14
Gmchrm,ru c, (brnvnhimrcu 4clu Vol. 55. pp. 1435-1448 Copyright Cm 1991 Pcrgamon Press pk. Printed m U.S.A. 0016.7037/91/$3.00 + .X? 40Ar/39Ar results for alkali feldspars containing diffusion domains with differing activation energy T. MARK HARRISON,’ OSCAR M. LOVERA,’ and MATTHEW T. HEIZLER~ ‘Department of Earth and Space Sciences and Institute of Geophysics and Planetary Physics, University of California, Los Angeles, Los Angeles, CA 90024, USA ‘Division of Geological and Planetary Sciences, California Institute of Technology, Pasadena, CA 91125, USA 3Department of Earth and Space Sciences, University of California, Los Angeles, Los Angeles, CA 90024. USA (Received June 29, 1990; accepted in revised form Fhwary 26, 1991) Abstract-Extension of the single diffusion domain/activation energy closure model of Dodson to apply to minerals with a discrete distribution of domain sizes appears to reconcile “0Ar/39Ar age spectra of alkali feldspars with their associated Arrhenius plots. However, small remaining discrepancies, particularly apparent in log (r/ro) plots, suggest that some, and perhaps most, alkali feldspars contain diffusion domains with activation energies that may vary by as much as 8 kcal/mol. An important consequence of even relatively small variations in activation energy between domains is that the shape of an age spectrum can change dramatically by varying the laboratory heating schedule. We find that Arrhenius and log (r/rO) plots have the potential to reveal even small differences in activation energy (-2 kcal/mol) between domains, at least in cases where the domains are well separated in size. Variations in activation energy of -5 kcal/mol can result in differences in calculated closure temperature of up to 30°C from that obtained assuming equal activation energies for all domains. Overestimates of apparent activation energy and other inconsistencies resulting from reversed heating experiments may reflect annealing of subgrain features which define the smallest diffusion domain size. Use of the diffusion compensation relationship may provide a way to assess the relative distribution of diffusion domain sizes in a sample containing domains with multiple activation energies. lNTRODUCTION Background MANY, AND PERHAPS MOST, 4oAr/39Ar age spectra for slowly cooled alkali feldspars are significantly different from model age spectra calculated assuming a single diffusion-domain size. In addition, Arrhenius plots calculated from the mea- sured loss of 3yAr during the step-heating experiment show departures from linearity that are inconsistent with diffusion from domains of equal size. LOVERA et al. (1989) extended the single diffusion-domain closure model of DODSON ( 1973) to apply to minerals with a discrete distribution of domain sizes. The discrete diffusion-domain distribution model offers an internally consistent explanation for the commonly ob- served features of alkali feldspar age spectra and their asso- ciated Arrhenius plots. Because these domains contain a range of closure temperatures, LOVERA et al. (1989) found that a single K-feldspar sample may reveal a broad segment of a cooling history, rather than the single datum usually expected. The excellent agreement among cooling history segments for coexisting samples with differing activation energies gives confidence that extrapolations of the kinetic results down- temperature are meaningful as nonlinear behavior in any one sample would displace that segment off the shared cooling history (HARRISON, 1990). Subsequently, LOVERA et al. (199 1) tested and confirmed the general predictions of the domain distribution theory by performing non-conventional 40Ar/3yAr extraction experiments involving isothermal heat- ing, short and long duration heating (50 set to 3 days), and cycled heating/cooling experiments. These studies showed that the diffusion domain distribution parameters (domain size, p, and volume fraction, 4) can be adequately obtained through fits to both the age spectrum (formed over millions of years) and the Arrhenius results (obtained in the laboratory over a matter of hours) assuming that the activation energies of all domain sizes are equal, or nearly so. Analyses by LOVERA et al. (199 1) of small (-400 pm) individual crystals indicate that the domains are an in- trinsic property of potassium feldspars that are not separable at the micron scale. Because the form of Arrhenius plots varies with laboratory heating schedule for samples with a domain size distribution, an alternative plot was proposed (RICHTER et al., 1989, 1991; LOVERA et al., 1991) in which the log of the deviation from the diffusion law for the earliest released argon (ro) is plotted against cumulative % 39Ar re- leased. The log (r/ro) plot yields domain size data independent of laboratory heating schedule, provided all domains share the same activation energy. Equations for deriving diffusion coefficients (D/r2), p, $, and log (r/ro) are given in LOVERA etal. (1989, 199l)andR1~~~~~etal. (1991). Scientific Rationale Although this new approach has been successful in recon- ciling problematic features of alkali feldspar age spectra and their associated Arrhenius plots, it seems appropriate to revisit the assumption that all domains share a common activation energy for two reasons: 1) The most surprising aspect of our recent work (LOVERA et al., 1989, 1991; LOVERA, 1989; RICHTER et al., 1989, 1991; RYERSON and HARRISON, 1990) is the suggestion that vir- tually all low temperature potassium feldspars contain sub- grain diffusion domains and that their distribution is discrete 1435

40Ar/39Ar results for alkali feldspars containing ...sims.ess.ucla.edu/argonlab/pdf/harrison_etal_GCA_1991.pdf · 40Ar/39Ar dating of alkali feldspars 1437 12+ 1 ‘* m~‘~* r 0.5

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: 40Ar/39Ar results for alkali feldspars containing ...sims.ess.ucla.edu/argonlab/pdf/harrison_etal_GCA_1991.pdf · 40Ar/39Ar dating of alkali feldspars 1437 12+ 1 ‘* m~‘~* r 0.5

Gmchrm,ru c, (brnvnhimrcu 4clu Vol. 55. pp. 1435-1448 Copyright Cm 1991 Pcrgamon Press pk. Printed m U.S.A.

0016.7037/91/$3.00 + .X?

40Ar/39Ar results for alkali feldspars containing diffusion domains with differing activation energy

T. MARK HARRISON,’ OSCAR M. LOVERA,’ and MATTHEW T. HEIZLER~

‘Department of Earth and Space Sciences and Institute of Geophysics and Planetary Physics, University of California, Los Angeles, Los Angeles, CA 90024, USA

‘Division of Geological and Planetary Sciences, California Institute of Technology, Pasadena, CA 91125, USA 3Department of Earth and Space Sciences, University of California, Los Angeles, Los Angeles, CA 90024. USA

(Received June 29, 1990; accepted in revised form Fhwary 26, 1991)

Abstract-Extension of the single diffusion domain/activation energy closure model of Dodson to apply to minerals with a discrete distribution of domain sizes appears to reconcile “0Ar/39Ar age spectra of alkali feldspars with their associated Arrhenius plots. However, small remaining discrepancies, particularly apparent in log (r/ro) plots, suggest that some, and perhaps most, alkali feldspars contain diffusion domains with activation energies that may vary by as much as 8 kcal/mol. An important consequence of even relatively small variations in activation energy between domains is that the shape of an age spectrum can change dramatically by varying the laboratory heating schedule. We find that Arrhenius and log (r/rO) plots have the potential to reveal even small differences in activation energy (-2 kcal/mol) between domains, at least in cases where the domains are well separated in size. Variations in activation energy of -5 kcal/mol can result in differences in calculated closure temperature of up to 30°C from that obtained assuming equal activation energies for all domains. Overestimates of apparent activation energy and other inconsistencies resulting from reversed heating experiments may reflect annealing of subgrain features which define the smallest diffusion domain size. Use of the diffusion compensation relationship may provide a way to assess the relative distribution of diffusion domain sizes in a sample containing domains with multiple activation energies.

lNTRODUCTION

Background

MANY, AND PERHAPS MOST, 4oAr/39Ar age spectra for slowly cooled alkali feldspars are significantly different from model age spectra calculated assuming a single diffusion-domain size. In addition, Arrhenius plots calculated from the mea-

sured loss of 3yAr during the step-heating experiment show departures from linearity that are inconsistent with diffusion from domains of equal size. LOVERA et al. (1989) extended the single diffusion-domain closure model of DODSON ( 1973) to apply to minerals with a discrete distribution of domain

sizes. The discrete diffusion-domain distribution model offers

an internally consistent explanation for the commonly ob- served features of alkali feldspar age spectra and their asso-

ciated Arrhenius plots. Because these domains contain a range of closure temperatures, LOVERA et al. (1989) found that a single K-feldspar sample may reveal a broad segment of a cooling history, rather than the single datum usually expected. The excellent agreement among cooling history segments for coexisting samples with differing activation energies gives confidence that extrapolations of the kinetic results down- temperature are meaningful as nonlinear behavior in any one sample would displace that segment off the shared cooling history (HARRISON, 1990). Subsequently, LOVERA et al. (199 1) tested and confirmed the general predictions of the domain distribution theory by performing non-conventional 40Ar/3yAr extraction experiments involving isothermal heat- ing, short and long duration heating (50 set to 3 days), and cycled heating/cooling experiments.

