Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of...

162
Experimental Constraints on Lithium Exchange between Clinopyroxene, Olivine and Aqueous Fluid at High Pressures and Temperatures by Natalie Caciagli Warman A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy Department of Geology University of Toronto © Copyright by Natalie Caciagli Warman 2010

Transcript of Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of...

Page 1: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

Experimental Constraints on Lithium Exchange between Clinopyroxene, Olivine and Aqueous Fluid at High

Pressures and Temperatures

by

Natalie Caciagli Warman

A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy

Department of Geology University of Toronto

© Copyright by Natalie Caciagli Warman 2010

Page 2: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

ii

Experimental Constraints of Lithium Exchange between

Clinopyroxene, Olivine and Aqueous Fluid at High Pressures and

Temperatures

Natalie Caciagli Warman

Doctor of Philosophy

Department of Geology University of Toronto

2009

Abstract

Clinopyroxene, olivine, plagioclase and hydrous fluid lithium partition coefficients have been

measured between 800-1100oC at 1 GPa. Clinopyroxene-fluid partitioning is a function of

temperature (ln DLicpx/fluid = -7.3 (+0.5) + 7.0 (+0.7) * 1000/T) and appears to increase with

increasing pyroxene Al2O3 content. Olivine-fluid partitioning of lithium is a function of

temperature (ln DLiol/fluid = -6.0 (+2.0) + 6.5 (+2.0) * 1000/T) and appears to be sensitive to olivine

Mg/Fe content. Anorthite-fluid lithium partitioning is a function of feldspar composition, similar

to the partitioning of other cations in the feldspar-fluid system. Isotopic fractionation between

clinopyroxene and fluid, Licpx-fluid, has been measured between 900-1100oC and ranges from -

0.3 to -3.4 ‰ (±1.4 ‰).

Lithium diffusion has been measured in clinopyroxene at 800-1000oC and in olivine at 1000oC.

The lithium diffusion coefficient is independent of the diffusion gradient as values are the same

if the flux of lithium is into or out of the crystal and ranges from -15.19 ± 2.86 m2/s at 800oC to -

11.97 ± 0.86 m2/s at 1000oC. Lithium diffusion in olivine was found to be two orders of

magnitude slower than for clinopyroxene at similar conditions. Closure temperatures calculated

Page 3: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

iii

for lithium diffusion in clinopyroxene range from ~400 to ~600oC. These results demonstrate

that lithium equilibration between fluids and minerals is instantaneous, on a geological

timescales.

The confirmation of instantaneous equilibration, combined with min-fluid partition coefficients

and values for Licpx-fluid, permits quantitative modeling of the evolution of lithium concentration

and isotopic composition in slab-derived fluids during transport to the arc melt source. Our

results indicate that fluids migrating by porous flow will rapidly exchange lithium with the

mantle, effectively buffering the fluid composition close to ambient mantle values, and rapidly

attenuating the slab lithium signature. Fluid transport mechanisms involving fracture flow are

required to maintain a slab-like lithium signature (both elemental and isotopic) from the slab to

the melt source of island arc basalts.

This study demonstrates that mineral-fluid equilibration is rapid, and as a result the lithium

content of minerals can only reliably represent chemical exchange in the very latest stages of the

sample’s history.

Page 4: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

iv

Acknowledgments

This thesis dissertation marks the conclusion of work that began in 2001. There were many who

helped, supported, and cheered me on my way and I fear that to be able to acknowledge everyone

who assisted me would be an impossible task. I am certain that once this work has been

submitted I will realize that I have left out several important people.

First, I must thank Dr. James Brenan, my mentor and supervisor, who never gave up on me

despite everything. I am indebted to Dr. Lesley Rose Weston (my lab partner in crime) and Dr.

Boris Foursenko for all their time and invaluable mechanical, technical, and moral support. I am

grateful to my collaborators at Lawrence Livermore National Laboratory, Dr. Doug Phinney, Dr.

Ian Hutcheon and Dr. Rick Ryerson for access to and assistance with the SIMS. Thanks to Dr.

Bill McDonough and his lab at University of Maryland for the isotopic analyses. I also wish to

thank Dr. Grey Bebout at Lehigh University, for his patience and understanding when, to quote

Edison, “I found 10,000 ways that won’t work.” before I found one way that did work. I would

also like to thank Dr. Paul Tomascak, who pointed me in the direction that would bear the most

fruit and Dr. Jon Davidson for his unwavering encouragement for always having time for a ‘little

chat’.

Mainly I am indebted to my husband Tim, who never once let me give up on myself.

This work must be dedicated to my children, Olivia and Henry, whose existence shaped this

project more than any other thing.

Page 5: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

v

Table of Contents

Acknowledgments.......................................................................................................................... iv

Table of Contents............................................................................................................................ v

List of Tables ................................................................................................................................. ix

List of Figures ................................................................................................................................. x

List of Appendices ....................................................................................................................... xiii

1 Introduction ................................................................................................................................ 1

1.1 Chemical Properties of Lithium.......................................................................................... 1

1.2 Sources and Concentrations of Lithium in the Earth .......................................................... 3

1.2.1 Subducted Materials (AOC and Seafloor Sediments) ............................................ 3

1.2.2 Eclogites.................................................................................................................. 4

1.2.3 Upper Mantle .......................................................................................................... 5

1.2.4 Convergent Margin Magmas .................................................................................. 6

1.3 The scale of lithium heterogenities ..................................................................................... 7

1.4 Previous experimental work ............................................................................................... 8

1.5 Focus of Thesis and Distribution of Work.......................................................................... 9

2 Lithium Partitioning and Isotopic Fractionation ...................................................................... 14

2.1 Introduction....................................................................................................................... 14

2.2 Methods............................................................................................................................. 16

2.3 Analytical Techniques ...................................................................................................... 18

2.3.1 Major Element Analyses....................................................................................... 18

2.3.2 Lithium Analyses .................................................................................................. 19

2.3.2.1 MC-ICPMS............................................................................................. 19

2.3.2.2 SIMS....................................................................................................... 19

Page 6: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

vi

2.3.2.3 LA-ICPMS ............................................................................................. 20

2.4 Results............................................................................................................................... 21

2.4.1 Major Element Chemistry..................................................................................... 21

2.4.2 Mineral – Fluid Lithium Partitioning.................................................................... 21

2.4.2.1 Clinopyroxene ........................................................................................ 23

2.4.2.2 Olivine .................................................................................................... 24

2.4.2.3 Plagioclase .............................................................................................. 25

2.4.3 Olivine-Clinopyroxene Pair Experiments............................................................. 26

2.4.3.1 Experiments at variable fO2.................................................................... 27

2.4.3.2 Experiments with added REE................................................................. 27

2.4.4 Lithium Isotope Fractionation............................................................................... 28

2.4.4.1 Clinopyroxene- fluid .............................................................................. 28

2.4.4.2 Olivine-Clinopyroxene Lithium Isotope Fractionation .......................... 29

2.5 Discussion ......................................................................................................................... 29

2.5.1 Controls on Partitioning........................................................................................ 29

2.5.1.1 Clinopyroxene ........................................................................................ 29

2.5.1.2 Olivine .................................................................................................... 31

2.5.1.3 Plagioclase .............................................................................................. 31

2.5.1.4 Intermineral Partitioning......................................................................... 32

2.5.2 Controls on Isotopic Fractionation........................................................................ 33

2.5.3 Lithium Incorporation into the Mantle ................................................................. 34

2.5.4 The Mantle Wedge as a Chromatograph .............................................................. 35

2.5.5 Isotopic Evolution of Lithium-Bearing Fluids in the Mantle ............................... 39

2.5.5.1 Percolation and Rayleigh Distillation..................................................... 39

2.5.5.2 Generation of 6Li-rich fluids................................................................... 41

2.5.5.3 Generation of 6Li-rich zones in the mantle............................................. 42

Page 7: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

vii

2.6 Conclusions....................................................................................................................... 44

3 Lithium Diffusion..................................................................................................................... 66

3.1 Introduction....................................................................................................................... 66

3.2 Experimental Methods ...................................................................................................... 68

3.3 Analytical Techniques ...................................................................................................... 69

3.3.1 Major Element Analyses....................................................................................... 69

3.3.2 Lithium Analyses .................................................................................................. 70

3.3.2.1 LA-ICPMS ............................................................................................. 70

3.3.2.2 Secondary Ion Mass Spectrometry (SIMS)............................................ 70

3.3.3 Data Reduction...................................................................................................... 71

3.4 Results............................................................................................................................... 72

3.4.1 Diffusion in Clinopyroxene .................................................................................. 72

3.4.2 fO2 Series Experiments ......................................................................................... 72

3.4.3 Diffusion in Olivine .............................................................................................. 73

3.4.4 Kinetic Fractionation of 7Li/6Li ............................................................................ 73

3.5 Discussion ......................................................................................................................... 74

3.5.1 Effect of fO2 on lithium diffusion in clinopyroxene ............................................. 74

3.5.2 Comparison with other lithium diffusion studies.................................................. 76

3.5.3 Comparison with diffusion of other cations in clinopyroxene.............................. 76

3.5.4 Geological Implications ........................................................................................ 77

3.5.4.1 Preservation of Lithium Signatures ........................................................ 77

3.5.4.2 Closure Temperature .............................................................................. 79

3.5.4.3 The Potential for Re-Equilibration of Lithium Composition ................. 79

3.5.4.4 Diffusion-Induced Isotopic Fractionation .............................................. 82

3.6 Conclusions....................................................................................................................... 84

4 Technique Development to Study Muscovite-Fluid Partitioning of Nitrogen....................... 104

Page 8: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

viii

4.1 Introduction..................................................................................................................... 104

4.2 Theoretical Considerations ............................................................................................. 106

4.2.1 N-speciation and Isotopic Fractionation ............................................................. 106

4.2.2 Buffering pH ....................................................................................................... 108

4.3 Experimental Methodology ............................................................................................ 109

4.4 Analytical Methods......................................................................................................... 109

4.5 Results............................................................................................................................. 110

4.5.1 Nitrogen Partitioning .......................................................................................... 110

4.5.2 Nitrogen Isotopic Fractionation .......................................................................... 111

4.6 Discussion ....................................................................................................................... 112

4.6.1 Utility of NH4Cl as Nitrogen Source .................................................................. 112

4.6.2 Analytical Considerations................................................................................... 112

4.6.3 Experimental Considerations .............................................................................. 113

4.6.4 Isotopic Fractionation Experiments and Atmospheric Contamination............... 114

4.7 Suggestions for Future Work .......................................................................................... 114

5 Summary of Results and Conclusions.................................................................................... 125

References................................................................................................................................... 128

Appendix 1.................................................................................................................................. 140

6 Summary of Boron Work....................................................................................................... 140

Page 9: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

ix

List of Tables

Table 2.1 Composition of Starting Material ................................................................................. 46

Table 2.2 Experimental Conditions .............................................................................................. 47

Table 2.3 Standards and Reference Material ................................................................................ 48

Table 2.4 Run Product Composition............................................................................................. 49

Table 2.5 Run Product Lithium Concentration............................................................................. 50

Table 2.6 Isotopic Composition of Starting Materials and Run Products .................................... 51

Table 3.1 Composition of Starting Material ................................................................................. 85

Table 3.2 Measurements of Standards .......................................................................................... 86

Table 3.3 Summary of Experiments ............................................................................................. 87

Table 4.1 Experiments and Results............................................................................................. 116

Table 4.2 Percentage of Nitrogen Contribution from Air........................................................... 116

Page 10: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

x

List of Figures

Figure 1.1 Sources and Concentration of Lithium in the Earth .................................................... 11

Figure 1.2 Li/Y ratio and 7Li in Arc Lavas................................................................................. 12

Figure 1.3 Lithium Diffusion Coefficients ................................................................................... 13

Figure 2.1 Internal Reference Materials ....................................................................................... 52

Figure 2.2 Standards and Reference Material............................................................................... 53

Figure 2.3 Photomicrographs of Starting Material and Run Products.......................................... 54

Figure 2.4 Time Resolved Spectra................................................................................................ 55

Figure 2.5 lnDLi cpx/fluid vs 1000/T ........................................................................................... 56

Figure 2.6 lnDLi ol/fluid vs 1000/T.............................................................................................. 57

Figure 2.7 Anorthite/Fluid Lithium Partitioning .......................................................................... 58

Figure 2.8 Olivine/Clinopyroxene Lithium Partitioning .............................................................. 59

Figure 2.9 Lithium Partitioning From Mantle Xenoliths and Experimental Studies.................... 60

Figure 2.10 Mineral/Fluid Isotopic Fractionation of Lithium ...................................................... 61

Figure 2.11 Time for Li and B Transport to Top of Column........................................................ 62

Figure 2.12 Evolution of the Slab Derived Fluid by due to Rayleigh Distillation ....................... 63

Figure 2.13 Lithium Coordination and P-T Paths......................................................................... 64

Figure 2.14 Evolution of 7Li of Mantle Wedge due to Hydrofractures ...................................... 65

Figure 3.1 Li Elemental and Isotopic Gradients in San Carlos Opx............................................. 88

Figure 3.2 Effect of fO2 Anneal .................................................................................................... 89

Page 11: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

xi

Figure 3.3 Zero time Experiment.................................................................................................. 90

Figure 3.4 Results for Experiment Kcpx-12 ................................................................................. 91

Figure 3.5 X-Ray Maps of Run Product ....................................................................................... 92

Figure 3.6 Time Series .................................................................................................................. 93

Figure 3.7 Measured Lithium Diffusion Coefficients................................................................... 94

Figure 3.8 fO2 Experiment Series ................................................................................................. 95

Figure 3.9 Lithium Diffusion Profile in San Carlos Olivine ........................................................ 96

Figure 3.10 Lithium Diffusion and Isotopic Fractionation in Kcpx-2.......................................... 97

Figure 3.11 Comparison of Lithium Diffusion Coefficients ........................................................ 98

Figure 3.12 Comparison of Diffusivities Measured in Pyroxene ................................................. 99

Figure 3.13 Retention of Lithium Composition.......................................................................... 100

Figure 3.14 Comparison of Closure Temperature of Li and Sr in Clinopyroxene ..................... 101

Figure 3.15 Lithium Isotopic Compositions of Kilauea Iki Lava Lake Rocks........................... 102

Figure 3.16 Li Isotopic Gradient in San Carlos Opx and Modeled Profile ................................ 103

Figure 4.1 Summary of N Concentration and Isotopic Composition ......................................... 117

Figure 4.2 Calculated N2-, and NH3-NH4+ Fractionation ........................................................... 118

Figure 4.3 Relationship of fH2, fN2, and fNH3............................................................................ 119

Figure 4.4 Scanning Electron Micrograph of Muscovite Texture .............................................. 120

Figure 4.5 Nitrogen contents of run products ............................................................................. 121

Figure 4.6 Nitrogen isotopic compositions of run products ....................................................... 122

Page 12: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

xii

Figure 4.7 Isotopic shifts of run products ................................................................................... 123

Figure 4.8 Puncturing Device ..................................................................................................... 124

Page 13: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

xiii

List of Appendices

6.1 11B notation ................................................................................................................... 140

6.2 Evidence of Boron Mobility from Arc Lavas ................................................................. 140

6.3 Evidence of Boron Mobility from Eclogites................................................................... 142

6.4 Summary of Experimental Methodology........................................................................ 143

6.5 Details of Boron Study.................................................................................................... 143

6.6 Boron Analyses............................................................................................................... 146

6.7 References for Boron Study............................................................................................ 148

Page 14: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

1

1 Introduction

According to Davidson (1996), three fundamental questions remain unanswered in the study of

island arc magmagenesis:

“1. What is the (presubduction) composition of the mantle wedge source of arc magmas? 2. To what extent does it melt, and by what process? 3. What is the composition and amount of slab-derived component added to the wedge?”

