C) as a function of alpha dose. In figure 8, we plot Tc as a function of alpha dose, an alternative, but equally effective visualization of the damage-diffusivity relationship as Tc combines both kinetic parameters into a single value that is more intuitive to practitioners of thermochronology. For completeness and comparison, we also include previously published results from unoriented zircons in figures 7 and 8 and calculate alpha doses from each sample's zircon He date (FCT, 98PRGB18, and 98PRGB4 from Reiners and others, 2002; 1CS15 and M146 from Reiners and others, 2004; and ZKTB4050 and ZKTB1516 from Wolfe and Stockli, 2010). Unfortunately, these previous studies did not control for the degree of radiation damage in each sample and we can only estimate damage when plotting these results. For samples with either simple thermal histories (FCT and 1CS15) or independent constraints on radiation damage (M146), alpha dose values derived from a zircon He date adequately describe the accumulated self-irradiation damage since damage was last annealed. However, for one sample, 98PRGB18, we used the U-Pb date to calculate alpha doses, which is appropriate given its thermal history.

Fig. 7.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 7.

Plot of He diffusivity versus alpha dose for samples described in figures 4 and 6, as well as samples whose kinetic parameters are already published. Temperature is held constant at 180 degrees Celsius. Gray triangle represent single sample from Wolfe and Stockli (2010) (ZKTB1516) that does not fall along our observed trend.

Fig. 8.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 8.

Plot of closure temperature versus alpha dose for samples described in figures 4 and 6, as well as samples whose kinetic parameters are already published. Closure temperatures were calculated using a spherical geometry, diffusion domain size of 60 microns, and cooling rate of 10 °C/my. Gray triangle represents single sample from Wolfe and Stockli (2010) (ZKTB1516) that does not fall along our observed trend.

Sample 98PRGB18 comes from a relatively shallow part of the Gold Butte block in Nevada, an ∼15 km section of Mesoproterozoic crystalline rock that was rapidly exhumed by normal faulting at 15 to 16 Ma. This sample resided at only ∼90 °C prior to Miocene exhumation (Reiners and others, 2000) and most likely never experienced temperatures high enough to fully anneal its radiation damage. Furthermore, the U-Pb derived alpha dose values for 98PRGB18 are generally consistent with some (but certainly not all) of the Raman spectra measured on different zircons from the same sample. Given internal zonation, radiation damage in this crystal is heterogeneous and we do not report a single value for FWHM. Instead, these values range from 3.6 to 16.5 cm−1 with a mean of 7.5 cm−1 (1001.5-1007.0 cm−1 shift). The higher values in this range are consistent with high amounts of radiation damage and it is possible that the diffusion data for 98PRGB18 comes from similar heavily damaged zircons.

Despite a less than ideal spread, these Raman data provide at least some estimate for damage in 98PRGB18, and can be compared to the Raman spectra from another sample from the same crustal block, 98PRGB4. This sample resided at much higher temperatures (likely >300 °C) prior to exhumation at 16 Ma (Reiners and others, 2000), and this date—coupled with U-Th concentrations in Reiners and others (2000)—yields an alpha dose similar to that of Mud Tank. Its FWHM values range from 2.8 to 10.9 cm−1 (1003.8-1007.5 cm−1 shift) with a mean of 5.7 cm−1 and are generally lower than 98PRGB18. Again, significant spread due to 98PRGB4's heterogeneous composition prevents us from assigning a single value for FWHM. Despite the complexity of these samples, Raman spectra give a best approximation of damage levels in 98PRGB18 and 98PRGB4 and some of these spectra are generally consistent with a zircon He date calculated alpha doses in 98PRGB4, and a zircon U-Pb date calculated alpha doses in 98PRGB18.

With these calculated alpha doses, our new diffusion data, as well as almost all previously published kinetics (with the exception of ZKTB1516), define the following relationship between He diffusion (at any T) and alpha dose: Between ∼1 × 1016 to ∼5 × 1017 α/g, diffusivity decreases by nearly three orders of magnitude. Diffusivity then increases again by as much as roughly 10 orders of magnitude at damage extents of N17 (fig. 7).

DISCUSSION

In the following sections, we develop a model for He diffusion in zircon that explains both our diffusion experiment results and the date-eU correlations in the context of the alpha dose-diffusivity relationship. We first describe an hypothesis that accounts for the physical significance of both types of date-eU correlations and provides the theoretical framework for our subsequent model derivation and parameterization. For this hypothesis we interpret positive correlations as a consequence of isolated radiation damage zones acting as impediments to He diffusion by increasing the tortuosity of diffusion pathways. In contrast, we interpret negative date-eU correlations as a result of interconnection of damage zones at moderate to high alpha doses (>2 × 1018 α/g), as is qualitatively consistent with previous observations of increased He diffusion in highly damaged zircon (Holland, 1954; Hurley, 1954; Nasdala and others, 2004a).

Positive Date-eU Correlations

Due to similarities between our positive correlations and those observed with the apatite (U-Th)/He thermochronometer (for example, Flowers and others, 2007; Flowers and others, 2009; Flowers and Kelley, 2011), the effects of radiation damage on He diffusion in apatite provide context for interpreting zircon He positive correlations. Shuster and others (2006) showed that He diffusivity in apatite decreased with increasing damage and subsequent studies (for example, Flowers and others, 2007) supported this conclusion with observations of positive date-eU correlations. We suggest that similar behavior occurs in zircon and leads to positive date-eU correlations. For example, in the context of a thermal history like that depicted in figure 9, grains with different amounts of radiation damage, as well as an initial span of uniform dates, are reset to varying degrees during a reheating event (fig. 9A). If eU is a proxy for the total accumulated radiation damage (that is, all the zircons in a sample have experienced the same t-T history), then those grains with low eU lose a larger fraction of their He, resulting in a younger date, than grains with high eU, leading to a positive correlation. Samples that have undergone slow, monotonic cooling may also exhibit positive correlations. If a sample spends a significant amount of time at temperatures low enough for damage accumulation without annealing, but high enough to be in the PRZ, then arrays of zircon He dates may form a positive correlation (fig. 9A). Both scenarios demonstrate that this correlation results from a sample with grains that span a range in eU, and have resided in the PRZ after disparate amounts of damage have accumulated in those grains.

Fig. 9.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 9.

Schematic of possible time-temperature paths that produce date-eU relationships. The arrows on each plot's axes indicate the direction in which the given value increases. Varying eU contents lead to differential accumulations of radiation damage, which in turn lead to differential He diffusivities. (A) If accumulation is relatively low, and a sample experiences a thermal pulse, or cools slowly through the PRZ so that damage in-growth and diffusion happen simultaneously, then a positive correlation results. (B) Conversely, if damage accumulation is relatively high (due to older zircons), then a thermal pulse results in a negative correlation, or significant He loss may occur at low temperatures.

A key difference between our interpretation of the damage-diffusivity relationship in zircon and the interpretation for the apatite system, though, is the mechanism that we suggest causes diffusivity to decrease with increasing damage. In the apatite He system, the decrease in He diffusivity is hypothesized to result from accumulation of crystal defects caused by alpha recoil damage that act as He traps. These traps sequester He (governed by an equilibrium partition coefficient) and prevent or slow its diffusive migration out of the crystal (Farley, 2000). In contrast to He diffusion in apatite, various authors (Farley, 2007; Reich and others, 2007; Saadoune and others, 2009) have demonstrated that He diffusion in an ideal or defect-free zircon should occur almost solely along c-axis parallel channels [0 0 1]. Any disruption of these pathways would force He through c-axis orthogonal openings ([1 0 0], [0 1 0], and [1 0 1]), which are much less energetically favorable (Reich and others, 2007). We argue that decreases in He diffusivity in zircon are largely due to the increasing disruption of diffusion fast-paths (c-axis parallel channels) by radiation damage, an effect similar to road blocks being placed on a major highway. As these barriers are erected inside the zircon, a He atom's path becomes more tortuous and the effective diffusivity of the grain decreases. Evidence for this increasing disruption of c-axis parallel channels comes from figure 5. The D0 values for orthogonal oriented slabs (diffusion predominantly in the c-axis parallel direction) decrease across the damage spectrum whereas the D0 values for the parallel oriented slabs (diffusion predominantly in the c-axis orthogonal direction) remain the same, with both sets of D0 values becoming similar at high damage. In other words, diffusion kinetics in the orthogonal direction begin to more closely resemble diffusion kinetics in the parallel direction with increasing damage, and this increasing similarity can be explained by tortuosity. Tortuosity may also contribute to lowering He diffusivities in apatite, but we envision this phenomenon is more important for He diffusion in zircon due to its strongly anisotropic behavior in specimens with little or no accumulated damage. We do not rule out that damage zones in zircon may also trap some amount of He, but we suggest that this effect is secondary compared to the closing of preferred diffusion directions, which are probably not present in apatite.

Negative Date-eU Correlations

Like the positive date-eU correlations, previous research provides some context for interpreting our negative date-eU correlations. Various authors (Holland, 1954; Hurley, 1954; and Nasdala and others, 2004a) have suggested that He diffusivities increase abruptly once zircon reaches a certain threshold of radiation damage. As damage increases, Reiners (2005) proposed that zircon He dates scale with the remaining crystalline fraction of the zircon as determined by the double-overlapping cascade model (Gibbons, 1972). This further suggests that above a threshold, interconnected damage zones form through-going channels in the zircon lattice and create fast diffusion pathways for He. In order to reach this interconnection or percolation threshold, zircons must sustain long-term damage accumulation at temperatures low enough to prevent annealing. Zircons may achieve a heavily damaged state in which either significant He loss and resetting occurs at surface temperatures, or some brief, low-temperature reheating event may cause resetting (fig. 9B). In detail, some amount of less diffusive material must remain in heavily damaged zircons as most negative correlations are gradual (that is Minnesota River Valley, Bighorn suite) and not so abrupt. However, in general, both scenarios could result in negative correlations and are plausible explanations for the datasets plotted in figure 2. Two of these samples, the Minnesota River Valley and Bighorn suites, come from Archean rocks that have likely been within a few kilometers of the surface for 108 to 109 years and have consequently accumulated large amounts of radiation damage. The Sri Lankan zircons are not as old but some do have high eU, which in some cases resulted in the complete breakdown of crystal structure (for example N17, Nasdala and others, 2004a).

The Sri Lankan dataset also shows a potential percolation threshold effect whereby dates are fairly reproducible up to a critical eU concentration, in this case about 2000 ppm eU, above which they decrease with increasing eU. The Cooma date-eU trend reinforces the notion of a transition in diffusion behavior as it displays both a positive and negative correlation with an abrupt transition between each. Damage in-growth since granulite facies metamorphism at ∼433 Ma (Williams, 2001) and a large disparity in eU concentration (∼100-1300 ppm) produced a suite of zircons in which two different processes affected He diffusion in different parts of the eU spectrum. In this particular dataset, the transition between the two types of He diffusion occurs at an eU of ∼1000 ppm. Similarly, nearly all zircons in our positive correlations contain eU concentrations below 1000 ppm. The Cooma dataset serves as an important demonstration of how radiation damage may have contrasting effects on He diffusivity in a suite of grains from a single sample, and shows the approximate concentration of eU over which a transition from one process to the other may occur. Because both types of diffusion mechanism may operate on the same sample, these samples underscore the necessity for understanding the damage-diffusivity relationship across the entire damage spectrum.