These studies showed that the diffusion domain distribution

parameters (domain size, p, and volume fraction, 4) can be adequately obtained through fits to both the age spectrum

(formed over millions of years) and the Arrhenius results (obtained in the laboratory over a matter of hours) assuming that the activation energies of all domain sizes are equal, or

nearly so. Analyses by LOVERA et al. (199 1) of small (-400 pm) individual crystals indicate that the domains are an in- trinsic property of potassium feldspars that are not separable at the micron scale. Because the form of Arrhenius plots varies with laboratory heating schedule for samples with a domain size distribution, an alternative plot was proposed (RICHTER et al., 1989, 1991; LOVERA et al., 1991) in which

the log of the deviation from the diffusion law for the earliest

released argon (ro) is plotted against cumulative % 39Ar re- leased. The log (r/ro) plot yields domain size data independent of laboratory heating schedule, provided all domains share

the same activation energy. Equations for deriving diffusion coefficients (D/r2), p, $, and log (r/ro) are given in LOVERA etal. (1989, 199l)andR1~~~~~etal. (1991).

Scientific Rationale

Although this new approach has been successful in recon- ciling problematic features of alkali feldspar age spectra and their associated Arrhenius plots, it seems appropriate to revisit the assumption that all domains share a common activation energy for two reasons:

1) The most surprising aspect of our recent work (LOVERA et al., 1989, 1991; LOVERA, 1989; RICHTER et al., 1989, 1991; RYERSON and HARRISON, 1990) is the suggestion that vir- tually all low temperature potassium feldspars contain sub- grain diffusion domains and that their distribution is discrete

1435

Page 2: 40Ar/39Ar results for alkali feldspars containing ...sims.ess.ucla.edu/argonlab/pdf/harrison_etal_GCA_1991.pdf · 40Ar/39Ar dating of alkali feldspars 1437 12+ 1 ‘* m~‘~* r 0.5

1436 T. M. Harrison, 0. M. Lovera, and M. T. Heizler

rather than continuous. Clearly, a deeper appreciation of this

phenomenon would follow from the determination of the

features that control the loss of argon from feldspars. When examined by optical, X-ray, and transmission electron mi- croscopy (TEM) methods, an assortment of features in low

temperature feldspars are observed, such as grain boundaries, albite and pericline twins, tweed structure, dislocation

subgrains and tangles, stepped twins, orthoclase enclaves in microcline, exsolution lamellae of varying widths, anti-phase

boundaries, coherent and noncoherent perthite boundaries, micro-pores, healed cracks, and cleavage, that could be pos- sible candidates for argon permeable boundaries (EGGELTON

and BUSECK, 1980; MACLAREN, 1978; FITZ GERALD and

MACLAREN, 1982; ZEITLER and FITZGERALD, 1986; SMITH and MCLAREN, 1983; YUND, 1983a,b,c; YUND et al., 1981; HARRISON and MCDOUGALL, 1981; PARSONS et al., 1988).

However, no one of these features recommends itself to us

above others as capable of regulating the widespread occur- rence of a discrete distribution of diffusion domain sizes.

Further complicating this assessment are possible changes to the microstructure that occur during heating (e.g., SMITH et

al., 1987). The extent to which we mistakenly ascribe domain size qualities to other effects (e.g., multiple activation energies, point defects, etc.) will influence our assessment of the re- sponsible process(es) and their length scale. To identify which, if any, of the above features control argon diffusion first re-

quires a clear understanding of the relative contributions of frequency factor and apparent size to the observed variations.

2) If all domains shared a common activation energy and

our only goal was to reconstruct geological thermal histories, then the above discussion is moot as knowledge of the dif- fusion geometry (LOVERA et al., 199 1) and domain size dis- tribution is not required for this purpose. For example, a

possible approach to inverting this problem eliminates direct assessment of the size distribution. Indeed, providing that the diffusion behavior in the laboratory is the same as that operating in nature, thermal history results are relatively in- sensitive to the formulation of the diffusion model. However, if we are incorrect in our assumption that the various domains

are characterized by the same activation energy, significant errors in calculated cooling histories could result.

MULTI-DOMAIN SAMPLES WITH VARYING ACTIVATION ENERGY

Overview

In the course of testing the hypothesis that alkali feldspars contain discrete diffusion domains, several 4QAr/39Ar age

spectrum experiments were performed to assess the behavior of Arrhenius plots derived from step-heating data under un- usual temperature-time conditions (for example, stepping from high to low temperature). The resulting age spectra contained unexpected features that led us to first consider qualitatively what effect varying activation energy among domains could have on age spectra produced by different heating schedules, and later to explicitly model this behavior.

The Arrhenius parameters which arise from translation of step-heating data through a diffusion model are activation energy (E) and a parameter (Do/r2) that includes both the

frequency factor term, Do, and a measure of the effective

domain size, r. In samples that contain more than one domain

size, r maps out information on all domains present. It is

axiomatic that if all domains in a given sample contain the same activation energy, the Arrhenius relationships for all

domains are parallel and do not cross one another. However, if there are even relatively small differences in activation en-

ergy between domains, unusual behavior can result. Even

for conventional laboratory heating schedules (i.e., mono-

tonically increasing heating steps), complex age spectra can result from such mixtures. Moreover, for unconventional schedules, the shape of the age spectrum becomes a function

of the laboratory heating.

Two Domains

Numerical simulations are useful to illustrate effects which may result from samples having diffusion domains with dif-

ferent activation energies. Consider a synthetic sample con- taining two spherical domains (note that assumptions re-

garding diffusion geometry have little influence on the overall results; LOVERA et al., 199 1) of equal age and volume fraction with contrasting activation energies and Do/p2 as shown in

Fig. 1 a. By examining the behavior that results from simulated heating of the sample above and below the point of intersec- tion of these two lines on the Arrhenius diagram (hereinafter

called the intersection point), we can illustrate one possible

effect on the age spectrum. If this model sample is heated at 900°C for eight hours, irradiated, and then step-heated in

the laboratory using a schedule similar to that shown for Ar-

Ma-4b(a) (Fig. lOc), it yields an age spectrum (curve 1 in Fig. 1 b) that is typical of our expectation of episodic loss (TURNER,

1968). However, the same hypothetical sample heated at

500°C for 20 years, although experiencing exactly the same heating intensity (and argon loss) for the 45 kcal/mol domain as the 900°C experiment (i.e., Dt/p* = 0.0124, where the diffusion coefficient, D = Do exp (-EJRT), t is heating du-

ration, p is the diffusion domain size, R is the gas constant, and T is absolute temperature), has lost considerably more argon from the 35 kcal/mol domain than in the higher tem-

perature experiment. Moreover, using the identical laboratory

heating schedule as that which was used to simulate degassing of the first sample, the resulting age spectrum (curve 2 in Fig. 1 b) is remarkably different. This convex-upwards spectrum

results from the more highly outgassed lower activation energy domain dominating the late gas release.

In fact, 40Ar/39Ar age spectra of this kind have previously been observed for mixtures of white micas. WIJBRANS and MCDOUGALL (1986) reported such an example from white micas separated from metamorphic rocks on the island of Naxos in the Greek Cyclades. Two metamorphic events are recognized in the rocks of Naxos; the first metamorphism

(Ml) occurred at relatively high pressure and low temperature, and the second was related to the formation of an elongate metamorphic dome (M2). Far from the dome, phengite as- sociated with the MI metamorphism is widespread. In the core of the dome, only muscovite which grew during M2 is found. At intermediate distances, samples were only partially recrystallized during M2 and contain both phengite and mus-

Page 3: 40Ar/39Ar results for alkali feldspars containing ...sims.ess.ucla.edu/argonlab/pdf/harrison_etal_GCA_1991.pdf · 40Ar/39Ar dating of alkali feldspars 1437 12+ 1 ‘* m~‘~* r 0.5

40Ar/39Ar dating of alkali feldspars 1437

12+ 1 ‘* m~‘~* r 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 1.4 1.5 0 10 20 30 40 50 60 70 80 90 100

1000/K Cumulative % 3gAr released

FIG. 1. (a) Arrhenius plot oftwo diffusion domains with contrasting activation energies that intersect at about 800°C. Above this temperature, the domain with the 45 kcal/mol activation energy diffuses argon most rapidly, but below this temperature, the opposite is true. Synthetic heating experiments are performed on equal volume mixtures of these two domains at (I) 900°C for 8.5 hours and (2) 500°C for 20 years. (b) Theoretical age spectra of equal mixtures of the two diffusion domains shown in (a) following heating at 900°C for 8.5 hours (curve 1) and 500°C for 20 years (curve 2) using a step heating schedule similar to that shown for Ar-Ma-4b(a) in Fig. 10~. Heating the sample above the intersection point yields an age spectrum (curve 1) similar to that expected for episodic loss, while heating below the intersection results in a convex-upwards spectrum (curve 2).

covite. The pure muscovite yielded a flat age spectrum with an age of - 12 Ma (Fig. 2a), interpreted by WIJBRANS and MCDOUGALL (1986) as the time of argon closure following Mz. Pure phengite has a monotonically rising age spectrum with ages as old as 43 Ma (Fig. Za), considered to reflect original closure from the high pressure MI metamorphism. White mica from the intermediate zone revealed convex- upward age spectra (Fig. 2b). The key to explaining this be- havior was the observation that muscovite tended to degas preferentially both early and late in the heating relative to the phengite (Fig. 2c), and thus the young ages associated with the muscovite dominated both those portions of gas release. An artificial mixture of the two pure mineral separates in the same proportions to that observed in the intermediate zone revealed a similar spectrum (Fig. 2b) to the natural mix- ture. The question remains why one phase would degas pref- erentially relative to the other only at low and high temper- atures? One explanation is that muscovite has a lower acti- vation energy for argon loss relative to phengite, and the experiment was begun below their intersection point. The early enhanced gas loss from muscovite is due to the higher intrinsic transport rate from this lower activation energy do- main at low temperatures and the later dominance of mus- covite in argon release reflects the relatively small temperature window for degassing of phengite which results from its higher activation energy. Although we are not aware of diffusion or dehydration data for phengite to compliment the muscovite data (e.g., ROBBINS, 1972; ZIMMERMAN, 1972), the contrast-

ing retentivities for these phases in nature appear consistent with differing activation energies.