Despite the significant strides that have been made in both our knowledge of earth processes and

our technical ability to analyze earth materials with greater precision and accuracy, these

questions remain unsatisfactorily answered today.

Low abundance, or trace elements, can provide essential information to address these issues. For

example, both convergent margin basalts and mid ocean ridge basalts (MORB) have similar

major element compositions; however, convergent margin basalts are differentiated by high

LILE/REE; (large ion lithophile element - Rb, K, Cs, Ba, Sr to rare earth element - actinides; La

through Lu) and high LILE/HFSE (high field strength element – Nb, Ta, Zr, Hf, Ti) signature

(Davidson, 1996). This pattern is interpreted to mean that arc magmas are products of the

overlying mantle wedge melt plus a LILE-rich fluid or melt originating from the subducted slab.

Boron, lithium and nitrogen have often been employed to identify the composition and amount

of the slab component in island arc magmas. These elements are relatively fluid mobile and

somewhat incompatible in mantle minerals and are easily released from the slab and

concentrated into the magmas at the arc front (Ishikawa and Tera, 1999; Leeman, 1996; Ishikawa

and Nakamura, 1994; Ishikawa et al., 2001; Moriguti and Nakamura, 1998).

1.1 Chemical Properties of Lithium

Lithium belongs to the Group 1 elements of the periodic table. Like the other alkali metals it is

characterized by low ionization energy and low electronegativity, and commonly forms

hydroxides, nitrides, carbonates and chlorides. When octahedrally coordinated it has an effective

ionic radius of 0.59 Å which is comparable to octahedrally coordinated Mg2+ (0.72 Å) and Fe2+

Page 15: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

2

(0.78 Å), which allows it to substitute for these elements in olivine, pyroxenes, amphiboles and

clays (Brenan et al., 1998; Wenger and Armbruster, 1991). The oxygen co-ordination of lithium

can vary from 3 to 8; although lithium has a preference for tetrahedral co-ordination in melts and

fluids (Cormier et al., 1998; Majérus et al., 2003), it can be accommodated by octahedral co-

ordination, as is the case in many silicate minerals (Wenger and Armbruster, 1991). Lithium has

a single valence electron with a very low ionization potential, which makes it easily solvated.

Lithium has two stable isotopes, 6Li and 7Li, with a relative mass difference of ~16 % and

abundances of ~7.52 % and ~92.48 %, respectively. Enrichments in lithium isotopes are

described as either:

(1) 10001LiLi/

LiLi/Li

std76

smp76

6

or 10001

LiLi/

LiLi/Li

std67

smp67

7

where ‘smp’ refers to the sample and ‘std’ refers to the standard, typically the NBS lithium

carbonate L-SVEC (SRM #8545). The δ7Li notation is recommended by the International Union

for Pure and Applied Chemistry (Coplen et al., 1996) and will be used here.

Figure 1.1 shows the lithium abundance and isotopic composition of various geochemical

reservoirs. As with other isotopes, lithium isotopic fractionation between minerals and fluids

depends on the difference in the zero point potential energy (ZPE) between the phases of interest.

Heavier isotopes have lower vibrational frequencies, and therefore a lower ZPE than lighter

isotopes (Chacko et al., 2001). The molecule or phase that will undergo the greatest reduction in

ZPE with the substitution of the heavy isotope will become enriched in the heavier isotope

(Chacko et al., 2001). Ab initio calculations have demonstrated that during mineral-solution

reactions 6Li should be preferentially incorporated into octahedrally coordinated sites in the solid

(Yamaji et al., 2001). This appears to be the case in the formation of secondary minerals

produced during alteration of crustal rocks (Huh et al., 2001; Pistiner and Henderson, 2003;

Seyfried et al., 1998). During the formation of clays, 6Li is concentrated in the solid phase while

the resulting fluid becomes enriched in 7Li (Huh et al., 2001; Pistiner and Henderson, 2003).

Experimental measurements of lithium isotopic fractionation between spodumene and fluids also

confirm this behavior (Wunder et al., 2006). Interestingly, experiments measuring lithium

isotopic fraction between staurolite and fluids found that 6Li was preferentially enriched in the

Page 16: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

3

fluids, and 7Li was enriched in the solids (Wunder et al., 2007). Considering that lithium is in

tetrahedral coordination in staurolite, this result also appears to confirm the ab initio calculations.

1.2 Sources and Concentrations of Lithium in the Earth

The various components of the convergent margin system have significant differences in lithium

concentrations and lithium isotopic compositions (see Figure 1.1). Inputs such as continental

crust (as pelagic clays) and altered oceanic crust tend to be enriched in lithium with respect to the

mantle, but the isotopic composition can vary considerably depending on the type and degree of

alteration or weathering. Mantle inputs tend to be more uniform in terms of lithium concentration

and isotopic composition; however, the MORB-source mantle can potentially contain both

elemental and isotopic heterogeneities. Very low δ7Li (-11 ‰ to +5 ‰) values are found in

samples of metamorphosed oceanic crust, possibly reflecting low temperature dehydration of the

slab during subduction (Zack et al., 2003). Conversely some pyroxenites from the Zabargad

peridotite, which is considered to be a fragment of exhumed mantle wedge, have high δ7Li

values (+8.4 ‰ to +11.8 ‰) suggesting that 7Li-rich fluids or melts derived from subducted

altered oceanic crust (AOC) may also be transferred to the mantle (Brooker et al., 2004). The

variability exhibited by the components of the arc magmagenesis system is not always reflected

in the output. Some island arc magmas display higher ratios of fluid mobile elements, such as

lithium, to relatively immobile elements, such as yttrium in the front-arc regions, which decrease

towards the back-arc. However, few arc lavas have δ7Li significantly greater than MORB, and

correlations between δ7Li and fluid enrichment are not always clear or consistent (Chan et al.,

2002; Tomascak et al., 2002; Tomascak, 2004; Leeman et al., 2004).

1.2.1 Subducted Materials (AOC and Seafloor Sediments)

Lithium in seafloor sediments ranges from 5 to 80 ppm with δ7Li ranging from -5 ‰ to +20 ‰

(Marschall et al., 2007 and references therein). In oceanic crust lithium can range from 5 to 6

ppm with a δ7Li of +1.5 ‰ to +5.6 ‰ in fresh mid-ocean ridge basalts (Moriguti and Nakamura,

1998; Tomascak et al., 2008; Chan et al., 1992) whereas altered oceanic crust may have >75 ppm

lithium with δ7Li of +14.2 ‰ (Moriguti and Nakamura, 1998; Tomascak et al., 2008; Chan et al.,

1992). Low temperature alteration of basalt results in concentration of lithium, and preferentially

Page 17: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

4

6Li, into secondary minerals and enrichment of 7Li in seawater with a fractionation of up to +19

‰ with respect to the residual solid (Chan et al., 1992; Chan et al., 2002; Seyfried et al., 1998;

James et al., 2003). At temperatures greater than 350 oC, lithium is mobilized by saline fluids

and the extent of the isotopic fractionation decreases (Chan et al., 1992; Chan et al., 2002;

Seyfried et al., 1998; James et al., 2003).

The extent of solid-fluid fractionation has been inferred by various authors by measuring lithium

in sediment-derived pore fluids and serpentine diapirs (Chan et al., 1992; Chan et al., 2002; Chan

and Kastner, 2000; Benton et al., 2004). Pore fluids from the Costa Rican trench show a ~11 ‰

enrichment of 7Li with respect to the down-going sediments (Chan and Kastner, 2000). A larger

range of isotopic compositions, 7Li of -0.5 ‰ to +10 ‰ has been measured in so called

serpentinite diapers, which are super-hydrated mantle wedge extruded in the fore arc of the

Mariana trench (Benton et al., 2001; Benton et al., 2004). The variability in the lithium isotopic

composition of vent fluids (δ7Li ranging from +5 ‰ to +43.8 ‰) likely reflects differences in

temperature, reaction paths and fluid - rock ratios as well as source rock composition (Zhang et

al., 1998; Chan and Kastner, 2000; Foustoukos et al., 2004).

1.2.2 Eclogites

Alpine eclogites are thought to be exhumed remnants of subducted oceanic crust. The eclogites

at Trescolmen, Switzerland investigated by Zack et al. (2003) displayed extremely light 7Li

values ranging from -11 ‰ to +5 ‰. This study speculated that during subduction the slab was

progressively depleted in 7Li, via Rayleigh distillation during dehydration, and that the resulting 6Li enriched material was recycled into the mantle. A more comprehensive study of eclogites,

blueschists and other high pressure metamorphic rocks from classic European (Swiss Alps,

Münchberg, Aldalen and Greek Islands) and Asian localities (Qaidam, Dabieshan and Tianshan)

by Marschall et al. (2007) found an even larger range of lithium concentrations and isotopic

compositions. Lithium concentrations ranged from ~1 to ~50 ppm and 7Li values ranged from -

21.9 ‰ to > +6 ‰ (Marschall et al., 2007). More significantly, no correlation was found between

lithium content and 7Li within any single locality or within the full population. In fact, many of

the lightest samples had > 30 ppm lithium, contrary to what would be expected from a simple

Rayleigh-type distillation of the subducted slab. Marschall et al. (2007) concluded that many

exhumed eclogites have lithium compositions (both elemental and isotopic) that have been

Page 18: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

5

influenced by influx of retrograde fluids and kinetically induced isotopic fractionation during

exhumation, and therefore primary subduction-induced fractionation has been modified by more

recent processes.

1.2.3 Upper Mantle

From analyses of pristine peridotite xenoliths, the upper mantle is estimated to contain 1.5 ppm

lithium with an average δ7Li of +4 ‰ (Jagoutz et al., 1979; Tomascak, 2004; Jeffcoate et al.,

2007). However, the fact that OIB and MORB can sometimes display a range of values has led

to the suggestion that variable amounts of recycled crustal material are sometimes present in the

mantle sources of these lavas (Tomascak et al., 2008). A study of MORB lavas from different

ridge systems found δ7Li ranging from +1.6 ‰ to +5.6 ‰. This represents a 5 ‰ heterogeneity

in the samples with no consistent correlation of δ7Li with major and trace elements or radiogenic

isotopes (Tomascak et al., 2008). Analysis of lithium in glass inclusions from Hawaii showed

δ7Li varying from –10.2 ‰ to +8.4 ‰ (Kobayashi et al., 2004). Similarly, low δ7Li (-3.3 ‰ to

+1.2 ‰) from glass inclusions in Iblean (Sicilian) Plateau tholeiites are thought to reflect melting

of an isotopically light region in the mantle (Gurenko and Schmincke, 2002). In rare cases, δ7Li

correlations with abundances of other trace elements or isotopes can be found. For example,

samples from the East Pacific Rise (EPR) show a weak correlation of increasing δ7Li with

increasing Cl/K, which was interpreted to reflect mixing or assimilation of recycled crustal

material (Tomascak et al., 2008). Other EPR samples show a positive correlation between δ7Li

and 143Nd/144Nd, (Elliot, 2004), again reflecting a possible recycled component. That reservoirs

with variable δ7Li exist in the mantle is also suggested by studies of peridotite massifs and

ultramafic xenoliths. Nishio et al. (2004) report the lithium isotopic composition of

clinopyroxene from xenoliths from Japan, SE Australia and eastern Russia. The δ7Li values from

NE Japan and SE Australia were high (+4 to +7 ‰), whereas Russian and SW Japan samples

were significantly lower (-17 to -3 ‰). In some of the samples from eastern Russia, δ7Li could

be positively correlated with 143Nd/144Nd but those correlations did not apply to the other sample

populations or even to all the eastern Russia samples. Ionov and Seitz (2008) also reported

lithium concentrations and isotopic data from xenoliths from the Kamchatka arc and the Vitim

(Siberia) volcanic field (an intra-plate continental volcanic setting). They found a relatively small

range of lithium concentrations, ~1 to ~2 ppm, and lithium isotopic composition of -3.6 to +6 ‰.

Page 19: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

6

These studies demonstrate significant variations in the isotopic compositions of lithium in the

upper mantle, but suggest that such variation is fairly localized. Reports of significant correlation

between δ7Li and other elemental or isotopic tracer elements are rare within any given sample

locality and no global correlation has been found within a given tectonic setting.