Arrhenius Trends

There are a number of possible explanations for the non-linear portions of the Arrhenius plots in figure 4. Some of these plots show concave-up trends in the initial prograde steps, in which less than ∼1 to 2 percent of the He is released. In contrast, subsequent temperature steps show much less variation and are more nearly linear. These concave-up trends can either plot above or below the linear Arrhenius trends. This feature, or ones similar to it, has been observed previously in zircon (Reiners and others, 2002; Reiners and others, 2004) and several other minerals, including titanite (Reiners and Farley, 1999), goethite (Shuster and others, 2005), magnetite (Blackburn and others, 2007), and apatite (Farley, 2000). It has been attributed to inhomogeneous He distributions due to zonation, alpha ejection, or initial diffusion rounding; surface roughness; multiple diffusion domains; and radiation damage (for example, Reiners, 2005).

Despite the myriad possible causes, samples with concave-up curves that lie above the linear trends (RB140, BR231, and M127) are still difficult to explain. Anisotropy could produce non-Arrhenius behavior of this type in step-heating results if activation energies of the contrasting diffusion directions are different (Reich and others, 2007; Watson and others, 2010; Bengston and others, 2012). However, our data, as well as all other experimental He diffusion data on zircon or zircon-structure phases (Farley, 2007; Cherniak and others, 2009), indicate that anisotropy is manifest as differences in frequency factor, not activation energies, which would not lead to significant departure from linearity on an Arrhenius plot. Furthermore, if anisotropy in zircon was caused by differences in activation energy, then large changes in slope should always be evident at low temperatures regardless of how many prograde and retrograde cycles have been performed. For the most part, we do not observe this. We therefore rule out anisotropy as the source of non-Arrhenius behavior in our initial prograde temperature steps.

Reiners and others (2002) suggested that this type of non-Arrhenius behavior could be due to the interaction between radiation damage zones and the zircon surface, with the damage zones acting as grain boundaries or fast diffusion pathways. Recent evidence of two diffusion pathways for Ar in quartz (Clay and others, 2010), as well as investigations of fast path diffusion in other minerals (Yund and others, 1981; Yund and others, 1989; Yurimoto and others, 1989; Worden and others, 1990; Hacker and Christie, 1991) are consistent with this. For example, because lattice diffusion generally has a higher Ea than grain boundary diffusion (Chakraborty, 2008), lattice diffusion can be faster at high temperatures, so diffusion from grain-boundary-like domains could dominate gas release in early, prograde steps, yielding initially high diffusivities in Arrhenius plots. With higher temperature steps increasing fractions of gas would derive from lattice diffusion as a result of the difference in values of Ea, but also because surficial grain-boundary-like sites would be rapidly depleted in the initial few tenths to one percent of gas released (fig. 4). If grain-boundary-like sites were reoccupied with He during high temperature steps, this could also explain persistent release of gas via grain boundary diffusion in later steps. Arrhenius trends for M127, Mud Tank, RB140, and BR231 appear to exhibit slight curvature at the lowest temperatures of the post-high temperature heating steps (after initial 500 °C step), consistent with this explanation.

Concave-up curves that lie below the linear Arrhenius trends (Mud Tank, G3, and N17) probably result from a combination of the behavior described above and initially rounded He profiles. N17, with a closure temperature below 0 °C, likely possessed a rounded concentration profile prior to being step-heated, and, despite our selection of interior parts of the Mud Tank and G3 zircons, some portion from the rounded diffusion profile of these two samples was included as well. Although the initial He released from these zircons does not appear to conform to ideal expectations of simple Arrhenius behavior, its effect after the first few percent of gas release (except for N17) is negligible.

Functional Form for Damage-Diffusivity Relationship

For the remainder of our discussion, we directly relate alpha dose to structural damage, and derive a mathematical parameterization of the damage-diffusivity relationship that fits the data in figure 7. This requires a number of assumptions. As previously stated, we assume that alpha doses calculated from He ages sufficiently reflect the total accumulation of alpha-decay events since structural damage was last annealed. Because most of our diffusion samples have thermal histories involving a phase of relatively rapid cooling, and because we have direct and independent measurements of a proxy for structural damage (FWHM) for a subset of them, this is a reasonable assumption.

We must also assume that He diffusion in zircon scales with both alpha dose and structural damage in the same way. This is an important consideration as alpha dose is a calculated damage proxy and not a direct measure of damage (we detail below the reasons for using alpha dose). We again rely on observations from the apatite He system to support this assumption. Shuster and Farley (2009) showed that the amount of ionizing kinetic energy released into a sample (kerma) and the fission-track density had similar effects on He diffusivity. They demonstrated that He diffusivity in apatite changes systematically with both increasing and decreasing kerma (through annealing) caused by either artificial irradiation or natural, self-irradiation as monitored by fission track density. Flowers and others (2009) expanded on this and derived a term, effective spontaneous fission track density, that directly related alpha dose, accumulation and annealing of fission tracks, and He diffusivity. Unfortunately, experimental observations analogous to those of Shuster and Farley (2009) are currently lacking for zircon. But both of these studies suggest that, at least to first order, fission track density (which is itself a measure of one type of structural radiation damage) scales with “effective alpha dose,” and can therefore be related to bulk He diffusivity.

Finally, we are assuming that the negative correlation between alpha dose and diffusivity at low doses is not due to the effects of He (or Pb) concentration on He diffusivity. Shuster and Farley (2009) clearly showed that radiation damage, not He concentration, controls diffusivity in apatite. But as stated above, the types of experiments performed by Shuster and Farley (2009) do not yet exist for zircon, so we cannot completely rule out a concentration-dependency for He diffusion. However, the data of Shuster and Farley (2009) provide some confidence that similar processes are occurring in zircon, and this could be confirmed in the future by applying their methodology to zircon samples.

Although these assumptions are required, we suggest that linking alpha doses (preferentially those calculated using FT or He ages) to radiation damage is an appropriate choice for our purposes as these assumed “effective alpha doses” provide a straight-forward method for calculating damage accumulation through time. Furthermore, as we detail below, it provides a crucial link between equations describing He diffusivity and those describing damage annealing. The alpha dose is therefore the best measurement for integrating He diffusion, damage accumulation, and damage annealing over geologic timescales into an easily accessible thermochronometric modeling tool, which is our primary objective in this section. With this in mind, we first derive a parameterization for the damage-diffusivity relationship and then integrate this parameterization into a He diffusion and damage annealing numerical model.

A mathematical description of the damage-diffusivity relationship must account for both the initial decrease and ultimate increase in diffusivity across the spectrum of damage from at least Mud Tank to N17 (∼1016-1019 α/g). These two samples represent the damage range encountered for the vast majority of zircon crystals sampled at or near the Earth's surface. We desire a functional form consistent with the hypothesis that decreasing diffusivity at low damage is caused by accumulation of isolated damage zones that block crystallographically preferred He transport pathways and increase the tortuosity of He migration. At high alpha doses, increasing diffusivity would be due to decreasing effective domain size of undamaged zircon volumes, which are increasingly separated by interconnected fast-diffusing damage zones.

In order to explain decreasing diffusivity at low damage, we introduce an effective diffusivity (De) that represents He diffusion in a damage-free lattice modified by increasing tortuosity. Increasing radiation damage blocks or constricts easy He migration paths (for example, c-axis parallel channels) forcing He to take a more tortuous path out of the zircon. This is analogous to diffusion in a porous medium where effective diffusivity is expressed as (Cussler, 1984): Embedded Image Dz is the diffusion coefficient within the pores, or in our case, diffusion along c-axis pipes in a pristine zircon, and τ is the tortuosity. We represent Dz with the diffusion kinetics from a minimally damaged zircon. For this zircon's D0, we fit the ORT_C slabs in figure 5 with a power-law relationship that yields an equation of y = 134.89*x−1.578, where y is D0 and x is dose. Projecting this relationship down to 1 × 1014 α/g gives a D0 of 193188 cm2/s (all constants and their values described in the remaining text are listed in table 5). For the Ea, we average the activation energies from the samples with minimal amorphous fractions (all samples excluding G3 and N17) and the previously published results. Although these parameters are both extrapolations, the following equations could be easily modified to account for future diffusion data from less damaged or more appropriate zircon specimens.

View this table:
  • View inline
  • View popup
  • Download powerpoint
Table 5

Constants and values used in parameterization

Tortuosity τ could be represented in a variety of ways, and the atomic-scale processes by which radiation damage may affect migration pathway dimensions and diffusivity are complex. Damage may displace atoms into open channels that, depending on which species is displaced, could cause variable decreases in porosity. C-axis channels of initially high ionic porosity could be completely blocked and a larger fraction of the He migration path would be forced to occur orthogonal to the c-axis. Although the exact geometry or porosity of damage zones is hard to constrain, we predict that most zones should act as He barriers and τ should increase as the chance increases for diffusing He atoms to encounter these barriers. To model this behavior, we use a metric introduced by Ketcham and others (2013) for characterizing the undamaged portion of the lattice, mean intercept length, lint, which is the average distance a particle can travel in a single direction without encountering a damage zone. The expression for τ relates the calculated lint in a given zircon to the mean intercept length in our extrapolated, minimally damaged zircon, which also displays high diffusivity (lint 0): Embedded Image The right-hand side of equation (2) is squared in part to improve the fit to our diffusion data (see below), but tortuousity is often mathematically expressed as the square of pore spacing or geometry (for example, Epstein, 1989). Ketcham and others (2013) have derived an empirical relation for lint by modeling the accumulation and percolation of chains of connected, capsule-shaped alpha recoil tracks. They express lint as a function of fraction amorphous (fa): Embedded Image where SV corresponds to the surface to volume ratio of the capsules (1.669 nm−1). In turn, fa is described using the direct impact model (Gibbons, 1972): Embedded Image where Ba is the mass of amorphous material produced per alpha decay (5.48 × 10−19 g/α-event), and α is the alpha dose.

An explanation for the increase in diffusivity at high damage requires a derivation that accounts for interconnected amorphous zones. We hypothesize that at sufficiently high self-irradiation levels, amorphous zones caused by damage become interconnected and connected to the grain's surface. This represents an important shift from damage zones acting as barriers to damage zones acting as fast paths. The processes by which this might occur are not entirely apparent. Various macroscopic and long-range order properties change at high damage and track with amorphous fraction (for example, Ewing and others, 2003) and we expect that diffusivity should as well. However, several different atomistic processes may cause these changes. Devanathan and others (2006) showed that damage zones are characterized by an amorphous core surrounded by a high interstitial density rind. These amorphous cores could become connected at high damage and may form fast paths while the rinds cause increasing tortuosity at lower damage. Interconnected fission tracks, which only reach a percolation threshold at high damage (Ketcham and others, 2013) are another candidate for creating these fast paths (see below). Regardless of the exact mechanism, the net effect of this process is creation of an increasing number of progressively smaller, undamaged zones that are increasingly isolated from one another by increasingly interconnected and progressively larger damage zones with much higher diffusivity. These two effects can be accounted for by 1) modeling a decrease in the size of the diffusion domain of the undamaged portion of the grain, and 2) describing the bulk diffusivity as a harmonic average (as appropriate for an average of rates) in the undamaged and damaged portions of the grain. In detail, if the grain had a heterogeneous spatial distribution of U and Th, increasing damage could conceivably produce an apparent spectrum of diffusion domain sizes that might manifest itself in step-heating data as decreasing diffusivity or concave-up Arrhenius trends.

