You are on page 1of 37

Available online at www.sciencedirect.

com

Geochimica et Cosmochimica Acta 112 (2013) 251–287


www.elsevier.com/locate/gca

Systematic variations of argon diffusion in feldspars


and implications for thermochronometry
William S. Cassata ⇑, Paul R. Renne
Department of Earth and Planetary Science, University of California – Berkeley, 307 McCone Hall #4767, Berkeley, CA 94720-4767, USA
Berkeley Geochronology Center, 2455 Ridge Road, Berkeley, CA 94709, USA

Received 31 May 2012; accepted in revised form 27 February 2013; available online 7 March 2013

Abstract

Coupled information about the time-dependent production and temperature-dependent diffusion of radiogenic argon in
feldspars can be used to constrain the thermal evolution attending a host of Earth and planetary processes. To better assess
the accuracy of thermal models, an understanding of the mechanisms and pathways by which argon diffuses in feldspars is
desirable. Here we present step-heating Ar diffusion experiments conducted on feldspars with diverse compositions, structural
states, and microstructural characteristics. The experiments reveal systematic variations in diffusive behavior that appear clo-
sely related to these variables, with apparent closure temperatures for 0.1–1 mm grains of 200–400 °C (assuming a 10 °C/Ma
cooling rate). Given such variability, there is no broadly applicable set of diffusion parameters that can be utilized in feldspar
thermal modeling; sample-specific data are required. Diffusion experiments conducted on oriented cleavage flakes do not
reveal directionally-dependent diffusive anisotropy to within the resolution limits of our approach (approximately a factor
of 2). Additional experiments aimed at constraining the physical significance of the diffusion domain are presented and indi-
cate that unaltered feldspar crystals with or without coherent exsolution lamellae diffuse at the grain-scale, whereas feldspars
containing hydrothermal alteration and/or incoherent sub-grain intergrowths do not. Arrhenius plots for argon diffusion in
plagioclase and alkali feldspars appear to reflect a confluence of intrinsic diffusion kinetics and structural transitions that
occur during incremental heating experiments. These structural transitions, along with sub-grain domain size variations, cause
deviations from linearity (i.e., upward and downward curvature) on Arrhenius plots. An atomistic model for Arrhenius
behavior is proposed that incorporates the variable lattice deformations of different feldspars in response to heating and com-
pression. The resulting implications for accurately extrapolating laboratory-derived diffusion parameters to natural settings
and over geologic time are discussed. We find that considerable inaccuracies may exist in published thermal histories obtained
using multiple diffusion domain (MDD) models fit to Arrhenius plots for exsolved alkali feldspar, where the inferred Ar par-
tial retention zones may be spuriously hot.
Ó 2013 Elsevier Ltd. All rights reserved.

1. INTRODUCTION into an array of planetary processes that span immense


time and temperature regimes, from rapid and high tem-
Thermochronometry by the 40Ar/39Ar technique (see perature bolide impact events (e.g., Cassata et al., 2010a)
e.g., McDougall and Harrison, 1999) provides insights to mountain uplift occurring over plate tectonic timescales
at near surface temperatures (e.g., Richter et al., 1991).
Applications of thermochronometry continue to grow in
⇑ Corresponding author. Current address: Lawrence Livermore number and diversity in the Earth and planetary sciences,
National Laboratory, 7000 East Avenue, Livermore, CA 94550, and numerical models based on thermochronometric data
USA. have become increasingly sophisticated (e.g., Braun, 2005;
E-mail addresses: cassata2@llnl.gov (W.S. Cassata), prenne@ Ehlers, 2005; Ehlers et al., 2005; Reiners et al., 2005;
bgc.org (P.R. Renne). Shuster and Farley, 2005, 2009; Reiners and Shuster,

0016-7037/$ - see front matter Ó 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.gca.2013.02.030
252 W.S. Cassata, P.R. Renne / Geochimica et Cosmochimica Acta 112 (2013) 251–287

2009; Shuster et al., 2010; Simoes et al., 2007; Cassata interpreted in terms of sub-grain domain structure and
et al., 2010a). Feldspars, being rich in potassium and MDD theory applied to alkali feldspars is widely utilized
abundant in most terrestrial and extraterrestrial rocks, to constrain thermal histories in a variety of tectonic set-
are ideal targets for thermochronometry. Ar diffusion in tings (e.g., Richter et al., 1991; McLaren and Dunlap,
K-feldspar has been a topic of extensive research through- 2006; Karlstrom et al., 2010).
out the past 40 years, due, in part, to the ease with which Since the development of MDD theory, it has drawn
high-resolution age and kinetic data can be simultaneously considerable criticisms that challenge the validity of funda-
acquired (e.g., Foland, 1974; Lovera et al., 1989, 1991, mental assumptions underlying the approach (e.g., Parsons
1997, 2002; Harrison et al., 1991; Arnaud and Kelley, et al., 1999; Villa, 2010). In particular, the assumption that
1997; Quidelleur et al., 1997; Wartho et al., 1999). The sub-grain domains are non-interacting (i.e., separated by
kinetics of Ar diffusion in plagioclase, the most abundant infinite reservoirs) has been called into question, prompting
mineral in Earth’s crust, have been less extensively re- the development of a multi-path approach to diffusion
searched, although various samples have been analyzed modeling (volume plus pipe/short-circuit diffusion; Lee,
for specific thermochronometric applications (e.g., Berger 1995). It has also been suggested that non-linearity ob-
and York, 1981; Harrison and McDougall, 1981; Onstott served on alkali feldspar Arrhenius plots reflects the conflu-
et al., 1989; Renne et al., 1990, 1996; Boven et al., 2001; ence of complex chemical, microtextural, and structural
Garrick-Bethell et al., 2009; Cassata et al., 2009; Shuster changes that occur both over geologic time and during lab-
et al., 2009). In general, inferred activation energies (Ea) oratory heating, and that cooling histories cannot be de-
for Ar diffusion in feldspars vary between 100 and rived from age spectra and Arrhenius plots (Parsons
350 kJ/mole and frequency factors (ln(Do/a2)) span 14 or- et al., 1999). Moreover, the assumption that diffusive loss
ders of magnitude (see reviews by Lovera et al., 1997; of Ar is Fickian (i.e., Ar moves from a region of high to
Cassata et al., 2009; Forster and Lister, 2010). Despite low concentration with a magnitude proportional to the
years of research, an explanation for such variability has concentration gradient) has been disputed (Villa, 2006). Vil-
proven elusive, and our basic understanding of the mech- la (2010) and Villa and Hanchar (2013) have shown that in
anisms, pathways, and processes by which Ar diffuses some instances discordance observed on 40Ar/39Ar age
within feldspars remains, in many regards, fragmentary. spectra depends more on the availability of recrystalliza-
In a seminal paper on Ar diffusion in orthoclase, Foland tion-enhancing fluids than on a heat source.
(1974) established that physical grain dimensions define, at Efforts to resolve apparent inconsistencies between
least in some instances, the diffusive lengthscale, based on MDD theory and mineralogical observations are on-going.
equivalent values for the diffusion coefficient D obtained McLaren et al. (2007) found that alkali feldspars with pris-
from grains that differed in size by a factor of 4. Likewise, tine, strain-controlled cryptoperthite from the Klokken sye-
physical grain dimensions are inferred to be the effective dif- nite yielded thermal histories consistent with independent
fusive lengthscale for Ar transport in micas and amphiboles constraints, whereas alkali feldspars with patch (locally
(e.g., Phillips and Onstott, 1988) and He transport in apa- recrystallized) perthite did not, presumably due to domain
tite, titanite, and zircon (e.g., Harrison and Zeitler, 2005). modification and Ar loss resulting from fluid–rock interac-
Subsequent observations of non-linear Arrhenius arrays tions. Likewise, Lovera et al. (2002) cautioned against the
for Ar diffusion in alkali feldspar and tiered age spectra, use of recrystallized alkali feldspars in MDD modeling,
neither of which are consistent with a uniform distribution noting a lack of correlation between age spectra and diffu-
of Ar within a single diffusion domain, prompted the devel- sion data. By obtaining in situ 40K–40Ca ages using second-
opment of multiple diffusion domain (MDD) theory (Love- ary ion mass spectrometry, Harrison et al. (2010) confirmed
ra et al., 1989, 1991, 1993; Harrison et al., 1991). According that both K and Ca were mobilized by non-thermal, fluid-
to MDD theory, deviations from linearity manifest on assisted recrystallization in Klokken feldspars, rendering
essentially all alkali feldspar step-heating Arrhenius plots them poor candidates for MDD modeling. Increasingly, de-
reflect the exhaustion of sub-grain domains of increasing tailed microstructural investigations are being used in coor-
size, each of which is non-interacting and stable during dination with 40Ar/39Ar data to better constrain complex
and following cooling through the Ar partial retention thermal histories and determine which alkali feldspars are
zone. Such sub-grain domains are sometimes reconcilable suitable for thermochronometry (e.g., Sanders et al., 2006;
with microscopic characteristics of a sample (e.g., fractures Heizler et al., 2007, 2008; McLaren and Reddy, 2008; Par-
and/or cleavage planes), but in many instances are more sons et al., 2010; Short et al., 2010; Flude et al., 2012).
ambiguous (e.g., Fitzgerald and Harrison, 1993). Interest- Notwithstanding the aforementioned studies, many
ingly, “gem-quality” alkali feldspars that have been shown unanswered questions remain regarding the physical signif-
to diffuse at the grain-scale based on in situ observations of icance of the diffusion domain, the stability of the diffusion
surface to core concentration gradients obtained using a domain, both over geologic time and during laboratory
UV-laserprobe (Arnaud and Kelley, 1997; Wartho et al., heating experiments, the isotropy of diffusion, the potential
1999) also yield non-linear Arrhenius arrays when incre- importance of structural state on diffusivity, and the possi-
mentally degassed, which indicates that such behavior ble pressure-dependence of diffusivity. To address these
may not solely arise due to a range in sub-grain domain fundamental questions, we conducted diffusion experiments
sizes, but rather may reflect an artifact of structural re- on feldspars whose diverse compositions, structural states,
sponse to heating. Although a causal explanation remains and microstructural characteristics span the natural vari-
elusive, deviations from linearity have nonetheless been ability of ternary feldspars. In this paper we present the
W.S. Cassata, P.R. Renne / Geochimica et Cosmochimica Acta 112 (2013) 251–287 253

results of these diffusion experiments and, together with pyrometer output against a type-K thermocouple under
existing data on noble gas diffusion in silicates, develop a high vacuum (<108 torr) and under the same conditions
framework for interpreting Arrhenius non-linearity and as the diffusion experiments (i.e., same viewport, coverslide,
systematic inter-sample variability in laboratory-derived focal point, etc.). Additional information regarding the
diffusion parameters. We close with a discussion on accu- heating and temperature measurement methodology can
rately extrapolating laboratory-derived diffusion parame- be found in Cassata et al. (2009) and Shuster et al. (2010).
ters to natural settings and over geologic time. During experimental heating, the samples were held at
temperatures between 300 and 1500 °C for durations of
2. METHODS 180–15,000 s. Typical durations required to reach the set-
point temperatures were <10 s. Following heating, samples
Argon diffusion experiments were conducted on feldspars cool to significantly below set-point temperatures in
that (1) range in composition from nearly pure orthoclase to roughly equivalent durations. Precision on all temperature
nearly pure anorthite (Fig. 1), (2) have diverse microtextures, measurements is better than 5 °C. Accuracy on type-K ther-
(3) range in structural state from fully ordered to fully disor- mocouples is approximately 0.4% (Croarkin et al., 1993).
dered, and (4) derive from diverse lithologies, including gran- The precision of the pyrometer-to-thermocouple calibra-
ites, pegmatites, basalts, gabbros, meteorites, silicic tuffs, and tion is 10 °C (Cassata et al., 2009). The released argon
metamorphic rocks. Electron microprobe analyses were used was purified using two SAES getters (one hot and one cold),
to characterize the spatial distribution of K, Na, and Ca in analyzed using a Mass Analyzer Products 215c mass spec-
each sample. Many of the specific samples discussed in this trometer, and measured statically on a Balzers SEV-217
paper have been the topic of extensive mineralogical research electron multiplier using procedures similar to those de-
throughout the past 50 years and have been characterized scribed by Renne et al. (1998). Most crystals were heated
using transmission electron microscopy (TEM) and X-ray in excess of their nominal melting temperatures and were
diffraction (XRD) techniques. completely fused following incremental heating.
Individual equant to elongate crystal fragments and Raw isotope measurements were corrected for 37Ar and
39
cleavage-bounded sheet-like flakes, devoid of obvious Ar decay, spectrometer discrimination, and extraction
microfractures, were hand-picked under a binocular micro- line blanks. Using the fraction of degassed Ca-derived
37
scope and irradiated for 50–200 h at the Oregon State Uni- Ar for plagioclase or K-derived 39Ar for alkali feldspar
versity TRIGA reactor in the Cadmium-Lined In-Core and the duration of each extraction, we calculated the dif-
Irradiation Tube (CLICIT) facility. Following irradiation, fusion coefficient (D) normalized to the characteristic length
single grains were loaded into small packets (5  scale (D/a2) using the equations of Crank (1975) following
3  1 mm) made from high-purity Pt–Ir alloy tubes and the algorithm of Fechtig and Kalbitzer (1966). Uncertain-
incrementally degassed using feedback-controlled laser ties in D/a2 values were calculated based on the analytical
heating with either a 30 or 150 W diode laser (k = 810 ± precision of each step. Spherical geometry was used for
10 nm). Temperatures were monitored using a type-K ther- equant to elongate grains and infinite sheet geometry for
mocouple mounted within the Pt–Ir packets or using an cleavage flakes. Inaccurate or inappropriate choices of dif-
optical pyrometer coaxially aligned with the diode laser. fusion geometry result in erroneous vertical displacements
To account for the temperature dependent emissivity of of Arrhenius arrays, with maximum and typical errors of
the Pt–Ir packets, we routinely calibrated the single-color 2.2 and <0.5 natural log units, respectively (Huber et al.,

Alkali Feldspars Plagioclase Feldspars


T (oC) T (oC)
Melt Melt
1400 Melt + Crystal 1400

1200 I1 1200

C1 (high)
1000 C2m (high) 1000

800 800

C2m (low) e1
600 C1 (low) 600
T
C1 e2
400 400

200 Peristerite Bøggild Huttenlocher 200


P1
Or 80 50 30 10 Ab 10 30 50 70 90 An
K-sanidine sanidine Na-sanidine anorthoclase albite oligoclase andesine labradorite bytownite anorthite

Fig. 1. Phase diagram for alkali and plagioclase feldspars [redrafted and modified from Parsons (2010)]. Dashed and solid lines represent
displacive and order–disorder transitions, respectively. C1, I1, P1, T, e1, and e2 feldspars are triclinic. C2m feldspars are monoclinic. A
detailed description of T-anorthite is given in the text. The striped field above the I1 field represents a metastable extension of the C1 field.
254 W.S. Cassata, P.R. Renne / Geochimica et Cosmochimica Acta 112 (2013) 251–287

2011). The temperature dependence of the diffusion coeffi- ECCa is a Ca-rich anorthoclase (An10.8Ab71.8Or17.4)
cients was quantified from weighted linear regression mod- from Easy Chair Crater (ECC) in Lunar Craters Vol-
els of ln(D/a2) against T1. Incremental releases of 36Ar, canic Field, Nevada, USA (Righter and Carmichael,
38
Ar, and 40Ar were also measured, but will not be dis- 1993). ECC anorthoclase crystals occur as centime-
cussed in this paper. Systematic mass-dependent variations ter-sized mega-xenocrysts entrained in an alkali basalt
in 37Ar and 39Ar diffusivity were not observed (discussed in matrix (Righter and Carmichael, 1993) and are com-
Section 5.2), nor were systematic variations in diffusivity re- positionally homogeneous at the micron-scale. We
lated to irradiation duration. have no detailed information regarding the
microtexture or degree of Al–Si ordering in ECC
3. BRIEF SAMPLE DESCRIPTION anorthoclase.
FCs is a sanidine (An1Ab27Or72; Bachmann et al.,
Brief descriptions of each sample are listed here and 2002) from the 28 Ma Fish Canyon Tuff, a massive
compositional data are given in Table 1. rhyolitic ignimbrite in southwestern Colorado, USA
(Lipman, 2000).
3.1. Alkali feldspars GSs is a sanidine (An3.7Ab30.5Or65.8) from a 108 ka
rhyolitic ash deposited within layered marine sediments
in the Gulf of Salerno off the west coast of Italy.
BMk is an orthoclase (An0.0Ab3.0Or97.0) from a pegma- MADk is a cryptoperthite (An0.2Ab5.5Or94.3) from a
tite in Benson Mines near Star Lake, New York, USA shallowly emplaced pegmatite near Itrongay, Madagas-
(Foland, 1974). Benson Mines orthoclase was described car (Coombs, 1954). Detailed descriptions of the miner-
by Foland (1974) as clear and colorless, compositionally alogy, albite lamellae, and “tweed” microtexture
homogeneous, and without alteration, zoning, or twin- characteristic of the orthoclase lamellae in Madagascar
ning. Our sample of Benson Mines orthoclase has a cryptoperthite can be found in Arnaud and Kelley
pinkish hue due to small, pervasively distributed iron (1997), Wartho et al. (1999), and Parsons and Lee
oxide inclusions. (2005).
BV-8k is a perthitic alkali feldspar (An0.8Ab19.0Or80.2) XTALk is an orthoclase (An0.1Ab11.3Or88.6) from the
from a gabbro in the Bushveld Complex, South Africa. granodioritic Ecstall Pluton (Brownlee and Renne,
Plagioclase from this gabbro was studied by Cassata 2010) in British Colombia, Canada.
et al. (2009). Deuteric alteration is pervasive, with inco-
herent (microporous) patch and vein perthite (e.g., Par- We have no detailed information regarding the micro-
sons and Lee, 2005) replacing film perthite. texture or degree of Al–Si order in GSs and FCs.

Table 1
Feldspar sample information.
Sample Locality Phase Lithology Structure An Ab Or
BMk Benson Mines, NY, USA Orthoclase Granitic Pegmatite C2m 0.0 3.0 97.0
XTALk Ecstall Pluton, BC, Canada Orthoclase Granodiorite Pluton C2m/C1 0.1 11.3 88.6
MADk Itrongay, Madagascar Cyptoperthite Granitic Pegmatite C2m/C1 0.2 5.6 94.6
BV-8k Bushveld Complex, S. Africa Perthite Gabbro C2m/C1 0.8 22.6 76.6
FCs Fish Canyon Tuff, CO, USA Sanidine Rhyolitic Ignimbrite C2m 1.0 27.0 72.0
GSs Gulf of Salerno, Italy Sanidine Rhyolitic Ash C2m 3.7 30.5 65.8
ECCa Easy Chair Crater, NV, USA Anorthoclase Alkali Basalt Lava C1 10.8 71.8 17.4
GRAp Extraterrestrial (GRA 06128) Albite/Olig. Plutonic Achondrite C1 14.1 84.1 1.9
ML-15p Mono Lake, CA, USA Oligoclase Rhyolitic Ash (#15) C1 19.6 74.2 6.2
XTALp Ecstall Pluton, BC, Canada Oligoclase Granodiorite Pluton C1 20.8 77.7 1.5
FCp Fish Canyon Tuff, CO, USA Olig./Andesine Rhyolitic Ignimbrite C1 29.8 61.7 4.5
BG-1p Bushveld Complex, S. Africa Andesine Gabbro C1 38.9 59.4 1.7
PR-92p Paraná-Etendeka LIP, Brazil Andesine Rhyodacite Lava C1 41.1 54.6 4.3
NCp Nain Complex, NL, Canada Labradorite Anorthosite e1/e2 49.4 47.9 2.7
HMS-2p Klamath Mtns., CA, USA Labradorite Basaltic Sill C1 52.8 44.2 3.0
SURTp Surtsey, Iceland Labradorite Alkali Basaltic Lava I1 59.6 39.6 0.8
OREGp Plush, OR, USA Labradorite Basaltic Lava I1 63.9 35.4 0.7
BV-8p Bushveld Complex, S. Africa Labradorite Gabbro e1 67.4 31.3 1.3
DCp Duluth Complex, MN, USA Bytownite Anorthosite e1/P1 74.7 24.9 0.4
VZ699 Verzasca Valley, Switzerland Bytownite Granitoid e1/P1 86.2 13.5 0.3
SW-1 Stillwater Comlpex, MT, USA Bytownite Anorthosite e1/P1 86.7 13.0 0.3
GV-09 Grass Valley, CA, USA Anorthite Anorthosite P1/T 94.2 5.2 0.6
JAPp Miyake, Japan Anorthite Basaltic Lava T 96.2 3.7 0.1
TROCp Lunar (76535) Anorthite Plutonic Troctolite P1 96.2 3.5 0.3
Symmetry groups and phase names based on Fig. 1.
W.S. Cassata, P.R. Renne / Geochimica et Cosmochimica Acta 112 (2013) 251–287 255