Three Domains

Now consider the behavior of a slowly cooled model sample which contains three discrete domains of equal volume frac- tions with activation energies of 39.2, 49.2, and 59.2 kcal/ mol. Although in many real samples the intersection point appears to occur at (meaningless) negative temperatures, for convenience we have set this value to 560% The three very different laboratory heating schedules chosen to release the argon below, at, and above the intersection point (Fig. 3a: circles, squares, and triangles, respectively) yield remarkably different Arrhenius piots (same symbols in Fig. 3b) for the same sample. For a heating schedule in which most of the gas is evolved close to the mutual intersection point (squares, Fig. 3b), step-heating data yield an averaged apparent acti- vation energy of -50 kcal/mol and virtually no hint of the existence of the three domains. Comparing the Arrhenius plots of Fig. 3b with those lines corresponding to the single domains used to build the sample distribution, we note that the way each grain contributes to the argon release is complex and strongly dependent on the heating schedule, giving rise in some cases to artificial activation energies.

In our previous studies (RICHTER et al., 1989, I99 1; LOV-

ER4 et al., 199 1) we found that samples with a single activation energy and a distribution of domains yield log (r/r01 plots which are independent of the heating schedule. This is a con- sequence of the equal activation energy assumption, and this relationship is no longer maintained when the sample con- tains domains with different activation energies and, implic- itly, different values of I&. Figure 3c shows the different log

Page 4: 40Ar/39Ar results for alkali feldspars containing ...sims.ess.ucla.edu/argonlab/pdf/harrison_etal_GCA_1991.pdf · 40Ar/39Ar dating of alkali feldspars 1437 12+ 1 ‘* m~‘~* r 0.5

1438 T. M. Harrison, 0. M. Lovera, and M. T. Heizler

50 50

81-556 white mica

40 40

z

; 30 2 al 30

,” r

+J +

g 20 2 20 :: 81-556 muscovite ki

2 g Q

10 70

0 0 0 10 20 30 40 50 60 70 80 90 100 0 10 20 30 40 50 60 70 80 90 100

Cumulative % 39Ar released Cumulative % JgAr released

l.OJ ’

z

z: 0.25- muscovite _

400 500 600 700 800 900 1000 1100 1200

Temperature (“C)

FIG. 2, (a) ~Ar/39Ar age spectra of pure phengite (81-580) and muscovite (81-556) from metamo~hic rocks of Naxos, Greece. (b) The age spectrum of a naturally mixed white mica (8 l-556) separate from Naxos confining both phengite and muscovite is similar in most respects with an age spectrum from an artificial 75:25 mixture ofthe phengite and muscovite shown in (a). (c) Plot of cumulative argon loss with temperature for the pure phengite and muscovite analyses shown in (a). Note the enhanced degassing of muscovite at high temperature. All data are from WIJBRANS and MCLXPJGALL (I 986).

(r/r*) plots obtained from the three heating schedules shown in Fig. 3a. No indication of the existence of the three domains on the log (r/r*) plot is observed when the gas is released close to the intersection point. In contrast, the log (r/r*) plots as- sociated with the degassing entirely below or above the in- tersection point (circles and triangles in Figs. 3b and c, re- spectively) contain substantial variations. Note that the cal- culation of log (r/r,,) results from comparison of data with the line defined from the initial heating steps (i.e., ro). Any artifact, such as mixing between domains, which modifies this activation energy (as shown in Fig. 3b) will have consid- erable impact on the resulting shape of the log (r/To) plot. However, as we will show later, log (r/To) plots are still useful to assess the domain distribution and to identify the existence of differences in activation energy.

As previously mentioned, different heating schedules also affect the form of the age spectrum of a sample with different activation energies. We have calculated model age spectra,

shown in Fig. 4, using the heating schedules given in Fig. 3a and a cooling history similar to that determined for the Chain of Ponds pluton (HEIZLER et al., 1988; LOVERA et al., 1989). The heating schedule used in the numerical analysis, shown by the circles in Fig. 3a, is most similar to a conventional laboratory schedule in which most of the degassing occurs below the intersection point and yields an age spectrum (Fig. 4, circles) similar in most respects to analysis of natural sam- ples in which the age spectrum is characterized by apparent ages which, in general, progressively rise (LOVERA et al., 1989). The heating schedule confined above the intersection (tri- angles, Fig. 3a) yields a convex age spectrum (Fig. 4, triangles) similar in most respects to the earlier examples of convex age spectra resulting from episodic loss (Figs. 1 b and 2b). Ages rise from 300 Ma early in gas release cresting at about 350 Ma in the middle, and then fall to ages of about 320 Ma. In the case of argon release close to the intersection point, the sample yields an averaged age spectrum (Fig. 4, squares) with

Page 5: 40Ar/39Ar results for alkali feldspars containing ...sims.ess.ucla.edu/argonlab/pdf/harrison_etal_GCA_1991.pdf · 40Ar/39Ar dating of alkali feldspars 1437 12+ 1 ‘* m~‘~* r 0.5

40Ar/39Ar dating of alkali feldspars 1439

600

2 4

log t (minutes)

16 , . . , . , . 8.. ~‘9.’ 0.6 0.8 1.0 1.2 1.4 1.6 1.6 :

0.6

1000/K

0

+ 2.0

-0.2 b (c) Cumulative % 3gAr released

most ages falling between 330 and 350 Ma. However, we

again point out that the intersection point for natural alkali

feldspars is typically at much higher (or even negative) tem- peratures than that used for these calculations (i.e., 560”(Z), and the difference in activation energies of 20 kcal/mol is

probably extreme. The K-feldspar samples from the Chain of Ponds pluton

studied by LOVERA et al. (1989) were subsequently rerun several times using various heating schedules including re- peated monotonic heating steps and thermal cycling (LOV-

ERA, 1989; LOVERA et al., 1991). Figure 5 shows the Arrhenius results for one such analysis of MH- 10 K-feldspar, MH- lO.cf, which is believed to contain three diffusion domain sizes.

The principal characteristic of this heating schedule (shown in Fig. 10d and tabulated in LOVERA et al., 199 1) is that after

revealing the linear portions corresponding to the bulk of the

argon in a given domain, the transition to the next domain

size occurs by repeated steps at constant temperature. In this way, the apparent diffusivity can be seen to initially drop rapidly, but then slow as the next dominant domain is en-

countered (corresponding to the plateau on a log (r/rO) plot). For example, steps run subsequent to those which define the 46.1 kcal/mol activation energy of r. were maintained at 1000°C for about 10 min each. The log (r/ro) for this analysis, shown in Fig. 6a, is relatively well behaved until about 70%

“Ar release at which point temperatures exceeding the onset of incongruent melting of K-feldspar under dry conditions

(1150 f 20°C: SCHAIRER and BOWEN, 1955) were achieved. Because the sample is no longer stable, data obtained above

this temperature have no meaning on such a plot.

The fits to the MH-lO.cf Arrhenius and log (r/ro) plots calculated by LOVERA et al. (199 l), obtained assuming the

sample contains a single activation energy (Fig. 6a, dashed line; Fig. 6b, calculated (triangles) vs. actual (squares)), appear satisfactory, except for the few points in the .4rrhenius plot where the temperature is cycled to lower values (for example, at 35 and 70% 39Ar release in Fig. 6a). These points correspond to a second-order feature observed in the log (Y/Y~) plots of all samples we have cycled to lower temperature following

the isothermal transition. Although they contain less than

3% of the total argon in the sample, the increased leverage these data have on the slope of the line from the low tem-

FIG. 3. (a) Heating schedules for three numerical experiments which demonstrate the effects of laboratory heating below (circles), near (squares), and above (triangles) the mutual intersection point ofthree diffusion domains with differing activation energies. Note the heating steps involve cycling temperature and long durations. (b) Synthetic Arrhenius plots which result from equal proportions of three diffusion domains (represented by the three lines) and the heating schedules shown in (a). The various heating schedules produce extremely vari- able representations of the domain distribution. (c) Plot of log (Y/TO) for the three heating experiments shown in (a) and (b). Heating near the intersection point results in apparent homogenization of the three domains and a volume weighted average activation energy. As a result, the log (Y/I~) (shown as squares) is flat throughout degassing. Although heating above (triangles) and below (circles) the intersection point results in some structure on this plot, estimates of apparent domain sizes will be incorrect unless the variation of activation energies is known.