1.2.4 Convergent Margin Magmas

Magmatism at convergent margins is commonly believed to be due to hydrous fluids from the

subducting slab fluxing the mantle wedge and lowering the mantle solidus. As mentioned

previously, the resulting magmas are very often characterized by higher ratios of fluid mobile

elements to relatively insoluble elements, such as the high field strength elements (HFSE). In

order to distinguish between crystal/melt fractionation effects and fluid involvement, fluid

mobile elements (such as lithium and boron) are measured against relatively insoluble elements

with similar solid/melt partitioning (i.e. B/Be, Li/Yb or Li/Y). The Izu arc in Japan is the

locality showing the clearest indication of a Li-bearing, slab-derived component. This is shown

by the correlated decrease in both Li/Y and δ7Li from the arc front lavas to those erupted in the

back arc region (Figure 1.2a, data from Moriguti and Nakamura, 1998). This is suggestive of

continuing mobilization of lithium into the arc source region by fluids derived from dehydration

reactions in the down going slab (Leeman, 1996; Ishikawa and Tera, 1999; Ishikawa and

Nakamura, 1994; Ishikawa et al., 2001). However, the trend displayed at Izu appears to be the

exception and not the rule (Figure 1.2b, data from Tomascak et al., 2002). Variations in slab age

and angle of subduction, which would influence the thermal regime, and therefore the extent of

dehydration, have been cited to explain these differences (Moriguti et al., 2004). Yet, a

difference in the Li/Y and 7Li behavior between the Kurile arc and the Izu arc, or even between

the Izu arc and the Japan arc where age of the subducted slab is similar, still persists, despite the

similarity of subduction regime (Morguti et al., 2004; Tomascak et al., 2002). Another

suggestion is that slab-derived fluids are significantly modified during transport through the

mantle wedge to the melt source, and that the lithium signal is attenuated by interaction with

mantle minerals (Tomascak et al., 2002). The extent of modification of the slab-derived

component during transport through, and interaction with, the overlying mantle wedge is

unknown, but it remains unanswered as to why more arc magmas do not display the same trends

as clearly as the Izu arc.

Page 20: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

7

1.3 The scale of lithium heterogenities

Recent in situ micron scale analyses of lithium isotopes in geological samples have yielded

unexpected results. High temperature equilibrium is expected to impose minimal differences in

the lithium isotopic composition of individual minerals; however, samples from some magmatic

environments have revealed significant isotopic heterogeneity at the grain-scale. For example,

Rudnick and Ionov (2007) reported highly variable δ7Li in clinopyroxene and olivine grains in

peridotite xenoliths from eastern Russia. δ7Li values ranged from -0.8 to -14.6 ‰ for

clinopyroxene and -1.7 to +11.9 ‰ for corresponding olivine, and olivine/clinopyroxene

distribution coefficients varied from 0.2 to 1.0, which is lower than previously estimated for

equilibrium partitioning. Analyses of olivine and clinopyroxene pairs from a xenolith from the

Vitim volcanic field found δ7Li to range from -17 to -18 ‰ in the pyroxenes with a δ7Li of +6 ‰

in the corresponding olivine (Ionov and Seitz, 2008). Bulk measurements of olivine phenocrysts

in primitive magmas from a variety of localities found a relatively uniform δ7Li of +3.2 to +4.9

‰; however, measurements of clinopyroxene yielded highly variable δ7Li (+6.6 ‰ to -8.1 ‰;

Jeffcoate et al., 2007). Both the olivine and clinopyroxene phenocrysts from Solomon Island

lavas are zoned with respect to lithium and δ7Li (Parkinson et al., 2007). Rims of phenocrysts are

enriched in lithium compared to the cores and the δ7Li decreases from core to rim by as much as

20 ‰ in a W-shaped profile (Parkinson et al., 2007). This pattern was also observed by Jeffcoate

et al. (2007) who measured a 40‰ variation in a single orthropyroxene crystal from a San Carlos

xenolith. The extreme grain-scale variability exhibited by lithium and lithium isotopes is not

limited to terrestrial samples. The basaltic lunar meteorite NWA 479 examined by Barrat et al.

(2005) contains olivine and pyroxene phenocrysts that also display a wide range of δ7Li values

(+2.4 to +15.1 ‰ in olivine and -0.2 to 16.1 ‰ in pyroxene). Beck et al. (2004) examined

pyroxenes in the shergottite meteorite NWA 480 and found extreme zoning of δ7Li from -17 ‰

in the cores to +10 ‰ in the rims and an absence of lithium compositional variation.

The extreme fractionation of lithium isotopic values documented in these high temperature

samples suggests kinetic rather than equilibrium processes (Lundstrom et al., 2005; Beck, 2006;

Jeffcoate et al., 2007; Parkinson et al., 2007; Rudnick and Ionov, 2007; Marchall et al., 2007).

This kinetic effect has been experimentally demonstrated by Richter et al. (2003) for diffusion of

Page 21: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

8

lithium in molten silicate where fractionation occurs due to slightly faster transport of 6Li than 7Li. The time scale for the development of diffusion-controlled isotopic fractionation is likely to

be quite short as documented by the rapid lithium exchange in natural samples. Berlo et al.

(2004) reported rapid mobilization of lithium in a study of plagioclase phenocrysts from the 1980

eruption of Mount St. Helens in Washington, USA. Plagioclase phenocrysts erupted prior to the

degassing event contained ~14 ppm lithium, whereas those erupted immediately after contained

~5 ppm. The implication is that the magma lost a significant amount of lithium in a seven-day

period, which was recorded in the lithium content of the plagioclase phenocrysts. Kent et al.

(2007) also found that the lithium contents of plagioclase phenocrysts from the Mount St. Helens

2004 dome lavas had increased due to the addition of a pre-eruptive lithium rich vapour phase.

Based on the lithium contents of plagioclase phenocrysts, melt inclusions, and plagioclase

encapsulated within gabbroic inclusions Kent et al. (2007) were able to estimate that the influx of

the lithium-rich volatile phase occurred within ~1 yr of the dome lava eruptions.

1.4 Previous experimental work

Previous experimental work has found lithium to be moderately incompatible in clinopyroxene

co-existing with either fluid or melt, and that partitioning is a function of clinopyroxene major

element composition, DLicpx/melt increasing with increasing Ca/Al ratio (Hart and Dunn, 1993;

Brenan et al., 1998a; Blundy et al., 1998; Blundy and Dalton, 2000; Hill et al., 2000; Bennett et

al., 2004). Lithium was also found to be moderately incompatible in olivine relative to melt

(Brenan et al., 1998a; Brenan et al., 1998b, Taura et al., 1998; Zanetti et al., 2004). To date,

lithium partitioning between olivine and fluids has not been measured. Coogan et al. (2005)

measured lithium partitioning between plagioclase and clinopyroxene and found DLiplag/cpx

increases with increasing temperature (900oC to 1200oC). In this case partitioning was not

investigated with respect to plagioclase or clinopyroxene major element chemistry.

Little work has been done to determine lithium isotope partitioning and diffusion at pressures

and temperatures corresponding to crustal and mantle processes. A study examining isotopic

fractionation between spodumene and hydrous fluids measured an enrichment of 7Li in the fluid

from +3.5 ‰ at 500oC to ~ +1.0 ‰ at 900oC and 2.0 GPa (Wunder et al., 2006). Isotopic

fractionation between fluids and mica (from 300oC to 500oC, 2.0 GPa) and staurolite (from

Page 22: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

9

670oC and 880oC, 3.5 GPa) found fluids to be preferentially enriched in 7Li relative to the mica,

and staurolite to be slightly enriched in 7Li relative to the fluid (Wunder et al., 2007). These

studies are consistent with ab initio calculations where 6Li is preferentially fractionated in sites

with octahedral coordinations (e.g. mica, spodumene; Yamaji et al., 2001, Wunder et al. 2006)

and 7Li is preferentially fractionated in to sites with tetrahedral coordination (e.g. staurolite;

Wunder et al., 2007). Fractionation of lithium isotopes was also measured in the quartz-

muscovite-fluid system, from 400-500oC (Lynton et al., 2005). Lynton et al. (2005) found both

the quartz and the mica to be preferentially enriched in 7Li, with fractionation factors ranging

from +8 to +12 ‰ for quartz and +18 to +20 ‰ for mica. For reasons that are unclear, these

results are inconsistent with the subsequent studies of Wunder et al. (2006, 2007) or the results of

this study. Because mica has lithium in octahedral coordination the expectation is that the fluids

would be preferentially enriched in 7Li with respect to the solid.

Experimental studies of kinetic isotopic fractionation (Richter et al., 2003) found that 7Li could

be fractionated from 6Li by tens of per mil during diffusion between molten basalt and rhyolite or

when diffusing through fluids. Although estimates of lithium diffusion coefficients have been

made from natural samples, to date there have been few laboratory measurements of lithium

diffusion in rock forming minerals (Figure 1.3). Giletti and Shanahan (1997) measured the

diffusion rates of various alkali elements in plagioclase feldspar. They found that diffusion in

feldspars is a function of ionic radius and cation charge, and as a result of its small size, lithium

diffusion rates are very rapid. Coogan et al. (2005) measured the diffusion coefficient for 6Li in

clinopyroxene between 800oC and 1100oC by using SIMS analysis and found lithium to be

similarly rapid.

1.5 Focus of Thesis and Distribution of Work

Knowledge of lithium elemental partitioning and isotopic fractionation between fluids and

common rock forming minerals is essential to evaluate the variations seen in natural samples

adequately. Information on lithium diffusion in minerals can be used to more accurately assess

the time-scales of magmatic and hydrothermal processes and account for intermineral isotopic

differences.

Page 23: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

10

This study examines the lithium, nitrogen and boron isotope fractionation that occurs in mineral-

fluid reactions during slab and mantle interaction. The existing experimental database is

insufficient to properly evaluate the degree of isotopic fractionation that occurs during fluid-

mineral partitioning of lithium, boron, and nitrogen. The results of this work provide the

essential input for modeling the behavior of lithium in the mantle.

Chapter 2 of this study examines the partitioning and isotopic fractionation that occurs between

clinopyroxene, olivine, plagioclase and aqueous fluids and the intermineral fractionation between

olivine and clinopyroxene. All experiments were conducted by N. Caciagli in the High Pressure

Laboratory at the University of Toronto. The elemental lithium was analyzed by laser ablation

inductively coupled plasma mass spectroscopy (LA-ICPMS) at the University of Toronto by N.

Caciagli. The multi collector inductively coupled plasma mass spectroscopy (MC-ICPMS)

analyses of the clinopyroxene starting material and bulk analyses of the run product

clinopyroxene were done at University of Maryland by W. F. McDonough. The lithium isotopic

composition of starting material anorthite and olivine, and in situ lithium isotopic compositions

were analyzed by secondary ionization mass spectroscopy (SIMS) at Lawrence Livermore

National Laboratory by N. Caciagli with the assistance of D. Phinney.

Chapter 3 of this study measures the diffusion coefficient of lithium in clinopyroxene and

olivine. All experiments were conducted by N. Caciagli in the High Pressure Laboratory at the

University of Toronto. The elemental lithium was analyzed by LA-ICPMS at the University of

Toronto by N. Caciagli and lithium isotopic analyses were done by SIMS at Lawrence

Livermore National Laboratory by N. Caciagli with the assistance of D. Phinney.

Chapter 4 of this study outlines the technique development for experimental measurements of

nitrogen partitioning and isotopic fractionation between fluids and muscovite. All experiments

were conducted by N. Caciagli in the High Pressure Laboratory at the University of Toronto. The

nitrogen elemental and isotopic analyses were done at Lehigh University with the assistance of

G. Bebout.

The exploratory work on techniques to measure boron partitioning and isotopic fractionation

between muscovite and fluid is summarized in Appendix 1.

Page 24: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

11

0 0.01 0.1 1 10 100 1000

ave river water

sea water

est mean c.c.

marine seds

AOC

MORB

OIB

IAB

eclogites

xenoliths

est mean mantle

ppm

A

-30 -20 -10 0 10 20 30 40

river water

sea water

est mean c.c.

marine seds

AOC

MORB

OIB

IAB

eclogites

xenoliths

est mean mantle

7Li

B

Figure 1.1 Sources and Concentration of Lithium in the Earth

Lithium concentration (A) and isotopic composition (B) of various terrestrial reservoirs. River water: Huh et al. (1998);

seawater: Millot et al. (2004); estimated continental crust (c.c): Teng et al. (2004); marine sediments: Bouman et al.

(2004); AOC: Chan et al. (1992); MORB: Moriguti and Nakamura (1998); Tomascak et al. (2008); Nishio et al. (2002);

OIB: Kobayashi et al. (2004); IAB: Moriguti and Nakamura (1998); Tomascak et al. (2002); eclogites: Marschall et al.

(2007); xenoliths: Nishio et al. (2004); est mean mantle: Jagoutz et al., (1979) and Tomascak (2004).

Page 25: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

12

0 0.2 0.4 0.6 0.8 1 1.2 1.41

2

3

4

5

6

7

8

Li/Y

7Li

A

0

1

2

3

4

5

6

7

8

0 0.2 0.4 0.6 0.8 1 1.2 1.4

7Li

Li/Y

B

Figure 1.2 Li/Y ratio and 7Li in Arc Lavas

A plot of 7Li as a function of Li/Y ratio in (a) Izu arc basalts and (b) basalts from other Sunda and Aleutian arcs.

Lavas from the Izu arc display a trend of increasing 7Li with increasing Li/Y ratio, and show an inverse relationship

with depth to the slab (Benioff zone) suggestive of progressively decreasing amounts of fluid being mobilized during

subduction. The trend of increasing 7Li with increasing Li/Y ratio is not consistently observed in other island arcs.

Izu arc data from Moriguti and Nakamura (1998), Sunda and Aleutian arc data from Tomascak et al. (2002).

Page 26: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

13

-22.0

-20.0

-18.0

-16.0

-14.0

-12.0

-10.0

-8.0

-6.0

4 8 12 16 20

DLi (

m2 /s

)

10,000/T (K)

Si-crystal

albite & anorthite

cpx, Coogan et al. 2005

20040060080010001400T (oC)

Figure 1.3 Lithium Diffusion Coefficients

Plot of log DLi vs. 10,000/T (K) for lithium diffusion in geologically significant minerals. Lithium diffusion in a p-type Si-crystal data are from Pell (1960), feldspar data are from Giletti and Shanahan (1997), and cpx data are from Coogan et al. (2005)

Page 27: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

14

2 Lithium Partitioning and Isotopic Fractionation

2.1 Introduction

Lithium and lithium isotopes are increasingly used as tracers of surface inputs to the mantle

during subduction. With a strong affinity for fluids, an incompatible nature during mantle

melting and a high relative mass difference (~16 %) between the two stable isotopes, (6Li and 7Li), lithium has the potential to serve as a robust indicator of fluid-rock interaction in a variety

of geological settings. Enrichments in lithium isotopes are described as either:

(2) 10001LiLi/

LiLi/Li

std76

smp76

6

or 10001

LiLi/

LiLi/Li

std67

smp67

7

where ‘smp’ refers to the sample and ‘std’ refers to the standard, typically the NIST lithium

carbonate L-SVEC. The δ7Li notation is recommended by the International Union for Pure and

Applied Chemistry (Coplen et al., 1996) and will be used here.