3.2. Plagioclase feldspars SURTp is a labradorite (An59.5Ab39.6Or0.9) from the


1963 eruption of Surtsey volcano, Iceland (Wenk,
BG-1p is an andesine (An38.9Ab59.4Or1.7) from a gabbro 1966). Surtsey labradorite crystals occur as large,
in the Bushveld Complex, South Africa. Some crystals glass-coated phenocrysts from a phreatic alkali-olivine
have a smoky appearance due to the presence of magne- basalt eruption, are compositionally homogeneous at
tite inclusions. the micron-scale, and retain Al-Si order, as constrained
BV-8p is labradorite/bytownite (An70) from a gabbro by the presence of e-reflections in X-ray diffraction pat-
in the Bushveld Complex, South Africa (Cassata et al., terns typical of the e- and I1-plagioclase structures
2009). Crystals contain sub-micron e-plagioclase and (Wenk, 1978; Wenk et al., 1980; Wenk and Nakajima,
anorthite lamellae associated with slow-cooling beneath 1980).
the Huttenlocher solvus and exhibit normal zoning (dis- SW-1 is a bytownite (An86.7Ab13.5Or0.3) from an anor-
cussed in detail in Section 4.1). thosite from the Stillwater Complex, a layered mafic
DCp is a bytownite (An74.7Ab24.9Or0.4) from a plutonic intrusion in Montana, USA. Comparable feldspars from
anorthosite of the Duluth Complex from Crystal Bay, the Stillwater Complex (Nord et al., 1974) comprise sub-
Minnesota, USA. micron (125 Å) e-plagioclase and anorthite exsolution
FCp is an oligoclase/andesine (An29.8Ab64.7Or4.5) from lamellae associated with slow-cooling beneath the Hut-
the 28 Ma Fish Canyon Tuff. tenlocher solvus (Fig. 1). The plagioclase crystals are
GRAp is an ordered oligoclase (An14.1Ab84.1Or1.8; extensively altered, with veins of muscovite and porous,
Shearer et al., 2010) from the ungrouped, plutonic microcrystalline Na-rich feldspar replacing bytownite.
achondrite GRA 06128. TROCp is a compositionally homogeneous anorthite
GV-09 is an anorthite (An94.2Ab5.2Or0.6) from a plutonic (An96.2Ab3.5Or0.3) from the lunar troctolite 76535.
anorthosite near Grass Valley, California, USA (Rainey VZ699 is a bytownite (An86.2Ab13.5Or0.3) from a meta-
and Wenk, 1978). The feldspar comprises P1- and I1-anor- morphosed Alpine granitoid from Verzasca Valley in
thite (Fig. 1) domains separated by APB’s (McLaren and Switzerland (Wenk, 1979). The granitoid was subjected
Marshall, 1974) and contains large muscovite inclusions to regional-scale, amphibolite facies metamorphism
and plagioclase sub-grains. The sample is extensively (Wenk, 1979). Plagioclase crystals comprise sub-micron
altered, with patches of Na-rich feldspar replacing anorthite. intergrowths of anorthite and e-plagioclase (Wenk,
HMS-2p is a compositionally homogeneous labradorite 1979) associated with slow-cooling beneath the Huttenl-
(An52.8Ab44.2Or3.0) from a basaltic sill in the eastern ocher solvus (Fig. 1) and contain quartz sub-grains and
Klamath Mountains, USA (Renne and Scott, 1988). alkali feldspar intergrowths.
JAPp is a compositionally homogeneous, gem-quality XTALp is an oligoclase (An20.8Ab77.7Or1.5) from the
anorthite (An96.2Ab3.7Or0.1) from Miyake, Japan. Ecstall Pluton.
ML-15p is a compositionally homogeneous oligoclase
(An19.6Ab74.2Or6.2) from a 32 ka rhyolitic ash from We have no detailed information regarding the micro-
Mono Lake, California, USA (Cassata et al., 2010b). texture or degree of Al–Si order in BG-1p, FCp, HMS-
NCp is a labradorite (An49.5Ab47.9Or2.6) from the type 2p, JAPp, ML-15p, and PR-92p.
locality near Nain, Canada (e.g., Wenk and Nakajima,
1980). The labradorite crystals comprise 100 nm thick 4. RESULTS AND DISCUSSION
exsolution lamellae with compositions of An35 and
An65, each of which has a modulated e-plagioclase 4.1. What defines a diffusion domain?
structure with albite-like and anorthite-like domains
separated by anti-phase boundaries (APBs; Wenk and 4.1.1. Evidence for grain-scale diffusion in compositionally
Nakajima, 1980). The lamellar interfaces are mostly homogeneous and cryptoperthitic feldspars
coherent, but contain some dislocations at misaligned To determine whether or not the physical crystal dimen-
e-plagioclase fringes (Wenk and Nakajima, 1980). The sions define the diffusion domain in unaltered, composition-
sample also contains sparsely distributed, 1 lm thick, ally homogeneous and cryptoperthitic feldspars and to test
K-rich exsolution lamellae and zircon and magnetite for diffusive isotropy, we conducted detailed Ar diffusion
inclusions. experiments on thin cleavage flakes of two extensively stud-
OREGp is a labradorite (An63.9Ab35.4Or0.7) from a lava ied alkali feldspars: Madagascar cryptoperthite (An0.2Ab5.5
flow near Plush, Oregon, USA (Stewart et al., 1966; Or94.3) and Benson Mines orthoclase (An0.0Ab3.0Or97.0). If
Wenk et al., 1980). Plush labradorite crystals occur as sub-grain features such as “tweed” domain boundaries or
large (cm-sized) phenocrysts in a massive basalt, are cryptoperthite lamellar interfaces define the diffusion do-
compositionally homogeneous at the micron-scale, and main (i.e., a in D/a2), then individual fragments of a micro-
contain regularly spaced, sub-micron e-plagioclase and texturally homogeneous megacryst should yield the same
anorthite intergrowths (Wenk et al., 1980; Wenk and value of D/a2 at a given temperature irrespective of macro-
Nakajima, 1980) consistent with slow-cooling beneath scopic grain dimensions. Conversely, if physical grain
the Huttenlocher solvus (Fig. 1). dimensions define the diffusion domain, then the fractional
PR-92p is an andesine (An41.1Ab54.6Or4.3) from a Cha- loss of Ar (i.e., the calculated D/a2 value) associated with a
pecó-type rhyodacite lava of the Paraná-Etendeka large given temperature extraction should vary inversely with the
igneous province exposed in Brazil. square of grain size. Thus by analyzing cleavage flakes that
256 W.S. Cassata, P.R. Renne / Geochimica et Cosmochimica Acta 112 (2013) 251–287

Fig. 2. Images of typical cleavage flakes analyzed for Ar diffusion kinetics. (a) Binocular microscope image and thickness measurements of a
Plush labradorite {0 0 1} cleavage flake. (b) Binocular microscope image of a Benson Mines orthoclase {0 0 1} cleavage flake. (c) Secondary
electron microscope image of the same Benson Mines orthoclase cleavage flake shown in (b). Cleavage flake orientations were determined
using electron backscatter diffraction patterns.

differ in thickness, we can assess whether or not the diffu- Foland and Xu (1990) inferred grain-scale diffusion from
sion domain is defined by sub-grain features. We analyzed fragments of Benson Mines orthoclase based on bulk loss
individual cleavage flakes of Benson Mines orthoclase and experiments. It follows that the non-linearity observed at
Madagascar cryptoperthite that varied in thickness by a high-temperature on the Madagascar cryptoperthite Arrhe-
factor of 2. Care was taken to select grains with high aspect nius plot (Fig. 3) is conclusively not due to a distribution of
ratios (e.g., >10:5:1) so as not to bias D/a2 values calculated sub-grain domain sizes (discussed in detail in Section 4.3).
using analytical solutions for infinite sheet geometry.1 Fol-
lowing grain selection, the width of each cleavage flake was 4.1.2. Evidence for grain-scale diffusion in exsolved
measured using a binocular microscope (Fig. 2). Half the plagioclase feldspars
thickness of each grain was assumed to be the relevant dif- In some instances, plagioclase crystals with sub-micron
fusive lengthscale for infinite sheet geometry. The results of to micron-sized exsolution lamellae associated with slow
the diffusion experiments are shown both normalized and cooling beneath the Huttenlocher and Bøggild solvi
un-normalized to the measured half-widths in Fig. 3. (Fig. 1) appear to diffuse Ar at the grain-scale. To illustrate
By inspection of Fig. 3a and b it is evident that thinner this point we use data obtained from a labradorite/bytow-
cleavage flakes yield higher values of Do/a2 at a given tem- nite (An70) from the Bushveld Complex, South Africa
perature and vice versa. When multiplied by the square of (Cassata et al., 2009). SEM images of HF-etched plagio-
the cleavage flake half-width, each aliquot yields equivalent clase crystals reveal sub-micron intergrowths (Fig. 4a),
values of D at a given temperature (Fig. 3c and d). Thus dif- which, based on the composition, are e-plagioclase and
ferences in Do/a2 values can be explained by differences in anorthite lamellae associated with slow-cooling beneath
macroscopic grain dimensions, which indicates that the the Huttenlocher solvus, consistent with observations of
physical crystal dimensions define the diffusion domain in other plagioclase crystals from the Bushveld Complex
these unaltered cryptoperthite and orthoclase feldspars. (e.g., McLaren, 1974). They exhibit normal zoning, with
Wartho et al. (1999) reached an equivalent conclusion significant (up to a factor of 2) K enrichment and subtle
based on in situ laser ablation observations of grain-scale, (<15%) Ca depletion near grain margins (Fig. 4b; Cassata
rim to core concentration gradients in preheated fragments et al., 2009). The 37ArCa/39ArK spectrum obtained from
of Madagascar cryptoperthite. Likewise, Foland (1974) and the step-wise degassing of an 1 mm crystal fragment
(Fig. 4c) mirrors the rim to core compositional variations
1 observed in situ by EPMA (Fig. 4b). If this sample con-
Grains with aspect ratios that deviate from ideal infinite sheet
tained a distribution of sub-grain domains, the Ca/K spec-
geometry yield erroneously high D/a2 values, where the maximum
error is 2.2 natural log units for a perfect sphere (Huber et al.,
tra would homogenize rim to core compositional variations
2011). in Ca and K, which is not observed. Likewise, 40Ar/39Ar
W.S. Cassata, P.R. Renne / Geochimica et Cosmochimica Acta 112 (2013) 251–287 257

Temperature (oC) Temperature (oC)


1400 1000 800 700 600 500 400 1400 1000 800 700 600 500 400
-5 -20
a 65 μm [001] c
110 μm [001]
-10 165 μm [010] -25
205 μm Bulk

-15 -30
ln (D/a2)

ln (D)
-20 Ea = 223.5 kJ/mole -35
ln(Do/a2) = 11.1 ln(s-1)
Ea = 216.0 kJ/mole
ln(Do/a2) = 9.5 ln(s-1)
-25 Ea = 213.5 kJ/mole -40 Benson Mines
ln(Do/a2) = 8.4 ln(s-1)
Ea = 219.7 kJ/mole Orthoclase
ln(Do/a2) = 9.0 ln(s-1) An0.0Ab3.0Or97.0
-30 -45
-5 -20
b 100 μm [001] d
100 μm [001]
115 μm [010]
-10 -25
145 μm [001]
175 μm [001]
185 μm Bulk
Ea = 277.5 kJ/mole
-15 -30
ln (D/a2)

ln(Do/a2) = 20.1 ln(s-1)


ln (D)
Ea = 275.0 kJ/mole
ln(Do/a2) = 19.2 ln(s-1)
-20 Ea = 278.2 kJ/mole -35
ln(Do/a2) = 19.1 ln(s-1)
Ea = 270.0 kJ/mole
ln(Do/a2) = 17.9 ln(s-1)
-25 Ea = 264.1 kJ/mole -40 Madagascar
ln(Do/a2) = 16.6 ln(s-1)
Ea = 273.4 kJ/mole Cryptoperthite
ln(Do/a2) = 17.4 ln(s-1) An0.2Ab5.6Or94.2
-30 -45
5 6 7 8 9 10 11 12 13 14 15 5 6 7 8 9 10 11 12 13 14 15
4 -1
10 /T (K ) 104/T (K-1)

Fig. 3. Ar Arrhenius plots for prograde heating of (a, c) Benson Mines orthoclase (39Ar) and (b, d) Madagascar cryptoperthite (39Ar) crystals,
calculated using equations for infinite sheet geometry for cleavage flakes and spherical geometry for equant (bulk) grains. Uncertainties in
D/a2 values are generally smaller than the symbols, but are not shown. Cleavage flake thickness and equant grain radii are listed in (a) and (b).
Data shown in (c) and (d) were calculated by multiplying the D/a2 values shown in (a) and (b) by half the thickness of the cleavage flake or the
equant grain radius. Anisotropy is not resolvable between {0 0 1} and {0 1 0} cleavage flakes at low-temperature (<800 °C). Apparent
discrepancies between equant grain and cleavage flake D values can be explained by our use of spherical geometry to model tetragonal prisms
(see text for discussion; Huber et al., 2011).

age spectra conform to single-domain diffusion theory (i.e., variations (0.8 natural log units), which may be due to dif-
they are not tiered; see Fig. 4 in Cassata et al., 2009) and ferences in the aspect ratios of the individual crystals (dis-
in situ laser ablation 40Ar/39Ar dating experiments reveal cussed above; Huber et al., 2011). If exsolution lamellae
young apparent ages at grain margins that increase toward defined the diffusion domain, each fragment, regardless of
the center (Fig. 4d). Thus it appears that the physical grain size, would yield the same D/a2 value at a given temperature
dimensions of this sample define the diffusion domain and because they have the same lamellar microstructure. There-
therefore micron-scale, strain-controlled intergrowths of fore, our experiments corroborate the conclusion of Par-
e-plagioclase and anorthite do not provide pathways for sons et al. (1999) that strain-controlled, coherent
efficient loss of Ar to grain surfaces. microstructures do not provide fast-pathways for Ar diffu-
A lack of correlation between lamellar microstructure sion, but rather define interfaces within complex domains.
and diffusive loss of Ar is further illustrated by diffusion We will discuss the curvilinear structure of the Arrhenius
experiments conducted on Schiller labradorite crystal frag- arrays shown in Fig. 5 at greater length in Section 4.3.
ments (An49.5Ab47.9Or2.6) from Nain, Canada, which com-
prise 100 nm thick exsolution lamellae with compositions 4.1.3. Evidence for sub-grain domains in altered, fractured,
of An35 and An65. Arrhenius arrays obtained from and inter-grown feldspars
equant crystal fragments that differ in size by a factor of To contrast with the previous sections, we now present
approximately 1.5 have nearly identical structure but are results of Ar diffusion experiments (Fig. 6a) conducted on
displaced vertically, where the larger grain yields lower fragments of deuterically altered alkali feldspar from the
D/a2 values at a given temperature (Fig. 5). At low- Bushveld Complex, South Africa (An0.8Ab19.0Or80.2). The
temperature, the offset between arrays (1.4 natural log alkali feldspar crystals contain strain-controlled, coherent
units) is slightly larger than expected based on grain size to semicoherent film perthite with 1 um thick Na-rich
258 W.S. Cassata, P.R. Renne / Geochimica et Cosmochimica Acta 112 (2013) 251–287

Fig. 4. Ar isotope data, compositional data, and a microscope image of single crystals of labradorite from the Bushveld Complex. (a)
Scanning electron microscope image of an HF-etched crystal. (b) Compositional data from an electron microprobe grain traverse (data from
Cassata et al., 2009). (c) 37ArCa/39ArK spectrum derived from an incremental degassing experiment (data from Cassata et al., 2009). (d) In situ
40
Ar/39Ar dating results from UV-laserprobe grain traverses (grain is 1 mm in diameter). The 37ArCa/39ArK spectrum (c) exhibits similar
compositional variations to those observed in situ by EPMA (b), indicating low-temperature extractions sampled gas from grain margins.
Likewise, in situ 40Ar/39Ar ages (d) revealed young apparent ages at grain margins, consistent with grain-scale diffusion.

lamellae and 3 um thick K-rich lamellae (Fig. 6b), with tron image of altered bytownite (An86.7Ab13.5Or0.3) from
incoherent patch and vein perthite replacing film perthite. the Stillwater Complex, which contains veins of muscovite
Arrhenius arrays obtained from crystals that differ in size and microcrystalline Na-rich feldspar. The non-linear
by a factor of approximately 2.5 do not reveal size-depen- Arrhenius array obtained from this sample (Fig. 7d) is con-
dent vertical displacements (Fig. 6a). The similarity of the sistent with alteration veins defining diffusion domains. In
low-temperature portions of the Arrhenius plots, irrespec- Fig. 7e we show a backscattered electron image of altered
tive of the macroscopic grain dimensions, suggests that each anorthite (An94.2Ab5.2Or0.6) from Grass Valley, which con-
crystal fragment contains comparably sized sub-grain do- tains large muscovite inclusions, plagioclase sub-grains, and
mains. These experiments corroborate the conclusions of patches of microcrystalline Na-rich feldspar. Again, the
Parsons et al. (1999) that deuteric microtextures provide Arrhenius array obtained from this sample (Fig. 7f) is
efficient pathways for Ar loss, and thus define diffusion sub-horizontal and consistent with a range in diffusive
domains. lengthscales. While plagioclase crystals from the Alpine
Arrhenius arrays consistent with a range of diffusive granitoid may represent viable candidates for thermal mod-
lengthscales were also obtained from microstructurally eling because the timing of formation and subsequent integ-
complex plagioclase crystals that contain sub-grains and/ rity of the diffusion domains can be assessed, the altered
or alteration. For example, in Fig. 7a we show a backscat- plagioclase feldspars from the Stillwater Complex and
tered electron image of a bytownite (An86.2Ab13.5Or0.3) Grass Valley are poor candidates for thermochronometry,
from Verzasca Valley, which contains quartz sub-grains as the domain structure may have been modified during
and alkali feldspar intergrowths. The Arrhenius array ob- or following cooling through the Ar partial retention zone.
tained from this sample (Fig. 7b) is sub-horizontal and con-
sistent with incoherent sub-grain boundary interfaces 4.1.4. Summary
between quartz, plagioclase, and alkali feldspar defining Ar diffuses in feldspar crystals that are devoid of hydro-
diffusion domains. In Fig. 7c we show a backscattered elec- thermal alteration and incoherent sub-grain intergrowths at
W.S. Cassata, P.R. Renne / Geochimica et Cosmochimica Acta 112 (2013) 251–287 259

Temperature (oC) but rather represent complex interfaces within composite


1400 1000 800 700 600 500 400 domains. Minor inclusions within otherwise coherent crys-
-5
Nain Complex tals appear not to affect the Ar domain structure, even when
Labradorite present in fairly large quantities. There is certainly some
-10 An49.4Ab47.4Or2.7 upper size limit beyond which feldspar crystals are invari-
ably fractured or cleaved. In our experience, grains larger
than 1 mm usually yield non-linear Arrhenius arrays typ-
-15 ical of material containing multiple diffusive length-scales.
ln (D/a2)

320 μm
490 μm Routine examination of samples using an electron micro-
-20 probe or SEM is generally sufficient to assess the nature
of the diffusion domain and determine the suitability of a
given mineral for thermochronometry.
-25 Ea = 276.6 kJ/mole
ln(Do/a2) = 2.0 ln(s-1)
Ea = 252.3 kJ/mole 4.2. Is diffusion isotropic?
ln(Do/a2) = 4.6 ln(s-1)
-30
5 6 7 8 9 10 11 12 13 14 15 To determine whether Ar diffusion is isotropic in feld-
4 -1
10 /T (K ) spars, we conducted detailed Ar diffusion experiments
(Figs. 3 and 8) on cleavage-bounded sheet-like flakes with
Fig. 5. Ar Arrhenius plots for prograde heating of Nain Complex
labradorite (37Ar) crystals, calculated using equations for spherical
high aspect ratios (>10:5:1). The lattice orientations of
geometry. Uncertainties in D/a2 values are generally smaller than these crystal fragments were determined by electron back-
the symbols, but are not shown. Approximate grain radii are based scatter diffraction (EBSD) with a Zeiss EVO scanning elec-
on estimated grain volumes calculated from the total yield of 37Ar, tron microscope (SEM) using TSL-OIM software. We
assuming spherical geometry. The vertical black bar shown at low- isolated {0 0 1} and {0 1 0} cleavage flakes from five coher-
temperature represents the expected offset between the arrays based ent, unaltered feldspars in which, as shown previously
on differences in grain size. The vertical gray bar represents and below, Ar diffuses at the grain-scale: Benson Mines
additional discrepancies that might arise due to geometric varia- orthoclase, Madagascar cryptoperthite, Easy Chair Crater
tions. Grains with aspect ratios that deviate from an ideal spherical anorthoclase, Surtsey labradorite, and Plush labradorite.
geometry yield erroneously low D/a2 values, where the maximum
In addition to the cleavage flakes used to examine diffu-
error is 2.2 natural log units (Huber et al., 2011).
sion normal to the {0 0 1} and {0 1 0} crystal faces (approx-
imately equivalent to the b and c axes), experiments were
conducted on equant grains, which include diffusion normal
the grain-scale, except where broken by cleavage planes or to the {1 0 0} crystal faces (approximately equivalent to the
fractures. Coherent lamellar interfaces typical of e-plagio- a axis). For the cleavage flakes, the relevant diffusion
clase intergrowths and strain-controlled perthite and anti- dimension for infinite sheet geometry is half of the total
perthite do not define Ar diffusion domain boundaries, thicknesses normal to the {0 0 1} and {0 1 0} crystal faces,

Fig. 6. (a) 39Ar Arrhenius plots for prograde heating of Bushveld Complex perthite crystals, calculated using equations for spherical
geometry. Uncertainties in D/a2 values are generally smaller than the symbols, but are not shown. Approximate grain radii (listed in the lower,
left corner) are based on estimated grain volumes calculated from the total yield of 39Ar, assuming spherical geometry. (b) BSE images
illustrating the perthitic microtexture. The crystals comprise strain-controlled film perthite with 1 lm thick Na-rich lamellae (darker) and
3 lm thick K-rich lamellae (lighter) and deuteric patch and vein perthite crosscutting and obscuring the film perthite. The Arrhenius arrays
do not reveal size-dependent vertical displacements, which indicates that these crystals comprise sub-grain domains that are probably defined
by the porous, microcrystalline deuteric alteration.
260 W.S. Cassata, P.R. Renne / Geochimica et Cosmochimica Acta 112 (2013) 251–287

Fig. 7. Ar Arrhenius plots for prograde heating and BSE images of (a, b) bytownite from a metamorphosed Alpine granitoid, (c, d) bytownite
from the Stillwater Complex, and (e, f) anorthite from Grass Valley. D/a2 values were calculated using equations for spherical geometry.
Uncertainties in D/a2 values are generally smaller than the symbols, but are not shown. The Alpine granitoid contains quartz sub-grains and
alkali feldspar intergrowths. The Stillwater Complex bytownite contains veins of muscovite and porous, microcrystalline Na-rich feldspathic
alteration. The Grass Valley anorthite contains large muscovite inclusions, plagioclase sub-grains, Na-rich feldspathic alteration, and Fe-rich
alteration. The Arrhenius arrays obtained from these samples are sub-horizontal and consistent with alteration surfaces and incoherent sub-
grain boundary interfaces between quartz, plagioclase, and alkali feldspar defining diffusion domains.

which we measured using a binocular microscope (Fig. 2 800 °C (Figs. 3c, d, 8b, d, and f). Regardless of size and
and Table 2). For the bulk crystals, grain volumes were cal- orientation, each grain yields D values that agree to within
culated based on the total yield of 37Ar or 39Ar (for plagio- approximately 0.5 natural log units or better (Table 2).
clase and K-feldspar, respectively) and grain geometries These minor differences can be explained by subtle varia-
were assumed to be spherical. tions in grain geometry (Huber et al., 2011) and thus limit
Anisotropy is not resolvable in the [1 0 0], [0 1 0], and the precision of the approach. Because Ar diffusion in feld-
[0 0 1] crystallographic directions at temperatures below spars appears to be isotropic, the appropriate geometry for
W.S. Cassata, P.R. Renne / Geochimica et Cosmochimica Acta 112 (2013) 251–287 261

Temperature (oC) Temperature (oC)


1400 1000 800 700 600 500 400 1400 1000 800 700 600 500 400
-5 -25
a 105 μm [001] b
105 μm [001]
-10 140 μm [010] -30
370 μm Bulk

-15 -35
ln (D/a2)

ln (D)
-20 Ea = 278.3 kJ/mole -40
ln(Do/a2) = 19.9 ln(s-1)
Ea = 300.0 kJ/mole
ln(Do/a2) = 22.2 ln(s-1)
-25 Ea = 281.7 kJ/mole -45 Easy Chair Crater
ln(Do/a2) = 19.0 ln(s-1)
Ea = 280.0 kJ/mole
Anorthoclase
ln(Do/a2) = 17.6 ln(s-1) An10.8Ab71.8Or17.4
-30 -50
-5 -25
c 95 μm [001] d
100 μm [010]
-10 120 μm [001] -30

-15 -35
ln (D/a2)

ln (D)