Page 6: 40Ar/39Ar results for alkali feldspars containing ...sims.ess.ucla.edu/argonlab/pdf/harrison_etal_GCA_1991.pdf · 40Ar/39Ar dating of alkali feldspars 1437 12+ 1 ‘* m~‘~* r 0.5

1440 T. M. Harrison, 0. M. Lovera, and M. T. Heizler

250 IO 0 20 40

Cumulative % 3eAr released

FIG. 4. Age spectra for the three synthetic experiments represented in Fig. 3 (a and b). The experiment in which heating was entirely restricted below the intersection point (circles) yields a spectrum sim- ilar in most respects to that observed using a conventional heating schedule (e.g., MH-lO.cf, Fig. 9b). However, heating near the inter- section (squares) results in a spectrum which tends towards a plateau. Heating above the intersection (triangles) produces a convex shaped spectrum similar in some respects to those shown in Figs. 1 b and 2b.

perature cycling make them sensitive monitors of changes in activation energy. A considerably better fit to all the data is

obtained using domains with slightly varying activation

2

4

G z

\6 c

5

F8 -

I

10

12 /

E

MH-1 O.cf

E = 54 kcal/m

(“C) 1200 1000 600 600 500 400

0.6 0.8 1.0 1.2 1.4 1.6

1000/K

FIG. 5. Arrhenius plot for MH-IO K-feldspar obtained using a combination of thermal cycling and repeated steps at constant tem- perature (MH-lO.cf). These results reveal the presence of different activation energies between the smallest (46 kcal/mol) and inter- mediate (54 kcal/mol) size domains. There is a suggestion of a higher activation energy for the largest domain, but the data is insufficient for a clear assessment. Solid squares are steps run above the incon- gruent melting temperature of K-feldspar.

Cumulative % 3gAr released 1.4

n

-0.2 2p ,’ 20, ’ , ” 1 “” , ’ ,sb_ ‘, ‘f

‘;; 4 1 MH-1 O.cf

z \ c 6 t

0.4 0.6 0.8 1.0 1.2 1.4 1.6

1000/K

FIG. 6. (a) Plot of log (r/rO) for the Arrhenius data of MH-lO.cf together with a model fit obtained assuming equal activation energies in all domains. Note that data obtained above 70% 39Ar release (thick lines) correspond to heating temperatures above the incongruent melting point of K-feldspar and thus contain no information regarding the natural domains. The theoretical fit assuming three diffusion do- main sizes (dashed line) is generally good with the exception of de- viations at 35% and 70% 39Ar release. (b) Calculated Arrhenius plot assuming a uniform (triangles) activation energy among diffusion domains. The agreement with the empirical results (squares) is ad- equate but deviations are apparent in the steps cycled to low tem- perature.

energies (Fig. 7a, dotted line; Fig. 7b, triangles). The synthetic data were calculated using activation energies of 46, 54, and 54 kcal/mol for the smallest, intermediate, and largest domain

sizes, respectively. Comparison of the two log (r/ra) plots (Fig. 7a, solid and dotted lines) reveals that the anomalous peak seen in the empirical data at 35% 39Ar released is reproduced in the synthetic pattern. This generally good correspondence gives us confidence in believing that, at least in cases where

the domains are well separated in size, this analysis can reveal even small (-2 kcallmol) relative differences in activation

energies among coexisting domains. The distribution of domains calculated for this sample us-

ing these different activation energies can be used to estimate

a new segment of cooling history by fits to the age spectrum and Arrhenius plot, much in the way done by LOVERA et al. (1989). Thus, we can compare this segment with our previous estimate obtained from assuming equivalent activation ener- gies for all domains. Figure 8 shows both fits to the age spec- trum (Fig. 8a) along with the different segments of calculated cooling history (Fig. 8b). We find that even relatively small differences in activation energy (-5 kcal/mol) produce dif- ferences of -30°C in the estimated cooling history, perhaps accounting for the small dispersion observed among the cooling history segments estimated from all experiments on samples from the Chain of Ponds pluton (LOVERA, 1989).

Page 7: 40Ar/39Ar results for alkali feldspars containing ...sims.ess.ucla.edu/argonlab/pdf/harrison_etal_GCA_1991.pdf · 40Ar/39Ar dating of alkali feldspars 1437 12+ 1 ‘* m~‘~* r 0.5

40Ar/39Ar dating of alkali feldspars 1441

1.0

SO.6 %I 2

0.2

Cumulative % 3gAr released

4 . . . I . I I . * I- O 20 40 60 80 100

$? 4-

c - ^d MH-lO.cf .

5 6- +4 Q

$ a- : QR 4

B ag

H lo- o e

’ D

,?- (“C) 1200 1000 600 600 500 0 (6) _

* I I ‘r )I . I 1 . I I I I I 7

-0.4 0.6 0.8 1.0 1.2 1.4 1.6

1000/K

FIG. 7. (a) Plot of log (r/r,,) for the Arrhenius data for MH-IO.cf (solid line) together with a model fit (dotted line) obtained using the activation energies shown in Fig. 5. The deviations observed in the actual results at 35% and 70% 39Ar release are preserved in the model calculations. Data obtained above 70% 39Ar release (thick lines) cor- respond to heating temperatures above the incongruent melting point of K-feldspar and thus contain no information regarding the natural domains. (b) Calculated Arrhenius plot assuming variable (triangles) activation energies (as shown in Fig. 5) among diffusion domains. The empirical data (squares) agree much better with the results cal- culated using multiple activation energies compared with the as- sumption of a constant value (i.e., Fig. 6b).

For this reason, it is important to design the laboratory heating schedule to optimize resolution of the activation energies and

distribution parameters of all domains. In summary, argon extracted from samples containing dif-

fusion domains of differing activation energies degassed en- tirely within one of three regimes (below, near, and above the intersection point on the Arrhenius plot) may yield age spectra and Arrhenius plots which differ markedly in their

content of information related to age and domain distribution parameters. In general, laboratory heating schedules can be

designed to circumvent this potential problem and reveal the

activation energies of the various domains (see Appendix).

RELATIONSHIP OF THEORY TO NATURAL SAMPLES

Reversed Heating Experiments

We performed an experiment “in reverse” on sample Ar- Ma4b by beginning the sample heating at high temperature (1000°C) for a short duration (50 set) followed by progres- sively lower temperatures (see Appendix for analytical details). The resulting age spectrum (Fig. 9a) differed substantially from that produced using a conventional heating schedule (i.e., monotonically increasing temperature steps) in that the early released gas gave older ages and the later released gas

gave younger ages. Despite the significant change in character

of the age spectrum, the integrated 40Ar/39Ar age was un-

changed (Ar-Ma-4b(a) conventional = 221.5 Ma; Ar-Ma- 4b(b) reversed = 222.0 Ma). The generality of this effect on age spectra was confirmed when other alkali feldspars, such as MH- 10 (Fig. 9b), were subsequently run in the same fash-

ion and revealed similar results. Arrhenius parameters and

heating schedules for both samples are shown in Figs. 10a through d. Although these age spectra share similarities with

the results of our modelling (Figs. 3 and 4) recall that we

selected an intersection point of 560°C for the three domains.

As mentioned earlier, in many real samples such as Ar-Ma- 4b and MH- 10 (Fig. 9), this intersection appears to occur at

very high or even negative temperatures. Although the mod- elling remains useful for making general predictions about

the effects of varying activation energy, we must appeal to an additional effect to explain why the age spectra run in

reverse fashion (i.e., from initially high to low temperature) share features in common with our model calculations.

In Fig. 10, we have compared the Arrhenius plot for MH-

lO.cf K-feldspar, run in the forward cycling fashion, to the

result for the reversed heating schedule (MH- lO.cb). The fact

that only the first step of the reversed heating schedule falls on the line defined for r. in the conventional experiment is

understood to be a consequence of the large amount of 39Ar released (-22%) during this high temperature step. However, comparison of both age spectra (Fig. 9b) present intriguing differences. Both samples yield essentially identical integrated

Cumulative % 3gAr released

400+, ’ ’ 3 3 ’ ’ * 1 *

+ c (a)

200 0 20 40 60 80 100

G & 400.

,,-’ - multiple-E,,,=’ .

e ,/ ,, ,... .‘. . $ 300. __--

,A /..” __-- -- _, ,....... ..” ..-...‘slng,e-E -

(u

E” zoo-

,,,( ;“_,d . ,.... .,........ . . . . 1’1:_ _____--- F

fbl . ,__ .

280 300 3;o 340 360 380

Age (Ma)

FIG. 8. (a) Age spectrum for MH-lO.cf together with calculated results assuming constant (dotted curve) and varying (dashed curve) activation energies. The similarity between the two results suggests that, in general, the form of age spectra is not a sensitive indicator of variations in activation energy. (b) Temperature histories resulting from the two fits to the age spectrum shown in (a). Note that the resulting histories differ by as much as 30°C underscoring the im- portance of recognizing the existence of multiple activation energies.