Several reservoirs of lithium, which are isotopically distinct from the mantle and each other, are

present within Earth. Seawater has 0.18 ppm lithium with δ7Li of +32 ‰ (James and Palmer,

2000) and the continental crust contains an average of 35 + 11 ppm lithium with a δ7Li that

ranges from –5 to +5 ‰ (Teng et al., 2004). Lithium in fresh mid-ocean ridge basalts (MORB)

can range from 5 to 6 ppm, with a δ7Li of +1.5 ‰ to +5.6 ‰ (Moriguti and Nakamura, 1998;

Tomascak et al., 2008; Chan et al., 1992). Altered oceanic crust (AOC) has a significantly

greater concentration of lithium, >75 ppm lithium, and is isotopically heavier than pristine

MORB with a δ7Li of up to +14.2 ‰ in the most altered oceanic crust (Moriguti and Nakamura,

1998; Tomascak et al., 2008; Chan et al., 1992). The mantle is estimated to contain 1.6 ppm

lithium with an average δ7Li of +4 ‰ (Jagoutz et al., 1979; Moriguti and Nakamura, 1998;

Tomascak, 2004; Tomascak et al., 2008; Teng et al., 2004). However, studies of mantle xenoliths

suggest that reservoirs with variable δ7Li exist in the mantle (Seitz et al., 2004; Nishio et al.,

2004; Brooker at al., 2004; Lundstrom et al., 2005). Very low δ7Li (-11 ‰ to +5 ‰) values are

found in orogenic eclogites which are thought to reflect low temperature dehydration of the slab

Page 28: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

15

during subduction (Zack et al., 2003). Conversely some pyroxenites from the Zabargad

peridotite, which is considered to be a fragment of an exhumed mantle wedge, have high δ7Li

values (+8.4 ‰ to +11.8 ‰) suggesting that 7Li-rich fluids or melts derived from subducted

AOC are also transferred to the mantle (Brooker et al., 2004).

In an investigation of lithium isotopes from the Kilauea Iki lava lake, Tomascak et al. (1999)

demonstrated that neither partial melting nor low pressure differentiation results in significant (>

+2 ‰) variations in δ7Li. This has led to the interpretation that the variability evident in some

mantle-derived lavas is due to melting of a heterogeneous source. The Izu arc shows a trend with

the greatest lithium and 7Li enrichment occurring at the arc front where δ7Li in lavas ranges from

+7.6 ‰ to +1.1 ‰ (see Figure 1.2a) suggesting enrichment of the arc melt source by fluids

derived from the down going slab (Moriguti and Nakamura, 1998). A study of lithium in glass

inclusions from Hawaii showed the lithium isotopic composition to vary from –10.2 ‰ to +8.4

‰ (Kobayashi et al., 2004). Similarly, low δ7Li (-3.3 ‰ to +1.2 ‰) has been measured in glass

inclusions from the Iblean (Sicilian) Plateau tholeiites (Gurenko and Schmincke, 2002). A study

of δ7Li values in OIBs from Antarctica did not show any significant deviations from MORB

values (Ryan and Kyle, 2004); however, a study of MORB lavas from different ridge systems

found a 5 ‰ heterogeneity in the samples, and no significant correlation of δ7Li with major

elements, trace elements or radiogenic isotopes (a slight apparent correlation with Cl/K was

observed; Tomascak et al., 2008). The fact that both OIB and MORB can sometimes display a

range of values has prompted many researchers to suggest that variable amounts of recycled

material with modified δ7Li is transported into the mantle sources of these lavas (Tomascak et

al., 2008). But the origin and scale of these mantle heterogeneities are not well defined.

The extent of the modification of the down going slab by fluid-mineral exchange during

subduction remains ambiguous, as is the extent of modification of the slab derived fluid during

transport though the mantle to the arc source. Previous experimental work has found lithium to

be moderately incompatible in clinopyroxene relative to fluid or melt with DLicpx/fluid increasing

with increasing Ca/Al ratio (Hart and Dunn, 1993; Brenan et al., 1998a; Blundy et al., 1998;

Blundy and Dalton, 2000; Hill et al., 2000; Bennett et al., 2004). Lithium was also found to be

moderately incompatible in olivine relative to melt (Brenan et al., 1998a; Brenan et al., 1998b;

Taura et al., 1998; Zanetti et al., 2004) and fluid (Blundy and Dalton, 2000). Coogan et al. (2005)

measured lithium partitioning between plagioclase and clinopyroxene and found DLiplag/cpx

Page 29: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

16

increasing with increasing temperature (900oC to 1200oC). In this case, partitioning was not

correlated with plagioclase or clinopyroxene major element composition.

Little experimental work has been done to examine lithium isotope fractionation at pressures and

temperatures relevant to crustal and mantle processes. Comparisons of δ7Li from metasomatized

and pristine peridotite xenoliths suggest that some olivine – clinopyroxene fractionation, (up to

3.5 ‰ enrichment in 7Li), may occur at mantle temperatures (950oC; Seitz et al., 2004). Fluid-

mineral partitioning will also fractionate lithium isotopes as documented in a single study which

measured an enrichment of 7Li in fluids relative to a Li-pyroxene (spodumene) by +3.5 ‰ at 500 oC to ~ +1.0 ‰ at 900oC and 2.0 GPa (Wunder et al., 2006).

To date, a systematic investigation of the degree of isotopic fractionation and the extent of

partitioning that occurs during mantle processes has been lacking. Both the isotopic

fractionation, Li, and the lithium partitioning between major mantle phases need to be known to

determine the extent to which a slab signal can propagate to the IAB source. This study presents

lithium partitioning and isotopic fractionation measurements between fluids and common rock

forming minerals. With this information, more accurate models can be developed to constrain the

origins of lithium anomalies in the mantle.

2.2 Methods

This study measured the partitioning of lithium between aqueous fluids and clinopyroxene,

olivine and plagioclase at pressures and temperatures corresponding to lower crustal and upper

mantle conditions (800oC to 1200oC; 1 GPa). Additional experiments were done to measure

olivine – clinopyroxene isotopic fractionation at similar conditions. Starting materials were

natural single crystals of: olivine (Fo82) from San Carlos, Arizona; plagioclase (bytownite) from

Crystal Bay, Minnesota; clinopyroxene (diopside) from Dekalb, New York; and plagioclase

(albite) from Mont St. Hilare, Quebec. Table 2.1 gives the composition of the starting materials.

In all cases, the mineral samples were first crushed to 1-3 mm grain size, after which grains free

of inclusions and alteration were hand picked and cleaned in dilute HNO3 and rinsed with ultra-

pure water in an ultrasonic cleaner. For the olivine and plagioclase experiments the minerals

were ground to a fine powder under ethanol with an additional SiO2 + Al2O3 (1:1) mixture (~3

Page 30: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

17

wt. % of total) added to approximate natural mantle fluid compositions (Brenan et al., 1998b;

Holloway, 1971). Experiments containing olivine were not buffered for Fe loss to the noble

metal capsule; therefore, run product compositions are shifted to more magnesium rich

compositions (from Fo82 to Fo97 – 99). Experiments with bytownite as a starting mineral

composition were not buffered for Na2O loss to the fluid and as a result run product

compositions are shifted from bytownite (An75) to anorthite (An98-99). A few plagioclase

experiments contained additional albite to stabilize more sodic compositions of plagioclase. With

one exception all experiments with Dekalb diopside as starting material had ~3 wt. % SiO2 added

(no Al2O3) since clinopyroxene dissolution buffers the aluminum content of the fluid. To

minimize compositional zoning, a large fluid to solid ratio (4:1 by mass) was utilized for all

experiments. One clinopyroxene experiment was carried out with the addition of 3 wt. % albite

(+3 wt. % SiO2) to encourage compositional zoning with respect to the aluminum content of the

clinopyroxene.

A series of mineral pair experiments were run with clinopyroxene and olivine to constrain their

inter-mineral isotopic fractionation. These experiments used the same starting materials prepared

as above and mixed 80:20 cpx-olivine by mass.

Isotopically labeled solutions were made from ultra pure water with lithium added as either

Li2CO3 (LSVEC, SRM#8545) or some combination of LSVEC and a 6Li spike. For each

experiment a sample of either clinopyroxene, olivine or plagioclase powder (+/- SiO2, Al2O3 or

albite) and lithium bearing solution were added to a large volume Ni capsule with a Pt insert

(Ayers et al., 1992). To promote crystal dissolution and re-precipitation, the bottom of the

capsule was centered in the hotspot of the furnace. The temperature gradient over the length of

the capsule is less than 10oC (Ayers et al., 1992).

The experiments were conducted in an end-loaded piston–cylinder apparatus (Boyd and England,

1960) using a 1.9 cm bore pressure vessel, employing a cylindrical graphite heater and pressure

cells consisting of crushable MgO, Pyrex and NaCl. Samples were initially cold pressurized to

~0.5 GPa and then heated to 300oC to generate sufficient internal pressure to prevent capsule

deformation (Brenan et al. 1995). Temperature and pressure were then increased simultaneously

with the maximum pressure being achieved by the time the sample reached 600oC. Temperature

was monitored with W26% Re-W5% Re thermocouples uncorrected for the effect of pressure on

Page 31: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

18

EMF. Experiments were run for 72 to 144 hours and quenched by cutting power to the sample

heater which resulted in temperatures dropping to < 300oC in 20 seconds. The capsules were then

recovered, punctured and dried. Fluid masses determined by weight loss were usually > 70 % of

the initial fluid mass; however, during the course of the experiment the capsule material became

work hardened; consequently, an undetermined amount of capsule material was sheared off

during puncturing. As a result, the fluid masses used in the mass balance calculations are the

initial fluid masses. In the case where a watertight seal was not maintained throughout an

experiment, a drop in pressure and a collapsed capsule would result.

One additional experiment was carried out in a cold seal vessel at 0.2 GPa and 800oC. In this

case the sample powder and lithium solution were loaded into a 5 mm O.D. Au capsule. The

capsule was then crimped, weighed, sealed by arc welding and reweighed to check for fluid loss.

The sample was loaded into a vertically mounted pressure vessel, first pressured to 0.2 GPa and

then externally heated. Temperatures were monitored with an internal type K thermocouple.

Experiments were quenched by removing the furnace and cooling the pressure vessel with

compressed air, which resulted in temperatures dropping to < 300oC in ~3 minutes. Table 2.2

provides a summary of experimental conditions.

2.3 Analytical Techniques

2.3.1 Major Element Analyses

Samples of starting material and splits of run products were mounted in epoxy, ground, polished

to 0.3 m and carbon coated. The major element compositions of the starting materials and run

product clinopyroxene, olivine and plagioclase were then obtained using the University of

Toronto’s Cameca SX50 Electron Probe X-ray Microanalyzer (EPMA). An accelerating voltage

of 15 kV and a focused 20 nA beam was used for all samples. The standards were albite for Na,

anorthite for Al, diopside for Ca, Mg, Si, basalt for Fe, and bustamite, (Mn,Ca)3Si3O9, for Mn.

X-ray intensities were converted to concentrations using modified ZAF or Phi-Rho-Z correction

schemes. The reported errors are the 1σ variations of (n) analyses.

Page 32: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

19

2.3.2 Lithium Analyses

2.3.2.1 MC-ICPMS

Bulk lithium isotopic composition and element concentrations were determined for run product

clinopyroxene using a Multi-Collector Inductively Coupled Plasma Mass Spectrometry (MC-

ICPMS) at the University of Maryland. Samples were first rinsed in ultra-pure water to remove

any water-soluble Li-bearing residue and then analyzed following the method described in Teng

et al. (2004). The samples, 4-10 mg of run-product clinopyroxene, were digested in a mixture of

HF and HNO3 and dissolved in a 4M HCl solution. The lithium from the samples was then

separated from the dissolved matrix by thrice processing the solutions in cation exchange

columns. The solutions were then introduced to the Nu-Plasma MC-ICP-MS in a 2% HNO3

solution, and isotopic compositions were obtained by measurement of 7Li and 6Li simultaneously

on two high and low mass Faraday cups. Each sample analysis was bracketed by measurement of

the L-SVEC standard. The isotopic values are reported as δ7Li (equation 2.1) in which the

lithium isotopic standard is NIST L-SVEC Li2CO3. The 2σ precision of each analysis is ±1 ‰.

Table 2.3 lists measurements of standards and reference materials.

2.3.2.2 SIMS

In situ analyses of the lithium abundance and isotopic composition of the starting materials, run

product clinopyroxene, olivine, and plagioclase were obtained using the Cameca IMS 3f ion

microprobe at Lawrence Livermore National Laboratory. Secondary ions were generated by

bombardment with a 5-12 nA negatively charged 16O primary beam, accelerated through –12.5

kV and focused to ~20 μm. The positive secondary ions were accelerated through 4.5 kV. 6Li

and 7Li were measured with a mass resolving power of 1011, and no energy offset was applied.

The background, mass 5.8, 6Li and 7Li were counted on an electron multiplier for 2 s, 10 s and 2

s respectively over 120-400 counting cycles, depending on count rate. Figure 2.1 shows the

‘uncorrected’ δ7Li values measured by SIMS of the internal reference materials; Dekalb diopside

(measured in this study by MCICP-MS), San Carlos olivine (Magna et al., 2006), and

clinopyroxene from experiment NCDL6 (analysed by MC-ICPMS), plotted against the δ7Li

values measured by MC-ICPMS. The δ7Li values measured by MC-ICPMS of the internal

reference materials range from -3 ‰ to +10 ‰ whereas the corresponding ‘uncorrected’ δ7Li

values measured by SIMS range from +15 ‰ to +30 ‰. The discrepancy between values is due

Page 33: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

20

to mass fractionations that are the result of both instrumental parameters and matrix effects

(Decitre et al., 2002, Tomascak, 2004). It is important to note that all the values plot on a single

line with a slope of ~1, suggesting that any matrix effect on the lithium instrumental isotopic

fractionation is still within the 2σ precision of the analysis (± 4‰). The 7Li/6Li ratios can be

corrected for this fractionation using the instrumental correction factor, ∆i (Decitre et al., 2002);

(3) ∆i = δ7LiSIMS – δ7LiMC-ICPMS

Because all of the fractionation measured is relative to the same starting material, the 7Li/6Li

ratio of Dekalb diopside was used as the internal reference material to determine ∆i. As a check

on the calibration of this correction factor, ∆i, was determined using San Carlos olivine, ∆i = -22

‰, and was found to be within 2σ (± 4 ‰) error of that for the Dekalb diopside, ∆i = -20 ‰.