We first parameterize our effective diffusivity using an harmonic average and to do this, we recast equation (1) as: Embedded Image where DN17 corresponds to the diffusivity of amorphous N17, and f′c represents fraction crystalline and is equal to 1 − f′a. We again use the direct impact model to describe f′a and f′c, but to improve the fit to our diffusion data we include an additional term (φ) within the exponential: Embedded Image A value greater than 1 for φ causes f′a to increase more rapidly at lower alpha doses. As we demonstrate below, a value of 3 for φ is necessary for a proper fit to the data, however, we currently have no physical explanation for why this term should be required. It is possible that, while the direct impact model may adequately describe the buildup of amorphous zones in zircon, it does not fully account for the interconnection of certain parts of the zones (for example high-vacancy damage core vs. damage rim), or for the contribution of an additional interacting effect such as accumulation of fission tracks (Ketcham and others, 2013).

In order to account for the reduction of the domain size of the undamaged portion of the grain, we modify equation (5) by dividing each D by domain size (a): Embedded Image In this set-up, a is equal to the initial grain size in an undamaged zircon, which effectively decreases as f′a increases following equation (6).

The φ term in equation (6) and the scaling of domain size with f′a in equation (7) are admittedly somewhat empirical and heuristic, respectively, and their relationship to actual physical phenomena are tenuous. Recent work by Ketcham and others (2013) may offer some additional insight into this issue. These authors suggest that fission track interconnection may play an important, but until recently, underappreciated role in creating the macroscopic and long-range order qualities typically attributed to fa. Furthermore, they derive an equation for a term that seems to more accurately reflect the effects of radiation damage on decreasing effective domain size: mean distance to nearest fission track dn. This number decreases with increasing fission track percolation and replacing (a*f′c)2 with dn in equation (7) results in a similar functional shape. Unfortunately, doing so does not provide a better fit to the real data. The current calculations do not account for whether the nearest fission track is part of a network connected to the outside of the grain, and improving the model in this respect may result in a better fit. For our present discussion, though, we proceed with equation (7) as it is currently derived.

Our combined equation for effective diffusivity is: Embedded Image where lint 0 is equal to 45920 nm (the value of lint calculated from equation (3) at an alpha dose of 1 × 1014 α/g). Figure 10A shows equation (8) with our step-heating results. Our parameterization adequately captures the decrease in diffusivity by ∼3 orders of magnitude at low damage and the subsequent increase in diffusivity by ∼11 orders of magnitude at high damage. We also show a comparison between the Tc data from figure 8 and Tc values calculated using equation (8) and an initial grain radius of 60 microns (fig. 10B). To obtain effective Ea and D0 values for this curve, we use a method that relies on pseudo-Arrhenius trends calculated at discrete doses with equation (8). These trends provide the kinetic parameters necessary to then calculate the Tc at the corresponding dose.

Fig. 10.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 10.

(A) Diffusivity as a function of alpha dose calculated at temperatures that roughly bracket the nominal zircon PRZ. Data points are the diffusivities of the various samples included in figure 7 (calculated with a domain size of 100 microns) and curves represent effective diffusivity as defined by equation (8). The various parameters used for this plot and their values are listed in table 5. (B) Closure temperature as a function of alpha dose calculated in a manner similar to figure 8. Gray points in both plots represent single sample that does not fall along our observed trend.

Although diffusional anisotropy may manifest itself in our model through the importance of increasing tortuosity at low damage, our model does not directly account for anisotropic diffusion. As figure 6 suggests, the effect of anisotropy on He diffusivity is minimal compared to the effect of radiation damage and we have not focused on it in this section. However, our derivation makes several predictions for a relationship between anisotropy and radiation damage. With increasing damage, c-axis parallel channels might be expected to become increasingly blocked and, correspondingly, anisotropy should decrease (Farley, 2007). Specifically, Mud Tank—our least damaged, oriented zircon—should be more anisotropic than all other samples. Figure 5 shows a large disparity between the D0 values for the two Mud Tank slabs, and the two trends (constant for PAR_C slabs, decreasing for ORT_C slabs) seem to converge at sample M127. This suggests that, at least in terms of D0, anisotropic differences decrease with increasing damage. However, the results in figure 6 complicate this observation. At high temperatures (500 °C), the difference in diffusivity between the Mud Tank slabs is ∼2 orders of magnitude, but at lower temperatures the degree of anisotropy between these two slabs is comparable to the anisotropy between slabs in the “moderate to high damage” group (∼1 order of magnitude difference, note that the Mud Tank lines have different slopes in figure 6). This suggests that anisotropy is relatively invariant across the damage spectrum from Mud Tank to M127, which seems to contradict our hypothesis that damage decreases the degree of anisotropy. Perhaps the anisotropy in Mud Tank is similar to the anisotropy in M127 because Mud Tank has a large number of other defects that also contribute to decreasing anisotropy. But why then would these defects not also lower diffusivity? We do not have a satisfactory answer for this yet, but point defects may have a relatively small effect on diffusivity compared to radiation damage, due to damage's more chaotic nature. Actual characterization of this difference requires more work. For now we focus our remaining discussion on using equation (8) to forward model date-eU correlations.

Implementation of Damage-Diffusivity Parameterization

As a demonstration of the utility of our parameterization, we link equation (8) to another equation describing damage annealing in zircon and forward model date-eU correlations in a manner similar to the RDAAM (Flowers and others, 2009). Model inputs consist of an effective diffusion domain lengthscale (assumed to be radius of a sphere with an equivalent surface-area to volume ratio as the grain), and U and Th concentrations for each grain, and a discretized t-T history for the entire dataset. With these, our model calculates the total He production, damage accumulation, damage annealing, He diffusivity, and He loss at each time step.

Damage accumulation and annealing are quantified with a series of equations similar to those described by Flowers and others (2009), and rely upon the kinetics of fission track annealing in zircon. Although we treat alpha damage as the primary factor in creating tortuosity and interconnections, using a fission track annealing model means we must assume that alpha damage anneals in a fashion similar to fission tracks. This is a potential problem in the apatite system as well, but the RDAAM's ability to predict date-eU correlations from reasonable t-T histories (for example, Flowers and others, 2007; Flowers and Kelley, 2011) shows that, to first order, modeling damage annealing with fission track kinetics is valid. What works well in apatite, though, may not be suitable for zircon. Especially problematic is the well-documented observation that alpha damage annealing in zircon occurs via two disparate processes—epitaxial recrystallization and ZrO2 nano-crystal formation—at different temperatures and at different initial damage concentrations (Meldrum and others, 1998; Capitani and others, 2000; Zhang and others, 2000; Nasdala and others, 2002; Zhang and others, 2010). Furthermore, Garver and others (2005) suggested that zircon fission track annealing processes are most likely damage-dependent as well. A possible solution would be to use an alpha damage annealing model that accounts for multiple annealing processes, but despite an extensive literature on alpha damage accumulation and annealing in zircon, to the best of our knowledge, no kinetic model has been parameterized to describe alpha damage annealing at time scales beyond hours or days. If such a model is developed, equation (8) could be easily coupled to it due to our modular model design. More importantly, our main objective here is not to contrast and compare various annealing models. Rather, we simply desire a quantitative approximation of damage annealing kinetics in zircon that seems reasonable for geologic time scales. Given the currently available annealing models, a fission track model best satisfies this requirement.

In order to combine the total alpha damage produced during a series of discrete time steps with a fission track annealing model, we introduce αe or equivalent alpha dose (α/g): Embedded Image where the initial factor serves to convert from nmol to decays, α is the number of alpha decays (in nmol) per gram produced in each time step, and ρr is reduced (normalized) fission-track density of fission tracks that formed during that time step. The alpha dose from each time step (t2, t1) is calculated as: Embedded Image where [238U], et cetera are in nmol/g. Our derivation of ρr starts with length reduction r, for which we use a simplified version of the fanning curvilinear fit (for example, Ketcham and others, 2007) to the ZFT annealing data of Yamada and others (2007): Embedded Image where β = −0.05721, C0 = 6.24534, C1 = −0.11977, C2 = −314.937, and C3 = −14.2868. To convert from reduced length to reduced density (ρr), we use the relation based on data reported by Tagami and others (1990), which begins at one and is truncated at a ρ/ρ0 value of 0.36, below which there are no data: Embedded Image

Once calculated, values of αe are linked to a diffusion model via equation (8). With an estimate of the diffusion coefficient from equation (8), we then solve the diffusion equation numerically for a spherical geometry with the Crank-Nicholson finite difference scheme used in the thermal modeling software package HeFTy (Ketcham, 2005). Although this is not the best representation of He diffusion in zircon (it implies isotropic diffusion), a spherical model captures the first-order features of the damage-diffusivity relationship. Future versions could incorporate the cylindrical finite element scheme of Watson and others (2010), which accounts for the effects of anisotropy. Alternatively, a spherical calculation scheme can be employed using the “active radius” method introduced by Gautheron and Tassan-Got (2010), who described how to determine the radius of an isotropic sphere that replicates diffusive loss from a prism with a given aspect ratio and diffusive anisotropy.

Our model demonstration consists of 6 different thermal histories, each designed to capture aspects of the date-eU correlations in figures 1 and 2. As inputs, we use eU ranging from ∼50 to ∼5000 ppm, grain radii of 60 μm, and t-T paths that begin at 600 Ma and end at the present. For comparison, we also model the single zircon He date that results from using the kinetics of Reiners and others (2004) with each of our thermal histories (black diamonds in fig. 11B). These t-T paths encapsulate 6 representative scenarios: 1) slow, monotonic cooling from 600 Ma to the present, 2) early cooling followed by a pulse of early-stage reheating, 3) early cooling followed by a pulse of late-stage reheating, 4) early cooling followed by prolonged time spent in the PRZ and then subsequent late cooling, 5) long term early heating and subsequent late-stage cooling, and 6) early cooling and prolonged exposure to low temperatures (fig. 11). Model outputs that result from scenarios 5 and 6 show mostly flat date-eU correlation, which are typical of most zircon He datasets. These two scenarios also result in He dates that are almost identical to dates obtained using the kinetics of Reiners and others (2004). For thermal histories with a single, rapid pulse of cooling from high temperatures, our new model does not alter interpretations of zircon He datasets made with previously published kinetics as this style of cooling will most likely not result in a date-eU correlation. Interestingly though, scenario 6 shows a steep negative correlation at high damage, despite having never been reheated above 20 °C post-500 Ma. These zircons have entered the PRZ without changing temperature and we note that a similar process may have occurred in the Sri Lankan dataset to produce the young dates for zircons K6 and N17.

Fig. 11.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 11.

Representative thermal histories (A) with corresponding forward modeled date-eU correlations (B). Each numbered t-T path in (A) corresponds to the similarly numbered correlation in (B). Thermal history number 6 is partially obscured by 2 and 3 because it follows the same path but lacks a post-500 Ma reheating event. For reference, we also plot the zircon He dates that result from each thermal history in (A) if the kinetics of Reiners and others (2004) are used. These are represented by the numbered black diamonds on the y-axis of (B).