Page 8: 40Ar/39Ar results for alkali feldspars containing ...sims.ess.ucla.edu/argonlab/pdf/harrison_etal_GCA_1991.pdf · 40Ar/39Ar dating of alkali feldspars 1437 12+ 1 ‘* m~‘~* r 0.5

T. M. Harrison, 0. M. Lovera, and M. T. Heizler 1442

250

200

A 0

2 150

?k 400 u

u

k ;7 350

CL

-? 300

250

MH-lO.cf n

4 / I >

0 10 20 30 40 50 60 70 80 90 100

Cumulative % 3gAr released

FIG. 9. (a) Age spectrum of Ar-Ma-4b(a) K-feldspar from the study of HARRISON et al. (1989) obtained using progressively increasing temperature steps. Also shown is the age spectrum (Ar-Ma-4b(b)) obtained using a heating schedule which monotonically dropped from an initially high temperature (1000°C). In the age spectrum obtained using the unconventional temperature history, the tendency is for early ages to be older, and later ages to be younger, compared to the original age spectrum. The integrated ~~Ar/39Ar ages of both samples are 222 Ma. The high degree of outgassing in the first step is due to the initially high temperature, even though it was only maintained For about a minute. (b) Age spectra of MH-lO.cf K-feldspar obtained using a conventional heating schedule (monotonically increasing temperature steps beginning at 400°C) and MH-lO.cb using an un- conventional (temperature cycling beginning at IOOO’C) heating schedule (see LOVERA et al., I99 1 for tabulated results). As with (a), the age gradient observed in the original analysis (MH-lO.cf integrated 40Ar/39Ar age = 348.6 Ma) is obscured in the reversed run (MH- lO.cb) with most steps approaching the integrated 40Ar/39Ar age of 345.2 Ma.

40Ar/39Ar ages (OH-lO.cf = 348.6 Ma; MH-lO.cb = 345.2 Ma), which, together with the observed Arrhenius plots, strongly suggests similar domain distributions. However, the reverse experiment shows older ages early, and younger ages late, in gas release. We can successfully model (squares, Fig. 11) details of the observed Arrhenius plot (squares, Fig. lob) for the reversed heating schedule using the domain parameters (Do/p2, E, and (p) shown in Fig. 11 (lines 1 through 4). When we applied the heating schedule used to degas MH-lO.cf to this synthetic domain distribution, the result (Fig. 1 I, aster- isks) appears very similar to the actual MH- 1 O.cf results (Fig. IObf. Note that even when we have used domains with dif- ferent activation energies, the conventional heating schedule still produces a linear array at low tem~rature with an ap- parent activation energy of about 54 kcal/mol. This is a sim- ilar observation to that made earlier with reference to Fig. 3b and reflects the degassing occurring near the intersection point. Despite the apparent success of modelling details of the Arrhenius plots for the reversed and conventional heating schedules using domains with different activation energies, this success was not translated to the modelling of the cor- responding age spectra. Figure 12 shows the age spectra for both heating schedules using the cooling history for the Chain

of Ponds pluton determined by LCWERA (1989). Since the domain distribution used in Fig. 11 was chosen to resemble peculiar features observed in the Arrhenius plots and age spectrum of MH- lO.cb, it is not surprising that the age spec- trum of the reversed heating schedule (Fig. 12, solid line) shows the details of the actual sample (MH-lO.cb). However, the synthetic age spectrum obtained from the conventional heating schedule (Fig. 12, dashed line) also shows some of these features (e.g., younger ages late in gas release) which were not observed in the actual age spectrum obtained from MH-lO.cf (Fig. 9b). The large difference in shape observed between the synthetic age s~ctrum (Fig. 12, dashed line) and the age spectrum of MH-lO.cf (Fig. 9b) suggests that our relatively simple model of multiple activation energy (Figs. 3 and 4) does not completely describe the behavior of the real samples heated in the reversed fashion.

Apparently the unorthodox heating experiment has in- duced a behavior that has overriden the activation and spatial controls extant during monotonic heating. We speculate that this behavior is the result of a phase change occurring in this alkali-feldspar between -750 and 900°C (depending on timescale) which mimics the effect of a much lower intersec- tion temperature.

Direct TEM imaging of this sample indicates that the only microstructuml features not present in this sample following heating in the range of 750 to 950°C are albite exsolution lamellae, generally separated by a characteristic spacing of -0.1 pm (FIX! GERALD and HARRISON, in prep.). Results of experiments in which splits of MH-10 K-feldspar are crushed until the largest diffusion domains are no longer ob- served (LOVEKA et al., in prep.) suggest that the scale of the largest domain is approximately 50 pm.

A hypothesis consistent with these various lines of evidence is that the smallest domain size in MH-10 alkali feldspar is defined by subgrain defects. During conventional monotonic heating, these features remain intact until the smallest do- mains are virtually degassed, but disappear by about 750 to 900°C (depending on heating duration). When heating begins at high temperatures (>850”C), the degassing from the smallest domains is accompanied by the rapid disappearance of the diffusion boundaries that define these features causing disproportionately lower gas loss in each subsequent step. Anomalously old ages are observed early in the age spectra of samples heated in reverse fashion (Fig. 9a and b) because at high temperatures the higher activation energy (and larger) domains with the oldest ages have a diminished contrast in retentivity compared to the smaller, and younger domains, and become a significant fraction of the released argon. This effect is coupled with the “capture” of young argon ages into what become highly retentive domains after the features that originally defined the smallest domains are annealed. This results in the younger ages observed in the later portions of gas release.

Multiple Activation Energies

Our suggestion that different size domains of an alkali feldspar have different activation energies for argon diffusion (and that this variation is domain size dependent) appears

Page 9: 40Ar/39Ar results for alkali feldspars containing ...sims.ess.ucla.edu/argonlab/pdf/harrison_etal_GCA_1991.pdf · 40Ar/39Ar dating of alkali feldspars 1437 12+ 1 ‘* m~‘~* r 0.5

40Ar/39Ar dating of alkali feldspars 1443

Oi o Ar-Ma-4b(a) D Ar-Ma-4b(b)

P 6.

n 7-

5 8-

I Q-

10 .

11 .

13

1200 1000 800 700 600

t

c 0

0

5oo(ocja) i 12c 1 ‘eL Sf*a’ I’ I’# : 1 c

0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 1.4 1.5

1000/K

1600

1500

1400

1300

1200

1100

1000

900

800

700

600

500

400

t

(cl -

d 2;o 500 740 lOi0 lZi0 f&O 17&I 2000

Heating duration (minutes)

aMH-lO.cf n MH-I O.cb

3.

4-

5-

6 -

7-

8-

9-

IO -

11 - _^

A

i3 *

An

1200 1000 800 700 600

A

A

500 (“qo() ILid r aI"rvl' IS rC, t 1 b

0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 1.4 1.5

1000/K

1600

1500

1400

1300

G e 1200

2 1100

J 1000

2 a 900

F 800

F 700

600

500

400

f t

MH-lO.cb

MH-lO.cf

I- O 500 9000 1500 2000 2500 3000 3500 4000

Heating duration (minutes)

FIG. 10. (a) Arrhenius plot showing data for both Ar-Ma-4b(a) (forward schedule, circles) and Ar-Ma-4b(b) K- feldspar (reversed schedule, squares) analyses. (b) Arrhenius plot showing data for both MH-IO.cf and MH-lO.cb K- feldspar analyses. The reversed heating schedule is represented as squares and the fonvard experiment is shown as triangles. (c) Heating schedules for Ar-Ma-4b(a) and Ar-Ma-4b(b). (d) Heating schedules for MH-lO.cf and MH-lO.cb.

to have little precedent in theory or experiment. Our initial response was to look for an alternate mechanism or experi- mental artifact that could lead to the false appearance of multiple-E’s. These efforts are reviewed below.

One possible explanation is that the contrast in activation energy reflects a change in diffusion mechanism between an intrinsic and extrinsic regime (e.g., BUENING and BUSECK,

1973). However, the cycling to low temperature during release of what is dominantly the intermediate domain (Fig. 5) pro- duces significant overlap in temperature to that used for r. ruling out an effect of this kind.

It has been observed (HORN et al., 1975; NORGETT and LIDIARD, 1968) that very high neutron doses result in in- creased retentivity of argon in solids (i.e., lowered D/r’),

probabiy due to an increased density of defect related traps. However, we are unable to incorporate this observation into a convincing explanation of our results. Firstly, at the very much lower neutron doses typical of our experiments, we have not observed a dependence of D/r2 on J (see also FO- LAND and XV, 1990). Secondly, models based on the trapping of inert gas atoms into irradiation produced defects predict high activation energies at low temperature and low activation energies at high temperature (NORGETT and LIDIARD, 1968), the opposite of what we have observed. Lastly, if all inert gas atoms are trapped into defects created during irradiation, we should not see the correlations that exist between Arrhenius plots and age spectra (e.g., LOVERA et al., 1989).