However, the concentration of lithium in Dekalb diopside and the NCDL6 experiment are both

greater than that of San Carlos olivine (i.e. 8.6 ppm and 67 ppm vs. 2.5 ppm), therefore the

Dekalb diopside and NCDL6 were primarily used as internal reference materials.

2.3.2.3 LA-ICPMS

In situ analyses of lithium abundances in clinopyroxene, olivine, and plagioclase were also

determined by laser ablation inductively coupled plasma mass spectrometry (LA–ICP–MS) at

the University of Toronto. The system employs a frequency quintupled Nd:YAG laser operating

at 213 nm, coupled to a VG PQExcell quadrupole ICP-MS. The laser was operated at 10 Hz,

with He flushing the ablation cell to enhance sensitivity (Eggins et al. 1998), and produced spot

sizes ~50 μm in diameter and ~ 50 μm deep.

The torch position and lens settings were adjusted prior to each analytical session to optimize the

signal intensity while ablating NIST 610 with a spot size of approximately 75 μm and a laser

beam energy of less than 3 mJ, so that the sensitivity to 7Li was maximized. Data were collected

as time-resolved spectra with background levels determined by counting for 20 s prior to the 60 s

of sampling by laser ablation. Analyses were collected in blocks of >20, with the first and last

two spectra acquired on standard reference materials. Count rates were collected and exported as

CSV (comma-delimited values) files by “ThermoElectron Plasmalab” (TRA) software. All

subsequent data reduction was performed off-line using the GLITTER 5.3 software package,

supplied by Macquarie Research, Ltd. Ablation yields were corrected by referencing to the

Page 34: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

21

known concentration of 43Ca or 55Mn as determined previously by electron microprobe analyses.

Lithium concentrations in clinopyroxene, olivine and anorthite were quantified using the “in

house” standard Kunlun diopside, which contains 42.6 ppm lithium. This standard was used

routinely because it generated a lower lithium background over the course of the analysis

compared to that produced by NIST 610. Kunlun diopside was, in turn, characterized using NIST

610 silicate glass, which contains 470.5 ppm lithium (Pearce et al., 1997). Table 2.3 compares

the accepted values for several standard and reference materials (fused into glass in sealed Pt

capsules at 1 GPa and 1200oC) which were also measured by LA-ICPMS using NIST 610 as a

standard. The precision for concentration measurements is better than ± 10 %. Figure 2.2 shows

lithium abundance of the cross-referenced samples plotted against lithium abundance determined

by LA-ICP-MS.

2.4 Results

2.4.1 Major Element Chemistry

Figure 2.3 shows the size and morphology of run product clinopyroxene, olivine and plagioclase.

The run products display considerable coarsening compared to starting materials, and well-

developed crystal faces. Crystals grew as aggregates of grains nucleating on the lid and/or upper

walls of the capsule. Table 2.4 lists the major element composition of the clinopyroxene, olivine

and plagioclase run products produced in this study. The minimum detection limits are less than

the lowest value cited for each element and the number in parentheses refers to 1σ of the

standard deviation for n analyses and reflects the degree of sample heterogeneity.

2.4.2 Mineral – Fluid Lithium Partitioning

Table 2.5 lists the lithium contents of experimentally produced crystals as analyzed by LA-

ICPMS as well as the calculated distribution coefficient (DLi) values for mineral/fluid

partitioning or where applicable the D-values for mineral/melt partitioning. Mineral/fluid

distribution coefficients (Dmin/fluid) were determined from final mass balance calculations:

(4) CLitotal = Cinitial

min Xinitial

min + Cinitialfluid X

initialfluid

(5) CLitotal = Cfinal

min Xfinal

min + Cfinalfluid X

finalfluid

Page 35: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

22

Where Cmin is the lithium concentration of the mineral in ppm, Cfluid is the lithium concentration

of the fluid in ppm and Xmin is the mass fraction of the mineral, and Xfluid is the mass fraction of

the fluid. The initial concentration of fluid, Cinitialfluid, and the initial mass fraction of fluid,

Xinitialfluid, is assumed to be equal to the final concentration and mass fraction, Cfinal

fluid and

Xfinalfluid. Mineral solubility is also assumed to be negligible such that Xfinal

min is equal to

Xinitialmin.

(6) CLitotal/ C

finalmin = Xfinal

min + (Cfinalfluid X

finalfluid)/ C

finalmin

The Nernst distribution coefficient is defined as,

(7) Dmin/fluid = Cfinalmin/ C

finalfluid

Then,

(8) [CLitotal/ C

finalmin] – Xfinal

min = 1/ Dmin/fluid Xfinalfluid

(9) Dmin/fluid = Xfinalfluid/( [C

Litotal/ C

finalmin] – Xfinal

min)

The fluid-mineral ratio was such that the volume of the solution would serve as an infinite

reservoir and the lithium concentration would remain constant throughout the experiment.

Mineral/mineral distribution coefficients (Dmin/min) were calculated from:

(10) Dmin/min = CfinalminA/CfinalminB

where Cfinal is the lithium concentration of mineral A or mineral B in ppm.

The range of lithium content in the run products was 5 ppm to 7 ppm in the clinopyroxene, 13

ppm to 466 ppm in the olivines, and 20 ppm to 70 ppm in the plagioclase. Mineral – fluid

equilibrium was assessed in terms of run-product homogeneity. The concentration of lithium

within a single experiment typically varied by 14 % relative to the mean concentration from

grain to grain and in a few cases, lithium contents varied as much as 30 %. The 50 μm spot size

and 80-second ablation time often meant that the laser analysis consumed the whole grain. The

run product olivine grains typically had diameters < 75 m, the clinopyroxene > 75 m, and the

anorthite grains > 100 m. Figure 2.4 shows the time resolved spectra for 7Li and 43Ca, measured

by LA-ICPMS, for clinopyroxene produced in the lowest temperature partitioning experiment

Page 36: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

23

(800oC). Time resolved spectra for all experiments, with a few exceptions thought to be the

result of fluid inclusions, display level spectra, which are interpreted as homogenous equilibrated

grains. The slight sloping of the spectra is due to signal decay as a result of the geometry of the

ablated pit and does not reflect sample heterogeneity as both the 7Li and 43Ca signals remain

consistent with respect to each other.

An additional test for equilibration between mineral samples and fluids was attempted by

measuring partition coefficients in reversal experiments where crystals previously equilibrated

with solution ‘A’ were re-equilibrated with solution ‘B’, containing a lower concentration of

lithium and a differing isotopic composition. These reversal experiments confirm isotopic

equilibrium; however, changes in mineral assemblages (zoisite in anorthite reversal, Mg-

hydroxides in olivine reversal, monticellite in olivine + clinopyroxene reversal, undetermined

phase in clinopyroxene reversal) make the mass balance calculations, used to determine the

distribution coefficients, impossible to resolve. A simple calculation using the diffusion rates of

lithium in clinopyroxene measured by Coogan et al. (2005) determines that at 800oC lithium

diffusion should penetrate to a distance of 150 μm in 72 hrs, which is greater than the radius of

the largest run product crystal (Figure 2.3). An experiment was attempted at 0.2 GPa and 800oC

to ascertain the effect of pressure on the partitioning behavior of lithium, with no significant

effect of decreasing pressure noted.

Lithium values were always shifted with respect to the starting material. Any change in the

lithium composition of the fluid due to uptake by the growing mineral is insignificant compared

to the total amount of lithium in the fluid even when the DLimin/fluid is greater than one. Generally,

the lithium mineral/fluid distribution coefficients decrease in the order: olivine (2.51 – 0.17),

plagioclase (0.32 – 0.090) and clinopyroxene (0.32 – 0.07).

2.4.2.1 Clinopyroxene

Run products produced in these experiments typically contained no phases other than

clinopyroxene. One experiment, NCDL3, consisted of approximately 50 % (by volume) olivine

crystals and 50 % clinopyroxene crystals, which was most likely a result of magnesium

contamination from the ceramic pressure cell during sample assembly or loading. Electron

microprobe traverses across individual clinopyroxene grains shows that the resulting crystals are

homogenous with respect to major elements. As shown in Table 2.4, these clinopyroxene grains

Page 37: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

24

have much lower Al2O3 (~0.5 wt. % to ~0.2 wt. %) and FeO (~0.9 wt. % to ~0.4 wt. %) contents

than natural upper mantle clinopyroxene (~3 wt. % Al2O3, ~2 wt. % FeO; Lundstrom et al.,

2005). However, their MgO concentrations (~17 wt. % to ~21 wt. %) and Na2O concentrations

(~0.02 wt. % to ~0.30 wt. %) are similar to those in clinopyroxene from mantle xenoliths (~17

wt. % MgO and ~0.30 wt. % Na2O; Lundstrom et al., 2005). The FeO content of all the

experimentally produced clinopyroxene is low (< 1 wt. %) due to loss of Fe to the platinum

capsule. Additionally because the starting clinopyroxene material contained very low (below

detection limits) abundances of chromium and titanium these elements are absent in the run

products.

The range of DLi (0.07 to 0.613) for clinopyroxene/fluid measured in this study is similar to the

clinopyroxene/fluid DLi (0.08 to 0.25) measured by Brenan et al. (1998b) and the

clinopyroxene/silicate melt DLi (0.14 to 0.27) measured by Brenan et al. (1998a).

Lithium partitioning between clinopyroxene and fluid decreases from 0.32 to 0.09 with

increasing temperature from 800oC to 1100oC at 1 GPa. The temperature dependence of lithium

partitioning between clinopyroxene and hydrous fluids can be demonstrated on a plot of ln

DLicpx/fluid versus 1000/T, (Figure 2.5). A linear regression of the data yields the relationship:

(11) ln DLicpx/fluid = -7.3 (+0.5) + 7.0 (+0.7) * 1000/T (R2=0.98)

where T is temperature in Kelvins.

2.4.2.2 Olivine

Olivine partitioning experiments occasionally produced some oxide grains and in the reversal

experiment (NCOR), an unidentified magnesian phase. All are likely due to incongruent

dissolution of olivine. Electron microprobe traverses of individual crystals show the run product

olivines to be homogenous with respect to major element chemistry. Iron loss to the platinum

capsule resulted in considerably more magnesian (Fo# 98 to 99) olivines than those naturally

occurring in the mantle. Reversal experiments were attempted; however, re-equilibrating run

product material, from experiment NCOL2, caused Fe and water soluble elements (i.e. Na, Ca,

etc.) to become further depleted, such that the final solid composition was no longer in the

stability field of olivine. The same effect occurred with the plagioclase reversal run, NCAR,

which resulted in the crystallization of zoisite.

Page 38: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

25

The range of DLi (0.17 to 0.57) measured in this study for olivine/fluid is similar to the

olivine/silicate melt DLi (0.13 to 0.35) measured by Brenan et al. (1998a) with the exception of

the experiment with added albite (OlAb10) which produced a value of 1.34. The temperature

dependence of lithium partitioning between olivine and hydrous fluids can be demonstrated on a

plot of ln DLiol/fluid versus 1000/T, (Figure 2.6). A linear regression of the data (excluding the

reversal, NCOLR) yields the relationship:

(12) ln DLiol/fluid = -6.0 (+2) + 6.5 (+2) * 1000/T (R2=0.82)

where T is temperature in Kelvins.

2.4.2.3 Plagioclase

The starting material for the anorthite experiments was bytownite (An76), but the experiments

that were not buffered for Na loss to the solution resulted in run product compositions very close

to end member anorthite (An96-99). The reversal run, NCAR, which consisted of re-equlibrating

material from NCA3, resulted in zoisite + unidentified Al-rich phase. At 1000oC run products

consisted of 30 % melt, 70 % anorthite crystals, at 900oC the amount of melt was negligible, and

runs at 800 oC were melt free. Electron microprobe traverses of these grains show that they are

homogenous with respect to major elements. In two of the experiments, AnAb10 and AnAb20,

the Na content of the fluid was buffered by addition of albite. This resulted in homogenous

anorthite crystals with only slightly more sodic compositions (see Table 2.4).

The range of DLi for anorthite – fluid partitioning measured in this study is 0.09 to 0.32. The

range of DSr and DBa for plagioclase with similar An content is 1 -3 and 0.1 to 0.2 respectively

(Blundy and Wood, 1991). Similar to the results from a study of Sr and Ba partitioning in

plagioclase (Bludy and Wood, 1991), the data show a linear relationship with a negative slope on

a plot of ln DLi versus XAn suggesting that lithium is more compatible in albite than in anorthite

(Figure 2.7). Linear regression of the six partitioning experiments yields the relationship, in

Jmol-1:

(13) RTlnDLi = 162,000 (+26,000)– 188,000 (+28,000) (XAn) (R2=0.96)

Page 39: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

26

where R is the gas constant, T is temperature in Kelvins, and XAn is the anorthite content of the

plagioclase. Following Blundy and Wood (1991), RTlnDLi is used rather than lnDLi to minimize

the effect of temperature in the linear regression.

2.4.3 Olivine-Clinopyroxene Pair Experiments

A series of experiments were done with olivine + clinopyroxene + fluid to investigate inter-

mineral partitioning and isotopic fractionation. The run products from these experiments

typically consisted of coarse-grained intergrowths of olivine and clinopyroxene and in the case

of experiments 2m-lo and Yb-1, molybdenum oxide (from outer capsule material) and ytterbium

oxide, respectively. Due to the uncertainties in the mass balance of each phase after

equilibration, mineral-fluid D values have not been calculated for these runs. Experiment 2m-1 at

900oC contained only clinopyroxene; however, NCDL3 at 900oC stabilized both clinopyroxene

and olivine. Experiment 2m-hi, investigating partitioning at log fO2 of –5, stabilized only

enstatite. The reversal experiment, 2m-R, resulted in olivine + clinopyroxene + monticellite

(CaMgSiO4; see NCOR and NCAR above).