-20 -40
Ea = 220.2 kJ/mole
ln(Do/a2) = 11.7 ln(s-1)
-25 Ea = 216.9 kJ/mole -45 Plush, Oregon
ln(Do/a2) = 10.0 ln(s-1)
Labradorite
Ea = 241.8 kJ/mole
ln(Do/a2) = 13.6 ln(s-1) An63.9Ab35.4Or0.7
-30 -50
-5 -25
e 65 μm [010]
f
85 μm [001]
-10 85 μm [001] -30

-15 -35
ln (D/a2)

ln (D)

-20 -40
Ea = 173.5 kJ/mole
ln(Do/a2) = 3.6 ln(s-1)
-25 Ea = 170.4 kJ/mole -45
ln(Do/a2) = 2.6 ln(s-1) Surtsey, Iceland
Ea = 182.2 kJ/mole Labradorite
ln(Do/a2) = 3.7 ln(s-1) An59.5Ab39.6Or0.9
-30 -50
5 6 7 8 9 10 11 12 13 14 15 5 6 7 8 9 10 11 12 13 14 15
4 -1 4 -1
10 /T (K ) 10 /T (K )

Fig. 8. Ar Arrhenius plots for prograde heating of (a, b) Easy Chair Crater anorthoclase (39Ar), (c, d) Plush labradorite (37Ar), and (e, f)
Surtsey labradorite (37Ar) crystals, calculated using equations for infinite sheet geometry for cleavage flakes and spherical geometry for equant
(bulk) grains. Uncertainties in D/a2 values are generally smaller than the symbols, but are not shown. Cleavage flake thickness and equant
grain radii are listed in (a), (c), and (e). Data shown in (b), (d), and (f) were calculated by multiplying the D/a2 values shown in (a), (c), and (e)
by half the thickness of the cleavage flake or the equant grain radius. Anisotropy is not resolvable between {0 0 1} and {0 1 0} cleavage flakes at
low-temperature (<800 °C). Apparent discrepancies between equant grain and cleavage flake D values can be explained by our use of spherical
geometry to model tetragonal prisms (see text for discussion; Huber et al., 2011).

thermal modeling is one that most closely resembles the in the [0 1 0] and [1 0 0] directions, although the magnitude
physical grain dimensions, or the shape of sub-grain do- of anisotropy is variable. We hypothesize that anisotropic
mains if applicable. diffusion develops at high-temperature due to anisotropic
Above 800–1000 °C, anisotropy is resolvable in Benson changes in thermal expansion. In particular, a slowing or
Mines orthoclase, Madagascar cryptoperthite, Surtsey lab- cessation of c-axis thermal expansion at high-temperature
radorite, and Plush bytownite (Figs. 3c, d, 8b, d, and f). In (e.g., Brown et al., 1984), with continued expansion along
each instance, diffusion in the [0 0 1] direction is faster than the a- and b-axes, would result in slower Ar diffusion
262 W.S. Cassata, P.R. Renne / Geochimica et Cosmochimica Acta 112 (2013) 251–287

Table 2
Summary of feldspar cleavage flake diffusion parameters.
Sample Thickness Orientation Ea ± 1r ln(Do/a2) ± 1r ln(Do) ± 1r
Phase (lm) (kJ/mole) (ln(s1)) (ln(m2s1))
Easy Chair Crater 105 [0 0 1] 278.3 ± 19.5 19.9 ± 2.8 0.2 ± 2.8
Anorthoclase 140 [0 1 0] 281.7 ± 20.0 19.0 ± 2.9 0.1 ± 2.9
105 [0 0 1] 300.0 ± 21.2 22.2 ± 3.1 2.5 ± 3.1
286.0 ± 11.5 0.8 ± 1.7

Surtsey, Iceland 85 [0 0 1] 182.2 ± 7.1 3.7 ± 0.9 16.4 ± 0.9


Labradorite 85 [0 0 1] 170.4 ± 5.8 2.6 ± 0.8 17.5 ± 0.8
65 [0 1 0] 173.5 ± 5.8 3.6 ± 0.8 17.1 ± 0.8
174.5 ± 3.6 17.1 ± 0.5

Plush, Oregon 95 [0 0 1] 220.2 ± 17.9 11.7 ± 2.6 8.2 ± 2.6


Labradorite 120 [0 0 1] 241.8 ± 26.6 13.6 ± 4.0 5.8 ± 4.0
100 [0 1 0] 216.9 ± 24.1 10.0 ± 3.6 9.8 ± 3.6
224.0 ± 12.5 8.1 ± 1.9

Benson Mines 65 [0 0 1] 223.5 ± 3.3 11.1 ± 0.4 9.6 ± 0.4


Orthoclase 110 [0 0 1] 216.0 ± 3.4 9.5 ± 0.4 10.1 ± 0.4
165 [0 1 0] 213.5 ± 3.3 8.4 ± 0.4 10.4 ± 0.4
217.7 ± 1.9 10.0 ± 0.2

Madagascar 145 [0 0 1] 270.0 ± 7.9 17.9 ± 1.0 1.2 ± 1.0


Cyptoperthite 175 [0 0 1] 264.1 ± 7.5 16.6 ± 1.0 2.1 ± 1.0
100 [0 0 1] 277.5 ± 9.1 20.1 ± 1.2 0.3 ± 1.2
100 [0 0 1] 275.0 ± 6.2 19.2 ± 0.8 0.6 ± 0.8
115 [0 1 0] 278.2 ± 7.9 19.1 ± 1.0 0.4 ± 1.0
272.8 ± 3.4 0.8 ± 0.4
39
Easy Chair Crater, Benson Mines, and Madagascar results based on Ar.
Plush and Surtsey results based on 37Ar.

normal to the c–b and c–a planes (approximately equivalent multiple-domain feldspars. Thus while many feldspars
to [1 0 0] and [0 1 0]) than the a–b plane (approximately undeniably contain sub-grain domains (e.g., Figs. 6 and
equivalent to [0 0 1]). A systematic review of the thermal 7), it may not be possible to infer meaningful domain distri-
expansion behavior of feldspars and a detailed discussion bution parameters from their Arrhenius plots. In this sec-
of the proposed relationship between thermal expansion tion we review the evidence that curvature observed in
and Ar diffusivity is given in Section 4.3. Arrhenius plots obtained from unaltered, unfractured feld-
spars is unrelated to a range in diffusive lengthscales. We
4.3. Non-linear Arrhenius arrays then speculate on the crystallochemical causes of such
behavior and discuss the resulting implications for
In Sections 4.1 and 4.2 we showed that physical crystal thermochronometry.
dimensions define the Ar diffusion domain of unaltered
feldspars and that Ar diffusion is isotropic to within our 4.3.1. Evidence that Arrhenius non-linearity reflects
experimental resolution (approximately a factor of 2). We structural or mechanistic changes
further showed that Arrhenius arrays obtained from sin- Arrhenius arrays obtained from crystal fragments de-
gle-domain grains are nonetheless non-linear, exhibiting gassed using the same heating schedule yield qualitatively
curvature or other departures from linearity at high temper- identical results that are displaced vertically according to
atures (Figs. 3, 5, and 8). Such non-linearity, when not grain size variations (Figs. 3 and 8). For example, ECC
attributable to multiple Ar diffusion domains, indicates that anorthoclase (Figs. 8a and b) yields linear Arrhenius arrays
(1) the relative importance of different diffusion mecha- between 400 and 600 °C that give way to downward curva-
nisms (i.e., interstitial, cation exchange, vacancy migration, ture between 650 and 800 °C, and ultimately re-establish
etc.) varies with temperature, and/or (2) the diffusive med- linear Arrhenius relationships between 800 and 1100 °C
ium (i.e., the structure of the feldspar lattice) varies with melting temperature. In each of four aliquots, the offset be-
temperature. This observation has significant implications tween the low- and high-temperature linear Arrhenius ar-
for the application of MDD theory to feldspar rays is approximately 2 natural log units. Smaller grains
thermochronometry. If deviations from linearity arise in yield higher ln(Do/a2) values and vice versa (Fig. 8a). If
single-domain feldspars due to structural or mechanistic we initially assume that the curvature observed between
changes, then similar behavior presumably occurs in 650 and 800 °C reflects the progressive exhaustion of
W.S. Cassata, P.R. Renne / Geochimica et Cosmochimica Acta 112 (2013) 251–287 263

Temperature (oC) Temperature (oC)


1400 1000 800 700 600 500 400 700 600 500 400 300 250
-5 0
a b Madagascar
Cryptoperthite
-10 -5 An0.2Ab5.6Or94.2

-15 -10
ln (D/a2)

ln (D/a2)
-20 -15

Argon Neon
-25 -20
Ea = 264.1 kJ/mole Ea = 98.2 kJ/mole
ln(Do/a2) = 16.6 ln(s-1) ln(Do/a2) = 4.9 ln(s-1)
-30 -25
5 6 7 8 9 10 11 12 13 14 15 10 11 12 13 14 15 16 17 18 19 20
4 -1 4 -1
10 /T (K ) 10 /T (K )

Fig. 9. Arrhenius plots for (a) 22Ne and (b) 39Ar diffusion in Madagascar cryptoperthite crystals, calculated using equations for infinite sheet
(Ar) and spherical (Ne) geometries. Uncertainties in D/a2 values are generally smaller than the symbols, but are not shown. Linear regressions
fit to the Ar and Ne Arrhenius arrays are shown in both plots in green and blue, respectively. The temperature at which Ne is exhausted
(600 °C) is denoted with a vertical, dashed gray line. The Ar Arrhenius array exhibits curvature at >800 °C, whereas Ne yields a linear
Arrhenius array, which indicates that crystallochemical changes occurring at high-temperature may contribute to Arrhenius non-linearity. Ne
diffusion data are from Gourbet et al. (2012). (For interpretation of the references to color in this figure legend, the reader is referred to the
web version of this article.)

sub-grain diffusion domains of increasing size, then the con- (Fig. 9b) confirms that crystal fragments of this feldspar
sistent 2 natural log unit offset between low- and high- represent coherent, single diffusion domains (discussed
temperature arrays requires a consistent factor of 2.7 above in Section 4.1). The curvature observed at high-
variation in sub-grain domain size. The vertical displace- temperature on the Ar Arrhenius plot (Fig. 9a) therefore
ments between arrays [i.e., differences in ln(Do/a2)] require unambiguously reflects a crystallochemical change.
that smaller grains have smaller sub-grain domains and vice While almost all feldspars yielded Arrhenius arrays
versa. Clearly this is illogical, as individual fragments of this that curve downward at some temperature above 600 °C,
compositionally homogeneous megacryst a priori differ only plagioclase crystals with compositions between An40
in size. Not surprisingly, all aliquots yield similar values of and An100 (the I1-, P1-, T-, or e-plagioclase structures;
Do when Do/a2 is multiplied by the square of the grain ra- Fig. 1) are unique in that Arrhenius arrays invariably
dius (Fig. 8b). Thus the 2 natural log unit displacement curve upward prior to higher temperature downward cur-
of the Arrhenius array must reflect an intrinsic change in vature (i.e., the diffusivity increases significantly faster
the diffusive medium or diffusion mechanism(s). The same than expected from upward extrapolation of a linear, low-
argument can be applied to each of the samples shown in er-temperature Arrhenius relationship; Fig. 10). Such up-
Figs. 3, 5, and 8. ward curvature cannot be explained by a range of
The results of 22Ne and 39Ar diffusion experiments in diffusive lengthscales, unless the original diffusive length-
Madagascar cryptoperthite further illustrate the point scale was reduced during the course of incremental heating
(Fig. 9). Like 39Ar, 22Ne was synthetically produced during by fracturing or cleaving. While we know of no reason
neutron irradiation, by the reaction 23Na(n,np)22Ne (Cass- that Ca-rich plagioclase feldspars should cleave or fracture
ata, 2011). The Arrhenius array for Ar diffusion exhibits a more readily than other feldspars, we obtained SEM
pronounced downward curvature at 800 °C (Fig. 9a). Due images of a selection of samples subjected to a typical
to the significantly higher diffusivity of Ne in this alkali incremental heating schedule to confirm that these crystals
feldspar (Gourbet et al., 2012), it was possible to completely remain intact, and no cleaving or fracturing was observed.
degas 22Ne at temperatures below which we observe non- Furthermore, retrograde heating subsequent to upward
linearity on the Ar Arrhenius plot, and a single, linear curvature (not shown in Fig. 10) yields ln(D/a2) values
Arrhenius array was obtained (Fig. 9b). Because K and that approach those observed in prior prograde heating
Na (and therefore 39Ar and 22Ne) are uniformly distributed cycles at lower-temperature (i.e., upward curvature is
in this sample, Ar and Ne a priori share the same domain reversible; discussed below). Although upward, reversible
distribution.2 The linearity of the Ne Arrhenius array curvature is consistent with temperature-dependent anisot-
ropy (i.e., different apparent activation energies for diffu-
sion along different crystallographic axes; Huber et al.,
2
Subtle variations in the relative concentrations of 39Ar and 22Ne
2011), such an explanation can be dismissed on the
may exist at grain margins due to differences in recoil distance, but grounds that both equant grain fragments and {0 0 1}
such localized, sub-micron variations are volumetrically and {0 1 0} cleavage flakes yield indistinguishable
insignificant. Arrhenius arrays (Figs. 5, 8d, and 8f). Therefore, these
264 W.S. Cassata, P.R. Renne / Geochimica et Cosmochimica Acta 112 (2013) 251–287

Temperature (oC) 4.3.2. Molecular dynamics considerations


1400 1000 800 700 600 500 400 In this section we couple observations of Ar diffusion in
-5
Ea = 196.1 kJ/mole feldspars with molecular dynamics (MD) simulations of so-
ln(Do/a2) = 5.4 ln(s-1) lid–gas interactions and experimental data on the physical
a
-10 properties of feldspars, over a range of temperatures, to
better understand the causes of non-linear Arrhenius ar-
rays. At this point it is worth reviewing the atomic motions
-15
ln (D/a2)

that facilitate Ar diffusion through the feldspar lattice,


which include the direct exchange, vacancy, direct intersti-
-20 tial, and interstitialcy mechanisms (e.g., McDougall and
Harrison, 1999). During direct exchange, an Ar atom occu-
pying a cation or anion vacancy exchanges position with a
-25 Lunar (76535) neighboring cation or anion. Similarly, during diffusion by
Anorthite the interstitialcy mechanism, an Ar atom occupying an
An96.2Ab3.5Or0.3
-30
interstitial site displaces a cation or anion to a neighboring
-5 interstitial site. Thus the direct exchange and interstitialcy
b Ea = 196.4 kJ/mole
ln(Do/a2) = 7.0 ln(s-1)
mechanisms require cationic and/or anionic movements
(i.e., the breaking of bonds). The observation that noble
-10
gas diffusion is orders of magnitude faster than cation
and anion diffusion, particularly at low temperatures (see
-15
reviews by Baxter, 2010; Cherniak, 2010), seems to require
ln (D/a2)

that the direct interstitial mechanism, wherein an Ar atom


moves from one interstitial site to the next without net mo-
-20 tion of the cations or anions, is the dominant mode of
transport within the lattice (e.g., Lagerwall, 1962; Schmel-
ing, 1965; Voltaggio, 1985). Diffusion through vacancies,
-25 Duluth Complex
as well as along extended linear and planar defects (e.g., fis-
Bytownite
An74.9Ab24.7Or0.4 sion tracks in zircon), is presumably also important,
-30 although these features are volumetrically minor.
-5 Molecular dynamics simulations (e.g., Watanabe et al.,
c Ea = 160.4 kJ/mole
1995; Reich et al., 2007; Du et al., 2008; Saadoune et al.,
ln(Do/a2) = 0.6 ln(s-1)
2009) provide insights into the potential energy barriers
-10
that must be overcome for diffusion to proceed by the direct
interstitial mechanism (e.g., Born–Mayer or Lennard–Jones
-15 potentials). The energy required to traverse a given intersti-
ln (D/a2)

tial lattice configuration depends on quantum mechanical


interactions (e.g., electron shell repulsions and dispersion
-20
forces) between noble gas atoms and relevant cations
and/or anions (e.g., Dove, 2001). Thus the properties of
-25 Bushveld Complex the atoms (e.g., van der Waals radii, electron densities,
Labradorite charges, etc.), as well as their crystallographic arrangement
An65-70Ab30-35Or1-2 (i.e., bond lengths and angles), determine the potential en-
-30 ergy barrier to diffusion. At a given temperature, a bulk
5 6 7 8 9 10 11 12 13 14 15
104/T (K-1) Ar diffusion coefficient reflects the migration of hundreds
of billions of atoms through a variety of differing interstitial
Fig. 10. 37Ar Arrhenius plots for prograde heating of (a) lunar lattice configurations arranged in a regular and repeating
anorthite (from troctolite 76535) (b) Duluth Complex bytownite, manner defined by the crystal structure. Different intersti-
and (c) Bushveld Complex labradorite crystals, calculated using
tial environments have different potential energy barriers
equations for spherical geometry. Uncertainties in D/a2 values are
to be overcome and impart different vibrational frequencies
generally smaller than the symbols, but are not shown. Feldspar
samples with compositions between An40 and An100 (i.e., having upon Ar atoms (Reich et al., 2007; Saadoune et al., 2009).
the I1-, P1-, T-, or e-plagioclase structures) yield Arrhenius arrays Interstitial configurations with low potential energy barriers
that curve upward between 600 and 800 °C, and then downward at are most easily overcome, provided that their vibrational
higher-temperature. frequencies are similar, and the connectivity and availabil-
ity of such interstitial lattice configurations within the crys-
tal structure determines the pathway of most efficient
escape. The greatest potential energy barrier that is tra-
versed along a given crystallographic axis should, in theory,
observations can only be explained by a change in the nat-
approximate the Ea for diffusion in that direction if the lat-
ure of Ar diffusion, resulting from either structural modi-
tice is static (i.e., bond lengths and angles do not undergo
fications to the feldspar lattice, or a significant variation in
changes) when subjected to heating or cooling. The
the mechanism(s) of diffusion.
W.S. Cassata, P.R. Renne / Geochimica et Cosmochimica Acta 112 (2013) 251–287 265

apparent Ea for bulk crystal diffusion is, a priori, intermedi- between noble gases and lattice atoms (i.e., the potential en-
ate between the highest and lowest values of the individual ergy barriers to diffusion) are exceedingly sensitive to subtle
crystallographic directions, and need not be continuous changes in bond lengths and angles. For example, the Len-
(i.e., a crystal with directionally-dependant potential energy nard–Jones potential (U), which describes the distance-
barriers may not yield a linear Arrhenius array, but the dependence of gas–gas and gas–solid repulsive forces
slope will always be intermediate between the Ea for (e.g., Du et al., 2008), is given by
diffusion along the individual crystallographic directions). $   6 %
12
In Fig. 11 we show Arrhenius arrays that would result from rij rij
U ðrij Þ ¼ 4eij  ; ð1Þ
the degassing of two hypothetical cubic grains that have rij rij
directionally-dependent potential energy barriers. Note that
the apparent Ea’s obtained are not equivalent to any of where i and j are two interacting atoms, rij is the distance
three individual crystallographic axes. When the contrast between them, rij is the Lennard–Jones size parameter
in potential energy barriers is modest (Fig. 11a), linear based on van der Waals radii, and eij is the Lennard–Jones
Arrhenius arrays are obtained. When the contrast is energy parameter determined empirically or through molec-
significant (Fig. 11b), the apparent Ea approximates the ular dynamics simulations (see Du et al. (2008) for various
pathway of most efficient escape at low-temperature, but noble gas values). Because of the exponential distance-
contributions from alternative pathways may become dependence, small (2–4%) changes in bond lengths that
significant at high-temperature and Arrhenius arrays may occur due to thermal expansion of feldspars during incre-
exhibit upward curvature. If the lattice is static over the mental heating experiments (e.g., Brown et al., 1984) can
range of temperatures employed in a diffusion experiment, result in significant (>50%) reductions in potential energy
then in theory, downward curvature on an Arrhenius plot barriers to diffusion (Fig. 12). Therefore an Arrhenius array
would never occur, as deviations from linearity only obtained by incrementally degassing a feldspar crystal
arise due to an increasing proportion of diffusion reflects diffusion through a dynamic crystal with tempera-
proceeding along a previously less significant crystallo- ture-dependent potential energy barriers and vibrational
graphic direction. frequencies. For this reason, molecular dynamics simula-
Feldspars, and most minerals for that matter, are not tions of noble gas diffusion that do not incorporate temper-
static during incremental heating experiments, and there- ature-dependent mineral structures (e.g., Reich et al., 2007;
fore more complex Arrhenius arrays are predicted by Saadoune et al., 2009) are not expected to yield apparent
molecular dynamics considerations. Thermal expansion diffusion parameters consistent with incremental degassing
and concomitant flexures of the feldspar tetrahedral frame- experiments.
work, as well as changes in the location, coordination, and Consider a hypothetical primitive isometric crystal with
vibrational modes of cations, produce anisotropic changes oxygen (O) atoms at its lattice nodes and unit cell dimen-
in bond lengths and angles upon heating. Repulsive forces sions a = b = c = 3.00 Å at 0 °C (Fig. 13a and Table 3).