We recognize analytical and interpretive pitfalls that could

Page 10: 40Ar/39Ar results for alkali feldspars containing ...sims.ess.ucla.edu/argonlab/pdf/harrison_etal_GCA_1991.pdf · 40Ar/39Ar dating of alkali feldspars 1437 12+ 1 ‘* m~‘~* r 0.5

1444 T. M. Harrison, 0. M. Lovera, and M. T. Heizler

2

12

14 I 3.4

(“C) 1200 1000 800 600 500’

0.6 0.6 1.0 1.2 1.4 1

1000/K

+

-r

I.6

during the laboratory heating, is it possible that the activation energy observed for r. is artificially lowered (with respect to the larger and apparently higher activation energy domains) by progressive annealing and that the activation energy given by the intermediate domain is the sole value? We tend to reject this possibility for two reasons. The first is that any proposed annealing effect (e.g., alkali interdiffision) will likely have a higher activation energy that Ar diffusion (e.g., YUND,

1983~) which would result in a rapid transition from a meta- stable to an unstable state, rather than result in a gradual drop in r’. The second is that, as already mentioned, the r. activation energy for MH-lO.cf (Fig. 5) was produced using

FIG. 1 I. Arrhenius plots for a synthetic sample having a distribution of diffusion domains with different activation energies. Solid lines show the corresponding Arrhenius relationship for each domain used in the distribution. The distribution itself was built to resemble pe- culiar features observed in the Arrhenius plot and age spectrum of sample MH-lO.cb (Fig. 9b), in which a reversed heating schedule beginning at 1000°C resulted in degassing of 22% of the 39Ar in the first step (corresponding to perhaps >90% of the gas contained in the least retentive domain). The Arrhenius plot for the synthetic sample obtained using the same reversed heating schedule used for MH-lO.cb are shown as squares and correspond well to the actual results (see Fig. lob). A forward heating schedule equivalent to that used for MH-lO.cf was also applied to this synthetic sample and the Arrhenius results, shown as asterisks, reveal a similar pattern to the empirical data (Fig. lob). Note that this Arrhenius plot does not allow a clear distinction between distributions with equal activation energies and those that contain different activation energies. Although these kind of heating schedules are apparently not sufficient to resolve small differences in activation energies, resolution can be obtained from their respective age spectra.

lead to the appearance of multiple activation energies in samples that contain no such contrast. In the case where the smallest domain is volumetrically low, a shallow apparent activation energy can result from monotonic heating due to the influence of the larger domains (e.g., HEIZLER and HAR- RISON, 199 1). A consequence of referencing this incorrect activation energy for r. to the larger domains is that a larger contrast in activation energy (or the appearance of a contrast in a sample that contains none) will result. This effect can be identified by cycling temperature while degassing ro. If r. is robust, then a single line will be defined. If not, then a zig- zag pattern on the Arrhenius plot will result. Likewise, an intermediate domain that is mapped out on an Arrhenius plot solely by dropping temperature can yield an apparently steeper slope to that actually present. This is due to the steadily increasing influence of the larger domain(s) in drawing down the value of D/r2. Although temperature cycling will also identify this effect, so will simple examination of the results on a log (r/ro) plot. We are confident that the results for MH- 1 O.cf presented in this paper are unaffected by these artifacts.

If the smallest domains are defined by defects which anneal

0 20 40 60 60 100

Cumulative %3sAr released

FIG. 12. Age spectra obtained from the synthetic sample of Fig. I 1 using both forward and reversed heating schedules (i.e., MH-lO.cf and MH-lO.cb). The reversed schedule (heavy line) yields a first step containing 22% of the cumulative % 39Ar release with an older age than those obtained in the subsequent few steps; a pattern similar to that observed in the empirical result for MH-lO.cb (Fig. 9b). In ad- dition, a decrease in age is observed over the last 15% of gas release, a feature previously noted for samples MH- I O.cb and Ar-Ma-4b(b) (Fig. 9). This decrease is a consequence of the crossing of the lines that correspond to the most retentive domains in the distribution (see Fig. 11). These two features (older ages early and younger ages later) are independent and due to the crossing lines of the most and least retentive domains. This is more clearly illustrated in the age spectrum of the same synthetic sample using the forward heating schedule (dashed line). Even when the first part of the age spectrum proceeds in the conventional way, the last part ofthe spectrum shows the same decrease in ages observed for the spectrum heated in the reversed manner (see Fig. 9b). No indication of a decrease in age in the latter part of gas release was observed in the real sample when we applied the forward heating schedule (Fig. 9b), and thus no in- dication of intersecting lines appear in this spectrum. Although we appear to be able to model unusual features such as those appearing in MH- IO.cb (Fig. 9b) using different activation energies for the var- ious domains, it is unlikely that we have revealed the correct inter- pretation as no evidence for these parameters were observed in the forward experiment, and a balance between older ages obtained in the first step and the younger ages in the last portion of gas release appears to be maintained in the integrated *Ar/39Ar ages of all the samples. These two features cannot be completely explained by a distribution with different activation energies as both effects appear to be independent. Annealing of sub-grain defects which define the smallest domain size may be responsible for this effect (see text).

Page 11: 40Ar/39Ar results for alkali feldspars containing ...sims.ess.ucla.edu/argonlab/pdf/harrison_etal_GCA_1991.pdf · 40Ar/39Ar dating of alkali feldspars 1437 12+ 1 ‘* m~‘~* r 0.5

40Ar/39Ar dating of alkali feldspars 1445

log Do (cm2/sec)

FIG. 13. Compensation plot for diffusing species in feldspars from the compilation of HART (198 1) (solid line) together with recent results of GILETTI (1990). The dashed line reflects constraints on the position and slope of the compensation law for Ar in feldspar from the ob- servation of how extraneous argon is accommodated in feldspars as a function of temperature and the maximum value of the ordinate permitted by the physical size of our analyzed aggregates.

temperature cycling. Any irreversible change in the diffusion dimension for the smallest domain would clearly plot below

the r. line when the temperature was dropped. In the case of MH- 1 O.cf, the opposite was observed as the step cycled to

lower temperature actually plots slightly above the r. line, albeit containing very little gas.

By demonstrating a link between total ionic porosity and diffusivity in silicates, FORTIER and GILETTI (1989) proposed

a systematic model from which the magnitude of activation energy for diffusion can be predicted. Although largely em-

pirical, the success of their model in providing an internally consistent framework for oxygen and argon diffusion mea-

surements suggests that the approach has merit. However, because the unit cell volumes (and thus ionic porosities) of high sanidine and maximum microcline are essentially iden-

tical, their model predicts no contrast in activation energy between even these endmember structural states.

Despite the lack of precedent for multiple activation ener- gies of a diffusing species within a single phase, our results appear to indicate the existence of just such a system. At the very least, we are currently unable to identify another expla-

nation that satisfactorially explains the behavior of the Ar- rhenius (and log r/ro) plots and the age spectra produced by reversed heating.

Domain Size and the Compensation Relationship

Use of the compensation law of HART (198 1) together with our experimental data results in overestimates of the largest domain size, in some cases exceeding the dimension of the

analyzed grains. Satisfactory agreement is obtained if the slope is retained, but the compensation curve is shifted on the or- dinate by only 2 kcal/mol (Fig. 13). The dashed line shown in Fig. 13 reflects constraints on the position and slope of the compensation law for Ar in feldspar arising from the

observation of how extraneous argon is accommodated in feldspars as a function of temperature (HARRISON and MCDOUGALL, 1981; ZEITLER, 1987; FOSTER et al., 1990)

and the maximum value of the ordinate permitted by the largest domain size of our analyzed aggregates. There is cer- tainly sufficient latitude in these data to permit our minor

modifications. It may appear that we have introduced one too many de- A simple method to account for the effect of varying fre-

grees of freedom (i.e., multiple E) to still permit assessment quency factor in our calculation of relative domain sizes is of the true relative domain sizes, as deviations from r. can to use the difference in temperature between the crossover now result from both differing diffusion domain size and fre- point and that used to assess the largest domain (1100°C in quency factor. However, a possible approach to separating the case of MH-lO.cf). If the domain assessment could be these two effects is to invoke the compensation relationship. made entirely at -450°C (the temperature at which all spe- Molecular diffusion in virtually all solids and melts appears cies in feldspar apparently diffuse at the same rate), then the to obey the Meyer-Neldel rule, a compensation between ac- only effect that would displace data from r. would be true tivation energy and logarithm of the pre-exponential factor differences in domain size. However, to complete the exper- for Arrhenius-type functions (MEYER and NELDEL, 1937). iment in a reasonable period of time requires that we go to Diffusion compensation relationships have been observed much higher temperatures. Thus, we must apply an appro- between different elements in a common host (melt or solid) priate correction to account for the change in D/r2 due to

and also result by varying pressure and water content

(WINCHELL, 1969; WATSON, 1979a,b; HART, 1981). A con-

sequence of diffusion compensation is that at one tempera-

ture, given by the slope of the line (the crossover point), all species diffuse at the same rate. Thermodynamic justifications for compensation have been proposed based on possible cor-

relations between reaction enthalpy and entropy (e.g., CRINE,

1984). However, there is also great concern that the relation-

ship could be simply an artifact arising from the limited ex-

perimental approaches available to reveal E and Do (e.g.,

NORWISZ et al., 1989a,b). The relatively abundant diffusion literature for feldspars

allowed HART (198 1) to propose a compensation law of the

form E = 50.7 (kcal/mol) + 3.4 log Do (cm*/sec), which cor- responds to a crossover temperature of -450°C (Fig. 13,

solid line). Arguments based on the siting of extraneous argon in geological samples place this value at about 400°C (e.g.,

FOSTER et al., 1990) generally confirming the results of HART (198 1). If this is the case, we have the basis to separate the

relative contribution of varying frequency factor from true

size variations within the Do/r2 parameter. If a domain is observed to have a higher activation energy than that observed

for ro, the resulting AE then translates, via a shift up the compensation line, to a different value of the intrinsic fre-

quency factor from that inferred for r. [slope = 0.3 log Do (cm2/sec)/E (kcal/mol)]. Note that since r. maps out all do- mains present in a sample, this estimate of E for the smallest domain may be influenced to higher values by the more re- tentive domains. However, because of the relatively small

variations in activation energy observed thus far and the gen- eral dominance of the smallest domain size in definition of

ro, we believe that little error results from this assumption.