Scenarios 1, 2, 3, and 4 demonstrate the various thermal histories that may produce positive, negative, or both types of correlations in the same sample. The thermal history for scenario 1 is characterized by slow cooling through the PRZ and results in a broad and relatively confined positive date-eU correlation (dates increase from 161-210 Ma). Damage in-growth and He diffusion occur simultaneously over a prolonged time span, which causes the damage amounts for the zircons in this scenario to be relatively similar while He is diffusing. Given high enough eU concentrations, the t-T path for scenario 1 also produces negative correlations in the same sample. Although this particular negative correlation is subtle, an older initial age for the start of slow cooling could produce a more pronounced negative correlation, as we further demonstrate with our Minnesota dataset below. Scenario 4 is somewhat similar to scenario 1 as both samples spend prolonged time periods in the PRZ. However, for most of their history, the zircons in scenario 4 are held at a lower temperature (180 °C) than those in scenario 1. At this temperature, damage in-growth outpaces annealing and the thermal history produces a negative correlation at relatively low eU concentrations once the sample is finally cooled.

In contrast, relatively short-lived pulses of reheating and cooling result in more marked positive or negative correlations. The thermal histories for both scenarios 2 and 3 contain reheating events that take place after the zircons have been held at low temperatures (20 °C) for at least 100 my. During this time span, the zircons in each scenario have acquired large differences in damage. In scenario 2, this results in a positive correlation that spans a date range from 428 to 538 Ma, but also a negative correlation that drops to zero at the highest eU. Like scenario 6, the PRZ for zircons with eU in excess of ∼3500 ppm is at very low temperatures and they no longer retain He. In scenario 3, we have chosen both a later start time (100 Ma) and a lower maximum temperature for the reheating event than scenario 2 (130 °C as opposed to 180 °C). The resulting plateau of dates at ∼550 Ma followed by a steep negative correlation is somewhat similar to scenario 6, except the steep drop off in dates occurs over much lower eU concentrations (from ∼1500 to 2000 ppm). The late-stage reheating event also produces a small plateau of ∼50 Ma dates at eU concentrations greater than ∼2000 ppm. These dates correspond to the initiation of cooling and illustrate that, given a large span in radiation damage, multiple pulses of reheating and subsequent cooling could be recorded in zircons from the same hand sample.

As a final demonstration, we forward model the thermal history of two real datasets shown in figures 1 and 2. We use one of the Apennines datasets (AP54B) as a representative positive date-eU correlation, and the Minnesota dataset as a representative negative date-eU correlation. For each dataset, we present a realistic thermal history that could have produced the observed data and show the resulting model output in figure 12. In the Apennines example, we rely on the thermochronometric data from previous publications to guide our t-T path construction. Bernet and others (2001) obtained ZFT dates on the same grains shown in figures 1 and 12A, which yield a peak date of 20.7±3.6 Ma (these author's P1 population). This translates to a lag time of 6.8 my as the depositional age of this particular unit in the Marnoso-arenacea Formation is 13.9 Ma. Zattin and others (2002) conducted a detailed regional study of this Formation and found that fully and partially reset AFT dates from their hinterland samples (most internal to the thrust belt) suggest maximum burial temperatures of 120 to 125 °C and post-depositional exhumation between 4 and 6 Ma. If we consider these multiple t-T constraints as model inputs, then we can produce the black line in figure 12A, a positive date-eU correlation that is in good agreement with the real dataset. In figure 12A, we also show the results from a simpler thermal history (dotted lines in t-T history and date-eU correlation) in order to demonstrate that our preferred thermal history (black line) produces a distinctive date-eU trend that best fits the data. Our new model therefore tightly constrains the thermal history of a given sample and discriminates between potential t-T paths.

Fig. 12.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 12.

Forward modeled date-eU correlations matched to datasets shown in figures 1 and 2. (A) Positive date-eU correlation (detrital Apennines sample AP54B) with forward modeled correlation resulting from thermal history shown in the left-hand panel. This thermal history is characterized by a period of rapid exhumation in the source terrain (as constrained by ZFT dates from Bernet and others, 2001), deposition and burial in the Apennine foreland basin, and final exhumation to the surface in the latest Miocene. We used a grain radius of 48 microns for each model zircon, which is the average for this dataset. Dotted light gray lines in both panels represent an alternative thermal history and the corresponding date-eU correlation. (B) Negative date-eU correlation (igneous Minnesota River Valley sample) with forward model correlation resulting from the thermal history shown in the left-hand panel. This thermal history begins at the approximate end of tectonism associated with the Penokean Orogeny in this area and proceeds at a cooling rate of .06 °C/my until 1100 Ma. This time corresponds with the initiation of the Keweenaw Rift System (the Minnesota River Valley was situated near the rift shoulder) and we model this as an increase in the long-term cooling rate to ∼.17 °C/my. Again, we used the average grain radius from this dataset (52 microns) as an input for each model zircon. Dashed and dotted lines in both panels represent two alternative thermal histories and the corresponding date-eU correlations.

Our Minnesota dataset has fewer t-T constraints and our modeled t-T path is slightly more speculative. A negative date-eU correlation and an oldest date of ∼925 Ma though suggests that these zircons have experienced very slow cooling rates since the Proterozoic. With this in mind, we explored several possible slow cooling paths since the Penokean orogeny at 1870 to 1820 Ma, which is the most recent episode of regional metamorphism to affect this area (Holm and others, 1998). Our best fit to the data consists of an initial cooling event beginning at 1850 Ma and 250 °C that proceeds at a rate of .06 °C/my. In order to reproduce both the pseudo-plateau of dates at low eU concentrations and the steep negative correlation at high eU concentrations, we accelerate our cooling rate from .06 °C/my to ∼.17 °C/my at 1100 Ma, which matches the age of opening for the failed Keweenawan Rift system. Again, we include the results from a couple of simpler histories (dotted and dashed lines in fig. 12B) to show that our choice for the Minnesota thermal history is not arbitrary. Slow cooling rates with a single value yield date-eU correlations that match either the high eU or low eU trends in the real data, but not both. We find a good fit to the data only by changing the cooling rate at a specific time (in this case, 1100 Ma). Despite being somewhat speculative, our model t-T constraints for the Minnesota dataset agree with the regional geologic history, produce a reasonable fit to the data, and demonstrate that the Minnesota sample has experienced very slow cooling at low temperatures for the past 1.8 by.

Impact of eU Zonation on Zircon Date-eU Correlations

In the above models, we assumed that both our real and model zircons are homogenous in their U and Th (hence eU) concentrations. Typical zircons, though, usually possess some degree of U and Th zonation. Although previous studies have discussed the effects of parent zonation on apatite (U-Th)/He dates in depth (for example, Farley and others, 1996; Meesters and Dunai, 2002; Hourigan and others, 2005; Farley and others, 2011; Ault and Flowers, 2012; Gautheron and others, 2012), the effects of strong parent zonation in zircon may be significantly different because of the reversal in damage-diffusivity relationship with progressive damage accumulation. This means that, for some thermal histories, different parts of the same zircon grain may have extremely different behavior, both of which may be very different from that expected from a grain with equivalent bulk eU homogeneously distributed.

Farley and others (2011) detailed three ways in which heterogeneous eU in zoned apatite (lacking the damage-diffusivity reversal potential) can affect measured ages and interpreted t-T histories. If an homogenous alpha ejection correction factor (FTH) is used for a zoned grain, then the resulting He date will be either too young or too old depending on where the majority of eU is concentrated (rim or core, respectively) (Farley and others, 1996; Hourigan and others, 2005). This issue can be dealt with by using a zoned alpha ejection correction (FTZ) that accounts for alpha particle redistribution, which is an option in HeFTy. Zonation also affects He diffusivity by altering the He concentration profile, and, because He diffusion is damage dependent, by creating distinct domains with different diffusion kinetics inside the crystal. In the apatite He system, these three factors can contribute to He date scatter (Flowers and Kelley, 2011). Ault and Flowers (2012) suggested, however, that for typical apatites eU concentrations between zones do not vary by more than a factor of ∼2, and for typical thermal histories the resulting relative date difference between zoned and homogenous apatites with the same bulk eU is no more than ∼10 percent.

We expect these zonation issues to result in greater fractional date differences for the zircon He system. Order of magnitude differences in eU zonation are not uncommon in zircons (for example, fig. 13 in Reiners and others, 2004), and these can cause large discrepancies in both damage and He concentration. Furthermore, unlike apatite, He diffusivity in zircon will either decrease or increase depending on the eU concentration of a given zone and the specific t-T path of the zircon. The interplay amongst the alpha ejection correction factors, He concentration profiles, and damage profiles in zoned zircons with different thermal histories is therefore complicated and a detailed examination of real He datasets with zoned grains is beyond the scope of our current study. However, we discuss below the results from several HeFTy models to show how zonation affects differential damage accumulation within a zircon and sample date-eU correlations.

Fig. 13.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 13.

Time-temperature histories and corresponding date-eU correlations for unzoned zircons, zircons with high eU rims, and zircons with high eU cores. The numbering for each t-T path is similar to figure 11, except we have omitted number 6 and replaced number 5 with a different thermal history. Core eU concentrations are either enriched or depleted by a factor of 7 relative to the bulk concentration of the whole grain. For example, if the bulk concentration is 500 ppm, then the core concentration in the high eU core zircon is 3500 ppm and it is ∼71 ppm in the high eU rim zircon. Both the concentration in the rim and the radial position of the rim are determined by maximizing the zircon zonation factor (see text for details). All zircons have a radius of 60 microns and the core is composed of either the inner 20 or 40 microns of the grain. Large bold symbols connected by horizontal curves represent FTH corrected dates, small bold symbols represent FTZ corrected dates, and small transparent symbols represent uncorrected dates. Lightly colored vertical bars have been added to aid in connecting corresponding FTH and FTZ corrected dates. See text for details on uncorrected, FTH, and FTZ corrected zircons. Note that the scale for each y-axis is different.

These model simulations consist of five representative thermal histories and the resulting date-eU correlations for a suite of unzoned zircons and zircons with simple concentric zonation of high eU cores or high eU rims (fig. 13). Both the zoned and unzoned zircons in each plot have bulk eU concentrations ranging from 250 to 1250 ppm with radii of 60 microns. The zonation profiles for the zircons in our models have two main inputs, the core eU concentration and the core's radial position (rim concentration is set by the bulk eU). In order to provide some uniformity to our choice of variables, we use a zonation impact index (mass of eU for the whole grain divided by the mass difference between the core and rim), which provides a rough estimate of magnitude of the effect that eU zonation has on a zircon's He concentration and damage profile. For a given bulk and core eU concentration, this value reaches a maximum at a certain core radial position and rim eU concentration. In turn, this particular radial position and rim eU maximizes the difference in diffusion kinetics between rim and core that result from He concentration and radiation damage disparities. Because we want to show a worst-case scenario for each date-eU correlation, we have chosen the rim eU concentration and core radial position that correspond to the maximum zonation impact number. For core eU concentration, we pick values that differ from the bulk eU concentration by a factor of seven (a bulk eU of 500 ppm leads to a core eU of either 71.4 or 3500 ppm), which results in a core radial position that is 20 microns from the center (total radius of 60 microns) for the enriched core zircons, and a core radial position that is 40 microns from the center for the enriched rim zircons. In zircons with enriched rims, the ratio between eU concentrations in the rim and core is 9.52:1, while the same ratio is 1:9.1 for zircons with enriched cores. A final consideration is the choice of a FTH or a FTZ alpha ejection correction. For unzoned zircons, we model both the uncorrected date at a given eU and the FTH corrected date (FTH = FTZ for the unzoned case). For each zoned zircon, we model the uncorrected date, the FTZ corrected date using the “redistribution” option in HeFTy, and the FTH corrected date. The FTH for each zoned zircon is calculated by taking the ratio between the unzoned FTH corrected date at equivalent eU and the unzoned uncorrected date at equivalent eU. For zoned grains, the FTH date is equivalent to measuring a raw grain date on a zoned zircon and applying a naive alpha ejection correction assuming no parent zonation.