In all the experiments run at Ni-NiO, with the exception of the reversal experiment, 2m-R,

olivine preferentially incorporated lithium relative to clinopyroxene. The range of DLi for

olivine/clinopyroxene measured in this study is 1.2 to 6.7. Figure 2.8 shows how

olivine/clinopyroxene partitioning increases with increasing temperature between 800oC and

900oC. The lithium content of the run product olivine in the 800oC experiment, 2m-2, is 101 ± 59

ppm; this large standard deviation suggests that the run product olivine may be heterogeneous

with respect to lithium and may have formed lithium rich fluid inclusions. The reversal

experiment, which equilibrated a split of the 2m-2 experiment at 800oC with a solution of 96

ppm lithium, did result in a lower lithium content in the run products than in the starting

material. This experiment is complicated by the formation of monticellite, which contains 39

ppm lithium, more than either the forsteritic olivine or clinopyroxene (11 ppm and 26 ppm,

respectively). The constant Dol/cpxLi versus temperature further suggests that

olivine/clinopyroxene partitioning of lithium is independent of temperature. Also shown in

Figure 2.8 is the ratio of DLiolivine/fluid/ DLi

cpx/fluid calculated from single-phase experiments, which

at 1000oC is the same as that determined from the two-phase experiment.

Page 40: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

27

2.4.3.1 Experiments at variable fO2

Three experiments were conducted to investigate the effect of oxygen fugacity on the lithium

partitioning behavior between olivine and clinopyroxene. Experiment, 2m-hi, which was run at

1000oC in a Pt + Re lined nickel capsule to generate an oxygen fugacity of log fO2 of –5,

stabilized enstatite. The enstatite had a lithium content of 5.28 ppm and resulted in a DLient/fluid of

0.02, which is much lower than the DLimin/fluid for either clinopyroxene (0.17) or olivine (0.48 at

Fo#98, or 0.17 at Fo#63) at the same temperature. Experiment 2m-lo was run at 1000oC in a Pt

lined Mo capsule to generate an oxygen fugacity of log fO2 of –15 and resulted in olivine with 30

ppm lithium and clinopyroxene with 42 ppm lithium which results in a Dol/cpxLi of 0.7. The

oxygen fugacity generated by the Ni lined Pt capsule at 1000oC is log fO2 = -10.3 and resulted in

olivine with 52 ppm lithium and clinopyroxene with 13 ppm lithium which results in a Dol/cpxLi of

4.0.

Higher oxygen fugacity results in higher abundances of Fe3+ relative to Fe2+. Incorporating a

higher proportion of Fe3+ into the olivine structure should generate more charge balancing

opportunities for the Li+ ion in the crystal structure; resulting in a coupled substitution where:

(14) Li1+ M1 + X3+

M2 (Mg2+, Fe2+)M1 + (Mg2+, Fe2+)M2

This is consistent with the results between the NNO run 2m-3, (higher fO2 and Dol/cpxLi = 4.0) and

the 2m-lo run (lower fO2 and Dol/cpxLi = 0.7).

2.4.3.2 Experiments with added REE

Experiment Yb-1, carried out at 1000oC with 0.25 mg of Yb2O3, was an attempt to determine the

effect of rare earth elements (REE) on the relative partitioning of lithium between olivine and

clinopyroxene. This experiment resulted in olivine with 175 ppm lithium and clinopyroxene with

17 ppm lithium, and resulted in a Dol/cpxLi of 10, which is significantly higher than the Dol/cpx

Li of

4.0 at 1000oC that results from no addition of REE. The increased partitioning of lithium into the

olivine is most likely a result of a coupled substitution with Yb3+, analogous to that in Equation

14 for Fe3+.

Page 41: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

28

2.4.4 Lithium Isotope Fractionation

2.4.4.1 Clinopyroxene- fluid

Table 2.6 gives the isotopic ratios of the starting materials and run products from the isotopic

fractionation experiments as well as the calculated ∆7Licpx-fluid, where;

(15) ∆7Licpx-fluid = δ7Licpx (‰)- δ7Lifluid (‰)

With the exception of the reversal runs, all experiments had Dekalb diopside (δ7Li of +9.7 ‰) as

starting material. Two sets of solutions were used: two L-SVEC based solutions, (A) with 243

ppm lithium and δ7Li of 0 ‰ and (B) 96 ppm lithium and δ7Li -2.7 ‰ for the reversal and two 6Li doped solutions one (C) with 306 ppm lithium and δ7Li of -88.4 ‰ and (D) 180 ppm lithium

and δ7Li of –46.1 ‰ for the reversal. In all the experiments, the crystals were preferentially

enriched in 6Li with respect to the fluid. Duplicate experiments at 900oC and run times of 72 hrs

and 142 hrs had ∆7Licpx-fluid within + 2 ‰, which is within the precision of the analysis,

indicating that run times were sufficient for isotopic equilibrium. As a further test, a reversal

experiment, Ldi-12, was conducted in which a split of sample Ldi-10 with δ7Li of –90.9 ‰ was

reacted with solution B (δ7Li –46.1 ‰). The run product clinopyroxene from Ldi-12 had a δ7Li

of –49.5 ‰, an enrichment of 45 ‰ in the heavier isotope from its initial value, and resulted in a

∆7Licpx-fluid of –3.4 ‰, which is within the precision of the other experiments. The lithium

isotopic fractionation at high temperatures (T>900oC) is +2.5 ‰, which is just at the limit of the

analytical precision of this study. The range of ∆7Licpx-fluid measured in this study is from –0.3 ‰

to –3.4 ‰, and follows the trend of ∆7Licpx-fluid decreasing with increasing temperature. When the

run products were not rinsed prior to sample digestion and analysis, the data produced scattered

results, most likely due to the precipitation of lithium as the remainder of the solution was dried

down after sample recovery.

Figure 2.10 is a plot of the ∆7Licpx-fluid (‰) from this study, as well as the ∆7Lispodumene-fluid (‰)

from Wunder et al (2006), the ∆7Libasalt-fluid (‰) measured between basalt and seawater at 350oC

(Chan et al., 1993) and 2oC (Chan et al., 1992) versus 1000/T (K). Despite the fact that Wunder

et al. (2006) used both OH- and Cl-bearing fluids and this study was chlorine free, with the

lithium introduced as Li2CO3, all of the data plot on the same regression line empirically

Page 42: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

29

determined by Wunder et al. (2006). There appears to be no difference in fractionation behavior

with pressure (seafloor to 2 GPa) or complexing anion.

It should be noted that measurements of lithium isotopic fractionation in the quartz-muscovite-

fluid system, from 400-500oC, found the quartz and the mica to be preferentially enriched in 7Li

(Lynton et al., 2005), which is inconsistent with this study.

2.4.4.2 Olivine-Clinopyroxene Lithium Isotope Fractionation

Table 2.6 gives the isotopic ratios of the starting materials and run products from the isotopic

fractionation experiments as well as the calculated ∆7Liol-cpx, where;

(16) ∆7Liol-cpx = δ7Liol (‰) - δ7Licpx (‰)

All experiments, with the exception of the reversal runs, had 82 wt. % Dekalb diopside (δ7Li of

+9.7 ‰) + 18 wt. % San Carlos olivine (δ7Li of +1.0 ‰) as starting material. Two L-SVEC

based solutions were used: (A) with 243 ppm lithium and δ7Li of 0 ‰ and (B) 96 ppm lithium

and δ7Li -2.7 ‰ for the reversal. In all the experiments, except the reversal run, the isotopic

composition of the olivine grains did not shift significantly; whereas, the isotopic composition of

the clinopyroxene became as much as 15 ‰ lighter (i.e. from δ7Li of +9.7 ‰ to ~-5 ‰). ∆7Liol-

cpx measured in this study is ~5 ±5 ‰, which is not resolvable with the precision of this study.

2.5 Discussion

2.5.1 Controls on Partitioning

2.5.1.1 Clinopyroxene

Clinopyroxene has two sites, M1 and M2; M1 has six-fold coordination with respect to oxygen

and M2 has eight-fold coordination. Given that the M1 site is slightly smaller than M2 (optimal

site radius (ro) of ~0.7 Å versus M2 ro of ~1.1 Å) and has lower defect energies for univalent

cations, it is likely that the primary site for lithium in the clinopyroxene structure is M1 where it

can exchange for Mg2+ (Purton et al., 1997). Such an exchange should be coupled by a trivalent

cation such as Al in jadeite (NaAlSi2O6) component or spodumene (LiAlSi2O6) or Fe3+ in a

aegirine-like molecule, NaFe3+Si2O6.

(17) Li1+M1 + X3+

M2 (Mg2+)M1 + (Ca2+) M2

Page 43: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

30

Due to Fe loss to Pt capsule the total iron content of the run product clinopyroxene is low, less

than 2 wt. % and does not vary systematically with temperature. A Mössbauer study of natural

diopside crystals by De Grave (2003) found all to contain some component of Fe3+ in either M1

or M2 sites. Furthermore, increasing Fe3+ at the expense of Fe2+ also serves to increase the

Mg/(Mg+Fe2+) ratio (Luth and Canil, 1993) which should then increase the availability of sites

for lithium exchange. Experiments examining the effect of fO2 on lithium diffusion in

clinopyroxene appear to confirm this; lithium diffusion in clinopyroxene appears to increase with

decreasing oxygen fugacity (Caciagli, Chapter 3).

Previous work has shown lithium partitioning to increase slightly with increasing Al/Si ratio

(Brenan et al., 1998b). A compilation of the data collected in this study and other experimental

data displays a similar trend. In Figure 2.5, the partition coefficients determined from cpx/fluid

experiments in this study, and Brenan et al. (1998b), and the cpx/melt experiments of Hart and

Dunn (1993), Blundy (1998), Brenan et al. (1998a), and Blundy and Dalton (2000) are plotted as

a function of temperature and the data points are labeled with the wt. % Al2O3 content of the run

product clinopyroxene. It is important to note that the alumina contents of run product

clinopyroxene in this study are approximately constant. Generally, lithium-partitioning at a given

temperature increases with increasing Al2O3 content of the run product clinopyroxene, regardless

if the partitioning measured is min/melt or min/fluid. For example, the run product

clinopyroxene from the cpx/melt experiment of Blundy and Dalton (2000) at 1375oC (and 0.8

GPa) has a similarly low alumina content as the clinopyroxene from this study, and plots on the

regression line determined in this study. As the alumina content of the clinopyroxene increases,

lithium partitioning increases and the data fall above the regression line. Other elements, such as

Fe and Na, appear to influence the lithium partitioning as well. For example, the experiment of

Brenan et al. (1998b) labeled 1.5* was conducted at 900oC with 0.5 M aq NaCl, and has run

product clinopyroxene with 1.5 wt. % Al2O3, but resulted in a lower cpx/fluid partition

coefficient than measured in another 900oC experiment from that study containing 0.2 wt. %

Al2O3 and the 900oC experiment containing 0.3 wt. % Al2O3 from this study.

This relationship is consistent with Equation 17 where lithium substitution is coupled with Al3+.

The correlation that is observed here may exist because the total Al2O3 content is correlated with,

or somehow serving as an indicator of the amount of Al3+.

Page 44: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

31

2.5.1.2 Olivine

The solution energies calculated by Purton et al. (1997) for forsterite demonstrate that the lowest

energy pairing is a 3+ cation in the M2 site coupled with Li+ in the M1 site. Previous studies have

shown lithium partitioning between olivine and silicate melt to be coupled with Al3+ (Suzuki and

Akaogi, 1995; Taura et al., 1998), suggesting the following mechanism.

(18) Li1+M1 + X3+

M2 (Mg2+, Fe2+)M1 + (Mg2+, Fe2+) M2

Comparison with other experimental data suggests that Fe may be affecting the partitioning of

lithium into olivine. In Figure 2.6, the partition coefficients measured from ol/fluid experiments

of this study, and the ol/melt experiments of Brenan et al. (1998a), Taura et al. (1998), Blundy

and Dalton (2000) and Zanetti et al. (2004) are plotted as a function of temperature and the data

points are labeled with the wt. % FeO in the run product olivine. The olivine-melt partitioning

experiments with low or no FeO plot on the same regression line as those determined in this

study. Experiments with run product olivine containing 8-10 wt. % FeO have higher lithium

partition coefficients than experiments with lower FeO contents in run product olivine, at a given

temperature.

Assuming that at a given fO2, and temperature the total Fe3+ content of olivine scales with total

Fe content, the trend of increasing lithium partitioning in olivine with increasing FeO, suggests

that lithium may be coupling with Fe3+ as substitution mechanism (see Equation 18). The

experiment of Zanetti et al. (2004), labeled 19*, further supports this suggestion. The Zanetti et

al. (2004) experiment was conducted at an fO2 equivalent to QFM-2, which is almost four orders

of magnitude more reducing than the experimental conditions of this study (NNO), and resulted

in a much lower lithium partition coefficient, despite its high (19 wt. %) FeO content. More

reducing experimental conditions in the Zanetti et al. (2004) experiment would result in lower

Fe3+ contents, and therefore less favorable conditions for Li+ substitution, than in the experiments

from this study, despite the fact that the olivine has a high FeO content.

2.5.1.3 Plagioclase

The primary control on lithium partitioning between plagioclase and hydrous fluids is the

composition of the feldspar (Figure 2.7). This is similar to the Sr and Ba partitioning behavior

observed between plagioclase and silicate melts or hydrothermal solutions (Blundy and Wood,

Page 45: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

32

1991; Lagache and Dujon, 1987). In the case of Sr or Ba, both these cations are divalent and

based on size and charge balance considerations should be accepted more readily into the

anorthite structure in exchange for Ca2+ rather than Na+ in the albite structure. This apparent

discrepancy is explained by the highly elastic nature of the albite structure (Blundy and Wood,

1991). Albite has a lower bulk modulus and a lower shear modulus than anorthite, which results

in an increased “flexibility” of the albite crystal structure (Angel et al., 1988; Blundy and Wood,

1991). These results suggest that the albite crystal lattice would better accommodate Li+, despite

the fact that Na+ is larger than Ca2+, than the more rigid anorthite structure.