Page 12: 40Ar/39Ar results for alkali feldspars containing ...sims.ess.ucla.edu/argonlab/pdf/harrison_etal_GCA_1991.pdf · 40Ar/39Ar dating of alkali feldspars 1437 12+ 1 ‘* m~‘~* r 0.5

1446 T. M. Harrison, 0. M.

the activation energy contrast. For example, using 46 and 54

kcal/mol for the smallest and largest domain sizes, respec-

tively, the relative change in D/r’ between 450 and 1100°C is a factor of 14. Multiplying this by the total contrast in domain size calculated for MH-10 of 32 (Table 3, LOVERA et al., 1989) yields a corrected size difference between the smallest and largest domains of about 450. If the -50 pm scale inferred from crushing experiments for the largest do- main size in MH-10 is characteristic (LOVERA et al., in prep.),

then our prediction is that the smallest domain size is of the

order 0.1 km.

SUMMARY AND CONCLUSIONS

The prevailing view that there is a relatively simple rela-

tionship between intracrystalline argon distribution and age spectrum (i.e., TURNER, 1968; ALBAR&DE, 1978; McDou-

GALL and HARRISON, 1988) is challenged by our finding that the form of an age spectrum will be dependent on the lab- oratory heating schedule if the sample contains domains of varying activation energy. However, no matter how unusual

the heating schedule and resulting age spectrum are, the re- covery of thermal history information remains possible if the sample does not experience changes to its domain structure during heating. The results presented here provide guidance in designing more effective 40Ar/39Ar experiments and suggest

several lines of future investigation. Our current efforts are directed at testing the defect annealing hypothesis through

correlating changes in diffusion characteristics and sample microstructure as a function of temperature (FITZ GERALD and HARRISON, in prep.; LOVERA et al., in prep.).

The principal conclusions of this study are

1)

2)

3)

4)

5)

6)

Potassium feldspars appear to contain diffusion domains with activation energies that vary by as much as -8 kcal/ mol between adjacent domain sizes. A consequence of even relatively small differences in do- main activation energies is the potential for unusual be-

havior in the age spectrum, the most extreme being that the shape of the age spectrum changes with varying lab-

oratory heating schedules. Arrhenius and log (r/ro) plots have the potential to reveal small differences in activation energy (-2 kcal/mol) be- tween domains, at least in cases where the domains are well separated in size. Variations in activation energy of -5 kcal/mol can result in differences in calculated closure temperatures of up to 30°C from that obtained assuming equal activation ener- gies for all domains.

Anomalous age spectra that result from reversed heating experiments may reflect annealing of subgrain features which define the smallest diffusion domain size.

The diffusion compensation relationship provides a basis for assessing the relative diffusion domain size distribution for samples containing multiple activation energies.

Acknowledgments-This research was principally supported by DOE grant DE-FG03-89ER14049 with other funding from NSF grants.

-overa, and M. T. Heizler

OML was supported during the course of this research by a Texaco Fellowship in Geophysics. Continuing discussions with John Fitz Gerald, Frank Richter, Peter Zeitler, David Foster, Peter Copeland, and Suzanne Baldwin have helped shape our evolving views of K- feldspar age spectra. Suzanne Baldwin, Bob Drake, and Jeremy Richards are thanked for their reviews.

Editoriul handling: D. E. Fisher

REFERENCES

ALBAR~DE F. (1978) The recovery of spatial isotope distributions from stepwise degassing data. Earth Planet. Sci. Lett. 39,387-397.

BUENING D. K. and BUSECK P. R. (1973) Fe-Mg lattice diffusion in olivine. J. Geophys. Rex 78, 6852-6862.

CRINE J. P. (1984) A thermodynamic model for the compensation law and its physical significance for polymers. J. Macromol. Sci. Phys. B23,201-219.

DODSON M. H. (1973) Closure temperature in cooling geochrono- logical and petrological systems. Contrib. Mineral. Petrol. 40,259- 274.

EGGLETON R. A. and BUSEK P. R. (1980) The orthoclase-microcline inversion: A high resolution transmission electron microscope study and strain analysis. Contrib. Mineral. Petrol. 74, 123-133.

F~TZ GERALD J. D. and MACLAREN A. C. (1982) The microstructure of microcline from granitic rocks and pegmatites. Contrib. Mineral. Petrol. 80, 2 19-229.

FOLAND K. A. and XU Y. (1990) Diffusion of “‘Ar and “Ar in ir- radiated orthoclase. Geochim. Cosmochim. Acta 54, 3 141-3 158.

FORTIER S. M. and GILETTI B. J. (1989) An empirical model for predicting diffusion coefficients in silicate minerals. Science 245, 1481-1484.

FOSTER D. A., HARRISON T. M., COPELAND P. and HEIZLER M. T. (1990) Effects of excess argon within large diffusion domains on K-feldspar age spectra. Geochim. Cosmochim. Acta 54, 1699- 1708.

GILETTI B. J. (1990) Kinetics of Mg, Ca and Sr in albite and otthoclase feldspars. Eos 71, 1664.

HARRISON T. M. (1990) Some observations on the interpretation of feldspar 40Ar/39Ar results. Isotope Geosci. 80, 2 19-229.

HARRISON T. M. and FITZ GERALD J. D. (1986) Exsolution in horn- blende and its consequences for 40Ar/39Ar age spectra and closure temperature. Geochim. Cosmochim. Acta 50,247-253.

HARRISON T. M. and MCDOUGALL I. (198 I) Excess 40Ar in meta- morphic rocks from Broken Hill, New South Wales: Implications for 40Ar/39Ar age spectra and the thermal history of the region. Earth Planet. Sci. Lett. 55, 123-149.

HARRISON T. M. and MCDOUGALL I. (1982) The thermal significance of potassium feldspar K-Ar ages inferred from 4oAr/39Ar age spec- trum results. Geochim. Cosmochim. Acta 46, 18 1 l- 1820.

HARRISON T. M., SPEAR F. S. and HEIZLER M. T. (1989) Geochro- nologic studies in central New England: II Post-Acadian hinged and differential uplift. Geology 17, 185-189.

HART S. R. (198 I) Diffusion compensation in natural silicates. Geo- chim. Cosmochim. Acta 45, 279-29 1,

HEIZLER M. T. and HARRISON T. M. (199 1) The heating duration and provenance ages of rocks in the Salton Sea Geothermal Field, southern California. J. Volcano/. Geotherm. Rex (in press).

HEIZLER M. T., Lux D. R., and DECKER E. R. (1988) The age and cooling history of the Chain of Ponds and Big Island Pond Plutons and the Spider Lake Granite, west-central Maine and Quebec. Amer. J. Sci. 288, 325-352.

HORN P., JESSBERGER E. K., KIRSTEN T. and RICHTER H. (1975) 39Ar-40Ar dating of lunar rocks: Effects of grain size and neutron irradiation. Proc. Lunar Sci. Conf: 6th, 1563- 159 1.

LOVERA 0. M. (1989) 40Ar/39Ar thermochronometry: Implications of having samples with a distribution of diffusion domain sizes. Ph.D. thesis, Univ. Chicago.

LOVERA 0. M., RICHTER F. M., and HARRISON T. M. (1989) 4”Ar/ 39Ar thermochronometry for slowly cooled samples having a dis- tribution of diffusion domain sizes. J. Geophys. Res. 94, 179 I7- 17935.

Page 13: 40Ar/39Ar results for alkali feldspars containing ...sims.ess.ucla.edu/argonlab/pdf/harrison_etal_GCA_1991.pdf · 40Ar/39Ar dating of alkali feldspars 1437 12+ 1 ‘* m~‘~* r 0.5

40Ar/39Ar dating of alkali feldspars 1447

LOVERA 0. M., RICHTER F. M., and HARRISON T. M. (1991) Dif- fusion domains determined by 39Ar released during step heating. J. Geophys. Res. 96, 2057-2069.

MACLAREN A. C. (1978) Defects and microstructures in feldspars. In Chemical Physics of Solids and Their Surfnces. Vol. 7 (eds. M. W. ROBERTS and J. M. THOMAS), pp. l-30. A Specialist Pe- riodical Report, The Chemical Society, London.

MCDOUGALL I. and HARRISON T. M. (1988) Geochronology and Thermochronology by the 40Ar/.‘9Ar Method. Oxford University Press, New York.

MEYER W. V. and NELDEL H. (1937) Uber die Beziehungen zwischen der Energiekonstanten e und der Mengenkonstanten (Y der Leit- werts-Temperaturformel bei oxydischen Halbleitern. Z. techn. Physik 18, 588-593.

NORGETT M. J. and LIDIARD A. B. (1968) The migration of inert gases in ionic crystals. Phil. Mag. 18, 1193-1210.

NORWISZ J., SMIESZEK Z., and KOLENDA Z. (1989a) Apparent linear relationship, compensation law and others. Part I. Thermochem. Acta 156, 313-320.

NORWISZ J., SMIESZEK Z., and KOLENDA Z. (1989b) Apparent linear relationship, compensation law and others. Part II. Thermochem. Acta 156, 321-345.