The model results for all corrected and uncorrected zircons are shown in figure 13. Thermal histories 1 through 4 are the same as in figure 11. We have omitted t-T paths 5 and 6 from figure 11 as both of these yield nearly flat date-eU correlations for zoned and unzoned zircons. Instead, we have added a new t-T path 5 to figure 13 that could be appropriate for zircons from Laramide uplifts of the US Rocky Mountains (for example, our Bighorn sample). In all t-T scenarios, the FTH corrected zoned dates (large bold symbols connected by curves in fig. 13) for zircons with high eU rims are younger than their unzoned counterparts. Only some of this discrepancy is due to an improper alpha ejection correction, as the FTZ corrected zoned dates (small bold symbols) are also younger than unzoned zircons with the same bulk eU. The other causes for younger dates—which in scenarios 3, 4 and 5, are the predominant ones—are the combined effects of a heterogeneous He concentration profile and radiation damage. Rims that are enriched in eU relative to the core cause an increase in effective bulk diffusivity (and younger dates) because more He is located near the grain boundary, and, at high eU concentrations, the rims become heavily damaged. This damage effect is particularly apparent in scenario 3, where the grains with the highest bulk eU concentrations have accumulated enough damage such that the rim acts as a diffusion fast path. If we model the high-eU-rim zircon with a bulk eU of 1250 ppm using the kinetics of Reiners and others (2004), then this scenario yields a date of 544 Ma (as opposed to 256 Ma using the kinetics presented here), which further suggests that radiation damage is the primary cause of these younger dates.

Despite being systematically younger, most of the high-eU-rim dates have a similar style of date-eU correlation as the unzoned dates. One exception is scenario 5, where a date-eU correlation that is monotonically negative in the unzoned case is positive at low bulk eU concentrations. Although the rims of the high-eU-rim zircons have accumulated high degrees of damage prior to cooling, the cores range from low (∼5.5 × 1016 α/g at 50 Ma) to moderate (∼1.1 × 1017 α/g at 50 Ma) amounts of damage. This difference in damage is enough to cause a more retentive core and lower bulk diffusivity in the 500 ppm bulk eU grain than the 250 ppm bulk eU grain, which in turn results in a positive date-eU correlation over this eU range.

To explain the date-eU correlations for the high-eU-core zircons, we must similarly consider the degree of damage in both the core and the rim and how those two damage domains combine to affect bulk diffusivity. In scenario 1, dates are either older or younger than the unzoned dates at the same eU, and the style of date-eU correlation changes from a positive correlation in the unzoned grains to a positive-negative-plateau correlation in the high-eU-core grains. The onset of a negative correlation has shifted as the zircon core accumulates a high degree of damage (and therefore has a high diffusivity) at relatively low bulk eU concentrations. Interestingly though, a plateau at the highest bulk eU concentrations suggests that the high diffusivity of the core is somewhat mitigated by the degree of damage in the zircon's rim. A similar phenomenon occurs in scenario 2, and is especially apparent in scenario 3 if we compare the high-eU-rim grains with the high-eU-core grains. In scenarios 1 and 2, the rims of the high-eU-core zircons at high bulk eU concentrations have accumulated a moderate amount of damage (roughly 4 × 1017 α/g prior to any thermal event) such that the rim decreases diffusivity and acts to retard He diffusing out of the crystal, similar to the effects of RDAAM on zoned apatites with eU enriched rims (Farley and others, 2011; Ault and Flowers, 2012). For the rims of the high-eU-core grains in scenario 3, the damage accumulation is more substantial, but the final reheating event occurs at a relatively low temperature. This temperature is low enough to cause these zircon rims to be more retentive than the corresponding rims in the high-eU-rim zircons (and thus yield older dates).

In contrast to scenarios 1, 2, and 3, the thermal histories in 4 and 5 result in damage higher than ∼5 × 1017 α/g, and relatively high diffusivity in both the cores and rims for most bulk eU concentrations. In terms of damage, this places almost all of the domains in these high-eU-core zircons to the high-damage side of the point of lowest diffusivity in figure 10A. The rims of the high-eU-core zircons with 250 ppm bulk eU concentrations are the only domains with damage lower than 5 × 1017 α/g (both are ∼3 × 1017 α/g prior to final cooling at 50 Ma) in these two scenarios. The net result in both scenarios is a date-eU correlation for the high-eU-core grains that is similar in style to the unzoned zircons, but with systematically younger dates. An exception to this is the positive correlation for bulk eU concentrations of 250 ppm and 500 ppm in scenario 5. Among the high-eU-core grains, the rim for the 250 ppm bulk eU grain has a damage amount of ∼2.9 × 1017 α/g at 50 Ma (the age prior to final cooling), while the rim for the 500 ppm bulk eU grain has a damage amount of ∼5.8 × 1017 α/g at 50 Ma. This difference in damage causes the 500 ppm bulk eU grain to have a more retentive rim and thus lower bulk diffusivity relative to the 250 ppm bulk eU grain, which in turn produces a slight positive correlation.

These five t-T scenarios illustrate that date-eU correlations for zoned zircons may be complex because of reasons already discussed in previous studies, but also because of the reversal in He diffusion behavior with progressive damage accumulation. In the absence of a priori knowledge of parent zonation patterns, the effects of parent zonation on He diffusivity may complicate thermochronologic interpretations of date-eU correlations. We again note that we have purposely chosen worst-case zoning scenarios in order to convey the full scope of this issue. Real zircons may possess a less extreme degree or pattern of zonation, which will mitigate some of the date dispersion observed in figure 13. Consistent styles but variable degrees of zonation (for example, enriched rim but with varying rim thickness or enrichment factor relative to the core) may be expected among zircons from some rock types. This would produce dispersion in a band of date-eU correlations between the unzoned case (black curves) and an extreme zoned case (blue or red curves, depending on zonation style) in figure 13. However, suites of zircons from some samples may not possess systematic zonation patterns like those in figure 13. Especially in detrital settings, one may expect to date zircons with a range of zonation patterns: some may have high eU cores, some may have high eU rims, and others might be unzoned. In this context, the curves in figure 13 should be interpreted as bounding constraints for the total range of zonation variability. For a given t-T path, and no consistent zonation style, real date-eU correlations could potentially plot anywhere within the black, blue, or red curves.

Despite the apparent severity of the zonation problem, however, parent zonation may be characterized from laser ablation depth profiles or other techniques, prior to bulk grain dating. These observations also point to the potential for exploiting parent zonation to provide date-eU trends within individual grains, for example with in-situ laser ablation dating (Vermeesch and others, 2012). Under certain conditions, zircon grains with zoned parent concentrations may behave similarly to crystals with multiple He diffusion domains. In-situ measurements and/or step-heating analyses could potentially be used to interrogate the distribution of He and dates among these domains, providing powerful constraints on thermal histories from single grains.

CONCLUSIONS

Several suites of single-grain zircon (U-Th)/He dates from single rock samples show positive and negative correlations with eU. These correlations are a consequence of the two different ways that radiation damage affects He diffusion in zircon. Evidence for two contrasting effects of radiation damage (as related to alpha dose) on diffusion comes from zircon step-heating experiments, which show that between about 1 × 1016 and 5 × 1017 α/g, diffusivity decreases by about three orders of magnitude. Diffusivity then begins to increase rapidly with increasing damage, by up to 10 orders of magnitude at damage levels of sample N17 (∼8 × 1018 α/g). We hypothesize that decreases in diffusivity at low damage are caused by damage zones blocking preferred c-axis parallel pathways. As damage levels approach N17, these zones become increasingly interconnected and form through-going fast diffusion pathways that shrink the effective diffusion domain size. We parameterize the damage-diffusivity relationship with an equation that combines both of these effects. We also couple our parameterization to an equation for damage annealing in order to forward model date-eU correlations from specific t-T histories. Our model offers insight into some of the issues associated with He diffusion in natural zircons and provides other researchers with a tool for understanding and exploiting the significance of date dispersion in zircon He datasets.

ACKNOWLEDGMENTS

This work was supported by NSF grant EAR-0910577 to PWR as well as funding from the COSA2 Collaboration between UA Geosciences and ExxonMobil. L.N. acknowledges financial support by the Austrian Science Fund (FWF) through grants P20028-N10 and P24448-N19. Jiba Ganguly and Sumit Chakraborty provided helpful comments on some of the equations. Sri Lankan reference zircon samples were kindly made available by Allen K. Kennedy (RB140, BR231, G3) and Wolfgang Hofmeister (M127, N17). We are grateful to Andreas Wagner for the preparation of the doubly polished slabs, and to Uttam Chowdhury for analytical assistance. We thank Madalyn Blondes and Louise Miltich for allowing us to present some of their previously unpublished zircon He dates from the Apennines and Minnesota, respectively, and we thank Ian Campbell and Charlotte Allen for the Cooma samples. We appreciate helpful reviews by David Shuster, Rebecca Flowers, and Andrew Carter.