Temperature (oC) Temperature (oC)


1400 1000 800 700 600 500 400 1400 1000 800 700 600 500 400
-5
Ua = 160.0 kJ/mole Ua = 140.0 kJ/mole
Ub = 200.0 kJ/mole Ub = 280.0 kJ/mole
-10 Uc = 240.0 kJ/mole Uc = 340.0 kJ/mole
Ea = 172.7 kJ/mole

-15
ln (D/a2)

-20

-25

a b
-30
5 6 7 8 9 10 11 12 13 14 15 5 6 7 8 9 10 11 12 13 14 15
4 -1
10 /T (K ) 104/T (K-1)

Fig. 11. Modeled Arrhenius arrays (gray circles) that would result from degassing cubic grains with directionally-dependant diffusive
anisotropy. The red, blue, and green Arrhenius relationships describe diffusion in the a, b, and c crystallographic directions, respectively, and
the corresponding potential energy barriers to diffusion (U) are listed in upper, right corner of each plot. (a) Crystals with modestly
contrasting potential energy barriers yield linear Arrhenius arrays. The apparent Ea for bulk crystal diffusion (listed in the upper, right corner)
is intermediate between the highest and lowest values of the individual crystallographic directions. (b) Crystals with strongly contrasting
potential energy barriers yield curvilinear Arrhenius arrays. The apparent Ea approximates the pathway of most efficient escape at low-
temperature, and increases at higher-temperature as contributions from alternative pathways become significant. (For interpretation of the
references to color in this figure legend, the reader is referred to the web version of this article.)
266 W.S. Cassata, P.R. Renne / Geochimica et Cosmochimica Acta 112 (2013) 251–287

10
5
temperature, then upward and downward curvature may
Ar-O be observed on Arrhenius plots, coincident with the acceler-
Ar-Na ation or deceleration of the thermal expansion rate, respec-
Potential Energy Barrier (kJ/mole)

4
10 Ar-Ca tively (Fig. 13d and e). If thermal expansion proceeds
linearly with temperature but anisotropically (e.g., the lattice
expands more rapidly along the c axis), then downward cur-
3
10 vature may be observed on Arrhenius plots (Fig. 13f). There-
fore, it is not surprising that bulk diffusion coefficients
describing noble gas migration through complex triclinic or
2
10 monoclinic crystals that undergo structural changes when
heated do not define strictly linear Arrhenius arrays. In feld-
1
spars, thermal expansion rarely proceeds linearly or isotrop-
10 ically as a function of temperature.
With this in mind, we turn to a detailed description of
0
the feldspar structure, which is central to our discussion of
10 the relationship between Ar diffusion kinetics and temper-
1.4 1.6 1.8 2.0 2.2 2.4
ature-dependent changes in cell parameters, and therefore
Distance (Angstroms)
potential energy barriers. Oxygen atoms (O) form tetrahe-
Fig. 12. Plot of the distance-dependence of the Ar–Na, Ar–Ca, and dra centered by aluminum and silicon atoms (T-sites), the
Ar–O Lennard–Jones potentials, calculated from Eq. (1) using the corners of which are shared, creating four-member rings.
Lennard–Jones parameters listed in Table 3. These tetrahedral rings join together forming kinked
chains, or “crankshafts” (Fig. 14a). Linking chains to-
gether across mirror planes creates large interstices (M-
Assume that it undergoes linear, isotropic thermal expan- sites) that are filled by Na, K, and Ca atoms (Fig. 14a).
sion of 4% between 0 and 1200 °C (approximately equiva- Thus the feldspar structure comprises an aluminosilicate
lent in magnitude to albite thermal expansion; Brown framework with alkali and alkali earth cations occupying
et al., 1984). The potential energy barrier to Ar diffusion large structural cavities (Fig. 14b). The symmetry of feld-
at a given temperature can be calculated by solving Eq. spars is controlled by the degree of Al and Si ordering on
(1) for the minimum in Ar–O repulsive interactions using T-sites, composition, pressure, and temperature. All Na-
the Lennard–Jones parameters listed in Table 3. The asso- and Ca-rich feldspars are triclinic (unit cell angles
ciated vibrational frequencies (to) can be approximated by a – b – c – 90°) at room temperature, as the lattice is dis-
solving an equation (Glicksman, 2000) adapted from tran- torted about the small Ca and Na atoms (Fig. 14c). At
sition state theory (Vineyard, 1957): high temperatures, disordered Na-rich alkali and plagio-
 1 clase feldspars (e.g., anorthoclase, high albite, high oligo-
1 U 2 clase; Fig. 1) undergo a displacive transition to
to ¼ pffiffiffi ; ð2Þ
2 md 2 monoclinic symmetry (unit cell angles a – b = c = 90°)
where m is the atomic weight of Ar, d is the jump distance due to thermal vibrations of the cations and tetrahedral
between two minima and is equal to the relevant unit cell framework (Fig. 14d). Sanidine is monoclinic at all tem-
dimension, and U is the potential energy barrier to diffu- peratures because the disordered framework is held apart
sion. These vibrational frequencies can be translated into by large K atoms (Fig. 14d). Ordered Na- and K-rich feld-
tracer diffusion frequency factors (mo) using Einstein’s equa- spars (e.g., microcline and low albite) are triclinic at all
tion (see Reich et al., 2007): temperatures. Regardless of T-site ordering and tempera-
ture, Ca-rich plagioclase feldspars (e.g., andesine, labra-
1 dorite, bytownite, anorthite, and Ca-rich oligoclase;
mo ¼ d 2 to : ð3Þ
2 Fig. 1) are triclinic. Orthoclase is a metastable alkali feld-
Fig. 13b depicts the Arrhenius array that would result spar with pseudo-monoclinic symmetry (monoclinic aver-
from incrementally heating our hypothetical isometric crys- age symmetry), comprising submicroscopic triclinic
tal between 600 and 1100 °C. By inspection it is clear that domains in which Al is ordered on different T1-sites (Par-
the apparent Ea (268.9 kJ/mole) overestimates the 600 °C po- sons and Lee, 2005). Si and Al diffusion is exceedingly
tential energy barrier to diffusion (203.5 kJ/mole) by 32%, slow could be studied at typical laboratory temperatures
even as the potential energy barrier is reduced with increasing and timescales (e.g., Carpenter et al., 1985), and thus
temperature due to thermal expansion. If thermal expansion changes in symmetry due to ordering are not expected
is negative (Fig. 13c), the apparent Ea (198.1 kJ/mole) under- to occur during step-heating. In the following sections,
estimates the 600 °C potential energy barrier (339.0 kJ/mol) we discuss how structural state and changes in bond
by 42%. In this regard, it would be interesting to determine lengths (both T–O and M–O) and bond angles (O–T–O,
He diffusion kinetics in proton-irradiated cubic zirconium T–O–T, M–O–T, O–M–O) that occur during step-heating
tungstate, a mineral that undergoes remarkable isotropic experiments affect diffusion kinetics.
negative thermal expansion over laboratory temperatures
typical of He diffusion experiments (e.g., 100–500 °C). If 4.3.2.1. Na-rich feldspars. Disordered albite, anorthoclase,
thermal expansion proceeds non-linearly as a function of and Na-rich oligoclase provide the most straightforward
W.S. Cassata, P.R. Renne / Geochimica et Cosmochimica Acta 112 (2013) 251–287 267
Temperature (oC)
1400 1000 800 700 600 500 400
-10
U1100 = 165.0 kJ/mole
a U1000 = 172.2 kJ/mole
-15
b U900 = 179.5 kJ/mole
U800 = 187.1 kJ/mole
U700 = 195.2 kJ/mole
U600 = 203.5 kJ/mole
-20 Ea = 268.9 kJ/mole

ln (D/a2)
-25
Linear/Isotropic
T
a = b = c = 3.00 Å -30

Temperature ( C)
o
ΔV
-35
1400 1000 800 700 600 500 400
-30 -10
U1100 = 421.1 kJ/mole d
c U1000 = 402.8 kJ/mole
U900 = 385.9 kJ/mole
-35 -15
U800 = 369.7 kJ/mole
U700 = 353.8 kJ/mole
U600 = 339.0 kJ/mole
-40 Ea = 198.1 kJ/mole -20
ln (D/a2)

ln (D/a2)
-45 -25
Neg. Lin./Isotropic Non-linear/Isotropic
ΔV T
-50 -30

T ΔV
-55 -35
-10 -10
e f
a,b
-15 -15
c

-20 -20
ln (D/a2)

ln (D/a2)

-25 -25
Non-linear/Isotropic Linear/Anisotropic
T T
-30 -30
a,b
ΔV c ΔV
-35 -35
5 6 7 8 9 10 11 12 13 14 15 5 6 7 8 9 10 11 12 13 14 15
104/T (K-1) 104/T (K-1)

Fig. 13. (a) A hypothetical primitive isometric crystal with oxygen atoms at its lattice nodes (black) and unit cell dimensions a = b = c = 3 Å
at 0 °C. An Ar atom (red) is shown at the body center. (b) Modeled Arrhenius relationship (black line and colored circles) that would result
from incrementally heating the isometric crystal, assuming that it undergoes linear, isotropic thermal expansion of 4% between 0 and 1200 °C
(shown schematically in the inset: T = temperature, V = volume). The potential energy barriers to Ar diffusion (U) at various temperatures,
calculated by solving Eq. (1) for the minimum in Ar–O repulsive interactions, are listed in the upper-right corner. The associated vibrational
frequencies used to plot the temperature-dependent Arrhenius relationships (colored lines) were calculated by solving Eqs. (2) and (3). The
apparent Ea (listed in the upper-right corner) reflects the temperature-dependent changes in unit cell dimensions, and therefore potential
energy barriers and vibrational frequencies. (c) Modeled Arrhenius relationship calculated assuming that the isometric crystal undergoes
linear, isotropic thermal expansion of 4% between 0 and 1200 °C. (d) Modeled Arrhenius array calculated assuming that the isometric
crystal undergoes non-linear, isotropic thermal expansion of 3.6% between 0 and 900 °C and 0.4% between 900 and 1200 °C. Downward
curvature is observed when the thermal expansion coefficient decreases at 900 °C. (e) Modeled Arrhenius array calculated assuming that the
isometric crystal undergoes non-linear, isotropic thermal expansion of 1% between 0 and 800 °C and 3% between 800 and 1200 °C. Upward
curvature is observed when the thermal expansion coefficient increases at 800 °C. (f) Schematic Arrhenius array calculated assuming that the
isometric crystal undergoes non-linear, anisotropic thermal expansion of 2% between 600 and 1200 °C along the c-axis and 0% between 600
and 1200 °C along the a- and b-axes. The Arrhenius relationship for diffusion along the c-axis is temperature-independent because the a–b
plane does not undergo expansion (i.e., the interstitial lattice configurations traversed in the c-direction do not expand). Downward curvature
is observed at high-temperature as diffusion along the c-axis becomes increasingly less significant. (For interpretation of the references to color
in this figure legend, the reader is referred to the web version of this article.)
268 W.S. Cassata, P.R. Renne / Geochimica et Cosmochimica Acta 112 (2013) 251–287

Table 3
Definitions of thermal expansion model (Fig. 13) parameters and their values.
Parameter Notation Parameter value Parameter unit Equation
Crystal dimensions at 0 °C A0, B0, C0 100 lm
Unit cell dimensions at 0 °C a0, b0, c0 3.00 Å
Jump distance at 0 °C (i.e., unit cell width) d0 3.00 Å (1)
Interstitial distance between Ar and lattice atoms at 0 °C rAr–X 2.12 Å (1)
Ar–O Lennard–Jones energy parametera eAr–O 0.9143 kJ/mole (1)
Ar–Ca Lennard–Jones energy parametera eAr–Ca 1.7453 kJ/mole (1)
Ar–Na Lennard–Jones energy parametera eAr–Na 2.4929 kJ/mole (1)
Ar–O Lennard–Jones size parameterb rAr–O 3.0566 Å (1)
Ar–Ca Lennard–Jones size parameterb rAr–Ca 2.5845 Å (1)
Ar–Na Lennard–Jones size parameterb rAr–Na 2.5756 Å (1)
Interstitial Ar–O Lennard–Jones potential at 0 °C UAr–O 262.2 kJ/mole (1)
Vibrational frequency for Ar–O at 0 °C toAr–O 1.91  1011 s1 (2)
Tracer diffusion frequency factor for Ar–O at 0 °C moAr–O 8.59  109 m2 s (3)
Thermal expansion coefficientc a Variable Å/°C
a
Lennard–Jones energy parameters are from Du et al. (2008).
b
Lennard–Jones size parameters are from Watanabe et al. (1995).
c
Variable thermal expansion coefficients were employed (see Fig. 13).

illustration of the relationship between structural state and in Na-rich feldspars can thus been divided into two regimes
Ar diffusion kinetics, as they undergo a displacive transition of regular thermal expansion that do not significantly alter
during step-heating experiments wherein cell parameters the structure (associated with the triclinic and monoclinic
(unit cell lengths a, b, and c, and unit cell angles a, b, and symmetry states) separated by the triclinic–monoclinic tran-
c) change significantly over a narrow (200 °C) tempera- sition zone, wherein expansion in the b–c plane slows or
ture interval (e.g., Winter et al., 1979). At low-temperature, ceases (Fig. 15a and b).
all Na-rich feldspars are triclinic because the aluminosili- Likewise, the temperature-dependence of Ar diffusivity
cate framework is puckered around small Na cations. The in Na-rich feldspars that undergo a triclinic to monoclinic
coordination about the M-site is irregular, with 5–7 of the transition when heated can be divided into two dissimilar
nine nearest oxygen atoms having short M–O distances that regimes separated by a transition zone. For example, in
define the primary coordination environment (Fig. 14c) and Fig. 16 we show Arrhenius plots obtained from two such
2–4 having significantly longer M–O distances (Kroll and feldspars: anorthoclase from Easy Chair Crater and oligo-
Mueller, 1980; Harlow, 1982; Brown et al., 1984). Upon clase from Mono Lake. Both samples yield linear Arrhenius
heating from room temperature, thermal expansion of arrays at low- and high-temperature separated by a curvi-
Na-rich feldspars proceeds such that a, b, c, and the unit linear transition zone, wherein the rate at which diffusivity
cell volume (V) increase approximately linearly as a func- increases with temperature slows (i.e., the Arrhenius arrays
tion of temperature, with only minor changes to a, b, and exhibit downward curvature; Fig. 16). We suggest that cur-
c (Fig. 15a and b; Prewitt et al., 1976; Winter et al., 1979; vature observed on Arrhenius plots reflects the change in
McMillan and Brown, 1980; Brown et al., 1984). As heating the thermal expansion regime associated with the tri-
proceeds toward the triclinic–monoclinic transition temper- clinic–monoclinic transition. When expansion in the b–c
ature (the “critical temperature”, which depends on compo- plane slows or ceases and a approaches 90°, the rate at
sition and T-site order), the thermal expansion regime which Ar diffusivity increases with temperature appears to
changes: at least one component of expansion in the b–c slow, as predicted by molecular dynamics considerations
plane slows or ceases and a accelerates towards 90° (described above). We propose that linear arrays observed
(Fig. 15b) until monoclinic symmetry is attained at the crit- below and above the non-linear portions of the Arrhenius
ical temperature (Winter et al., 1979; Brown et al., 1984). plots correspond to the approximately linear thermal
This change in thermal expansion regime corresponds to expansion regimes below and above the triclinic–mono-
a shortening of longer M–O bonds (Fig. 14d), lengthening clinic transition zone, and therefore describe Ar diffusion
of shorter M–O bonds, and rotation (tilting) of the tetrahe- in the triclinic and monoclinic forms of the feldspars,
dral framework (Harlow, 1982). Thus over an 200 °C respectively. The high-temperature linear array observed
temperature range (that depends in detail on composition on the ECC anorthoclase Arrhenius plot (Fig. 16a) com-
and T-site order), thermal vibrations of the tetrahedral mences at the exact temperature at which monoclinic sym-
framework and cations cause the lattice to undergo rapid metry is expected to be attained in a feldspar of its
changes in unit cell parameters, bond lengths and angles, composition (750 °C; Kroll and Bambauer, 1981). The
and in the coordination environment of the M-site. Above Arrhenius plot obtained from Mono Lake oligoclase
the critical temperature, thermal expansion again proceeds (Fig. 16b) exhibits downward curvature approximately
in a regular manner, with a, b, c, and V increasing approx- 150 °C lower than predicted by Kroll and Bambauer
imately linearly with temperature (Fig. 15a; Brown et al., (1981), which may reflect a more disordered lattice. The
1984). The temperature-dependence of thermal expansion critical temperature for the monoclinic–triclinic transition
W.S. Cassata, P.R. Renne / Geochimica et Cosmochimica Acta 112 (2013) 251–287 269

a b

Al

Si

~c O
~a
Ca
~b ~a

c Triclinic M-site d Monoclinic M-site

c c

b b

Fig. 14. Stylized diagrams of various aspects of the feldspar structure. (a) Schematic illustration of the “crankshafts” formed by four-member
rings of tetrahedra (represented by chains of squares drawn with solid lines) running approximately parallel to the a-axis. These crankshafts
are linked across (0 1 0) mirror planes (approximately orthogonal to the b-axis; dashed lines) creating large interstices that are filled by Na, K,
and Ca atoms (M-sites). Modified after Megaw (1973). (b) Projection down [0 1 0] of anorthite showing details of the cation coordination
environment and tetrahedral framework. Crankshafts of Al and Si tetrahedra running approximately parallel to the a-axis are denoted with
different shades of blue (Al) and orange (Si). Six- and sevenfold coordinated M-sites are denoted by red and salmon Ca atoms, respectively.
Solid and dashed red and salmon lines represent the primary and secondary coordination spheres, respectively. Redrafted from Megaw et al.
(1962). (c, d) Coordination environment of Na atoms in the triclinic (sevenfold) and monoclinic (ninefold) albite structures. Solid and dashed
lines represent the primary and secondary coordination spheres, respectively. Redrafted from Winter et al. (1979). (For interpretation of the
references to color in this figure legend, the reader is referred to the web version of this article.)

can be lowered by hundreds of degrees in highly disordered structure, and a transition zone characterized by significant
feldspars (Barth, 1965; Stewart and von Limbach, 1967). changes in the unit cell angle a, associated with the transi-
Unlike the disordered Na-rich feldspars discussed tion toward monoclinic symmetry. In Fig. 17 we compare
above, ordered Na-rich and more Ca-rich feldspars of the Arrhenius plots obtained from four such compositionally
C1 symmetry group (e.g., ordered oligoclase, ordered and homogeneous feldspars: oligoclase from the achondrite
disordered andesine; Fig. 1) do not transform to a mono- GRA 06128, oligoclase/andesine from the Fish Canyon
clinic structure below their melting temperature, although Tuff, andesine from the Paraná-Etendeka large igneous
the rate of change in cell parameters nonetheless increases province, and labradorite from the Klamath Mountains.
at high temperatures (>900 °C; Prewitt et al., 1976; Winter Arrhenius plots obtained from these samples can be divided
et al., 1979; Brown et al., 1984). Thus the thermal expansion into two portions: (1) a low-temperature, linear array, and
of more ordered Na-rich feldspars can be divided into two (2) a high-temperature, downwardly curved array. We pro-
parts (Fig. 15c and d): a low temperature regime character- pose that portions (1) and (2) are associated with the tri-
ized by approximately linear changes in a, b, c, and V with clinic and transitional structures, respectively. Thus the
increasing temperature, associated with the triclinic transitional structure of these feldspars, like that of the
270 W.S. Cassata, P.R. Renne / Geochimica et Cosmochimica Acta 112 (2013) 251–287

a Triclinic Mono. b
a
b
β

c α
Cell Dimension Triclinic Monoclinic

Cell Angle
γ

c d
a
b
β

c α
Triclinic Triclinic
Cell Dimension

Cell Angle
γ

e f
a
b
β

Monoclinic c
Monoclinic
Cell Dimension

Cell Angle

α
γ

Temperature Temperature

Fig. 15. Schematic plots of the effects of thermal expansion on unit cell dimensions and angles in (a, b) Na-rich feldspars that undergo a
triclinic to monoclinic transition at high-temperature, (c, d) Na-rich plagioclase feldspars that do not undergo the aforementioned transition,
and (e, f) K-rich feldspars that are monoclinic at all temperatures. Trends are based on references given in Section 4.3.

Na-rich samples discussed above, is associated with a feldspars in that they exhibit monoclinic symmetry at room
downward curvature on Arrhenius plots. Further study of temperature. Because the tetrahedral framework is held
Ar diffusion in feldspars with well defined, low-temperature apart by large K atoms (i.e., it is not puckered about small
monoclinic–triclinic transition temperatures (e.g., Pb-feld- Na cations; Fig. 14d), a displacive transition does not occur
spar, Benna et al., 2000; synthetic ternary feldspars, Kroll during incremental heating experiments. Instead, unit cell
and Bambauer, 1981) would be productive in better delin- dimensions a, b, and c increase approximately linearly as
eating the relationship between symmetry changes and no- a function of temperature, with more expansion accommo-
ble gas diffusion. dated along the a-axis (Fig. 15e and f; Brown et al., 1984;
Hovis et al., 1999; Mackert et al., 2000). Arrhenius plots
4.3.2.2. K-rich feldspars. Unexsolved K-rich feldspars (e.g., obtained from such K-rich feldspars mirror the simplicity
sanidine and orthoclase) are unique among ternary of their thermal expansion. For example, Fig. 18 depicts
W.S. Cassata, P.R. Renne / Geochimica et Cosmochimica Acta 112 (2013) 251–287 271

Temperature (oC) Temperature (oC)


1400 1000 800 700 600 500 400 1400 1000 800 700 600 500 400
-25 -5
a b
-30 -10
Triclinic

-35 -15

ln (D/a2)
ln (D)

Monoclinic Triclinic
-40 -20
Monoclinic Adhering
Glass
-45 Easy Chair Crater -25 Mono Lake
Anorthoclase Oligoclase
An10.8Ab71.8Or17.4 An19.6Ab74.2Or6.2
-50 -30
5 6 7 8 9 10 11 12 13 14 15 5 6 7 8 9 10 11 12 13 14 15
104/T (K-1) 104/T (K-1)

Fig. 16. Ar Arrhenius plots for prograde heating of Na-rich feldspars. (a) Mono Lake oligoclase (37Ar) and (b) Easy Chair Crater
anorthoclase (39Ar). Uncertainties in D/a2 values are generally smaller than the symbols, but are not shown. The temperature-dependence of
Ar diffusivity these samples can be divided into two regimes of constant apparent Ea separated by a transition zone. The low- and high-
temperature linear Arrhenius arrays correspond to the triclinic and monoclinic structures of the feldspars, respectively. The intermediary
curvature reflects the change in cation coordination and thermal expansion regime (Fig. 15a) associated with the triclinic–monoclinic
transition. Lighter, low-temperature data obtained from Mono Lake oligoclase reflect Ar derived from minor adhering glass, and can be
distinguished on the basis of their Ca/K ratios.

the results of Ar diffusion experiments conducted on two 1980). As discussed previously, upward curvature does
disordered, K-rich feldspars: sanidine from the Gulf of Sal- not reflect a reduction in the grain size with heating (i.e.,
erno and sanidine from the Fish Canyon Tuff. The Arrhe- thermal fracturing) or temperature-dependent anisotropy
nius plots are linear at all temperatures below their (i.e., the increasing importance of diffusion along a given
melting points. Fig. 3a depicts the results of Ar diffusion crystallographic axis), and therefore presumably reflects a
experiments conducted on four aliquots of Benson Mines unique aspect of the structural response of Ca-rich feld-
orthoclase. Like the sanidine crystals, Benson Mines ortho- spars to heating.
clase yields robust, linear Arrhenius arrays (Fig. 3a), with Before we expound upon some potential crystallochem-
only modest departures from linearity observed near the ical causes of the unique behavior exhibited by Ca-rich feld-
melting point in bulk grain and {0 1 0} cleavage flakes. Such spars, here we review their crystal structure and changes
behavior may be indicative of a slowing or cessation of that occur when heated. We discuss the Ca-rich feldspars
expansion along the c-axis, which could cause downward in two groups: (1) anorthite (An>95) and (2) intermediate
curvature on Arrhenius plots obtained from grain frag- plagioclase (An45–95). At room temperature, the unit cell
ments that include diffusion normal to the c-axis (i.e., in of anorthite has four non-equivalent M-site locations due
the a and b directions). Unfortunately, we are unaware of to lattice distortions associated with its high Al concentra-
thermal expansion or detailed structural data on orthoclase tion (P1 symmetry; Megaw et al., 1962). The non-equiva-
at these temperatures. To summarize, monoclinic K-rich lent M-sites give rise to c-reflections in X-ray diffraction
feldspars (i.e., sanidine and orthoclase), unlike other ter- patterns (Foit and Peacor, 1973) and can be grouped into
nary feldspars, do not undergo structural or thermal expan- two dissimilar pairs based on their average M–O distances
sion regime changes when heated and yield linear Arrhenius (i.e., coordination environments; Megaw et al., 1962). At
arrays at temperatures below 1000 °C. room-temperature, the lattice distorts about the small, diva-
lent Ca atoms in one of two undifferentiated ways related
4.3.2.3. Ca-rich feldspars. Ca-rich plagioclase specimens by C2 symmetry (i.e., the lattice can distort “left” or
(An>45) belonging to the I1, T, P1, e1, and e2 structural “right”). Random nucleation will produce both left and
groups (Fig. 1) are unique among ternary feldspars in that right domains, each of which contain the four non-equiva-
Arrhenius arrays (Figs. 5 and 10) invariably curve upward lent M-sites. This process is somewhat analogous to the for-
at intermediate heating temperatures (600–800 °C), prior to mation of “left” and “right” domains in orthoclase, wherein
a downward curvature at higher temperatures (>900 °C). Al orders on equivalent T1-sites related by mirror symme-
This behavior occurs in Ca-rich plagioclase specimens of try, creating a tweed microtexture (Parsons and Lee,
both high (disordered) and low (ordered/exsolved) struc- 2005). Likewise, at room temperature intermediate plagio-
tural states (discussed below), and is therefore not related clase specimens (e1, e2, and I1) have non-unique M-site
to the modulated e-plagioclase structures associated with positions that give rise to e-reflections in X-ray diffraction
the Huttenlocher and Bøggild solvi (Fig. 1), wherein nano- patterns (Wenk et al., 1980; Wenk and Nakajima, 1980),
meter-sized albitic domains separate anorthitic domains in and the lattice can distort in “left” and “right” arrange-
an antiphase relationship (e.g., Wenk and Nakajima, ments related by 1/2(a + b + c) symmetry (Brown, 1984).
272 W.S. Cassata, P.R. Renne / Geochimica et Cosmochimica Acta 112 (2013) 251–287

Temperature (oC) Temperature (oC)


1400 1000 800 700 600 500 400 1400 1000 800 700 600 500 400
-5
Ea = 226.6 kJ/mole Ea = 209.4 kJ/mole
a ln(Do/a2) = 11.5 ln(s-1) b ln(Do/a2) = 8.0 ln(s-1)

-10

-15
ln (D/a2)

-20

-25 Klamath Mtns. GRA 06128


Labradorite Oligoclase / Albite
An52.8Ab44.2Or3.0 An14.1Ab84.1Or1.9
-30
-5
Ea = 206.9 kJ/mole Ea = 218.5 kJ/mole
ln(Do/a2) = 8.4 ln(s-1)
d ln(Do/a2) = 9.8 ln(s-1)
c
-10

-15
ln (D/a2)

-20

-25 Paraná Traps Fish Canyon Tuff


Andesine Oligoclase
An41.1Ab54.6Or4.3 An29.8Ab64.7Or4.5
-30
5 6 7 8 9 10 11 12 13 14 15 5 6 7 8 9 10 11 12 13 14 15
104/T (K-1) 104/T (K-1)

Fig. 17. 37Ar Arrhenius plots for ordered and disordered plagioclase feldspars, calculated using equations for spherical geometry. (a) Klamath
Mountains labradorite, (b) extraterrestrial (from achondrite GRA 06128) albite/oligoclase, (c) Paraná Traps andesine, and (d) Fish Canyon
Tuff oligoclase. Uncertainties in D/a2 values are generally smaller than the symbols, but are not shown. The downward curvature observed at
high-temperature is associated with the structural transition from triclinic toward monoclinic symmetry. Because monoclinic symmetry is not
achieved below the melting temperature, high-temperature linear Arrhenius arrays observed in more Na-rich feldspars (e.g., Fig. 16) are not
observed in more Ca-rich feldspars.