2.5.1.4 Intermineral Partitioning

Previous studies of the lithium content in mantle xenoliths have made a correlation between

lithium contents of olivine and clinopyroxene pairs and xenolith paragenesis, e.g. equilibrated,

metasomatised, etc., (Figure 2.9; Seitz and Woodland, 2000; Paquin and Altherr, 2002;

Woodland et al., 2002; Woodland et al., 2004). Specifically, olivine-clinopyroxene pairs that are

apparently equilibrated (are in chemical equilibration, have no major inhomogeneities or mineral

zoning; Seitz and Woodland, 2000) tend to fall on a linear ~1:1 trendline when lithium

abundance of the olivine is plotted against the lithium abundance of the clinopyroxene. The

xenoliths apparently metasomatised by silicate melt and hydrous fluids fall below the trend line

(depleted in olivine relative to clinopyroxene), and those altered by carbonatite melt fall on and

above the trend line (enriched in olivine relative to clinopyroxene). Figure 2.9 also includes

experimental data corresponding to similar metasomatic regimes. The higher concentrations of

lithium are due to experimental requirements and analytical detection limits. All the

experimental data fall in a relatively restricted range of Dol/cpxLi of ~1 or > 1. A silicate melt

equilibrated olivine-clinopyroxene pair, from Brenan et al. (1998b), plots slightly below the line

projected from the equilibrated mantle xenoliths; and carbonatite melt olivine-clinopyroxene

pairs, from Blundy and Dalton (2000), fall on the projected line, above the silicate melt

experiment. Olivine-clinopyroxene pairs equilibrated with hydrous fluids in this study fall above

the equilibrated mantle xenoliths trend, not below, where the hydrous fluid metasomatised

samples plot. Despite the diversity of experimental methods, the range of Dol/cpxLi exhibited in

natural samples is not reflected in experimental studies, as no experimental studies have shown

Dol/cpxLi

< 1.

Page 46: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

33

One possible explanation is that the clinopyroxene compositions resulting from the hydrous fluid

experiments are very low in Al2O3, (0.2 wt. % to 0.3 wt. %); much lower than the Al2O3 content

of clinopyroxene found in mantle most xenoliths (2 – 6 wt. %, Seitz and Woodland, 2000).

Clinopyroxene-fluid partitioning experiments of Brenan et al. (1998b) have shown that lithium

partitioning increases with increasing Al content of the pyroxene. The high Dol/cpxLi values from

this study may be a reflection of the low Al2O3 content of the pyroxene. In the case of

carbonatite melt equilibrated olivine-clinopyroxene pairs from Blundy and Dalton (2000), their

experiments contain similar Al2O3 compositions as those found in carbonatite melt

metasomatised xenoliths, and the discrepancy between experimentally determined Dol/cpxLi,,

(values of ~1 or > 1) and those measured in mantle xenoliths (D < 1) still exists. Recent studies

of lithium and lithium isotopes in mantle xenoliths and other rocks have found that intermineral

partitioning and fractionation of lithium often will not correlate with expected equilibrium

values. These apparent disequilibrium signatures are attributed to remobilization of lithium with

differing rates of diffusion between olivine and clinopyroxene (Parkinson et al. 2007; Rudnick

and Ionov, 2007; Jeffcoate et al., 2007).

2.5.2 Controls on Isotopic Fractionation

Lithium isotopic fractionation between minerals and fluids depends on the difference in the zero

point potential energy (ZPE) between the phases of interest. 7Li is heavier and has a lower

vibrational frequency, and therefore a lower ZPE than 6Li (Chacko et al., 2001). The phase that

will undergo the greatest reduction in ZPE will preferentially take 7Li over 6Li (Chacko et al.,

2001). This has been demonstrated by Ab initio calculations, which have predicted that during

mineral-solution reactions 6Li, is preferentially incorporated into octahedrally coordinated sites

in the solid, and 7Li is preferentially incorporated into the dominantly tetrahedrally coordinated

sites in the fluid (Yamaji et al., 2001).

The coordination state and bonding environment of lithium in both spodumene, a clinopyroxene

with up to 7 % lithium content, and Ca-clinopyroxene, in this case diopside with 6 to 60 ppm

lithium, is octahedral. The consistent fractionation between spodumene and aqueous fluids at 2

GPa (Wunder at al., 2006); clinopyroxene and aqueous fluids at 1 GPa (measured in this study);

and altered seafloor basalts (Chan et al., 1992; Chan et al., 1993) suggests that the coordination

state of lithium in aqueous fluids does not change throughout this range of conditions.

Page 47: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

34

2.5.3 Lithium Incorporation into the Mantle

Pristine mid-ocean ridge basalts (MORB) contain 5-6 ppm lithium, with an average δ7Li of +4

‰, and resemble the mantle with respect to lithium content and composition (Jagoutz et al.,

1979; Moriguti and Nakamura, 1998; Tomascak, 2004; Tomascak et al., 2008). The mantle

source that produces MORB is thought to also provide the source for IAB, after modification by

a slab-derived flux. Given that neither partial melting nor differentiation and crystallization will

cause significant fractionation of δ7Li (Tomascak et al., 1999), variations in lithium content and

isotopic composition in some arc lavas are believed to arise from the slab inputs to the melt

source regions. During prograde metamorphism the mineral assemblages in the subducted slab

become more anhydrous with increasing pressure and temperature. Fluids produced by

dehydration reactions in the subducting slab add fluid mobile elements to the overlying mantle

wedge; a signature that is believed to be reflected in the lavas derived from this re-hydrated

mantle. For example, Kamchatka arc lavas are most enriched in boron relative to Nb or Zr at the

arc front and the enrichment decreases to MORB values with increasing slab depth (Ishikawa et

al., 2001). This is suggestive of continuing mobilization of fluid mobile elements into the arc

source region by fluids derived from dehydration reactions in the down going slab (Leeman,

1996; Ishikawa and Tera, 1999; Ishikawa and Nakamura, 1994; Ishikawa et al., 2001). The Izu

arc is one of the few localities where a clear correlation between lithium content, boron content,

δ7Li, and distance to the arc front can be made. The δ7Li in the Izu lavas range from +7.6 ‰ in

the arc front, to +1.1 ‰ in the back arc; this is thought to reflect enrichment of the arc melt

source by fluids derived from the down going slab (Moriguti and Nakamura, 1998). Typically

the lithium isotopic composition of most arc lavas ranges from δ7Li approximately +5 ‰ to +1

‰ and shows a slight negative correlation with Li/Y ratio (Tomascak et al., 2002). It is not clear

why correlations between lithium content and δ7Li of arc lavas are not consistent among all arcs.

Variations in the extent of slab dehydration due to slab age, angle of subduction and overall

thermal regime have been suggested as possible factors (Moriguti et al., 2004). However,

differences in the Li/Y content and 7Li composition between the Kurile arc lavas and the Izu arc

lavas, or even the Japan arc lavas, persist despite similarities in the subduction regime and age of

the slab (Morguti et al., 2004; Tomascak et al., 2002). Another suggestion is that slab-derived

fluids are significantly modified during transport through the mantle wedge to the melt source,

and that the lithium signal is attenuated by interaction with mantle minerals (Tomascak et al.,

Page 48: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

35

2002). Subtle differences in the mode of fluid transport through the mantle wedge could lead to

significant differences in the overall behavior of lithium in subduction zones.

2.5.4 The Mantle Wedge as a Chromatograph

The fluids derived from the dehydrating slab during subduction are potentially very lithium rich

depending on the nature of the subducted sediments, in some cases containing as much as 2000

ppm lithium or more (Chan et al., 2002). However, consistent and clear correlations of lithium

with other fluid mobile elements, such as boron, are rare. More commonly, a slab-like lithium

signal cannot be correlated with other indicators of fluid involvement such as B/Be ratios or

depth to slab. For example, some of the calc-alkaline lavas belonging to the Panamanian Old

Group lavas have high B/Be contents, suggesting high fluid input, and MORB-like δ7Li (+4.7 to

+5.6 ‰; Tomascak et al., 2000). Similar behavior is found in other Central American lavas

(Chan et al., 2002), as well as lavas from the Aluetian and Kurile arcs (Tomascak et al., 2002). It

has been suggested that the relatively compatible mineral-fluid partition coefficients for lithium,

the rapid diffusion of lithium into mantle minerals, and the high rock/fluid ratio experienced by

the fluids in the mantle wedge can provide a mechanism by which the lithium signal is decoupled

from other fluid mobile trace elements in slab-derived fluids (Tomascak et al., 2000; Tomascak,

2004; Wunder et al., 2006).

The diffusion coefficients measured in this study provide constraints as to the time required for

fluid-mineral equilibrium. This information, coupled with measurements of lithium partitioning

and isotopic fractionation between fluids and mantle minerals, allows for quantitative modeling

of the interaction between slab-derived fluids and the mantle wedge during fluid transport from

the slab to the arc melt source.

The effect of mantle wedge and fluid interaction can be evaluated following the method of

Navon and Stolper (1987), who modeled the distance traversed by various elements flowing

through an ideal mantle column of fixed porosity. This ideal column contains solid rock with an

interconnected fluid network along grain edges. Assuming that partition coefficients are

constant, the fluid fraction is uniform, and the densities and diffusivities of the solid and fluid do

not vary across the column length then:

Page 49: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

36

(19) 0

z

CVX

t

C fff

f

where Cf is the trace element concentration in the fluid, Xf is the mass fraction of the trace

element in the fluid, Vf is the velocity of the fluid, t is time and, z is the distance traversed along

the column. Assuming solid-fluid equilibrium is maintained, the most incompatible trace

elements have fronts that travel farther than trace elements that are more compatible for any

given time. The transport velocity of a trace element relative to the transport velocity of the fluid,

(Vtr/Vfl), is equal to the mass fraction of the trace element in the fluid (Xf):

(20) Xf = f / (f +(1-)sD)

Where is the volume fraction of fluid in the column (assumed to be 0.03 by Navon and Stolper,

1987), f and s are the fluid density and mantle wedge density; (assumed to be 1 g/cm3 and 3

g/cm3 respectively) and, D is the bulk partition coefficient for the element of interest.

A bulk DLi of 0.42 was calculated, assuming 80 % olivine and 20 % clinopyroxene, using the DLi

of olivine-fluid and cpx-fluid measured in this study. The olivine-cpx boron data (from Brenan et

al., 1998a) and cpx-fluid boron (from Brenan et al., 1998b) were combined to estimate the bulk

DB for the same lherzolite assemblage and XfB was also calculated as above for comparison.

From Navon and Stolper (1987) the rate at which a point of constant concentration moves

through the column (Vtr) is:

(21) fftrCf

VXVt

z

So for a given column length, at the time that the fluid front reaches the melt source, the boron

front will be 91 % of the column length, and the lithium front will only be 2 % of the column

length. The maximum capacity of the column for each element can be determined by calculating

at what time the trace element front reaches the top of the column, relative to the time the fluid

reaches the top of the column.

(22) ff

ff

c XVL

VXL

t

t 1

Page 50: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

37

where tc is the time for the fluid front to reach the top of the column, t is the time for the trace

element front to reach the top of the column, L is the column length, Vf is the fluid velocity, and

Xf is the mass fraction of the trace element in the fluid. For a given column length and a fluid

velocity, the boron front will reach the melt source at approximately the same time as the fluid

front (1.08tc); however, for the lithium front to reach the top of the column requires the column

to be filled ~42 times. This is a strikingly large volume of fluid, requiring a total of 1.26 cm3 of

fluid for every 1 cm3 of rock for a column with 3 % porosity, which is equivalent to ~40 wt. %

fluid. The highest estimates of fluid involved in arc magmatism from the literature is ~20 wt. %

(Ayers, 1998) and most estimates range from 1-5 wt. % (Stolper and Newman, 1994). Even if the

subducting slab has the capacity to generate such a large quantity of fluid, the time required to

deliver this amount of fluid to the melt zone needs to be considered.

Measurements of U-series disequilibria provide constraints on the timescales of fluid-mobile

element transport from the slab to the melt source (Elliott et al., 1997). Young lavas from

subduction zones often contain an excess of 238U relative to 230Th, or [238U]/[230Th] >1 (Elliott et

al., 1997). Unlike Th, U readily partitions into oxidized fluids, therefore a [238U]/[230Th] ratio >1

is believed to be the result of the addition of a slab derived fluid containing both 238U and 234U

(the parent of 230Th; half-life ( ~250 kyr), to the melt source within the last 30, 000 years

(Elliott et al., 1997). Fluid velocities predicted from U-series disequilibria range from 4 to 10

m/yr. Given these velocities, the time required for the lithium signal to reach the melt source can

be compared with that for the boron signal.

Assuming that partial melting of peridotite occurs at depths of 100 km below intra-oceanic arcs

(Plank et al., 2009) and a Benioff zone of ~125 km, then a maximum column length will be ~25

km. As shown in Figure 2.11, it would take the boron signal 100-5000 years to travel 25 km to

the melt source, whereas the lithium signal will need between 10,000 years and 200,000 years to

reach the melt source. If fluid velocities are 4-10 m/yr then, only the boron signal will reach the

melt source while 238U and 230Th still maintain measurable isotopic disequilibrium. Other studies

have made similar estimates of flux rates using 226Ra-230Th (Sigmarsson et al., 2002), the large

excess of 226Ra over 230Th displayed in many young lavas requires fractionation, presumably due

to fluid transport, to occur within 8 ka, or ~5half-lives (Sigmarsson et al., 2002). These time

constraints give rise to fluid velocities of 10-100 m/yr (Sigmarsson et al., 2002). Only at the

highest estimated fluid velocity, the lithium signal will reach the melt source within 10,000

Page 51: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

38

years; this will satisfy the time constraints determined from measurements of U-series

disequilibria and Ra-Th disequilibria (~10,000 yr; Figure 2.11).

More recent studies have suggested that slab dehydration produces a zone of hydrated

lithosphere, mainly consisting of chlorite, which is down dragged by corner flow to depths where

melting may take place in the asthenosphere given the right subduction geometry (Grove et al.,

2009). In this scenario mantle melting occurs 50-100 km below intra-oceanic arcs depending on

the angle of slab-dip and the slab convergence rate. Assuming the Benioff zone is between 100

and 125 km, the minimum column length will be ~10 km. As shown in Figure 2.11, the boron

signal needs only 100-2000 years to travel 10 km, whereas the lithium signal will need 4000-

100,000 years to reach the melt source. Even if the column length is very short, in order for the

lithium signal to reach the melt source within the timescales constrained by U-series and Ra-Th

disequilibria, fluid velocities between 10 m/yr and 100 m/yr are necessary. It should be noted

that lithium partition coefficients for cpx/chlorite have been estimated from lithium

concentrations in the eclogites from Syros, Greece, and are ~1, therefore it is expected that

chlorite/fluid partitioning will be similar to clinopyroxene/fluid partitioning with respect to

lithium (Marschall et al., 2006).