REFERENCES

  1. ↵
    1. Ault A. K.,
    2. Flowers R. M.
    , 2012, Is apatite U-Th zonation information necessary for accurate interpretation of apatite (U-Th)/He thermochronometry data?: Geochimica et Cosmochimica Acta, v. 79, p. 60–78, doi:http://dx.doi.org/10.1016/j.gca.2011.11.037
    OpenUrlCrossRef
  2. ↵
    1. Bengston A.,
    2. Ewing R. C.,
    3. Becker U.
    , 2012, He diffusion and closure temperatures in apatite and zircon: A density functional theory investigation: Geochimica et Cosmochimica Acta, v. 86, p. 228–238, doi:http://dx.doi.org/10.1016/j.gca.2012.03.004
    OpenUrlCrossRef
  3. ↵
    1. Bernet M.,
    2. Zattin M.,
    3. Garver J. I.,
    4. Brandon M. T.,
    5. Vance J. A.
    , 2001, Steady-state exhumation of the European Alps: Geology, v. 29, n. 1, p. 35–38, doi:http://dx.doi.org/10.1130/0091-7613(2001)029〈0035:SSEOTE〉2.0.CO;2
    OpenUrlAbstract/FREE Full Text
  4. ↵
    1. Blackburn T. J.,
    2. Stockli D. F.,
    3. Walker J. D.
    , 2007, Magnetite (U-Th)/He dating and its application to the geochronology of intermediate to mafic volcanic rocks: Earth and Planetary Science Letters, v. 259, n. 3–4, p. 360–371, doi:http://dx.doi.org/10.1016/j.epsl.2007.04.044
    OpenUrlCrossRef
  5. ↵
    1. Capitani G. C.,
    2. Leroux H.,
    3. Doukhan J. C.,
    4. Rios S.,
    5. Zhang M.,
    6. Salje E. K. H.
    , 2000, A TEM investigation of natural metamict zircons: Structure and recovery of amorphous domains: Physics and Chemistry of Minerals, v. 27, n. 8, p. 545–556, doi:http://dx.doi.org/10.1007/s002690000100
    OpenUrlCrossRef
  6. ↵
    1. Chakraborty S.
    , 2008, Diffusion in solid silicates: A tool to track timescales of processes comes of age: Annual Review of Earth and Planetary Science, v. 36, p. 153–190, doi:http://dx.doi.org/10.1146/annurev.earth.36.031207.124125
    OpenUrlCrossRef
  7. ↵
    1. Hanchar J. M.,
    2. Hoskin P. W. O.
    1. Cherniak D. J.,
    2. Watson E. B.
    , 2003, Diffusion in zircon in Hanchar J. M., Hoskin P. W. O., editors, Zircon: Reviews in Mineralogy and Geochemistry, v. 53, p. 113–143, doi:http://dx.doi.org/10.2113/0530113
    OpenUrlCrossRef
  8. ↵
    1. Cherniak D. J.,
    2. Watson E. B.,
    3. Thomas J. B.
    , 2009, Diffusion of helium in zircon and apatite: Chemical Geology, v. 268, n. 1–2, p. 155–166, doi:http://dx.doi.org/10.1016/j.chemgeo.2009.08.011
    OpenUrlCrossRef
  9. ↵
    1. Clay P. L.,
    2. Baxter E. F.,
    3. Cherniak D. J.,
    4. Kelley S. P.,
    5. Thomas J. B.,
    6. Watson E. B.
    , 2010, Two diffusion pathways in quartz: A combined UV-laser and RBS study: Geochimica et Cosmochimica Acta, v. 74, n. 20, p. 5906–5925, doi:http://dx.doi.org/10.1016/j.gca.2010.07.014
    OpenUrlCrossRef
  10. ↵
    1. Cussler E. L.
    , 1984, Diffusion: Mass transfer in fluid systems: New York, Cambridge University Press, 525 p.
  11. ↵
    1. DeCelles P. G.
    , 1994, Late Cretaceous-Paleocene synorogenic sedimentation and kinematic history of the Sevier thrust belt, northeast Utah and southwest Wyoming: Geological Society of America Bulletin, v. 106, n. 1, p. 32–56, doi:http://dx.doi.org/10.1130/0016-7606(1994)106〈0032:LCPSSA〉2.3.CO;2
    OpenUrlAbstract/FREE Full Text
  12. ↵
    1. Devanathan R.,
    2. Corrales L. R.,
    3. Weber W. J.,
    4. Chartier A.,
    5. Meis C.
    , 2006, Molecular dynamics simulation of energetic uranium recoil damage in zircon: Molecular Simulation, v. 32, n. 12–13, p. 1069–1077, doi:http://dx.doi.org/10.1080/08927020600959929
    OpenUrlCrossRef
  13. ↵
    1. Dijkman F. G.,
    2. van der Maas J. H.
    , 1976, Dependence of bandshape and depolarization ratio on slitwidth: Applied Spectroscopy, v. 30, n. 5, p. 545–546, doi:http://dx.doi.org/10.1366/000370276774456859
    OpenUrlCrossRef
  14. ↵
    1. Epstein N.
    , 1989, On tortuosity and the tortuosity factor in flow and diffusion through porous media: Chemical Engineering Science, v. 44, n. 3, p. 777–779, doi:http://dx.doi.org/10.1016/0009-2509(89)85053-5
    OpenUrlCrossRef
  15. ↵
    1. Hanchar J. M.,
    2. Hoskin P. W. O.
    1. Ewing R. C.,
    2. Meldrum A.,
    3. Wang L.,
    4. Weber W. J.,
    5. Corrales L. R.
    , 2003, Radiation effects in zircon, in Hanchar J. M., Hoskin P. W. O., editors, Zircon: Reviews in Mineralogy and Geochemistry, v. 53, p. 387–425, doi:http://dx.doi.org/10.2113/0530387
    OpenUrlCrossRef
  16. ↵
    1. Farley K. A.
    , 2000, Helium diffusion from apatite: General behavior as illustrated by Durango fluorapatite: Journal of Geophysical Research—Solid Earth, v. 105, n. B2, p. 2903–2914, doi:http://dx.doi.org/10.1029/1999JB900348
    OpenUrlCrossRef
  17. ↵
    1. Farley K. A.
    , 2007, He diffusion systematics in minerals: Evidence from synthetic monazite and zircon structure phosphates: Geochimica et Cosmochimica Acta, v. 71, n. 16, p. 4015–4024, doi:http://dx.doi.org/10.1016/j.gca.2007.05.022
    OpenUrlCrossRef
  18. ↵
    1. Farley K. A.,
    2. Wolf R. A.,
    3. Silver L. T.
    , 1996, The effects of long alpha-stopping distance on (U-Th)/He ages: Geochimica et Cosmochimica Acta, v. 60, n. 21, p. 4223–4229, doi:http://dx.doi.org/10.1016/S0016-7037(96)00193-7
    OpenUrlCrossRef
  19. ↵
    1. Farley K. A.,
    2. Reiners P. W.,
    3. Nenow V.
    , 1999, An apparatus for high-precision helium diffusion measurements from minerals: Analytical Chemistry, v. 71, n. 10, p. 2059–2061, doi:http://dx.doi.org/10.1021/ac9813078
    OpenUrl
  20. ↵
    1. Farley K. A.,
    2. Shuster D. L.,
    3. Ketcham R. A.
    , 2011, U and Th zonation in apatite observed by laser ablation ICPMS, and implications for the (U-Th)/He system: Geochimica et Cosmochimica Acta, v. 75, n. 16, p. 4515–4530, doi:http://dx.doi.org/10.1016/j.gca.2011.05.020
    OpenUrlCrossRef
  21. ↵
    1. Schaeffer O. A.,
    2. Zahringer J.
    1. Fechtig H.,
    2. Kalbitzer S.
    , 1966, The diffusion of argon in potassium-bearing solids, in Schaeffer O. A., Zahringer J., editors, Potassium Argon Dating: New York, Springer-Verlag, p. 68–107.
  22. ↵
    1. Flowers R. M.,
    2. Kelley S. A.
    , 2011, Interpreting data dispersion and “inverted” dates in apatite (U-Th)/He and fission-track datasets: An example from the US midcontinent: Geochimica et Cosmochimica Acta, v. 75, n. 18, p. 5169–5186, doi:http://dx.doi.org/10.1016/j.gca.2011.06.016
    OpenUrlCrossRef
  23. ↵
    1. Flowers R. M.,
    2. Shuster D. L.,
    3. Wernicke B. P.,
    4. Farley K. A.
    , 2007, Radiation damage control on apatite (U-Th)/He dates from the Grand Canyon region, Colorado Plateau: Geology, v. 35, n. 5, p. 447–450, doi:http://dx.doi.org/10.1130/G23471A.1
    OpenUrlAbstract/FREE Full Text
  24. ↵
    1. Flowers R. M.,
    2. Ketcham R. A.,
    3. Shuster D. L.,
    4. Farley K. A.
    , 2009, Apatite (U-Th)/He thermochronometry using a radiation damage accumulation and annealing model: Geochimica et Cosmochimia Acta, v. 73, n. 8, p. 2347–2365, doi:http://dx.doi.org/10.1016/j.gca.2009.01.015
    OpenUrlCrossRef
  25. ↵
    1. Garver J. I.,
    2. Reiners P. W.,
    3. Walker L. J.,
    4. Ramage J. M.,
    5. Perry S. E.
    , 2005, Implications for timing of Andean uplift from thermal resetting of radiation-damaged zircon in the Cordillera Huayhuash, northern Peru: The Journal of Geology, v. 113, n. 2, p. 117–138, doi:http://dx.doi.org/10.1086/427664
    OpenUrlCrossRef
  26. ↵
    1. Gautheron C.,
    2. Tassan-Got L.
    , 2010, A Monte Carlo approach to diffusion applied to noble gas/helium thermochronology: Chemical Geology, v. 273, n. 3–4, p. 212–224, doi:http://dx.doi.org/10.1016/j.chemgeo.2010.02.023
    OpenUrlCrossRef
  27. ↵
    1. Gautheron C.,
    2. Tassan-Got L.,
    3. Barbarand J.,
    4. Pagel M.
    , 2009, Effect of alpha-damage annealing on apatite (U-Th)/He thermochronology: Chemical Geology, v. 266, n. 3–4, p. 157–170, doi:http://dx.doi.org/10.1016/j.chemgeo.2009.06.001
    OpenUrlCrossRef
  28. ↵
    1. Gautheron C.,
    2. Tassan-Got L.,
    3. Ketcham R. A.,
    4. Dobson K. J.
    , 2012, Accounting for long alpha-particle stopping distances in (U-Th-Sm)/He geochronology: 3D modeling of diffusion, zoning, implantation, and abrasion: Geochimica et Cosmochimica Acta, v. 96, n. 1, p. 44–56, doi:http://dx.doi.org/10.1016/j.gca.2012.08.016
    OpenUrlCrossRef
  29. ↵
    1. Gibbons J. F.
    , 1972, Ion implantation in semicondoctors—Part II: Damage production and annealing: Proceedings of the IEEE, v. 60, n. 9, p. 1062–1096, doi:http://dx.doi.org/10.1109/PROC.1972.8854
  30. ↵
    1. Green P. F.,
    2. Crowhurst P. V.,
    3. Duddy I. R.,
    4. Japsen P.,
    5. Holford S. P.
    , 2006, Conflicting (U-Th)/He and fission track ages in apatite: Enhanced He retention, not anomalous annealing behavior: Earth and Planetary Science Letters, v. 250, n. 3–4, p. 407–427, doi:http://dx.doi.org/10.1016/j.epsl.2006.08.022
    OpenUrlCrossRef
  31. ↵
    1. Hacker B. R.,
    2. Christie J. M.
    , 1991, Observational evidence for a possible new diffusion path: Science, v. 251, n. 4989, p. 67–70, doi:http://dx.doi.org/10.1126/science.251.4989.67
    OpenUrlAbstract/FREE Full Text
  32. ↵
    1. Faul H.
    1. Holland H. D.
    , 1954, Radiation damage and its use in age determination, in Faul H., editor, Nuclear Geology: New York, Wiley, p. 175–179.
  33. ↵
    1. Holm D. K.,
    2. Darrah K. S.,
    3. Lux D. R.
    , 1998, Evidence for widespread ∼1760 Ma metamorphism and rapid crustal stabilization of the early Proterozoic (1870-1820 Ma) Penokean Orogen, Minnesota: American Journal of Science, v. 298, n. 1, p. 60–81, doi:http://dx.doi.org/10.2475/ajs.298.1.60
    OpenUrlAbstract/FREE Full Text
  34. ↵
    1. Hourigan J. K.,
    2. Reiners P. W.,
    3. Brandon M. T.
    , 2005, U-Th zonation-dependent alpha-ejection in (U-Th)/He chronometry: Geochimica et Cosmochimica Acta, v. 69, n. 13, p. 3349–3365, doi:http://dx.doi.org/10.1016/j.gca.2005.01.024
    OpenUrlCrossRef
  35. ↵
    1. Hurley P. M.
    , 1952, Alpha ionization damage as a cause of low helium ratios: Transactions of the American Geophysical Union, v. 33, p. 174–183, doi:http://dx.doi.org/10.1029/TR033i002p00174
    OpenUrlCrossRef
  36. ↵
    1. Faul H.
    1. Hurley P. M.
    , 1954, The helium age method and the distribution and migration of helium in rocks, in Faul H., editor, Nuclear Geology: New York, Wiley, p. 301–329.
  37. ↵
    1. Hurley P. M.,
    2. Larsen E. S. Jr..,
    3. Gottfried D.
    , 1956, Comparison of radiogenic helium and lead in zircon: Geochimica et Cosmochimica Acta, v. 9, n. 1–2, p. 98–102, doi:http://dx.doi.org/10.1016/0016-7037(56)90060-6
    OpenUrlCrossRef
  38. ↵
    1. Reiners P. W.,
    2. Ehlers T. A.
    1. Ketcham R. A.
    , 2005, Forward and inverse modeling of low-temperature thermochronometry data, in Reiners P. W., Ehlers T. A., editors, Low-temperature thermochronology: Techniques, interpretations, and applications: Reviews in Mineralogy and Geochemistry, v. 58, p. 275–314, doi:http://dx.doi.org/10.2138/rmg.2005.58.11
    OpenUrlCrossRef
  39. ↵
    1. Ketcham R. A.,
    2. Carter A.,
    3. Donelick R. A.,
    4. Barbarand J.,
    5. Hurford A. J.