Temperature (oC) Temperature (oC)


1400 1000 800 700 600 500 400 1400 1000 800 700 600 500 400
-5
Ea = 220.4 kJ/mole Ea = 219.4 kJ/mole
a ln(Do/a2) = 8.4 ln(s-1)
b` ln(Do/a2) = 9.0 ln(s-1)

-10

-15
ln (D/a2)

-20

-25 Fish Canyon Gulf of Salerno


Sanidine Sanidine
An1Ab27Or72 An3.7Ab30.5Or65.8
-30
5 6 7 8 9 10 11 12 13 14 15 5 6 7 8 9 10 11 12 13 14 15
4 -1 4 -1
10 /T (K ) 10 /T (K )

Fig. 18. 39Ar Arrhenius plots for K-rich feldspars, calculated using equations for spherical geometry. (a) Fish Canyon Tuff sanidine and (b)
Gulf of Salerno sanidine. Uncertainties in D/a2 values are generally smaller than the symbols, but are not shown. These Arrhenius arrays,
along with those for Benson Mines orthoclase (Fig. 3), are the most linear observed in this study. Unlike other ternary feldspars, sanidine and
orthoclase do not undergo significant changes in unit cell angles and cation coordination when heated (Fig. 15e and f). The Arrhenius plots
obtained from these K-rich feldspars mirror the simplicity of their thermal expansion.
W.S. Cassata, P.R. Renne / Geochimica et Cosmochimica Acta 112 (2013) 251–287 273

At 300 °C, the mosaic of “left” and “right” domains in 600 °C, the temperature at which significant deviations
P1-anorthite become sufficiently small in size such that from linearity arise on Arrhenius plots. The rate of thermal
body-center (I1) average symmetry is attained at the P1– expansion appears to generally increase above 600 °C in
T transition (Fig. 1; Smith and Ribbe, 1969; Brown, 1984; Ca-rich specimens (Fig. 19a–c), while it appears to generally
van Tendeloo et al., 1989). Again, T-anorthite is somewhat slow in Na-rich specimens (Fig. 19d–f). We are not aware of
analogous to orthoclase in the sense that an average prop- detailed thermal expansion data at higher temperatures
erty is inferred (I1 symmetry) from sub-micron domains (>900 °C), but Ca-rich feldspars approach monoclinic sym-
having another property (P1 symmetry). This transition is metry at elevated temperatures (Phillips et al., 1997). Pre-
reversible upon cooling at a sufficiently slow rate, but can dicted triclinic–monoclinic transition temperatures are in
be quenched in rapidly cooled samples, giving rise to diffuse excess of Ca-rich feldspar melting points, but decrease with
(as opposed to sharp) c- or e-reflections at room tempera- increasing K and Na content or the incorporation of large
ture (Mueller et al., 1972; Wenk, 1978). At even higher tem- cations like Sr, Ba, and Pb (e.g., Benna et al., 2000; Benna
peratures yet (>800 °C), the framework becomes less and Bruno, 2001). We assume that the pronounced down-
puckered about the small Ca cations, only two non-equiva- ward curvature observed on Ar Arrhenius plots above
lent M-sites exist (Megaw et al., 1962; Brown, 1984), c- 800–900 °C reflects a slowing or cessation of expansion, pri-
reflections disappear in anorthite (Foit and Peacor, 1973; marily in the b–c plane of the crystal, along with a normal-
Wenk et al., 1973), e-reflections disappear in intermediate ization of interstitial M–O bond lengths, as in the approach
plagioclase (Foit and Peacor, 1967; Bown and Gay, toward monoclinic symmetry in Na-rich feldspars, which
1969), and true body-center (I1) symmetry is attained. results in significant changes to the potential energy barriers
As noted above, Arrhenius arrays obtained from Ca- to Ar diffusion. Interestingly, Foit and Peacor (1967) ob-
rich feldspars are considerably more complex than those served an increasing diffuseness of a- and b-reflections in
discussed previously, exhibiting upward curvature at inter- X-ray diffraction patterns between 600 and 1000 °C in
mediate temperatures (600–800 °C; Fig. 10). Because this Ca-rich feldspar, which Smith and Ribbe (1969) suggest re-
behavior only appears only to occur in specimens belonging flects a change in the vibrational mode of the framework.
to the I1, T, P1, e1, and e2 structural groups, it may reflect Bloss (1964) observed an inflection in a curve of optical
anisotropic changes in cation location or bond lengths and extinction angle vs. temperature at 800 °C. Kohler and Wei-
angles associated with the non-equivalent M-site environ- den (1954) observed an endothermic differential thermal
ments unique to these feldspars. Such a non-linear response analysis peak at 800 °C.
of the lattice to heating is not entirely unexpected given the Thus we can draw an analogy between Ca-rich and Na-
dissimilar bonding environments of the M-site pairs (dis- rich feldspars. In both feldspars, the structure is distorted
cussed above). If this behavior is typical of Ca-rich feld- about the small cations at low-temperature regardless of
spars, then the potential energy barriers to diffusion structural state. As Na-rich feldspars are heated, expansion
should change non-linearly with temperature, giving rise first proceeds approximately linearly (Fig. 15a), giving rise
to curvature on Arrhenius plots. to linear Arrhenius arrays (Fig. 16). Subsequently, the nor-
We are only aware of four studies on thermal expansion malization of M–O distances and changes in bond angles
in Ca-rich feldspars (Stewart et al., 1966; Grundy and are manifest as a downward curvature on Arrhenius plots
Brown, 1974; Tribaudino et al., 2010; Hovis et al., 2010), (Fig. 16). Likewise, as Ca-rich feldspars are heated and
two of which contain sufficiently detailed data to warrant expansion proceeds, a downward curvature is observed at
further consideration. Tribaudino et al. (2010) presented high-temperature (Fig. 10). However, in the case of Ca-rich
expansion data between 25 and 700 °C for plagioclases of feldspars, changes in the potential energy barriers to Ar dif-
compositions An100 and An60. Expansion in the An100 pla- fusion appear to be more complex, giving rise to upward
gioclase is approximately linear between 350 and 700 °C, a curvature at intermediate temperatures (Fig. 10). Given
temperature range over which linearity is observed on that upward curvature is only manifest on Arrhenius plots
Arrhenius plots for feldspars of similar composition obtained from feldspars that yield e- and c-reflections, such
(Fig. 10). A significant change in volume associated with behavior may reflect variable thermal expansion about the
the P1–T structural transition occurs at 225 °C, which non-equivalent M-sites pairs that have different coordina-
likely precludes extrapolating Ar Arrhenius relationships tion environments and, presumably, different vibrational
defined by higher-temperature data down to geologically responses to heating. Detailed structural measurements of
relevant conditions (discussed in more detail below). The M–O and O–O bond lengths and angles in Ca-rich feldspars
rate of expansion in the An60 plagioclase increases at tem- at high temperature would be constructive in better delin-
peratures above 500–600 °C, a temperature range over eating the relationship between structural changes and up-
which upward deviations from linearity commence on ward curvature observed on Ar Arrhenius plots.
Arrhenius plots (Fig. 10). Hovis et al. (2010) present expan-
sion data between 22 and 925 °C for several Ca-rich feld- 4.3.2.4. Exsolved alkali feldspars. We now return to Arrhe-
spar compositions. The data are significantly less detailed nius plots obtained from a “gem-quality”, cryptoperthitic
than that of Tribaudino et al. (2010), but generally define alkali feldspar from a pegmatite in Madagascar. This spec-
approximately linear to curvilinear arrays. For the purposes imen comprises micron-sized orthoclase lamellae with
of comparing thermal expansion and Ar diffusion, in “tweed” microtexture and sub-micron albite lamellae (Par-
Fig. 19 we show linear regressions of the thermal expansion sons and Lee, 2005). Arrhenius plots obtained from this
data of Hovis et al. (2010) acquired above and below sample exhibit a downward curvature at 800 °C
274 W.S. Cassata, P.R. Renne / Geochimica et Cosmochimica Acta 112 (2013) 251–287

1.0
Andesine a b c

Cell Dimension Expansion (%)


0.8 Oligoclase

0.6

0.4

0.2
a-axis b-axis c-axis
0.0
1.0
Andesine d e f
Cell Dimension Expansion (%)

0.8 Labradorite
Bytownite
Anorthite
0.6

0.4

0.2
a-axis b-axis c-axis
0.0
0 200 400 600 800 1000 0 200 400 600 800 1000 0 200 400 600 800 1000
Temperature (oC) Temperature (oC) Temperature (oC)

Fig. 19. Linear regressions of plagioclase thermal expansion data acquired at temperatures above and below 600 °C by Hovis et al. (2010).
Above 600 °C, thermal expansion coefficients decrease in Na-rich feldspars (a–c) and increase in Ca-rich feldspars (d–f).

(Fig. 3d), similar to Na-rich feldspars (Fig. 16), although Deviations from linearity manifest on the Arrhenius plots
dampened in magnitude. Such curvature presumably re- are substantially greater than those observed from the more
flects the triclinic to monoclinic symmetry transition in al- pristine cryptoperthite from Madagascar (Figs. 3d and 9b)
bite lamellae, as well as resulting changes in lamellar or anorthoclase from Easy Chair Crater (Fig. 16a). It is not
interface strain. Orthoclase specimens with high concentra- clear how much of the observed deviation from linearity is
tions of Ca and Na and/or high T-site order exhibit frame- due to a range in sub-grain domain size, and how much is
work distortions similar to Na-rich feldspars at room due to structural changes occurring above 600 °C. When fit-
temperature, with irregular M-sites comprising seven short- ting such an Arrhenius plot with a multiple diffusion do-
er and two longer M–O bond lengths (Brown, 1984). Thus main model, one must consider inaccuracies that may
it appears that a minor modal abundance of albite lamellae arise due to these structural changes before the apparent
(e.g., 5%), coupled with high T-site order, cause Madagas- diffusion parameters can be extrapolated down-temperature
car cryptoperthite to diffuse more similarly to Na-rich to geologically relevant conditions.
feldspars like ECC anorthoclase (Fig. 16a) than monoclinic
K-rich feldspars like Fish Canyon sanidine (Fig. 18a) or 4.3.3. Reversibility
Benson Mines orthoclase (Fig. 3c). We presumably ob- A key prediction of our hypothesis relating “at-temper-
tained a linear Arrhenius plot for Ne diffusion in Madagas- ature” structural changes to deviations from linearity ob-
car cryptoperthite (Fig. 9a) because 22Ne was completely served in Ar Arrhenius plots is that such curvature is
degassed between 200 and 600 °C, a range over which no reversible, insofar as the structural changes are reversible.
significant structural changes occur within albite lamellae. Simple angular changes of the framework associated with
In the above example we highlighted results from a thermal expansion only require a fraction of a second to oc-
“gem-quality”, exsolved alkali feldspar. Often alkali feld- cur (i.e., the time required for the passage of a thermal
spars analyzed for thermochronometry are more complex, wave), but structural transitions associated with changes
with deuteric alteration or fractures reducing the effective in M–O coordination have an activation energy related to
domain size to sub-grain dimensions. Under these circum- the breaking and reforming of bonds (Smith and Ribbe,
stances, curvature on Ar Arrhenius plots should result from 1969). Thus while simple thermal expansion is reversible,
both a range in sub-grain domain sizes and structural changes in unit cell dimensions and angles associated with
changes that occur in Na-rich lamellae. For example, structural changes are not purely second order (Brown
Fig. 6 depicts the results of diffusion experiments conducted et al., 1984; Salje et al., 1985). For example, Barth (1965)
on fragments of deuterically altered, exsolved alkali feld- and Stewart and von Limbach (1967) showed that the tri-
spar from the Bushveld Complex, South Africa. The crys- clinic–monoclinic transition in Na-rich feldspars is not
tals contain strain-controlled, coherent to semicoherent immediately reversible upon cooling. Furthermore, disor-
film perthite crosscut by incoherent (microporous) patch dered Na-rich feldspars attain monoclinic symmetry at
and vein perthite resulting from deuteric alteration. much lower temperatures (by 200 °C) than more ordered
W.S. Cassata, P.R. Renne / Geochimica et Cosmochimica Acta 112 (2013) 251–287 275

feldspars of identical composition (Stewart and von Lim- Ar Arrhenius plots obtained from feldspars subjected to
bach, 1967; Grundy and Brown, 1974; Kroll and Mueller, cycled (i.e., prograde and retrograde) heating schedules are
1980; Hovis, 1980; Brown et al., 1984), and the critical tem- consistent with the abovementioned description of struc-
perature changes even if minor T-site ordering is induced by tural changes that occur at temperature. Retrograde heat-
heating (Barth, 1965; Stewart and von Limbach, 1967). ing following temperatures at which downward curvature
Thus changes in the feldspar structure are themselves kinet- is manifest on Arrhenius plots obtained from Na-rich feld-
ically controlled, with an associated potential energy barrier spars yields ln(D/a2) values that approach those observed in
to be overcome when heated, and exhibiting hysteresis prior prograde heating cycles at lower-temperature (i.e.,
when cooled. downward curvature is reversible, but with some hysteresis;

Temperature (oC)
o
Temperature ( C)
1400 1000 800 700 600 500 400 1400 1000 800 700 600 500 400
-5
a Easy Chair Crater b Bushveld Complex
Anorthoclase Perthite
-10 An10.8Ab71.8Or17.4 An0.8Ab19.0Or80.2

-15
ln (D/a2)

-20 t (hour) t (hour)


0 4 8 12 16 20 0 2 4 6 8 10
1200 1200

900 900
-25
600 600
T(oC) T(oC)
-30
-5
c Madagascar Plush, Oregon
Cryptoperthite d Labradorite
-10 An0.2Ab5.6Or94.2 An63.9Ab35.4Or0.7

-15
ln (D/a2)

-20 t (hour) t (hour)


0 2 4 6 8 10 0 2 4 6 8 10
1300 1300

1000 1000
-25
700 700
o o
T( C) T( C)
-30
-5
Bushveld Complex f Surtsey, Iceland
e Labradorite Labradorite
-10 An65-70Ab30-35Or1-2 An59.5Ab39.6Or0.9

-15
ln (D/a2)

-20 t (hour) t (hour)


0 2 4 6 8 10 0 4 8 12 16 20
1300 1300

1000 1000
-25
700 700
o o
T( C) T( C)
-30
5 6 7 8 9 10 11 12 13 14 15 5 6 7 8 9 10 11 12 13 14 15
4 -1
10 /T (K ) 104/T (K-1)

Fig. 20. Ar Arrhenius plots for (a) Easy Chair Crater anorthoclase (39Ar), (b) Bushveld Complex perthite (39Ar), (c) Madagascar
cryptoperthite (39Ar), (d) Plush labradorite (37Ar), (e) Bushveld Complex labradorite (37Ar), and (f) Surtsey labradorite (37Ar) crystals
subjected to cycled (prograde and retrograde) heating schedules, calculated using equations for (a, f) infinite sheet and (b–e) spherical
geometries. Uncertainties in D/a2 values are generally smaller than the symbols, but are not shown. The laboratory heating schedules are
shown in the insets, and the colors correspond to different heating schedule cycles. These plots illustrate the reversibility of curvature observed
on feldspar Arrhenius plots.
276 W.S. Cassata, P.R. Renne / Geochimica et Cosmochimica Acta 112 (2013) 251–287

Fig. 20a–c). Likewise, retrograde heating following temper- extended linear and planar defects. The potential energy
atures at which upward curvature is manifest on Arrhenius barrier associated with a given interstitial lattice configura-
plots obtained from Ca-rich feldspars yields ln(D/a2) values tion is strongly dependent on the distance between an Ar
that approach those observed in prior prograde heating cy- atom and the lattice atoms, and therefore is highly sensitive
cles at lower-temperature (i.e., upward curvature is revers- to subtle changes in lattice dimension and geometry that oc-
ible, but with some hysteresis; Fig. 20d–f). In a prograde cur when feldspars are heated. Segments of linearity ob-
heating cycle, reducing the number, duration, and temper- served on Ar Arrhenius plots appear to coincide with
ature spacing of extractions appears to delay departures temperature intervals over which the structure is stable, or
from linearity associated with the triclinic to monoclinic the unit cell angles, dimensions, and volume change in a lin-
transition by 50 °C or more (Figs. 3a and 20a), which is ear to curvilinear manner with temperature. Deviations
consistent with the attendant structural changes having from linearity appear to coincide with structural changes in-
associated potential energy barriers. Conversely, such a duced by heating (i.e., they reflect changes in the diffusive
heating schedule does not appear to delay upward curva- medium). Thus Ar diffusion in feldspars cannot solely be de-
ture on Ca-rich feldspar Arrhenius plots (Fig. 20f), which scribed by simple, temperature-independent potential en-
is consistent with non-linearity due to simple changes in ergy barriers and associated vibrational frequencies.
thermal expansion regime. Prograde and retrograde heating Instead, the potential energy barriers to diffusion and asso-
within low-temperature portions of Arrhenius plots not ciated vibrational frequencies vary with temperature, in
associated with structural changes yield reproducible values some instances giving rise to linear Arrhenius arrays from
of ln(D/a2) at any given temperature (Figs. 18, 20a, and c; which apparent activation energies and frequency factors
Cassata et al., 2009). Thus Ar diffusion experiments are can be inferred, and in other instances yielding curved or
generally consistent with simple thermal expansion occur- otherwise non-linear Arrhenius arrays. Critical to 40Ar/39Ar
ring at low-temperature and non-second-order phase tran- and (U–Th)/He thermochronometry is an understanding of
sitions occurring at higher-temperature. For comparison, whether or not diffusivities at geologically relevant tempera-
in Fig. 21 we show Arrhenius plots obtained from plagio- tures differ significantly from down-temperature extrapola-
clase crystals from the Bushveld Complex, South Africa, tions of linear Arrhenius relationships established during
that contain quartz and alkali feldspar sub-grains, and thus higher temperature laboratory heating experiments.
a range in diffusive lengthscales. By inspection it is clear
that retrograde heating does not yield ln(D/a2) values that 5. IMPLICATIONS FOR THERMOCHRONOMETRY
approach those observed in prior prograde arrays, but AND NOBLE GAS DIFFUSION KINETICS
rather yields sub-parallel linear arrays consistent with the
MDD theory (i.e., consistent with the exhaustion of Ar 5.1. Geologic relevance of laboratory-derived diffusion
within smaller domains). parameters

4.3.4. Summary Because crystallochemical changes (e.g., thermal


Ar diffusion in feldspars proceeds primarily through me- expansion and structural transitions) alter the potential en-
tal–oxygen (M–O) and oxygen–oxygen (O–O) interstitial ergy barriers to diffusion, care must be taken in extrapolat-
lattice configurations, as well as through vacancies and ing laboratory-derived diffusion down-temperature to

o
Temperature ( C) Temperature (oC)
1400 1000 800 700 600 500 400 1400 1000 800 700 600 500 400
-5
Bushveld Complex Bushveld Complex
a Andesine b Andesine
An38.9Ab59.4Or1.7 with magnetite
-10

-15
ln (D/a2)

-20 t (hour) t (hour)


0 4 8 12 16 20 0 6 12 18 24 30
1300

1000 1000
-25
700 700
o o
T( C) T( C)
-30
5 6 7 8 9 10 11 12 13 14 15 5 6 7 8 9 10 11 12 13 14 15
4 -1 4 -1
10 /T (K ) 10 /T (K )