PARSONS I., REX D. C., GUISE P., and HALLIDAY A. N. (1988) Argon loss by alkali feldspars. Geochim. Cosmochim. Acta 52, 1097-I I 12.

RICHTER F. M., LOVERA 0. M., HARRISON T. M., and HEIZLER M. T. (1989) 39Ar release during step heating: A structure micro- probe. Eos 70, 489.

RICHTER F. M., LOVERA 0. M., HARRISON T. M., and COPELAND P. C. ( 199 1) Tibetan tectonics from 40Ar/39Ar analysis of a single K-feldspar sample.Earth Planet. Sci. Lett. (in press).

ROBBINS G. A. (1972) Radiogenic argon diffusion in muscovite under hydrothermal conditions. MS. thesis, Brown University, Provi- dence.

RYERSON F. J. and HARRISON T. M. ( 1990) Degassing of argon from microclines within the thermal aureole of the Obsidian Dome Conduit, Long Valley Caldera, California: Constraints on em- placement history. J. Geophys. Res. 95, 2781-2792.

SCHAIRER J. F. and BOWEN N. L. (1955) The system KzO-A1203- SiOt . Amer. J. Sci. 253, 68 l-746.

SMITH J. V. (1974a) Feldspar Minerals. I. Crystal Structure and Physical Properties. Springer-Verlag, Heidelberg.

SMITH K. L. and MCLAREN A. C. (1983) TEM investigation of a microcline from a nepheline syenite. Phys. Chem. Minerals 10, 69-76.

SMITH K. L., MCLAREN A. C., and O’DONNELL R. G. (1987) Optical and electron microscope investigation of temperature-dependent microstructures in anorthoclase. Canadian J. Earth Sci. 24, 528- 543.

STEIGER R. H. and JAGER E. (1977) Subcommission on geochro- nology: Convention on the use of decay constants in geo- and cosmochronology. Earth Planet. Sci. Lett. 36, 359-362.

TURNER G. (1968) The distribution of potassium and argon in chon- d&es. In Origin and Distribution of the Elements (ed. L. H. AH- RENS), pp. 387-398. Pergamon, London.

WATSON E. B. (1979a) Calcium diffusion in a simple silicate melt to 30 kbar. Geochim. Cosmochim. Acta 43, 3 13-322.

WATSON E. B. (1979b) Diffusion of cesium ions in HzO-saturated

STAUDACHER TH., JESSBERGER E. K., DORFLINGER D., and KIKO J. ( 1978) A refined ultrahigh-vacuum furnace for rare gas analysis. J. Phys. E. Sci. Instr. 11, 781-789.

granitic melt. Science 205, 1259- 1260. WIJBRANS J. R. and MCDOUGALL I. (1986) 40Ar/39Ar datina of white .

micas from an Alpine high-pressure metamorphic belt on Naxos (Greece): The resetting of the argon isotopic system. Contrib. Min- eral. Petrol. 93, 187-194.

WINCHELL P. (1969) The compensation law for diffusion in silicates. High Temp. Sci. 1, 200-215.

YUND R. A. (1983a) Subsolidus phase relations in the alkali feldspars. In Feldspar Mineralogy, 2nd edn. (ed. P. H. RIBBE); Reviews in Mineralogy 2, pp. 141-169.

YUND R. A. (1983b) Microstructure, kinetics and mechanisms of

alkali feldspar exsolution. In Feldspar Mineralogy, 2nd edn. (ed. P. H. RIBBE); Reviews in Mineralogy 2, pp. 177-202.

YUND R. A. (1983~) Diffusion in feldspars. In Feldspar Mineralogy, 2nd edn. (ed. P. H. RIBBE); Reviews in Mineralogy 2, pp. 203- 222.

YUND R. A., SMITH B. M. and TULLIS J. (198 1) Dislocation-assisted diffusion of oxygen in albite. Phys. Chem. Mineral. 7, 185- 189.

ZEITLER P. K. (I 987) Argon diffusion in partially degassed alkali feldspar: Insights from 40Ar/39Ar analysis. Chem. Geol. 65, 167- 181.

ZEITLER P. K. and FITZ GERALD J. D. (1986) Saddle-shaped 40Ar/ 39Ar age spectra from young, microstructurally complex potassium feldspars. Geochim. Cosmochim. Acta 50, 1 I85- 1199.

ZIMMERMAN J. C. (1972) Water and gas in the main silicate families: Distribution and application to geochronology and petrogenesis. Sci. Terre Mem. 2.

APPENDIX: ANALYTICAL METHODS

40Ar/39Ar isotopic measurements of alkali feldspar aggregates were obtained using automated Nuclide 4.5-60-RSS and VG 1200s mass spectrometers following the analytical procedures of HARRISON and FITZ GERALD (1986). Samples were heated in a Ta crucible within a double-vacuum furnace (e.g., STAUDACHER et al., 1978). A recent modification is to immerse in water the seals which separate the double-sided stainless-steel flange, to which the Ta crucible is electron- beam-brazed (HARRISON and FITZ GERALD, 1986) from the rest of the vacuum line. This arrangement allows extremely low blanks to be realized (e.g., 7 X lo-” mol 40Ar at 1000°C). Accuracy of tem- perature measurement and control is estimated to be better than +5”C for most conditions. Samples were irradiated in evacuated quartz vials along with Fe-mica biotite (307.3 Ma) flux monitors in the H-5 position of the Ford Reactor, University of Michigan. Un- certainties for ages are shown at the one sigma level and do not include the error in the J factor, estimated to be 0.5%. All ages have been calculated with the decay constants and isotope abundances recommended by STEICER and JAGER ( 1977). Tabulated results of the argon isotopic analyses corrected for neutron produced interfer- ences are available by request from the first author.

The multi-domain/multi-activation energy hypothesis places new demands on the 40Ar/39Ar analyst. Previously, the principle challenge in 40Ar/39Ar dating was to obtain uniform resolution across the age spectrum. For example, the rapid and variable onset of thermal de- terioration in amphiboles calls for considerable caution in step heating. Although this consideration has generally been unnecessary for alkali feldspars, we are now faced with two somewhat competing goals. The laboratory heating schedule which best brings out the Arrhenius and domain distribution parameters from alkali feldspars (i.e., cycled heating) invariably yields an age spectrum with non-uniform reso- lution, and vice versa. The following discussion suggests compromises which can lead to the clearest thermal history information being ob- tained from the sample.

Our initial response to this dilemma was to split the sample and run two heating experiments-one to optimize the Arrhenius plot and one to yield an asthetically pleasing age spectrum. However, the assumption implicit in this solution, that domain and activation pa- rameters are uniform from split to split (or even grain to grain), appears to be flawed. Our experience suggests that the only thing common to all grains of a basement feldspar sample is their thermal history. This requires that both the age and Arrhenius information be obtained on the same split.

The most fundamental requirement of the laboratory heating schedule is that extraction of the bulk of the gas occur below the temperature at which alkali feldspar reacts to form leucite plus melt. Our experience suggests that if temperature is maintained at - 1100°C no deterioration of the largest domain will occur. Under these con- ditions, even Al/Si ordering is unlikely to occur (see SMITH, 1974). The onset of melting appears to vary from sample to sample with some maintaining volume diffusion rates at temperatures as high as 1170°C. However, even at these temperatures most alkali feldspars require many days to degas 100% of their argon. Although the in- strumental performance requirements for these long heating durations

Page 14: 40Ar/39Ar results for alkali feldspars containing ...sims.ess.ucla.edu/argonlab/pdf/harrison_etal_GCA_1991.pdf · 40Ar/39Ar dating of alkali feldspars 1437 12+ 1 ‘* m~‘~* r 0.5

1448 T. M. Harrison, 0. M. Lovera, and M. T. Heizler

(e.g.. 39Ar blank, vacuum durability) are easily met by a modern argon extraction system (see MCDOUGALL and HARRISON, 1988), four to five day heatings are probably not a practical option for routine analysis. A goal of obtaining >85% of the argon is generally achievable with one overnight heating at 1140°C. In general, this will allow sufficient resolution on the log (r/rO) plot to reveal details of the largest domain present without consuming excessive amounts of laboratory time. Although this time requirement still exceeds the previous con- vention, the clarity of the thermal history information more than outweighs the apparent loss in productivity.

Compared with conventional monotonic heating, temperature cy- cling vastly improves the resolution of both the domain distribution parameters (LOVERA et al., I99 1) and activation energy variations. Our prefered approach is to plot diffusion coefficients (this requires only a rough estimate of/and potassium content of the sample) on

an Arrhenius plot during monotonically increasing sample heatings of 10 to I5 min (beginning at -450°C) to assess when the smallest domain is being exhausted (i.e., the shift off ro). At this point, a con- stant temperature (typically 900°C) is maintained, allowing the de- scent to the next dominant domain size to be observed. Once estab- lished on the next dominant domain, the temperature is then dropped to -7OO”C, either in stages or directly. Because the argon concen- tration gradient near the surface of the domains is now relatively shallow, heating at 700°C must be maintained for longer periods of time to yield measurable gas quantities. It is often convenient to heat the sample overnight to acquire sufficient signal. The temperature is then raised and the whole process begins over again for the next largest domain. As we document in this paper, resolution of the ac- tivation energy of the largest domain is often difficult as relatively little down-temperature leverage of the slope is possible.