    , 2007, Improved modeling of fission-track annealing in apatite: American Mineralogist, v. 92, n. 5–6, p. 799–810, doi:http://dx.doi.org/10.2138/am.2007.2281
    OpenUrlAbstract/FREE Full Text
  40. ↵
    1. Ketcham R. A.,
    2. Guenthner W. R.,
    3. Reiners P. W.
    , 2013, Geometric analysis of radiation damage connectivity in zircon and its implications for helium diffusion: American Mineralogist, doi:http://dx.doi.org/10.2138/am.2013.4249
    OpenUrlCrossRef
  41. ↵
    1. Meesters A. G. C. A,
    2. Dunai T. J.
    , 2002, Solving the production-diffusion equation for finite diffusion domains of various shapes: Part II. Application to cases with α-ejection and nonhomogeneous distribution of the source: Chemical Geology, v. 186, n. 1–2, p. 57–73, doi:http://dx.doi.org/10.1016/S0009-2541(01)00423-5
    OpenUrlCrossRef
  42. ↵
    1. Meldrum A.,
    2. Boatner L. A.,
    3. Weber W. J.,
    4. Ewing R. C.
    , 1998, Radiation damage in zircon and monazite: Geochimica et Cosmochimica Acta, v. 62, n. 14, p. 2509–2520, doi:http://dx.doi.org/10.1016/S0016-7037(98)00174-4
    OpenUrlCrossRef
  43. ↵
    1. Nasdala L.,
    2. Irmer G.,
    3. Wolf D.
    , 1995, The degree of metamictization in zircon: A Raman spectroscopic study: European Journal of Mineralogy, v. 7, n. 3, p. 471–478.
    OpenUrlAbstract/FREE Full Text
  44. ↵
    1. Nasdala L.,
    2. Wenzel M.,
    3. Vavra G.,
    4. Irmer G.,
    5. Wenzel T.,
    6. Kober B.
    , 2001, Metamictisation of natural zircon: Accumulation versus thermal annealing of radioactivity-induced damage: Contributions to Mineralogy and Petrology, v. 141, n. 2, p. 125–144, doi:http://dx.doi.org/10.1007/s004100000235
    OpenUrlCrossRef
  45. ↵
    1. Nasdala L.,
    2. Lengauer C. L.,
    3. Hanchar J. M.,
    4. Kronz A.,
    5. Wirth R.,
    6. Blanc P.,
    7. Kennedy A. K.,
    8. Seydoux-Guillaume A.
    , 2002, Annealing radiation damage and the recovery of cathodoluminescence: Chemical Geology, v. 191, n. 1–3, p. 121–140, doi:http://dx.doi.org/10.1016/S0009-2541(02)00152-3
    OpenUrlCrossRef
  46. ↵
    1. Nasdala L.,
    2. Reiners P. W.,
    3. Garver J. I.,
    4. Kennedy A. K.,
    5. Stern R. A.,
    6. Balan E.,
    7. Wirth R.
    , 2004a, Incomplete retention of radiation damage in zircon from Sri Lanka: American Mineralogist, v. 89, p. 219–231.
    OpenUrlAbstract/FREE Full Text
  47. ↵
    1. Beran A.,
    2. Libowitzky E.
    1. Nasdala L.,
    2. Gotze J.,
    3. Hanchar J. M.,
    4. Gaft M.,
    5. Krbetschek M. R.
    , 2004b, Luminescence techniques in earth sciences, in Beran A., Libowitzky E., editors, Spectroscopic methods in mineralogy: EMU Notes in Mineralogy, v. 6, p. 43–91.
    OpenUrl
  48. ↵
    1. Nasdala L.,
    2. Grambole D.,
    3. Gotze J.,
    4. Kempe U.,
    5. Vaczi T.
    , 2011, Helium irradiation study on zircon: Contributions to Mineralogy and Petrology, v. 161, n. 5, p. 777–789, doi:http://dx.doi.org/10.1007/s00410-010-0562-7
    OpenUrlCrossRef
  49. ↵
    1. Rahn M. K.,
    2. Brandon M. T.,
    3. Batt G. E.,
    4. Garver J. I.
    , 2004, A zero-damage model for fission-track annealing in zircon: American Mineralogist, v. 89, p. 473–484.
    OpenUrlAbstract/FREE Full Text
  50. ↵
    1. Reich M.,
    2. Ewing R. C.,
    3. Ehlers T. A.,
    4. Becker U.
    , 2007, Low-temperature anisotropic diffusion of helium in zircon: Implications for zircon (U-Th)/He thermochronometry: Geochimica et Cosmochimica Acta, v. 71, n. 12, p. 3119–3130, doi:http://dx.doi.org/10.1016/j.gca.2007.03.033
    OpenUrlCrossRef
  51. ↵
    1. Reiners P. W.,
    2. Ehlers T. A.
    1. Reiners P. W.
    , 2005, Zircon (U-Th)/He thermochronometry, in Reiners P. W., Ehlers T. A., editors, Low-temperature thermochronology: Techniques, interpretations, and applications: Reviews in Mineralogy and Geochemistry, v. 58, p. 151–179, doi:http://dx.doi.org/10.2138/rmg.2005.58.6
    OpenUrlCrossRef
  52. ↵
    1. Reiners P. W.,
    2. Farley K. A.
    , 1999, Helium diffusion and (U-Th)/He thermochronometry of titanite: Geochimica et Cosmochimica Acta, v. 63, n. 22, p. 3845–3859, doi:http://dx.doi.org/10.1016/S0016-7037(99)00170-2
    OpenUrlCrossRef
  53. ↵
    1. Reiners P. W.,
    2. Brady R.,
    3. Farley K. A.,
    4. Fryxell J. E.,
    5. Wernicke B.,
    6. Lux D.
    , 2000, Helium and argon thermochronometry of the Gold Butte block, south Virgin Mountains, Nevada: Earth and Planetary Science Letters, v. 178, n. 3–4, p. 315–326, doi:http://dx.doi.org/10.1016/S0012-821X(00)00080-7
    OpenUrlCrossRef
  54. ↵
    1. Reiners P. W.,
    2. Farley K. A.,
    3. Hickes H. J.
    , 2002, He diffusion and (U-Th)/He thermochronometry of zircon: Initial results from Fish Canyon Tuff and Gold Butte: Tectonophysics, v. 349, n. 1–4, p. 297–308, doi:http://dx.doi.org/10.1016/S0040-1951(02)00058-6
    OpenUrlCrossRef
  55. ↵
    1. Reiners P. W.,
    2. Spell T. L.,
    3. Nicolescu S.,
    4. Zanetti K. A.
    , 2004, Zircon (U-Th)/He thermochronometry: He diffusion and comparisons with 40Ar/39Ar dating: Geochimica et Cosmochimica Acta, v. 68, n. 8, p. 1857–1887, doi:http://dx.doi.org/10.1016/j.gca.2003.10.021
    OpenUrlCrossRef
  56. ↵
    1. Saadoune I.,
    2. Purton J. A.,
    3. de Leeux N. H.
    , 2009, He incorporation and diffusion pathways in pure and defective zircon ZrSiO4: A density functional theory study: Chemical Geology, v. 258, n. 3–4, p. 182–196, doi:http://dx.doi.org/10.1016/j.chemgeo.2008.10.015
    OpenUrlCrossRef
  57. ↵
    1. Salje E. K. H.,
    2. Chrosch J.,
    3. Ewing R. C.
    , 1999, Is “metamictization” of zircon a phase transition?: American Mineralogist, v. 84, n. 7–8, p. 1107–1116.
    OpenUrlAbstract
  58. ↵
    1. Shuster D. L.,
    2. Farley K. A.
    , 2009, The influence of artificial radiation damage and thermal annealing on helium diffusion kinetics in apatite: Geochimica et Cosmochimica Acta, v. 73, n. 1, p. 183–196, doi:http://dx.doi.org/10.1016/j.gca.2008.10.013
    OpenUrlCrossRef
  59. ↵
    1. Shuster D. L.,
    2. Vasconcelos P. M.,
    3. Heim J. A.,
    4. Farley K. A.
    , 2005, Weathering geochronology by (U-Th)/He dating of goethite: Geochimica et Cosmochimica Acta, v. 69, n. 3, p. 659–673, doi:http://dx.doi.org/10.1016/j.gca.2004.07.028
    OpenUrlCrossRef
  60. ↵
    1. Shuster D. L.,
    2. Flowers R. M.,
    3. Farley K. A.
    , 2006, The influence of natural radiation damage on helium diffusion kinetics in apatite: Earth and Planetary Science Letters, v. 249, n. 3–4, p. 148–161, doi:http://dx.doi.org/10.1016/j.epsl.2006.07.028
    OpenUrlCrossRef
  61. ↵
    1. Reiners P. W.,
    2. Ehlers T. A.
    1. Tagami T.
    , 2005, Zircon fission-track thermochronology and applications to fault studies, in Reiners P. W., Ehlers T. A., editors, Low-temperature thermochronology: Techniques, interpretations, and applications: Reviews in Mineralogy and Geochemistry, v. 58, p. 95–122.
    OpenUrlCrossRef
  62. ↵
    1. Tagami T.,
    2. Ito H.,
    3. Nishimura S.
    , 1990, Thermal annealing characteristics of spontaneous fission tracks in zircon: Chemical Geology: Isotope Geoscience Section, v. 80, p. 159–169, doi:http://dx.doi.org/10.1016/0168-9622(90)90024-7
    OpenUrlCrossRef
  63. ↵
    1. Vermeesch P.,
    2. Sherlock S. C.,
    3. Roberts N. M. W.,
    4. Carter A.
    , 2012, A simple method for in-situ U-Th-He dating: Geochimica et Cosmochimica Acta, v. 79, p. 140–147, doi:http://dx.doi.org/10.1016/j.gca.2011.11.042
    OpenUrlCrossRef
  64. ↵
    1. Watson E. B.,
    2. Wanser K. H.,
    3. Farley K. A.
    , 2010, Anisotropic diffusion in a finite cylinder, with geochemical applications: Geochimica et Cosmochimica Acta, v. 74, n. 2, p. 614–633, doi:http://dx.doi.org/10.1016/j.gca.2009.10.013
    OpenUrlCrossRef
  65. ↵
    1. Weber W. J.,
    2. Ewing R. C.,
    3. Wang L.-M.
    , 1994, The radiation-induced crystalline-to-amorphous transition in zircon: Journal of Materials Research, v. 9, n. 3, p. 688–698, doi:http://dx.doi.org/10.1557/JMR.1994.0688
    OpenUrlCrossRef
  66. ↵
    1. Williams I. S.
    , 2001, Response of detrital zircon and monazite, and their U-Pb isotopic systems, to regional metamorphism and host-rock partial melting, Cooma Complex, southeastern Australia: Australian Journal of Earth Sciences, v. 48, n. 4, p. 557–580, doi:http://dx.doi.org/10.1046/j.1440-0952.2001.00883.x
    OpenUrlCrossRef
  67. ↵
    1. Wolfe M. R.,
    2. Stockli D. F.
    , 2010, Zircon (U-Th)/He thermochronometry in the KTB drill hole, Germany, and its implications for bulk He diffusion kinetics in zircon: Earth and Planetary Science Letters, v. 295, n. 1–2, p. 69–82, doi:http://dx.doi.org/10.1016/j.epsl.2010.03.025
    OpenUrlCrossRef
  68. ↵
    1. Worden R. H.,
    2. Walker F. D. L.,
    3. Parsons I.,
    4. Brown W. L.
    , 1990, Development of microporosity, diffusion channels and deuteric coarsening in perthitic alkali feldspars: Contributions to Mineralogy and Petrology, v. 104, n. 5, p. 507–515, doi:http://dx.doi.org/10.1007/BF00306660
    OpenUrlCrossRef
  69. ↵
    1. Yamada R.,
    2. Murakami M.,
    3. Tagami T.
    , 2007, Statistical modeling of annealing kinetics of fission tracks in zircon; reassessment of laboratory experiments: Chemical Geology, v. 236, n. 1–2, p. 75–91, doi:http://dx.doi.org/10.1016/j.chemgeo.2006.09.002
    OpenUrlCrossRef
  70. ↵
    1. Yund R. A.,
    2. Smith B. M.,
    3. Tullis J.
    , 1981, Dislocation-assisted diffusion of oxygen in albite: Physics and Chemistry of Minerals, v. 7, n. 4, p. 185–189, doi:http://dx.doi.org/10.1007/BF00307264
    OpenUrlCrossRef
  71. ↵
    1. Yund R. A.,
    2. Quigley J.,
    3. Tullis J.
    , 1989, The effect of dislocations on bulk diffusion in feldspars during metamorphism: Journal of Metamorphic Geology, v. 7, n. 3, p. 337–341, doi:http://dx.doi.org/10.1111/j.1525-1314.1989.tb00601.x
    OpenUrl
  72. ↵
    1. Yurimoto H.,
    2. Morioka M.,
    3. Nagasawa H.
    , 1989, Diffusion in single crystals of melilite: I. oxygen: Geochimica et Cosmochimica Acta, v. 53, n. 9, p. 2387–2394, doi:http://dx.doi.org/10.1016/0016-7037(89)90360-8
    OpenUrlCrossRef
  73. ↵
    1. Zattin M.,
    2. Picotti V.,
    3. Zuffa G. G.
    , 2002, Fission-track reconstruction of the front of the northern Apennine thrust wedge and overlying Ligurian unit: American Journal of Science, v. 302. n. 4, p. 346–379, doi:http://dx.doi.org/10.2475/ajs.302.4.346
    OpenUrlAbstract/FREE Full Text
  74. ↵
    1. Zhang M.,
    2. Salje E. K. H.,
    3. Capitani G. C.,
    4. Leroux H.,
    5. Clark A. M.,
    6. Schluter J.,
    7. Ewing R. C.
    , 2000, Annealing of α-decay damage in zircon: A Raman spectroscopic study: Journal of Physics: Condensed Matter, v. 12, n. 13, p. 3131–3148, doi:http://dx.doi.org/10.1088/0953-8984/12/13/321
    OpenUrlCrossRef
  75. ↵
    1. Zhang M.,
    2. Salje E. K. H.,
    3. Ewing R. C.
    , 2010, OH species, U ions, and CO/CO2 in thermally annealed metamict zircon (ZrSiO4): American Mineralogist, v. 95, n. 11–12, p. 1717–1724, doi:http://dx.doi.org/10.2138/am.2010.3567
    OpenUrlAbstract/FREE Full Text
PreviousNext
Back to top