Fig. 21. 37Ar Arrhenius plots for (a) Bushveld Complex andesine and (b) Bushveld Complex andesine with magnetite inclusions subjected to
cycled (prograde and retrograde) heating schedules, calculated using equations for spherical geometry. Uncertainties in D/a2 values are
generally smaller than the symbols, but are not shown. The laboratory heating schedules are shown in the insets, and the colors correspond to
different heating schedule cycles. These samples contain quartz and alkali feldspar sub-grains, and illustrate that low-temperature curvature
observed on Arrhenius plots obtained from some C1 plagioclase feldspars that comprise multiple diffusion domains is not reversible.
W.S. Cassata, P.R. Renne / Geochimica et Cosmochimica Acta 112 (2013) 251–287 277

geologically relevant conditions. The problems that arise difference method (the Crank–Nicholson scheme) to model
are twofold. First, if changes in structure or thermal changes in 40Ar concentration as a function of time and
expansion regime occur below the temperatures at which temperature according to the cooling path denoted by the
laboratory diffusion experiments are conducted, then labo- dashed black line in Fig. 22c. The mass diffusion equation
ratory-derived diffusion coefficients are not applicable to with a production term was solved implicitly with boundary
lower-temperature thermal histories. Second, if Arrhenius conditions of zero concentration at the grain edge and zero
plots obtained from samples that comprise multiple diffu- flux at the center node for each of three spherical domains
sion domains exhibit curvature due to structural transi- [domain distribution parameters (Ea, ln(Do/a2), and percent
tions, then corrections must be applied to the apparent of the total K contained therein) are listed in black and red
domain size distribution to account for any such geologi- font in the inset to Fig. 22b]. A uniform distribution of 39Ar
cally irrelevant non-linearity. It is therefore important to was imparted across each domain to simulate 39Ar produc-
understand the temperatures at which a given composition tion by neutron irradiation prior to laboratory heating, and
40
of feldspar exhibits deviations from linearity on Arrhenius Ar and 39Ar were then incrementally degassed3 to yield
plots due to structural transitions, as well as the magnitude the model age spectrum [see Huber et al. (2011) for more
of such deviations. In this section we highlight relevant details regarding the code]. Under the assumption that devi-
examples of these problems and explore the extent to which ations from linearity would arise on an Arrhenius plot ob-
they bias forward modeled thermal histories. An expanded tained from the laboratory degassing of an exsolved
discussion of the implications of the relationship between feldspar at temperature above 600 °C due to both the
interatomic potentials and Arrhenius non-linearity for exhaustion of sub-grain domains and the monoclinic to tri-
40
Ar/39Ar thermochronometry using micas and amphiboles clinic transition in the albite lamellae, we applied an ad hoc
and (U–Th)/He thermochronometry using apatite, titanite, adjustment to the laboratory release of Ar such that the in-
and zircon is given in the Supplementary Files. ferred domain ln(Do/a2) values for the two larger domain
sizes (listed in blue font in the inset to Fig. 22b) are three
5.1.1. Ar diffusion in alkali feldspar natural log units lower than the true domain ln(Do/a2) val-
Arrhenius arrays obtained from alkali feldspars with ues (listed in red font in the inset to Fig. 22b), giving rise to
highly sodic compositions (e.g., albite and anorthoclase) the Arrhenius array shown in gray. We then forward mod-
or lamellae (i.e., perthite) begin to deviate from linearity be- eled 1000 monotonic cooling histories using the geologically
tween 600 and 800 °C due to structural changes associated irrelevant laboratory diffusion parameters. The resulting
with the triclinic to monoclinic transition. Likewise, if feld- model age spectra were compared to the target age spec-
spars comprise a range in sub-grain domain sizes, deviations trum (Fig. 22a) and a fit statistic was calculated for each.
from linearity typically begin to arise between 400 and Those t–T paths that yielded age spectra that best fit the
700 °C due to the exhaustion of smaller domains (e.g., Love- target spectrum are shown in red in Fig. 22c, the average
ra et al., 1997). It is generally not possible to completely de- of which is denoted by a solid black line. The thermal his-
gas domains with dimensions greater than a few microns at tory predicted using the laboratory-derived domain distri-
temperature below which structural transitions occur be- bution parameters is erroneously hot. At a given time the
cause of the unfeasibly long heating durations required. predicted temperature exceeds the “true” temperature (the
Thus deviations from linearity associated with both the black dashed line) by up 40 °C (Fig. 22c). Alternatively,
exhaustion of sub-grain domains and the triclinic to mono- the time at which the sample is at a given temperature is
clinic structural transition will inevitably occur during labo- erroneously late (young) by up to 15 Ma. Thus considerable
ratory heating experiments. As such, extracting meaningful inaccuracies may exist in published thermal histories in-
information regarding domain size distributions requires a ferred from MDD models fit to Arrhenius arrays obtained
method for apportioning non-linearity between these con- from exsolved alkali feldspars. Moreover, it may not be
current phenomena. Fortunately, no complex structural possible extract geologically relevant domain distribution
transitions or thermal expansion regime changes are known parameters from alkali feldspars comprising a range in
to occur in alkali feldspars at low-temperature (<400 °C), sub-grain domain sizes.
and therefore laboratory-derived Ar diffusion parameters A further concern in assessing the accuracy of MDD
can be extrapolated down-temperature with reasonable con- models is establishing the temporal relationship between
fidence, provided that the abovementioned corrections can the formation of alteration surfaces, fractures, and other
be made to apparent domain distribution parameters in- domain-defining features and the thermal history of interest
ferred from high-temperature extractions. (i.e., constraining the stability of the domain over geologic
To assess inaccuracies in modeled thermal histories cal- time). Alkali feldspars extracted from undeformed granites
culated using apparent domain distribution parameters in- whose alteration and crystallization features are character-
ferred from Arrhenius plots without consideration for ized only by comagmatic microtextures (e.g., deuteric
curvature that arises due to structural transitions, consider turbidity, strain-controlled exsolution, symplectically-
a hypothetical multiple-domain alkali feldspar composed of crystallized myrmekite, etc.) represent suitable candidates
albite and orthoclase lamellae. Assume that it was intruded for thermochronometry (e.g., McLaren and Reddy, 2008).
at 20 km depth 100 Ma ago and was slowly uplifted to the
surface. In Fig. 22a and b we show a hypothetical 40Ar/39Ar 3
The following heating schedule was used: 595 s at 300, 300, 400,
age spectrum and Arrhenius plot, respectively, for this
400, 500, 500, 600, 600, 700, 700, 800, 800, 900, 900, 1000, 1000,
sample. The age spectrum was calculated using a finite
and 1150 °C.
278 W.S. Cassata, P.R. Renne / Geochimica et Cosmochimica Acta 112 (2013) 251–287

Temperature (oC)
1400 1000 800 600 500 400
90 -5
a
b
80
-10

70
-15

ln (D/a2)
Age (Ma)

60
-20 Ea = 167.4 kJ/mole
50 ln(Do/a2) = 10 ln(s-1) w/ 15%
ln(Do/a2) = 9 ln(s-1) w/ 35%
-25 ln(Do/a2) = 8 ln(s-1) w/ 50%
40
ln(Do/a2) = 6 ln(s-1) w/ 35%
ln(Do/a2) = 5 ln(s-1) w/ 50%
30 -30
0.0 0.2 0.4 0.6 0.8 1.0 5 6 7 8 9 10 11 12 13 14 15
Cumulative Fraction 39Ar 104/T (K-1)

350
c
300

250
Temperature (oC)

200

150

100

50

0
0 20 40 60 80 100
Time (Ma)

Fig. 22. (a) Hypothetical 40Ar/39Ar age spectrum obtained from a multiple-domain alkali feldspar comprising albite and orthoclase lamellae,
intruded at 20 km depth 100 Ma ago and slowly uplifted to the surface. (b) Hypothetical Ar Arrhenius array (gray) obtained from this sample
(see text for calculation details), shown alongside the inferred domain Arrhenius relationships [black and blue lines, with domain ln(Do/a2)
values and percent of the total K contained therein listed in the lower, left corner] and true domain Arrhenius relationships (black and red
lines, with Ar domain distribution parameters listed in the lower-left corner). Differences exist between the true (blue) and inferred (red)
domain Arrhenius relationships because deviations from linearity arise during laboratory heating experiments due to both the exhaustion sub-
grain domains and structural transitions that occur upon heating. (c) Simulated thermal histories using the inferred (erroneous) domain
distribution parameters. Thermal histories that best fit the target age spectrum are shown in red (orange = moderate fit; blue = poor fit). The
inferred thermal history denoted by the black line (an average of the best fit solutions) is anomalously hot relative to the “true” solution
(shown as a dashed black line). (For interpretation of the references to color in this figure legend, the reader is referred to the web version of
this article.)

The presence of stress-induced fractures, complex grain linearity, further complications exist in relating labora-
boundary orientations (e.g., McLaren and Reddy, 2008), tory-derived diffusion kinetics to geologically relevant tem-
clay minerals, and/or dissolution-precipitation features peratures and timescales. In our experience, approximately
associated with post-magmatic tectonic and hydrothermal one out of every five carefully selected fragments of a pris-
activity render a given sample questionably reliable for tine and apparently single-domain feldspar sample yielded
thermochronometry. Likewise, pervasive low or high tem- anomalous results consistent with multiple diffusion do-
perature excess argon, intermediate age maxima (i.e., dou- mains, presumably due to the presence of microfractures
ble-humped age spectra), and/or a lack of correlation potentially associated with sample preparation and han-
between age and kinetic (i.e., ln(r/ro)) data render question- dling that were not identified. Ar diffusion parameters
able reliability (Lovera et al., 2002). Petrographic and inferred from furnace heating experiments on tens of milli-
microanalytical observations are necessary to determine grams of these pristine feldspars would therefore reflect the
the suitability of a given sample for MDD thermal model- release of Ar from both geologically relevant domains, as
ing (Parsons et al., 1999). well as laboratory-induced domains. Thus laser or furnace
Beyond understanding the suitability of a given alkali heating of individual crystals or crystal fragments, as
feldspar for thermochronometry and the corrections that employed in this study (and more generally by the
must be applied to apparent domain distribution parame- (U–Th)/He community), appears necessary, albeit unfeasi-
ters to account for structurally-related deviations from ble in some instances, to avoid biasing Arrhenius arrays.
W.S. Cassata, P.R. Renne / Geochimica et Cosmochimica Acta 112 (2013) 251–287 279

Additionally, because Ar diffusion in feldspars is dependent instances, be prohibitively difficult. One potential approach
on composition and the spatial distribution of the diffusant, to placing constraints on thermal histories using these feld-
grain-scale diffusivity may be spatially variable within ex- spars is to disregard correcting Arrhenius arrays for devia-
solved feldspars. The effects that exsolution lamellae of dif- tions from linearity arising due to structural changes. By
ferent shapes and sizes, spatial distributions, and diffusive fitting high-temperature Arrhenius arrays with apparent,
and compositional contrasts relative to the host have on albeit erroneous, domain distribution parameters, one can
apparent diffusion parameters and inferred thermal histo- at least place upper boundaries on temperatures at a given
ries has not been quantitatively assessed, despite the routine time.
analysis of perthitic alkali feldspars for 40Ar/39Ar One further complication unique to plagioclase feldspars
thermochronometry (e.g., Heizler and Harrison, 1998; with compositions of An>95 is a significant change in ther-
McLaren et al., 2007; Harrison et al., 2010; Parsons et al., mal expansion regime and unit cell volume associated with
2010). the low-temperature (300 °C) P1–T symmetry transition
(Fig. 1). Unfortunately, it is not currently feasible to deter-
5.1.2. Ar diffusion in plagioclase mine single crystal Ar diffusion kinetics at temperatures
Plagioclase with compositions of An<95 do not undergo <300 °C to quantify the effects of the P1–T transition due
structural transitions or thermal expansion regime changes to the prohibitively long heating or irradiation durations re-
at temperatures associated with the Ar partial retention quired to generate measurable signals. Because cooling be-
zone (100–400 °C; Tribaudino et al., 2010). Therefore, lab- low the transition temperature is accompanied by a
oratory-derived Ar diffusion parameters based on linear
Arrhenius arrays can be extrapolated down-temperature
with reasonable confidence, provided that Ar can be shown 400
to diffuse at the grain-scale in the sample of interest. Plagio- a
clase feldspars comprising a range in sub-grain domain sizes 350
pose unique problems for thermal modeling. It is generally Activation Energy (kJ/mole)
not possible to completely degas domains with dimensions 300
greater than tens of microns at temperature below which
either (1) upward curvature associated with the intermedi- 250
ate-temperature (600–700 °C) structural transition occurs
in Ca-rich feldspars (Fig. 10), or (2) downward curvature 200
associated with the high-temperature (>800–900 °C) struc-
tural transition toward monoclinic symmetry occurs in all
150
plagioclase feldspars. As with alkali feldspars, corrections
must be applied to apparent domain distribution parame-
100
ters extracted from multi-domain Arrhenius arrays to ac- 400
count for deviations from linearity that arise due to these b
structural changes. 375
In the case of Ca-rich feldspars, the manifestation on
Closure Temperature (oC)

350
Arrhenius plots of the 600–700 °C structural transition (up-
ward curvature; Fig. 10) is opposite that of a range in sub- 325
grain domain sizes (downward curvature; e.g., Fig. 21).
300 Sanidine
Thus apparent domain size distributions inferred from this Orthoclase
(600–700 °C) interval in Arrhenius plots underestimate the 275 Cryptoperthite
true variations in domain size. The Arrhenius plots are fur- Perthite
Disordered C1
ther complicated (or initially complicated in the case of Na- 250
Ordered C1
rich plagioclase feldspars) by a downward curvature at e-plagioclase
225 I1-plagioclase
high-temperature (>800–900 °C) due to the structural tran- Anorthite
sition toward monoclinic symmetry. Because monoclinic 200
symmetry is never achieved, the downward Arrhenius cur- 1.0 0.8 0.6 0.4 0.2 0.0 0.2 0.4 0.6 0.8 1.0
Or* Ab* An*
vature persists through the melting temperature (i.e., linear-
ity is never re-established; Fig. 17), rendering it difficult to Fig. 23. Summary of (a) apparent activation energies and (b)
deconvolve curvature due to a range in sub-grain domain closure temperatures as a function of feldspar composition (data
sizes from that due to the structural changes. Knowledge given in Table 4). Only results from samples in which Ar is inferred
of the temperature at which a pristine, single-domain crys- to diffuse at or near the grain-scale are shown. Closure tempera-
tal of a given feldspar exhibits downward curvature and the tures were calculated assuming spherical geometry and a 10 °C/Ma
magnitude of the deviation from linearity as function of cooling rate. Apparent ln(Do/a2) values obtained from Arrhenius
plots that were calculated using equations for infinite sheet
temperature are required to accurately correct apparent do-
geometry (i.e., from cleavage flakes) were corrected by 2.2 natural
main distribution parameters inferred from multiple-do-
log units following Huber et al. (2011), so that the inferred closure
main Arrhenius plots. Extracting geologically relevant temperatures are comparable to those obtained from equant grains
domain distribution parameters from plagioclase feldspars (see Table 4). Ternary compositions (An, Ab, Or; Table 4) are
comprising a range in sub-grain domain sizes may, in many represented on this figure as Ab*, where Ab* = An  Or.
Table 4

280
Summary of feldspar Ar diffusion parameters.
Sample Locality Phase Symmetry Composition Geom. Ea ± 1r ln(Do/a2) ± 1r Tc ± 1r
(kJ/mole) (ln(s1))
An Ab Or (°C)
BMk-1 Benson Mines Orthoclase C2m 0.0 3.0 97.0 IS 223.5 ± 3.3 11.1 ± 0.4 334.0 ± 4.0
BMk-2 Benson Mines Orthoclase C2m 0.0 3.0 97.0 IS 216.0 ± 3.4 9.5 ± 0.4 335.1 ± 3.9
BMk-3 Benson Mines Orthoclase C2m 0.0 3.0 97.0 IS 213.5 ± 3.3 8.4 ± 0.4 343.4 ± 3.6
BMk-4 Benson Mines Orthoclase C2m 0.0 3.0 97.0 IS 229.9 ± 3.0 11.1 ± 0.3 310.6 ± 4.4
BMk-5 Benson Mines Orthoclase C2m 0.0 3.0 97.0 S 219.7 ± 3.2 9.0 ± 0.4 322.7 ± 3.3

W.S. Cassata, P.R. Renne / Geochimica et Cosmochimica Acta 112 (2013) 251–287
XTALk-1 Ecstall Pluton Orthoclase C2m/C1 0.1 11.3 88.6 S 238.4 ± 8.3 12.2 ± 1.0 330.6 ± 8.4
MADk-1 Itrongay, Madagascar Cyptoperthite C2m/C1 0.2 5.6 94.6 IS 270.0 ± 7.9 17.9 ± 1.0 363.9 ± 6.1
MADk-2 Itrongay, Madagascar Cyptoperthite C2m/C1 0.2 5.6 94.6 IS 264.1 ± 7.5 16.6 ± 1.0 365.7 ± 5.3
MADk-3 Itrongay, Madagascar Cyptoperthite C2m/C1 0.2 5.6 94.6 IS 277.5 ± 9.1 20.1 ± 1.2 355.1 ± 6.9
MADk-4 Itrongay, Madagascar Cyptoperthite C2m/C1 0.2 5.6 94.6 IS 275.0 ± 6.2 19.2 ± 0.8 359.9 ± 5.1
MADk-5 Itrongay, Madagascar Cyptoperthite C2m/C1 0.2 5.6 94.6 IS 278.2 ± 7.9 19.1 ± 1.0 368.4 ± 5.6
MADk-6 Itrongay, Madagascar Cyptoperthite C2m/C1 0.2 5.6 94.6 IS 266.7 ± 7.9 17.7 ± 1.0 333.3 ± 6.7
MADk-7 Itrongay, Madagascar Cyptoperthite C2m/C1 0.2 5.6 94.6 S 273.4 ± 7.9 17.4 ± 1.0 351.7 ± 6.3
BV-8k-1 Bushveld Complex Perthite C2m/C1 0.8 22.6 76.6 S 232.6 ± 12.0 10.9 ± 1.7 332.1 ± 8.9
BV-8k-2 Bushveld Complex Perthite C2m/C1 0.8 22.6 76.6 S 247.6 ± 12.9 12.6 ± 1.8 348.4 ± 8.6
BV-8k-3 Bushveld Complex Perthite C2m/C1 0.8 22.6 76.6 S 236.0 ± 16.2 11.5 ± 2.2 333.1 ± 12.9
FCs-1 Fish Canyon Tuff Sanidine C2m 1.0 27.0 72.0 S 224.3 ± 4.1 9.4 ± 0.5 329.7 ± 4.3
FCs-2 Fish Canyon Tuff Sanidine C2m 1.0 27.0 72.0 S 220.4 ± 4.6 8.4 ± 0.5 332.3 ± 5.7
GSs-1 Gulf of Salerno Sanidine C2m 3.7 30.5 65.8 S 219.4 ± 2.5 9.0 ± 0.3 321.8 ± 3.2
ECCa-1 Easy Chair Crater Anorthoclase C1 10.8 71.8 17.4 IS 278.3 ± 19.5 19.9 ± 2.8 359.3 ± 10.7
ECCa-2 Easy Chair Crater Anorthoclase C1 10.8 71.8 17.4 IS 281.7 ± 20.0 19.0 ± 2.9 377.4 ± 9.6
ECCa-3 Easy Chair Crater Anorthoclase C1 10.8 71.8 17.4 IS 300.0 ± 21.2 22.2 ± 3.1 380.1 ± 9.9
ECCa-4 Easy Chair Crater Anorthoclase C1 10.8 71.8 17.4 S 280.0 ± 30.8 17.6 ± 4.4 364.1 ± 15.9
GRAp-1 ET (GRA 06128) Albite/Olig. C1 14.1 84.1 1.9 S 209.4 ± 5.8 8.0 ± 0.6 307.7 ± 8.3
ML-15p-1 Mono Lake Oligoclase C1 19.6 74.2 6.2 S 273.3 ± 12.9 14.8 ± 1.3 382.7 ± 13.3
ML-15p-2 Mono Lake Oligoclase C1 19.6 74.2 6.2 S 281.7 ± 13.1 16.3 ± 1.2 383.4 ± 14.6
XTALp-1 Ecstall Pluton Oligoclase C1 20.8 77.7 1.5 S 201.1 ± 4.1 7.5 ± 0.4 391.4 ± 6.6
FCp-1 Fish Canyon Tuff Olig./Andesine C1 29.8 61.7 4.5 S 218.5 ± 21.2 9.8 ± 2.7 309.4 ± 20.6
PR-92p-1 Paraná-Etendeka LIP Andesine C1 41.1 54.6 4.3 S 206.9 ± 16.6 8.4 ± 2.5 396.0 ± 13.0
NCp-1 Nain Complex Labradorite e1/e2 49.4 47.9 2.7 S 171.2 ± 12.0 2.0 ± 1.8 276.6 ± 11.4
NCp-2 Nain Complex Labradorite e1/e2 49.4 47.9 2.7 S 174.5 ± 7.9 4.6 ± 1.2 352.3 ± 7.7
HMS-2p-1 Klamath Mtns. Labradorite C1 52.8 44.2 3.0 S 226.6 ± 6.2 11.5 ± 0.8 309.4 ± 5.6
SURTp-1 Surtsey, Iceland Labradorite I1 59.6 39.6 0.8 IS 182.2 ± 7.1 3.7 ± 0.9 318.6 ± 8.4
SURTp-2 Surtsey, Iceland Labradorite I1 59.6 39.6 0.8 IS 170.4 ± 5.8 2.6 ± 0.8 297.3 ± 6.7
SURTp-3 Surtsey, Iceland Labradorite I1 59.6 39.6 0.8 IS 173.5 ± 5.8 3.6 ± 0.8 292.4 ± 6.6
SURTp-4 Surtsey, Iceland Labradorite I1 59.6 39.6 0.8 IS 199.0 ± 6.7 6.9 ± 0.9 322.8 ± 6.2
SURTp-5 Surtsey, Iceland Labradorite I1 59.6 39.6 0.8 IS 164.3 ± 5.8 2.8 ± 0.8 274.4 ± 7.6
OREGp-1 Plush, Oregon Labradorite I1 63.9 35.4 0.7 IS 220.2 ± 17.9 11.7 ± 2.6 317.5 ± 13.5
OREGp-2 Plush, Oregon Labradorite I1 63.9 35.4 0.7 IS 241.8 ± 26.6 13.6 ± 4.0 349.0 ± 15.0
OREGp-3 Plush, Oregon Labradorite I1 63.9 35.4 0.7 IS 216.9 ± 24.1 10.0 ± 3.6 330.6 ± 16.4
OREGp-4 Plush, Oregon Labradorite I1 63.9 35.4 0.7 IS 201.1 ± 10.0 6.5 ± 1.4 304.3 ± 9.7
BV-8p-1 Bushveld Complex Labradorite e1/P1 67.4 31.3 1.3 S 183.6 ± 9.5 3.3 ± 1.4 296.8 ± 35.2
W.S. Cassata, P.R. Renne / Geochimica et Cosmochimica Acta 112 (2013) 251–287 281

significant reduction (contraction) in unit cell volume (Tri-


309.4 ± 19.6
261.7 ± 10.3
287.5 ± 18.5

284.8 ± 12.2

304.6 ± 18.4
290.1 ± 3.1

277.5 ± 5.5
273.2 ± 7.8
baudino et al., 2010), one might suspect that diffusivity at
low-temperature is slower than that inferred from a
down-temperature extrapolation of a linear, higher-temper-
ature Arrhenius array. Therefore, using apparent diffusion
parameters inferred from laboratory heating experiments
to model thermal histories may yield erroneously cold re-
3.3 ± 2.5
0.6 ± 1.2
2.7 ± 2.6
3.6 ± 2.7
7 ± 2.0
16.6 ± 0.7
16.9 ± 0.9
5.4 ± 2.9

sults. Measurements of He diffusion in anorthite would be


productive in assessing potential deviations from linearity
that might arise due to the P1–T structural transition.

5.2. Summary of Ar diffusion in feldspars and an atomistic


Cleavage flake ln(Do/a2) values were corrected by 2.2 natural log units (Huber et al., 2011) for Tc calculations such that IS and S results are comparable.
187.8 ± 18.7

177.8 ± 18.3
182.8 ± 11.3
196.4 ± 13.7

196.1 ± 20.4

description
160.4 ± 8.3

236.8 ± 5.8
236.2 ± 7.5

In this section we summarize systematic differences in


feldspar Ar diffusion parameters (Ea and Do) and closure
temperatures (Tc) as a function of composition and struc-
tural state. We attempt to explain such variations in terms
of Lennard–Jones potentials, bond lengths and angles, ther-
mal expansion, and ionic porosity. Finally we discuss the
S
S
S
S
S
S
S
S

relationship between thermal and barometric expansion/


compression and the resulting implications for determining
the pressure-dependence of diffusivity.
1.3
1.3
0.4
0.4
0.4
0.1
0.1
0.3

5.2.1. Systematic differences in Ar diffusion parameters


Apparent activation energies (Ea) and closure tempera-
tures (Tc; Dodson, 1973) for Ar as a function of feldspar
IS is infinite sheet geometry (used for cleavage flakes) and S is spherical geometry (used for equant grains).
31.3
31.3
24.9
24.9
24.9
3.7
3.7
3.5

composition are shown in Fig. 23. Data included in this fig-


ure were acquired by regressing low-temperature, linear
Arrhenius arrays obtained from pristine, unaltered, and
Albite, oligoclase, andesine, labradorite, bytownite, and anorthite results are based on 37Ar data.

presumably single-domain grains. As discussed in Sec-


67.4
67.4
74.7
74.7
74.7
96.2
96.2
96.2

Sanidine, cryptoperthite, perthite, orthoclase, and anorthoclase results are based on 39Ar data.

tion 5.1, we believe these data can be extrapolated to geo-


logically relevant conditions with reasonable confidence.
Tc’s would be approximately 5 °C lower if calculated using equations for IS geometry.

A complete list of apparent diffusion parameters and sam-


ple localities, compositions, and structural states is given
in Table 4. To summarize, disordered Na-, K-, and Ca-rich
e1/P1
e1/P1
e1/P1
e1/P1
e1/P1

feldspars yield the highest, intermediate, and lowest appar-


P1
T
T

Tc calculated assuming spherical geometry and a 10 °C/Ma cooling rate.

ent activation energies and closure temperatures, respec-


tively, with Ea values between 170 and 285 kJ/mole and
Tc values between 200 and 400 °C (calculated assuming a
10 °C/Ma cooling rate). These trends are mirrored by differ-
Labradorite
Labradorite

ences in volumetric thermal expansion; Na-, K-, and Ca-


Bytownite
Bytownite
Bytownite
Anorthite
Anorthite
Anorthite

rich feldspars undergo changes in unit cell volume of


approximately 0.031, 0.018, and 0.015%/°C, respectively
Symmetry groups and phase names based on Fig. 1.

(Hovis et al., 1999; Tribaudino et al., 2010). A positive cor-


relation between thermal expansion and apparent activa-
tion energy is predicted by molecular dynamics
considerations (see Fig. 13 and related discussions) if the
Bushveld Complex
Bushveld Complex

room temperature potential energy barriers to diffusion


Duluth Complex
Duluth Complex
Duluth Complex

are comparable in the phases under consideration (i.e.,


Miyake, Japan
Miyake, Japan
Lunar (76535)

assuming that one phase does not have an intrinsically dif-


ferent potential energy barrier to diffusion). There also
seems to be a tendency for ordered plagioclase feldspars
to have lower apparent activation energies than disordered
feldspars of equivalent composition. We have not analyzed
a sufficient number of ordered alkali feldspars to make such
a comparison with disordered alkali feldspars. For experi-
TROCp-1

ments in which we were able to obtain comparably precise


BV-8p-2
BV-8p-3

JAPp-1
JAPp-2
DCp-1
DCp-2
DCp-3

37
Ar and 39Ar data, we did not observe systematic mass-
dependent variations in Ar diffusivity (Fig. S1).
282 W.S. Cassata, P.R. Renne / Geochimica et Cosmochimica Acta 112 (2013) 251–287

25 o o o
Tc = 250 C Tc = 300 C Tc = 350 C
Labradorite ðAn63:9 Ab35:4 Or0:7 Þ :
Ave. 1σ Error
  ð224:012:5ÞkJ=mole
ln D½m2 =s ¼ ð8:1  1:9Þeð RT Þ
; ð5Þ
20
Tc = 400 oC
Anorthoclase ðAn10:8 Ab71:8 Or17:4 Þ :
15   ð286:011:5ÞkJ=mole
ln(Do /a2) (ln(s-1))

ln D½m2 =s ¼ ð0:8  1:7Þeð RT Þ


; ð6Þ

10 Orthoclase ðAn0:0 Ab3:0 Or97:0 Þ :


Or
  ð217:71:9ÞkJ=mole
ln D½m2 =s ¼ ð10:0  0:2Þeð RT Þ
; ð7Þ
5
Cryptoperthite ðAn0:2 Ab5:6 Or94:6 Þ :
  ð272:83:4ÞkJ=mole
0 ln D½m2 =s ¼ ð0:8  0:4Þeð RT Þ
; ð8Þ
Ab An
-5
150 175 200 225 250 275 300 325 5.2.2. Atomistic description of diffusion and potential
Activation Energy (kJ/mole) predictive power
Several models based on ionic porosity (the fraction of
Fig. 24. Summary of kinetic parameters for Ar diffusion in Ca-rich
(red), Na-rich (green), K-rich (blue), and perthitic (blue–green
unit cell volume that is not occupied by ions) have been
gradient) feldspars. Lines correspond to closure temperatures proposed to explain relative differences in elemental diffu-
between 250 and 400 °C, calculated assuming spherical geometry sivity between various mineral phases (e.g., Dowty, 1980;
and a 10 °C/Ma cooling rate. Apparent ln(Do/a2) values obtained Fortier and Giletti, 1989; Dahl, 1994; Zhao and Zheng,
from Arrhenius plots that were calculated assuming infinite sheet 2007). In general, these models contend that diffusivity at
geometry (i.e., from cleavage flakes) were corrected by 2.2 natural a given temperature is inversely proportional to the total io-
log units following Huber et al. (2011) such that the inferred nic porosity (ZT) or anionic porosity (ZA), both of which
closure temperatures are comparable to those obtained from vary with temperature. If we assume that diffusivity is pro-
equant grains (see Table 4). The results of Lovera et al. (1997) and portional to total ionic porosity, then Ar should diffuse
Forster and Lister (2010), also corrected according to Huber et al.
more slowly in sanidine (ZT = 55.6) than in anorthite
(2011), are shown as gray circles. (For interpretation of the
references to color in this figure legend, the reader is referred to the
(ZT = 59.6) or albite (ZT = 59.7), wherein it should diffuse
web version of this article.) at an approximately equivalent rate. Such a relationship
is not observed, as Na- and Ca-rich feldspars have the high-
est and lowest Ar closure temperatures, respectively, with
In Fig. 24 we summarize our Ar diffusion parameters K-rich feldspars having intermediate closure temperatures
along with the results of two detailed studies on Ar diffu- (Figs. 23 and 24). Likewise, Fortier and Giletti (1989) ob-
sion in alkali feldspars (Lovera et al., 1997; McLaren served the slowest, intermediate, and fastest oxygen diffu-
et al., 2007). We observe an overlapping but more restricted sion in albite, K-feldspar, and anorthite, respectively, at
range in alkali feldspar closure temperatures, albeit from a both 500 and 700 °C. The total ionic porosity model is pre-
much smaller number of samples. However, it should be sumably inconsistent with these observations because inter-
noted that the data of Lovera et al. (1997) were obtained atomic potentials, bond dimensions, and bond angles are
from multi-grain samples that comprised sub-grain do- not considered. Ca- and Na-rich feldspars have similar tri-
mains with an inferred size of 10 lm corresponding to lin- clinic structures and nearly equivalent ionic porosities, yet
ear segments of Arrhenius arrays, which are one to two they yield different closure temperatures. The Ar-Na Len-
orders of magnitude smaller than the 100–1000 lm domain nard–Jones repulsive forces are 20% greater than Ar–Ca
size of the single-grain samples analyzed in our study. repulsive forces at a given distance (Fig. 12), thus explain-
Therefore, our results are not directly comparable. Because ing the lower mobility of Ar in Na-rich feldspars. Although
considerable inter-sample variability exists, we suggest that Ar–K repulsive forces at a given distance are expected to be
there is no broadly applicable set of Ar diffusion parameters greater than Ar–Ca and Ar–Na repulsive forces (e.g., Dü-
that can be utilized in thermal modeling, but rather that ren et al., 1982; Patil, 1991), the monoclinic structure of
sample specific data are required. Ar diffusion experiments K-rich feldspars comprises interstitial lattice configurations
conducted on oriented cleavage flakes (Figs. 3 and 8) that are not comparable to those of Ca- and Na-rich feld-
yielded apparent diffusion parameters (Table 2) that span spars. These structural differences appear to obscure the
the range in values shown in Fig. 24. As such, using the dif- importance of cationic interatomic potential variations.
fusion parameters of more than one of these samples whose To illustrate this point, consider Ar diffusion in the triclinic
compositions are similar to that of an unknown represents and monoclinic forms of ECC anorthoclase. The low- and
a reasonable approach to propagating uncertainties on Ar high-temperature Arrhenius arrays associated with these
diffusivities into thermal models. Ar diffusion parameters structural polymorphs are linear, sub-parallel, and offset
in these samples are constrained by the following Arrhenius vertically by approximately 2 natural log units (Fig. 16a).
relationships (Table 2 and Figs. 3 and 8): The significant differences in diffusivity highlight the impor-
tance of bond lengths and angles in determining noble gas
Labradorite ðAn59:6 Ab39:6 Or0:8 Þ :
  ð174:53:6ÞkJ=mole
mobility, as ionic porosity varies little and the composition
ln D½m2 =s ¼ ð17:1  0:5Þeð RT Þ
; ð4Þ remains unchanged across the structural transition.
W.S. Cassata, P.R. Renne / Geochimica et Cosmochimica Acta 112 (2013) 251–287 283

5.3. The equivalence of barometric and thermal expansion/ ments, can be gleaned from data obtained from higher tem-
compression perature and pressure extractions (i.e., one can exploit the
enhanced diffusivity that occurs at high-temperature by
Pressure-dependent reductions in Ar diffusivity have increasing pressure). Such an approach could be useful in
been observed in a variety of minerals, including biotite constraining the importance of geologically relevant poly-
(Harrison et al., 1985), phlogopite (Giletti and Anderson, morphic structural transitions on apparent diffusion param-
1975), and muscovite (Harrison et al., 2009). Such reduc- eters. For example, the diffusion consequences of the P1–T
tions in Ar diffusivity are attributed to the extra work re- structural transition in anorthite could be studied at typical
quired to migrate through a compressed medium (e.g. laboratory temperatures (>400 °C) if high pressures were
Harrison et al., 1985). A modified version of the Arrhenius employed.
equation is used to account for the pressure-dependence of
diffusivity (e.g., McDougall and Harrison, 1999), given by 6. CONCLUSIONS
Ea þPV
D ¼ D o eð RT Þ; ð9Þ We conducted Ar diffusion experiments on feldspars with
where V is the molar activation volume and P is pressure. diverse compositions, structural states, and microstructural
Historically (e.g., Harrison et al., 1985; Carroll and Stolper, characteristics to constrain the physical significance of the
1991), the activation volume has been inferred from isother- diffusion domain, the isotropy of diffusion, and the impor-
mal differences in diffusivity at two or more different pres- tance of composition and structural state in determining dif-
sures. The effect of increasing pressure, by implication of fusivity. We found that (1) Ar diffuses at the grain-scale in
Eq. (9), is assumed to linearly increase the potential energy unaltered, uncleaved feldspar crystals, whereas feldspars
barriers to diffusion. containing hydrothermal alteration and/or incoherent sub-
In the absence of the incorporation of enhanced concentra- grain intergrowths comprise multiple diffusion domains,
tions of gaseous species at high pressures (which could slow (2) coherent lamellar intergrowths (e.g., film perthite and
diffusion due to strong gas–gas Lennard–Jones repulsions; e-plagioclase) represent complex interfaces within diffusion
Du et al., 2008), pressure-dependent reductions in diffusivity domains, but do not appear to provide fast pathways for
can, as suggested above, be attributed to compression of the Ar diffusion, and (3) diffusive anisotropy is not resolvable
bond dimensions and angles, which in turn increases the po- within the precision limits of our methodology (approxi-
tential energy barriers to diffusion. Because such potential en- mately of factor of 2) at temperatures below 900 °C. It fol-
ergy barriers are exponentially sensitive to interatomic lows that the appropriate geometry for diffusion modeling is
distances (Fig. 12), a linear relationship between pressure one that most closely resembles the physical grain dimen-
and apparent activation volume at a given temperature is sions, or the shape of sub-grain domains.
not predicted by molecular dynamics considerations. More- Apparent Ar diffusion parameters appear intimately re-
over, the relationship between pressure and activation volume lated to composition and structural state. As such, there is
should vary with temperature, as thermal expansion alters the no broadly applicable set of Ar diffusion parameters that
unit cell dimensions and angles upon which barometric com- can be utilized in thermal modeling, and hence sample-spe-
pression acts. Thus a modified Arrhenius equation describing cific data are required. In determining sample-specific diffu-
the pressure- and temperature-dependence of diffusivity could sion parameters, laser or furnace heating of individual
be used, given by the following: crystals or crystal fragments appears necessary to avoid bias-
U ref þV ðP ;T Þ ing Arrhenius arrays toward anomalously low activation
D ¼ Do e RT ; ð10Þ energies due to the presence of microfractures associated
where Uref is the potential energy barrier to diffusion at a with sample preparation and handling (which affected
reference temperature and V(P,T) describes the pressure approximately one in five carefully selected grains in this
and temperature dependent deviations in diffusivity from study). In general, Na-, K-, and Ca-rich feldspars yield the
the reference Arrhenius relationship due to thermal or highest, intermediate, and lowest activation energies and clo-
barometric expansion/compression (i.e., due to changes in sure temperatures, respectively. Apparent closure tempera-
the interstitial configurations of lattice atoms; see Eqs. tures for 0.1–1 mm grains are between 200 and 400 °C,
(1)–(3), Fig. 13, and related discussions). assuming spherical geometry and 10 °C/Ma cooling rate,
In some crystal structures (e.g., albite; Brown et al., and activation energies vary between 170 and 285 kJ/mole.
1984), pressure and temperature are analogous variables Argon diffusion in feldspars appears to occur primarily
(i.e., increasing/decreasing pressure and decreasing/increas- through M–O and O–O interstitial lattice configurations,
ing temperature have equivalent affects on the crystal lat- vacancies, and extended linear and planar defects. Potential
tice). Thus a given unit cell configuration can be achieved energy barriers to diffusion associated with interstitial con-
at different temperatures by varying the pressure in accor- figurations of lattice atoms are exponentially sensitive to
dance with the equation of state. By implication the true, the distances between diffusing noble gas atoms and the
bulk crystal potential energy barrier to diffusion associated M–O interstitial constituents, and therefore exhibit signifi-
with a given unit cell configuration (Uref) can be inferred cant variability throughout an incremental degassing exper-
from a variable-pressure Arrhenius plot. It follows that iment due to changes in unit cell dimensions and angles that
the true potential energy barriers to diffusion in a given occur when feldspars are heated. Temperature intervals
mineral at geologically relevant temperatures, which are over which linearity is observed on Ar Arrhenius plots ap-
generally not accessible by laboratory diffusion experi- pear to coincide with intervals over which a given feldspar
284 W.S. Cassata, P.R. Renne / Geochimica et Cosmochimica Acta 112 (2013) 251–287

structure is stable and the thermal expansion rate is approx- Getty Foundation. W.S. Cassata was supported by a National Sci-
imately constant. Deviations from linearity, when not ence Foundation Graduate Research Fellowship.
attributable to multiple diffusion domains (i.e., when
observed on Arrhenius plots obtained from single-domain
APPENDIX A. SUPPLEMENTARY DATA
crystals), appear to coincide with structural transitions
brought about by heating, such as changes in unit cell an-
Supplementary data associated with this article can be
gles and cation coordination.
found, in the online version, at http://dx.doi.org/10.1016/
Knowledge of the stability of diffusion domains, both
j.gca.2013.02.030.
over geologic time and during laboratory heating experi-
ments, and an understanding of whether or not diffusivities
REFERENCES
at geologically relevant temperatures differ significantly
from down-temperature extrapolations of laboratory-
Arnaud N. O. and Kelley S. P. (1997) Argon behaviour in gem-
derived diffusion parameters are critical to obtaining quality orthoclase from Madagascar: experiments and some
accurate thermal histories. When structural or thermal consequences for 40Ar/39Ar geochronology. Geochim. Cosmo-
expansion regime changes occur below the temperatures chim. Acta 61, 3227–3255.
at which laboratory diffusion experiments are conducted Bachmann O., Dungan M. A. and Lipman P. W. (2002) The Fish
(e.g., P1-anorthite), corrections must be applied to extrap- Canyon magma body, San Juan volcanic field, Colorado:
olated, low-temperature diffusion coefficients. Likewise, rejuvenation and eruption of an upper-crustal batholith. J.
corrections must be made to apparent domain distribution Petrol. 43, 1469–1503.
parameters inferred from Arrhenius plots obtained by heat- Barth T. F. W. (1965) On the constitution of the alkali feldspars.
ing samples that comprise multiple diffusion domains and Contrib. Mineral. Petrol. 10, 14–33.
Baxter E. F. (2010) Diffusion of noble gases in minerals. Rev.
undergo structural transitions (e.g., perthite), both of which
Mineral. Geochem. 72, 509–557.
contribute to non-linearity. Routine examination of sam- Benna P. and Bruno E. (2001) Single-crystal in-situ high-temper-
ples using a petrographic microscope, electron microprobe, ature structural investigation on strontium feldspar. Am.
or SEM is generally sufficient to assess the nature of diffu- Mineral. 86, 690–696.
sion domains and the suitability of a given sample for Benna P., Tribaudino M. and Bruno E. (2000) I1-I2/c ferroelastic
thermal modeling (i.e., to establish the temporal relation- phase transition in the Ca0.2Pb0.8Al2Si2O8 feldspar as a function
ship between the formation of alteration surfaces, fractures, of temperature. Mineral. Mag. 64, 285–290.
and other domain-defining features and the thermal history Berger G. W. and York D. (1981) Geothermometry from dating
of interest). experiments. Geochim. Cosmochim. Acta 45, 795–811.
It is to be hoped that this work will stimulate further col- Bloss F. D. (1964) Optical extinction of anorthite at high
temperatures. Am. Mineral. 49, 1125–1131.
laboration between mineralogists and thermochronologists.
Boven A., Pasteels P., Kelley S. P., Punzalan L., Bingen B. and
Detailed “at-temperature” measurements of M–O and O–O Demaiffe D. (2001) 40Ar/39Ar study of plagioclases from the
bond lengths and angles, coupled with noble gas diffusion Rogaland anorthosite complex (SW Norway); an attempt to
experiments, would shed new light on the relationship be- understand argon ages in plutonic plagioclase. Chem. Geol. 176,
tween structural changes and Arrhenius non-linearity. 105–135.
Moreover, the simultaneous analysis of multiple noble Bown M. G. and Gay P. (1969) The effect of heat treatment on the
gas species released from individual grain fragments diffraction patterns of intermediate plagioclases. Z. Kristallogr.
would provide invaluable insights into the importance 129, 451–457.
of thermal expansion (i.e., changing interstitial bond Braun J. (2005) Quantitative constraints on the rate of landform
lengths and angles) in determining apparent diffusion evolution derived from low-temperature thermochronology.
Rev. Mineral. Geochem. 58, 351–374.
parameters. Such experiments may hold the key to a bet-
Brown W. L. (1984) Feldspars and Feldspathoids: Structures,
ter atomistic understanding of noble gas diffusion in Properties, and Occurrences. Springer.
minerals. Brown W. L., Openshaw R. E., McMillan P. F. and Henderson M.
B. (1984) A review of the expansion behavior of alkali feldspars:
ACKNOWLEDGEMENTS coupled variations in cell parameters and possible phase
transitions. Am. Mineral. 69, 1058–1071.
David Shuster, Tim Becker, Al Deino, and Greg Balco are Brownlee S. J. and Renne P. R. (2010) Thermal history of the
thanked for laboratory assistance, Sean Mulcahy and Kent Ross Ecstall pluton from 40Ar/39Ar geochronology and thermal
for electron microprobe assistance, Simon Kelley for providing ac- modeling. Geochim. Cosmochim. Acta 74, 4375–4391.
cess to his UV-laserprobe facilities and assistance acquiring in situ Carpenter M. A., McConnell J. D. C. and Navrotsky A. (1985)
data on the Bushveld Complex plagioclase crystals, Becky Smith Enthalpies of ordering in the plagioclase feldspar solid solution.
for orienting cleavage flakes using EBSD, and Darren Mark and Geochim. Cosmochim. Acta 49, 947–966.
Kevin Righter for generously providing samples. We are grateful Carroll M. R. and Stolper E. M. (1991) Argon solubility and
to Rudy Wenk for helpful discussions regarding feldspar structures diffusion in silica glass: implications for the solution behavior of
and for access to his collection of plagioclase samples. W. Hames, molecular gases. Geochim. Cosmochim. Acta 55, 211–225.
I. Villa, and an anonymous reviewer are thanked for thoughtful Cassata W. S. (2011) An isochron approach to 21Ne cosmic ray
and constructive reviews of the manuscript, and C. Hall is thanked exposure dating by activation with deuteron–deuteron fusion
for handling the manuscript. We acknowledge financial support neutrons. Chem. Geol. 284, 21–25.
from the U.S. National Science Foundation Petrology and Geo- Cassata W. S., Renne P. R. and Shuster D. L. (2009) Argon
chemistry Program (grant EAR-0838572) and the Ann and Gordon diffusion in plagioclase and implications for thermochronom-
W.S. Cassata, P.R. Renne / Geochimica et Cosmochimica Acta 112 (2013) 251–287 285