In this issue

American Journal of Science: 313 (3)
American Journal of Science
Vol. 313, Issue 3
1 Mar 2013
  • Table of Contents
  • Table of Contents (PDF)
  • About the Cover
  • Index by author
  • Ed Board (PDF)
Print
Download PDF
Article Alerts
Sign In to Email Alerts with your Email Address
Email Article

Thank you for your interest in spreading the word on American Journal of Science.

NOTE: We only request your email address so that the person you are recommending the page to knows that you wanted them to see it, and that it is not junk mail. We do not capture any email address.

Enter multiple addresses on separate lines or separate them with commas.
Helium diffusion in natural zircon: Radiation damage, anisotropy, and the interpretation of zircon (U-Th)/He thermochronology
(Your Name) has sent you a message from American Journal of Science
(Your Name) thought you would like to see the American Journal of Science web site.
CAPTCHA
This question is for testing whether or not you are a human visitor and to prevent automated spam submissions.
10 + 2 =
Solve this simple math problem and enter the result. E.g. for 1+3, enter 4.
Citation Tools
Helium diffusion in natural zircon: Radiation damage, anisotropy, and the interpretation of zircon (U-Th)/He thermochronology
William R. Guenthner, Peter W. Reiners, Richard A. Ketcham, Lutz Nasdala, Gerald Giester
American Journal of Science Mar 2013, 313 (3) 145-198; DOI: 10.2475/03.2013.01

Citation Manager Formats

  • BibTeX
  • Bookends
  • EasyBib
  • EndNote (tagged)
  • EndNote 8 (xml)
  • Medlars
  • Mendeley
  • Papers
  • RefWorks Tagged
  • Ref Manager
  • RIS
  • Zotero
Share
Helium diffusion in natural zircon: Radiation damage, anisotropy, and the interpretation of zircon (U-Th)/He thermochronology
William R. Guenthner, Peter W. Reiners, Richard A. Ketcham, Lutz Nasdala, Gerald Giester
American Journal of Science Mar 2013, 313 (3) 145-198; DOI: 10.2475/03.2013.01
Reddit logo Twitter logo Facebook logo Mendeley logo
  • Tweet Widget
  • Facebook Like
  • Google Plus One

Jump to section

  • Article
    • Abstract
    • INTRODUCTION
    • METHODS
    • RESULTS
    • DISCUSSION
    • CONCLUSIONS
    • ACKNOWLEDGMENTS
    • REFERENCES
  • Figures & Data
  • Info & Metrics
  • References
  • PDF

Related Articles

  • No related articles found.
  • Google Scholar

Cited By...

  • Thermochronological constraints on the exhumation history of the Carboniferous Katebasu gold deposit, western Tianshan gold belt, NW China
  • Post-orogenic exhumation in the western Pyrenees: evidence for extension driven by pre-orogenic inheritance
  • Late Mesozoic and Cenozoic thermotectonic history of eastern, central and southern Mexico as determined through integrated thermochronology, with implications for sediment delivery to the Gulf of Mexico
  • Timing of slip across the South Tibetan detachment system and Yadong-Gulu graben, Eastern Himalaya
  • Diachronous development of Great Unconformities before Neoproterozoic Snowball Earth
  • West Gondwana orogenies and Pangaea break-up: thermotectonic effects on the southernmost Mantiqueira Province, Brazil
  • Exhumation history of the Ayd&#x0131;n range and the role of the Bu&#x0308;yu&#x0308;k Menderes detachment system during bivergent extension of the central Menderes Massif, western Turkey
  • Thermal evolution in the exhumed basement of a stratovolcano: case study of the Miocene Matra Volcano, Pannonian Basin
  • Tracking low-temperature tectonism of the St. Lawrence Platform and Humber Zone, southern Quebec Appalachians, through apatite and zircon (U-Th)/He thermochronology
  • Phanerozoic low-temperature evolution of the Uruguayan Shield along the South American passive margin
  • Dry annealing of metamict zircon: A differential scanning calorimetry study
  • Cooling and exhumation of the southern Idaho batholith
  • Intrusive and depositional constraints on the Cretaceous tectonic history of the southern Blue Mountains, eastern Oregon
  • Investigating the Paleozoic-Mesozoic low-temperature thermal history of the southwestern Canadian Arctic: insights from (U-Th)/He thermochronology
  • Seeing is believing: Visualization of He distribution in zircon and implications for thermal history reconstruction on single crystals
  • Thermochronology constraints on Miocene exhumation in the Central Range Mountains, Trinidad
  • The Liuqu Conglomerate, southern Tibet: Early Miocene basin development related to deformation within the Great Counter Thrust system
  • Pleistocene onset of rapid, punctuated exhumation in the eastern Central Range of the Taiwan orogenic belt
  • Record of paleofluid circulation in faults revealed by hematite (U-Th)/He and apatite fission-track dating: An example from Gower Peninsula fault fissures, Wales
  • (U-Th)/He thermochronology records late Miocene accelerated cooling in the north-central Peruvian Andes
  • Reconnaissance thermochronology of southern Zealandia
  • Low-temperature thermochronologic constraints on the kinematic histories of the Castle Cliffs, Tule Springs, and Mormon Peak detachments, southwestern Utah and southeastern Nevada
  • Sevier belt exhumation in central Utah constrained from complex zircon (U-Th)/He data sets: Radiation damage and He inheritance effects on partially reset detrital zircons
  • Interplay between the thermal evolution of an orogenic wedge and its retro-wedge basin: An example from the Ukrainian Carpathians
  • Exhumation of the North American Cordillera revealed by multi-dating of Upper Jurassic-Upper Cretaceous foreland basin deposits
  • Incision into the Eastern Andean Plateau During Pliocene Cooling
  • Google Scholar

More in this TOC Section

  • Timing and Nd-Hf isotopic mapping of early Mesozoic granitoids in the Qinling Orogen, central China: Implication for architecture, nature and processes of the orogen
  • India in the Nuna to Gondwana supercontinent cycles: Clues from the north Indian and Marwar Blocks
  • Unravelling the P-T-t history of three high-grade metamorphic events in the Epupa Complex, NW Namibia: Implications for the Paleoproterozoic to Mesoproterozoic evolution of the Congo Craton
Show more Articles

Similar Articles

Keywords

  • zircon
  • He diffusion
  • Thermochronology
  • Radiation Damage

Navigate

  • Current Issue
  • Archive

More Information

  • RSS

Other Services

  • About Us

© 2023 American Journal of Science

Powered by HighWire