etry: a case study from the Bushveld Complex, South Africa. Foland K. A. (1974) Ar40 diffusion in homogenous orthoclase and
Geochim. Cosmochim. Acta 73, 6600–6612. an interpretation of Ar diffusion in K-feldspars. Geochim.
Cassata W. S., Shuster D. L., Renne P. R. and Weiss B. P. (2010a) Cosmochim. Acta 38, 151–166.
Evidence for shock heating and constraints on Martian surface Foland K. A. and Xu Y. (1990) Diffusion of 40Ar and 39Ar in
temperatures revealed by 40Ar/39Ar thermochronometry of irradiated orthoclase. Geochim. Cosmochim. Acta 54, 3147–
Martian meteorites. Geochim. Cosmochim. Acta 74, 6900–6920. 3158.
Cassata W., Singer B., Liddicoat J. and Coe R. (2010b) Reconciling Forster M. A. and Lister G. S. (2010) Argon enters the retentive
discrepant chronologies for the geomagnetic excursion in the zone: reassessment of diffusion parameters for K-feldspar in the
Mono Basin, California: insights from new 40Ar/39Ar dating South Cyclades Shear Zone, Ios, Greece. Geol. Soc. Lond. Spec.
experiments and a revised relative paleointensity correlation. Publ. 332, 17–34.
Quatern. Geochronol. 5, 533–543. Fortier S. M. and Giletti B. J. (1989) An empirical model for
Cherniak D. J. (2010) Cation diffusion in feldspars. Rev. Mineral. predicting diffusion coefficients in silicate minerals. Earth
Geochem. 72, 691–733. Planet. Sci. Lett. 245, 1481–1484.
Coombs D. S. (1954) Ferriferous orthoclase from Madagascar. Garrick-Bethell I., Weiss B. P., Shuster D. L. and Buz J. (2009)
Mineral. Mag. 30, 409–427. Early lunar magnetism. Earth Planet. Sci. Lett. 323, 356–359.
Crank J. (1975) The Mathematics of Diffusion, second ed. Claren- Giletti B. J. and Anderson T. F. (1975) Studies in diffusion: II.
don Press, Oxford. Oxygen in phlogopite mica. Earth Planet. Sci. Lett. 28, 225–
Croarkin M. C., Guthrie W. F., Burns G. W., Kaeser M. and 233.
Strouse G. F. (1993) Temperature-electromotive force reference Glicksman M. E. (2000) Diffusion in Solids: Field Theory, Solid-
functions and tables for the letter-designated thermocouple state Principles, and Applications. Wiley, New York.
types based on the ITS-90. National Institute of Standards and Gourbet L., Shuster D. L., Balco G., Cassata W. S., Renne P. R.
Technology Monograph 175, pp. 630. and Rood D. (2012) Neon diffusion kinetics in olivine, pyroxene
Dahl P. S. (1994) “Ionic porosity” as a predictor of diffusion and feldspar: retentivity of cosmogenic and nucleogenic neon.
parameters in thermochronometric minerals: evidence and Geochim. Cosmochim. Acta 86, 21–36.
tectonic implications. Mineral. Mag. 58, 205–206. Grundy H. D. and Brown W. L. (1974) A high-temperature X-ray
Dodson M. H. (1973) Closure temperature in cooling, geochrono- study of low and high plagioclase feldspars. In The Feldspars
logical, and petrological systems. Contrib. Mineral. Petrol. 40, (eds. W. S. MacKenzie and W. L. Brown). Manchester
259–274. University Press, Manchester, UK, pp. 162–173.
Dove M. T. (2001) Computer simulations of solid solutions. Solid Harlow G. E. (1982) The anorthoclase structures: the effects of
solutions in silicate and oxide systems of geological importance. temperature and composition. Am. Mineral. 67, 975–996.
EMU Notes Mineral. 3, 225–249. Harrison T. M., Celerier J., Aikman A. B., Hermann J. and Heizler
Dowty E. (1980) Crystal-chemical factors affecting the mobility of M. T. (2009) Diffusion of 40Ar in muscovite. Geochim.
ions in minerals. Am. Mineral. 65, 174–182. Cosmochim. Acta 73, 1039–1051.
Du Z., Allan N. L., Blundy J. D., Purton J. A. and Brooker R. A. Harrison T. M., Duncan I. and McDougall I. (1985) Diffusion of
40
(2008) Atomistic simulation of the mechanisms of noble gas Ar in biotite: temperature, pressure and compositional effects.
incorporation in minerals. Geochim. Cosmochim. Acta 72, 554– Geochim. Cosmochim. Acta 49, 2461–2468.
573. Harrison T. M., Heizler M. T., McKeegan K. D. and Schmitt A. K.
Düren R., Hasselbrink E. and Moritz G. (1982) On the interaction (2010) In-situ 40K–40Ca ‘double-plus’ SIMS dating resolves
of excited alkali atoms with rare gas targets in scattering Klokken feldspar 40K–40Ar paradox. Earth Planet. Sci. Lett.
processes. Z. Phys. A Atoms Nuclei 307, 1–11. 299, 426–433.
Ehlers T. A. (2005) Crustal thermal processes and the interpreta- Harrison T. M., Lovera O. M. and Matthew T. H. (1991)
40
tion of thermochronometer data. Rev. Mineral. Geochem. 58, Ar/39Ar results for alkali feldspars containing diffusion
315–350. domains with differing activation energy. Geochim. Cosmochim.
Ehlers T. A., Chaudhri T., Kumar S., Fuller C. W., Willett S. D., Acta 55, 1435–1448.
Ketcham R. A., Brandon M. T., Belton D. X., Kohn B. P. and Harrison T. M. and McDougall I. (1981) Excess 40Ar in
Gleadow A. J. W. (2005) Computational tools for low- metamorphic rocks from Broken Hill, New South Wales:
temperature thermochronometer interpretation. Rev. Mineral. implications for 40Ar/39Ar age spectra and the thermal history
Geochem. 58, 589–622. of the region. Earth Planet. Sci. Lett. 55, 123–149.
Fechtig H. and Kalbitzer S. (1966) The diffusion of argon in Harrison T. M. and Zeitler P. K. (2005) Fundamentals of noble gas
potassium-bearing solids. In Potassium–argon Dating (eds. H. thermochronometry. Rev. Mineral. Geochem. 58, 123–149.
Fechtig, S. Kalbitzer, O. Schaeffer and J. Zahringer). Springer- Heizler M. T. and Harrison T. M. (1998) The thermal history of the
Verlag, pp. 68–106. New York basement determined from 40Ar/39Ar K-feldspar
Fitzgerald J. D. and Harrison T. M. (1993) Argon diffusion studies. J. Geophys. Res. 103, 29795–29814.
domains in K-feldspar I: microstructures in MH-10. Contrib. Heizler M. T., Parsons I., Sanders R. E. and Heizler L. L. (2008) K-
Mineral. Petrol. 113, 367–380. feldspar microtexture and argon transport. In Joint Meeting of
Flude S., Lee M. R., Sherlock S. C. and Kelley S. P. (2012) Cryptic The Geological Society of America, Soil Science Society of
microtextures and geological histories of K-rich alkali feldspars America, American Society of Agronomy, Crop Science Society
revealed by charge contrast imaging. Contrib. Mineral. Petrol. of America, Gulf Coast Association of Geological Societies with
163, 983–994. the Gulf Coast Chapter of SEPM.
Foit F. F. and Peacor D. R. (1967) High temperature diffraction Heizler M. T., Parsons I., Sanders R. E., Heizler L. L. and
data on selected reflections of an andesine and anorthite. Z. Krlstrom K. E. (2007) K-Feldspar 40Ar/39Ar thermochronolo-
Kristallogr. 125, 147–156. gy of western USA precambrian lithosphere: improved under-
Foit F. F. and Peacor D. R. (1973) The anorthite crystal structure standings of systematics towards more accurate geological
at 410 and 830 °C. Am. Mineral. 58, 665–675. models. In GSA Annual Meeting.
286 W.S. Cassata, P.R. Renne / Geochimica et Cosmochimica Acta 112 (2013) 251–287

Hovis G. L. (1980) Angular relations of alkali feldspar series and McLaren S. and Dunlap W. J. (2006) Use of 40Ar/39Ar K-feldspar
the triclinic–monoclinic displacive transformation. Am. Min- thermochronology in basin thermal history reconstruction: an
eral. 65, 770–778. example from the Big Lake Suite granites, Warburton Basin,
Hovis G. L., Brennan S., Keohane M. and Crelling J. (1999) High- South Australia. Basin Res. 18, 189–203.
temperature X-ray investigation of sanidine–analbite crystalline McLaren S., Dunlap W. J. and Powell R. (2007) Understanding K-
solutions: thermal expansion, phase transitions, and volumes of feldspar 40Ar/39Ar data: reconciling models, methods and
mixing. Can. Mineral. 37, 701–709. microtextures. J. Geol. Soc. 164, 941–944.
Hovis G. L., Medford A., Conlon M., Tether A. and Romanoski McLaren S. and Reddy S. M. (2008) Automated mapping of K-
A. (2010) Principles of thermal expansion in the feldspar feldspar by electron backscatter diffraction and application to
40
system. Am. Mineral. 95, 1060–1068. Ar/39Ar dating. J. Struct. Geol. 30, 1229–1241.
Huber C., Cassata W. S. and Renne P. R. (2011) A lattice McMillan P. F. and Brown W. L. (1980) The unit-cell parameters
Boltzmann model for noble gas diffusion in solids: the of an ordered K–Rb alkali feldspar series. Am. Mineral. 65,
importance of domain shape and diffusive anisotropy and 458–464.
implications for thermochronometry. Geochim. Cosmochim. Megaw H. D. (1973) Crystal Structures: A Working Approach.
Acta 75, 2170–2186. W.B. Saunders Company, London.
Karlstrom K. E., Heizler M. and Quigley M. C. (2010) Structure Megaw H. D., Kempster C. J. E. and Radoslovich E. W. (1962)
and 40Ar/39Ar K-feldspar thermal history of the Gold Butte The structure of anorthite, CaAl2Si2O8: II. Description and
block: reevaluation of the tilted crustal section model. Geol. discussion. Acta Crystallogr. 15, 1017–1035.
Soc. Am. Spec. Pap. 463, 331–352. Mueller W. F., Wenk H. R. and Thomas G. (1972) Structural
Kohler A. and Weiden P. (1954) Vorlaufige versuche in der variations in anorthites. Contrib. Mineral. Petrol. 34, 304–314.
feldspat-gruppe mittels der DTA. Neues Jb. Mineral. Monat., Nord G. L., Heuer A. H. and Lally J. S. (1974) Transmission
249–252. electron microscopy of substructures in Stillwater bytownites.
Kroll H. and Bambauer H. U. (1981) Diffusive and displacive In The Feldspars (eds. W. S. MacKenzie and J. Zussman).
transformation in plagioclase and ternary feldspar series. Am. Manchester University Press, Manchester, pp. 522–535.
Mineral. 66, 763–769. Onstott T. C., Hall C. M. and York D. (1989) 40Ar/39Ar
Kroll H. and Mueller W. F. (1980) X-ray and electron-optical thermochronometry of the Imataca Complex, Venezuela.
investigation of synthetic high-temperature plagioclases. Phys. Precambrian. Res. 42, 255–291.
Chem. Mineral. 5, 255–277. Parsons I. (2010) Feldspars defined and described: a pair of posters
Lagerwall T. (1962) Diffusion of argon-41 in calcium fluoride (rare- published by the Mineralogical Society. Sources and supporting
gas diffusion in solids 9). Nukleonik 4, 158–161. information. Mineral. Mag. 74, 529–551.
Lee J. K. W. (1995) Multipath diffusion in geochronology. Contrib. Parsons I., Brown W. L. and Smith J. V. (1999) 40Ar/39Ar
Mineral. Petrol. 120, 60–82. thermochronology using alkali feldspars: real thermal history
Lipman P. W. (2000) The central San Juan caldera cluster: regional or mathematical mirage of microtexture? Contrib. Mineral.
volcanic framework. Geol. Soc. Am. Spec. Pap. 346, 9–69. Petrol. 136, 92–110.
Lovera O. M., Grove M. and Harrison T. M. (2002) Systematic Parsons I., Fitz Gerald J. D., Lee J. K. W., Ivanic T. and Golla-
analysis of K-feldspar 40Ar/39Ar step heating results II: Schindler U. (2010) Time-temperature evolution of microtex-
relevance of laboratory argon diffusion properties to nature. tures and contained fluids in a plutonic alkali feldspar during
Geochim. Cosmochim. Acta 66, 1237–1255. heating. Contrib. Mineral. Petrol. 160, 155–180.
Lovera O. M., Grove M., Mark Harrison T. and Mahon K. I. Parsons I. and Lee M. R. (2005) Minerals are not just chemical
(1997) Systematic analysis of K-feldspar 40Ar/39Ar step heating compounds. Can. Mineral. 43, 1959–1992.
results: I. Significance of activation energy determinations. Patil S. H. (1991) Adiabatic potentials for alkali-inert gas systems
Geochim. Cosmochim. Acta 61, 3171–3192. in the ground state. J. Chem. Phys. 94, 8089–8095.
Lovera O. M., Heizler M. T. and Harrison T. M. (1993) Argon Phillips D. and Onstott T. (1988) Argon isotopic zoning in mantle
diffusion domains in K-feldspar II: kinetic properties of MH- phlogopite. Chem. Geol. 16, 542–546.
10. Contrib. Mineral. Petrol. 113, 381–393. Phillips B. L., McGuinn M. D. and Redfern S. A. T. (1997) Si–Al
Lovera O. M., Richter F. M. and Harrison T. M. (1989) The order and the II-12/c structural phase transition in synthetic
40
Ar/39Ar thermochronometry for slowly cooled samples hav- CaAl2Si2O8–SrAl2Si2O8 feldspar: a 29Si MAS-NMR spectro-
ing a distribution of diffusion domain sizes. J. Geophys. Res. 94, scopic study. Am. Mineral. 82, 1–7.
17917–17935. Prewitt C. T., Sueno S. and Papike J. J. (1976) The crystal
Lovera O. M., Richter F. M. and Harrison T. M. (1991) Diffusion structures of high albite and monalbite at high temperatures.
domains determined by 39Ar released during step heating. J. Am. Mineral. 61, 1213–1225.
Geophys. Res. 96, 2057–2069. Quidelleur X., Grove M., Lovera O. M., Harrison T. M., Yin A.
Mackert J. R., Twiggs S. W. and Williams A. L. (2000) High- and Ryerson F. J. (1997) Thermal evolution and slip history of
temperature X-ray diffraction measurement of sanidine thermal the Renbu Zedong Thrust, southeastern Tibet. J. Geophys. Res.
expansion. J. Dent. Res. 79, 1590–1595. 102, 2659–2679.
McDougall I. and Harrison T. M. (1999) Geochronology and Rainey C. S. and Wenk H.-R. (1978) Intensity differences of
Thermochronology by the 40Ar/39Ar Method, second ed. Oxford subsidiary reflections in calcic plagioclase. Am. Mineral. 63,
University Press. 124–131.
McLaren A. C. (1974) Transmission electron microscopy of Reich M., Ewing R. C., Ehlers T. A. and Becker U. (2007) Low-
the feldspars. In The Feldspars (eds. W. S. MacKenzie and W. temperature anisotropic diffusion of helium in zircon: implica-
L. Brown). Manchester University Press, UK, pp. 378– tions for zircon (U–Th)/He thermochronometry. Geochim.
423. Cosmochim. Acta 71, 3119–3130.
McLaren A. C. and Marshall D. B. (1974) Transmission electron Reiners P. W., Ehlers T. A. and Zeitler P. K. (2005) Past, present,
microscope study of the domain structures associated with the and future of thermochronology. Rev. Mineral. Geochem. 58, 1–
b-, c-, d-, e- and f-reflections in plagioclase feldspars. Contrib. 18.
Mineral. Petrol. 44, 237–249.
W.S. Cassata, P.R. Renne / Geochimica et Cosmochimica Acta 112 (2013) 251–287 287

Reiners P. W. and Shuster D. L. (2009) Thermochronology and Smith J. V. and Ribbe P. H. (1969) Atomic movements in
landscape evolution. Phys. Today 62, 31–36. plagioclase feldspars: kinetic interpretation. Contrib. Mineral.
Renne P. R., Deckart K., Ernesto M., Feraud G. and Piccirillo E. Petrol. 21, 157–202.
M. (1996) Age of the Ponta Grossa dike swarm (Brazil), and Stewart D. B. and Von Limbach D. (1967) Thermal expansion of
implications to Parana flood volcanism. Earth Planet. Sci. Lett. low and high albite. Am. Mineral. 52, 389–413.
144, 199–211. Stewart D. B., Walker G. W., Wright T. L. and Fahey J. J. (1966)
Renne P. R., Onstott T. C., D’Agrella-Filho M. S., Pacca I. G. and Physical properties of calcic labradorite from Lake County,
Teixeira W. (1990) 40Ar/39Ar dating of 1.0–1.1 Ga magnetiza- Oregon. Am. Mineral. 51, 177–197.
tions from the Sao Francisco and Kalahari cratons: tectonic Tribaudino M., Angel R. J., Camara F., Nestola F., Pasqual D.
implications for Pan-African and Brasiliano mobile belts. Earth and Margiolaki I. (2010) Thermal expansion of plagioclase
Planet. Sci. Lett. 101, 349–366. feldspars. Contrib. Mineral. Petrol. 160, 899–908.
Renne P. R. and Scott G. R. (1988) Structural chronology, Van Tendeloo G., Ghose S. and Amelinckx S. (1989) A dynamical
oroclinal deformation, and tectonic evolution of the southeast- model for the PI–II phase transition in anorthite, CaAl2Si2O8.
ern Klamath Mountains, California. Tectonics 7, 1223–1242. Phys. Chem. Mineral. 16, 311–319.
Renne P. R., Swisher C. C., Deino A. L., Karner D. B., Owens T. Villa I. M. (2006) From nanometer to megameter: isotopes, atomic-
L. and DePaolo D. J. (1998) Intercalibration of standards, scale processes, and continent-scale tectonic models. Lithos 87,
absolute ages and uncertainties in 40Ar/39Ar dating. Chem. 155–173.
Geol. 145, 117–152. Villa I. M. (2010) Disequilibrium textures versus equilibrium
Richter F. M., Lovera O. M., Mark Harrison T. and Copeland P. modeling: geochronology at the crossroads. Geol. Soc. Lond.
(1991) Tibetan tectonics from 40Ar/39Ar analysis of a single K- Spec. Publ. 332, 1–15.
feldspar sample. Earth Planet. Sci. Lett. 105, 266–278. Villa I. M. and Hanchar J. M. (2013) K-feldspar hygrochronology.
Righter K. and Carmichael I. (1993) Mega-xenocrysts in alkali Geochim. Cosmochim. Acta 101, 24–33.
olivine basalts: fragments of disrupted mantle assemblages. Am. Vineyard G. H. (1957) Frequency factors and isotope effects in
Mineral. 78, 1230–1245. solid state rate processes. J. Phys. Chem. Solids 3, 121–127.
Saadoune I., Purton J. A. and de Leeuw N. H. (2009) He Voltaggio M. (1985) Estimation of diffusion constants by obser-
incorporation and diffusion pathways in pure and defective vations of isokinetic effects: a test for radiogenic argon and
zircon ZrSiO4: a density functional theory study. Chem. Geol. strontium. Geochim. Cosmochim. Acta 49, 2117–2122.
258, 182–196. Wartho J.-A., Kelley S. P., Brooker R. A., Carroll M. R., Villa I.
Salje E., Kuscholke B., Wruck B. and Kroll H. (1985) Thermo- M. and Lee M. R. (1999) Direct measurement of Ar diffusion
dynamics of sodium feldspar II: experimental results and profiles in a gem-quality Madagascar K-feldspar using the
numerical calculations. Phys. Chem. Mineral. 12, 99–107. ultra-violet laser ablation microprobe (UVLAMP). Earth
Sanders R. E., Heizler M. T. and Goodwin L. B. (2006) 40Ar/39Ar Planet. Sci. Lett. 170, 141–153.
thermochronology constraints on the timing of Proterozoic Watanabe K., Austin N. and Stapleton M. R. (1995) Investigation
basement exhumation and fault ancestry, southern Sangre de of the air separation properties of zeolite types A, X, and Y by
Cristo Range, New Mexico. Geol. Soc. Am. Bull. 118, 1489– Monte Carlo simulations. Mol. Simul. 15, 197–221.
1506. Wenk H.-R. and Nakajima Y. (1980) Structure, formation, and
Schmeling P. (1965) Influence of radiation damage upon the decomposition of APB’s in calcic plagioclase. Phys. Chem.
diffusion of argon in potassium chloride. Physica Status Solidi B Mineral. 6, 169–186.
11, 175–184. Wenk H.-R. (1966) Labradorite from Surtsey (Iceland). Schweiz.
Shearer C. K., Burger P. V., Neal C., Sharp Z., Spivak-Birndorf L., Mineral. Petrogr. Mitt. 46, 81–84.
Borg L., Fernandes V. A., Papike J. J., Karner J. and Wadhwa Wenk H.-R. (1978) Ordering of the intermediate plagioclase
M. (2010) Non-basaltic asteroidal magmatism during the structure during heating. Am. Mineral. 63, 132–135.
earliest stages of solar system evolution: a view from Antarctic Wenk H.-R. (1979) Superstructure variation in metamorphic
achondrites Graves Nunatak 06128 and 06129. Geochim. intermediate plagioclase. Am. Mineral. 64, 71–76.
Cosmochim. Acta 74, 1172–1199. Wenk H.-R., Joswig W., Tagai T., Korekawa M. and Smith B. K.
Short C., Heizler M. T., Heizler L. L. and Parsons I. (2010) (1980) The average structure of An62–66 labradorite. Am.
Advancing the 40Ar/39Ar MDD method through micoanalysis Mineral. 65, 8I–95.
of microtexturally characterized K-feldspar crystal fragments. Wenk H.-R., Mueller W. F. and Thomas G. (1973) Antiphase
In GSA Annual Meeting. domains in lunar plagioclase. Proc. Lunar Sci. Conf. 4, 909–
Shuster D. L., Balco G., Cassata W. S., Fernandes V. A., Garrick- 923.
Bethell I. and Weiss B. P. (2010) A record of impacts preserved Winter J. K., Okamura F. P. and Ghose S. (1979) A high-
in the lunar regolith. Earth Planet. Sci. Lett. 290, 155–165. temperature structural study of high albite, monalbite, and the
Shuster D. L. and Farley K. A. (2005) 4He/3He thermochronom- analbite–monalbite phase transition. Am. Mineral. 64, 409–
etry: theory, practice, and potential complications. Rev. Min- 423.
eral. Geochem. 58, 181–203. Zhao Z.-F. and Zheng Y.-F. (2007) Diffusion compensation for
Shuster D. L. and Farley K. A. (2009) The influence of artificial argon, hydrogen, lead, and strontium in minerals: empirical
radiation damage and thermal annealing on helium diffusion relationships to crystal chemistry. Am. Mineral. 92, 289–308.
kinetics in apatite. Geochim. Cosmochim. Acta 73, 183–196.
Simoes M., Avouac J. P., Beyssac O., Goffe B., Farley K. A. and
Associate editor: Chris M. Hall
Chen Y. G. (2007) Mountain building in Taiwan: a thermoki-
nematic model. J. Geophys. Res. 112, B11405.

You might also like