This suggests that for the lithium signal to be correlated with the boron signal, as well as other

fluid mobile elements, extremely high fluid volumes and velocities are required. The rather

improbably high fluid volume and extremely rapid fluid velocity required to transport the lithium

signal from the slab to the melt source is consistent with the lack of a slab-derived lithium signal

in many arc volcanics. The lithium signal will not reach the melt source because it will

preferentially partition into the mantle wedge, relative to other fluid mobile elements (such as

boron).

Given that subducting slabs are unlikely to generate the large fluid volumes required to transport

a lithium signal to the melt source, the occurrence of a slab-like lithium signal in arc lavas

implies a mode of fluid transport other than percolation. The rapid fluid velocities required by

Ra-Th disequilibria have led to the suggestion of fluid transport by hydrofracture rather than

percolation (Davies, 1997). Because fluid transport would be limited to fractures on the scale of

300-1500 m long and 10-200 mm wide (Davies, 1997), the fluid volume needed to generate very

high fluid/rock ratios is more reasonable because the fluids only interact with a small volume of

Page 52: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

39

rock. Since lithium partitions preferentially into the mantle relative to other fluid mobile

elements, minimal rock interaction would result in more lithium being transported to the melt

source.

2.5.5 Isotopic Evolution of Lithium-Bearing Fluids in the Mantle

2.5.5.1 Percolation and Rayleigh Distillation

Interaction of lithium bearing fluids with mantle minerals will result in changes to the isotopic

composition of both phases. If the shift in the isotopic composition of the fluid and the mantle

are known, then the amount of fluid: rock interaction can be estimated. The degree of

fractionation resulting from fluid-rock interaction can be modeled assuming a simple Rayleigh

distillation model:

(23) 3)1(377 10)10( fLiLi slabfluidfluid

Where 7Lifluid is the altered fluid, 7Lislabfluid is the initial composition of the slab-derived fluid,

and f is the fraction of the element in the fluid remaining after interaction with the mantle wedge.

In this case is calculated from the degree of cpx-fluid fractionation measured at 1100 oC in this

study and is defined as:

(24) = (7Limin + 1000)/(7Lifluid+1000)

The calculated here is 0.999, which is consistent with the calculated using data from the

study of Wunder et al. (2006). When is < 1, continued interaction of the fluid with mantle

minerals, i.e. distillation, will result in progressively heavier fluids. Figure 2.12 shows how the

isotopic composition of a fluid with an initial δ7Li of +9.7 ‰ (the slab input estimated by

Moriguti and Nakamura, 1998) would change during percolation through a mantle column. The

isotopic composition of the altered fluid increases and is heavier than the initial slab-derived

fluid and both the fore arc and back arc lavas of the Izu arc (δ7Li = +7.6 ‰ and +1.1 ‰,

respectively) for any amount of fluid/rock interaction.

It is important to note that the used in the above calculations is determined for cpx-fluid

fractionation at 1100oC. The temperature of the mantle will be lower near the subducted slab.

Depending on the age of the crust, the rate of subduction, and the degree of frictional shear

Page 53: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

40

heating the temperature in the mantle above the slab may be as low as 700-800oC (Peacock,

1993). Because fractionation of stable isotopes tends to increase with decreasing temperature

(Urey, 1947) the fractionation occurring at the base of the mantle column, close to the top of the

slab, may be even larger. Additionally, the degree of fractionation between fluids and the mantle

may be greater than what has been calculated above since the value used in the above

calculation was determined from cpx-fluid fractionation experiments and assumes that the

fractionation of lithium between olivine and fluids is the same. For reference, curves for values

of 0.998 and 0.996 are plotted on Figure 2.12, showing the effects of greater fractionation factors

on the fluid composition. As the fluids percolate through the mantle wedge the fraction of

lithium in the fluid decreases, and the δ7Li of the fluid becomes progressively greater. Because

lithium is readily taken up by mantle minerals, the fraction of lithium remaining in the fluid

becomes very small, Xf → 0.2 (see above), and depending on , the isotopic composition

becomes extremely fractionated with δ7Li ranging from +15 ‰ to +35 ‰ (depending on ;

Figure 2.12).

Because 7Li preferentially fractionates into fluids, interaction of slab fluids with mantle minerals,

i.e. percolation, will generate heavier, more 7Li-rich fluids. A fluid with an initial δ7Li of +9.7 ‰

percolating through the mantle wedge could not generate the isotopic signature observed in the

Izu arc lavas. Either the initial slab fluid is lighter, or the isotopic signature is the result of mixing

the 7Li-rich altered fluid and a lighter mantle reservoir. Interestingly, to generate the isotopic

composition of the Izu fore arc lavas using the Xf as calculated above, requires an initial slab

input with δ7Li = +4 ‰; essentially a fluid with MORB-like δ7Li. For the Izu lavas of the back-

arc region, with δ7Li = +1.1 ‰, it is unlikely that any component of the slab derived fluid has

made its way to the melt source by percolation, as even the least altered slab-derived fluid is

isotopically heavier than the unaltered MORB source mantle. The implication here is that this

isotopically light lithium is not due to percolation of the fluid through the mantle, but must be the

signature of a component derived from an isotopically light reservoir, such as the residual slab or

oceanic sediments (δ7Li ~ -2 ‰; Moriguti and Nakamura, 1998). Another possibility is that this

signal is due to an entirely different mechanism of isotopic fractionation and transport, which is

discussed in the following section.

Page 54: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

41

2.5.5.2 Generation of 6Li-rich fluids

Ab initio calculations have demonstrated that during mineral-solution reactions 6Li should be

preferentially incorporated into octahedrally coordinated sites in solid phases (Yamaji et al.,

2001). Recent work by Jahn and Wunder (2009) has examined how lithium speciation in hydrous

fluids affects isotopic fractionation. From Ab initio molecular dynamic (AIMD) calculations,

they have determined that during fluid-solid fractionation, 6Li will prefer sites with the higher

coordination. Lithium in pyroxene and olivine, the most abundant mantle minerals, is in

octahedral or six-fold coordination. When fluid densities are less than 1.0 g/cm3 coordination of

lithium in the fluid is mainly three-fold (Jahn and Wunder, 2009), and therefore 7Li will

preferentially partition into the fluid phase. As fluid density increases, the coordination of

lithium in the fluid also increases. When fluid density is greater than 1.2 g/cm3, the proportion of

5-fold and 6-fold coordinated lithium increases and the proportion of 3-fold and 4-fold

coordinated lithium decreases, and the overall average lithium coordination in the fluid is greater

than 4.5 (Jahn and Wunder, 2009). When lithium coordination in the fluid becomes greater than

lithium coordination in the mineral phase the sense of fractionation changes, and 6Li is predicted

to preferentially partition into the fluid phase. This change in sense of fractionation has been

observed during staurolite-fluid partitioning experiments at 3.5 GPa (Wunder at al., 2007).

Lithium is in tetrahedral coordination in staurolite and preferentially incorporates 7Li at 3.5 GPa.

Therefore, 6Li-rich fluids may be generated by mineral-fluid fractionation at high pressures.

Figure 2.13 is a plot of fluid density vs. temperature, with the average calculated Li-coordination

shown as degree of shading. Superimposed on this plot are the fluid densities calculated with the

CORK-EOS (Holland and Powell, 1991) using the Perple_X computer program (Connolly,

2005) for Franciscan and Alpine subduction zones (Ernst, 1988) as well as the most direct path

between the slab and a fore-arc volcano (Peacock, 1993). It is possible to generate 6Li-rich fluids

when mineral-fluid interaction occurs at depths greater than ~125 km. These results suggest that

as fluids percolate up through the mantle wedge, fluid density decreases and the average

coordination of lithium in the fluids will decrease; 6Li will once again preferentially fractionate

into the mineral phase and the fluids will become heavier. In order to preserve the 6Li-rich signal

generated at depth, fluids need to reach the melt source having undergone minimal interaction

with the mantle on their ascent path. Fluid transport by hydrofracture would satisfy this

Page 55: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

42

requirement, as it results in high fluid-rock ratios, which would transport the 6Li-rich fluids to the

melt source quickly, and with least amount of interaction with the mantle wedge.

2.5.5.3 Generation of 6Li-rich zones in the mantle

Hyrdofracture of the mantle by slab-derived fluids is an appealing mechanism to transport

lithium through the mantle, as channelized flow through hydrofractures would satisfy the high

fluid: rock ratios and rapid fluid velocities required by mineral-fluid partitioning. This transport

mechanism would also minimize isotopic fractionation by limiting mineral-fluid interaction,

thereby effectively propagating a slab signal all the way to the melt source region. The isotopic

composition of the mantle wall rock of the fractures would also shift, generating a local

isotopically light region in the mantle wedge. The isotopic shift of the wall rock depends on the

extent of reaction between the mantle and the slab-derived fluids. If mineral-fluid exchange is

fast, then local mineral-fluid isotopic equilibrium will occur, and the isotopic shift will depend

on the fluid-rock ratio.

This process can be modeled after the approach of Abart (1995; after Taylor 1977) by calculating

the progress of the reaction, which is defined as the ratio between the observed isotopic shift

in the rock and the maximum attainable isotopic shift:

(25) RF

RRiLiFR

iLi

iLi

fLi

In this case, LifR is the final isotopic composition of the mantle, Li

iR is the initial composition

of the mantle, LifF is the composition of the metasomatizing fluid and R-F is the fractionation

between the mantle and the fluid. The value of will be between 0 (no equilibration) and 1

(complete equilibration). Where mineral-fluid exchange is rapid, as is the case for lithium

exchange, then the degree of equilibration depends on the lithium atom equivalent fluid-rock

ratio, N (Taylor, 1977):

(26) N = - ln(1 - )

The final isotopic composition of the mantle wall rock is estimated here given an initial mantle

composition of δ7Li = +4 ‰ (Moriguti and Nakamura, 1998; Tomascak et al., 2008; Nishio et

al., 2002). This model assumes complete equilibration between the mantle wall rock and fluid

Page 56: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

43

will result in the mantle wall rock having a δ7Li‰ lower than the initial fluid composition

(cpx-fluid = -1 ‰ at 1100oC; this study, Wunder et al., 2005). Because the isotopic composition of

the initial slab fluid is not very well constrained, three different initial fluid compositions are

used in this illustration; δ7Li = +10 ‰ (estimate for the Izu arc fluids by Moriguti and

Nakamura; 1998), 0 ‰ and -10 ‰, the latter being arbitrary values reflecting generation of 6Li-

rich fluids at depth (Jahn and Wunder; 2009). Given a DLibulk = 0.42, N becomes (1/ DLi

bulk)*N in

weight units. Assuming the isotopic composition of the fluid does not change, which is the case

if the fluid/rock ratio is high, the isotopic composition of the vein/hydrofracture wall rock will

approach complete equilibrium with the fluid; equal to R-F of -1 ‰. Because the timescales of

lithium diffusion in olivine and clinopyroxene are rapid compared to the fluid transport times (2

m/1hr; see Ch.3, vs. 100 m/yr; Sigmarsson et al., 2002) the isotopic shift in the wall rock is a

function of the amount of fluid available.

Figure 2.14 shows the evolution of the final isotopic composition of the metasomatized mantle

wall rock with increasing fluid/rock ratio. The isotopic shift of the wall rock depends on the

extent of reaction between the mantle and the slab-derived fluids; if the fluid/rock ratio is high,

the isotopic composition of the vein/hydrofracture wall rock will approach complete equilibrium

with the fluid equal to R-F of -1 ‰. Also plotted are the δ7Li values of the Izu fore arc and back

arc lavas (which are typical of the range of δ7Li values found in many arc lavas; Tomascak et al.,

2002). The entire range of δ7Li values found in arc lavas can be achieved by metasomatizing the

mantle with fluids that have δ7Li between 0 ‰ and +10 ‰ and fluid/rock ratios >1.2. Both these

values are reasonable given the isotopic composition of subducted material and typical fluid/rock

ratios for hydrofractured zones. Values of δ7Li in seafloor sediments range from -5 ‰ to +20 ‰

(Marschall et al., 2007 and references therein) and altered oceanic crust has δ7Li of ~ +14 ‰

(Moriguti and Nakamura, 1998; Chan et al., 1992). Slab derived fluids with δ7Li between 0 ‰

and +10 ‰ could be achieved during dehydration of the slab, recalling that isotopic fractionation

in a cool slab at depths greater than ~125 km are likely to produce fluids that are isotopically

lighter than the solid (Jahn and Wunder, 2009). The fluid/rock ratios in vein systems can be very

high with typical values from calcite or quartz vein systems ranging from 70-100 cm3 of fluid per

100 cm3 of rock to as much as 1400 cm3 of fluid per 100 cm3 of rock (Spear, 1993).

Page 57: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

44

2.6 Conclusions

Lithium is moderately incompatible in the mantle during mineral – fluid exchange reactions. The

DLi measured in this study ranges from 1.34 – 0.14 in olivine, to 0.32 – 0.09 in plagioclase and,

0.32 – 0.07 in clinopyroxene. Lithium partitioning between clinopyroxene and hydrous fluids is a

function of temperature, decreasing with increasing temperature from 800oC to 1100oC at 1 GPa

and appears to increase with increasing Al2O3 content of the pyroxene. Olivine-fluid partitioning

of lithium is not a function of temperature, but appears to be sensitive to Mg/Fe content, although

this needs to be investigated more systematically. Lithium partitioning in anorthite is a function

of feldspar composition, similar to the partitioning of other cations in the feldspar-fluid system.

Lithium partitioning between olivine and clinopyroxene appears to be independent of

temperature; however, preliminary experiments examining the effect of REE content and fO2

suggest that DLiol/cpx may be a function of crystal chemistry. Isotopic fractionation between

clinopyroxene and fluid has been measured as well as between olivine and clinopyroxene. The

isotopic fractionation between clinopyroxene and fluid at 900oC is ~ +1 ‰ (±2 ‰) and the

measured isotopic exchange between olivine and clinopyroxene is ~ +5 ‰ (±4 ‰). Isotopic

fractionation between clinopyroxene and fluids is a function of temperature and consistent with

what has been observed in the spodumene – fluid system. The fractionation between spodumene

and hydrous fluids results in an enrichment of 7Li in the fluid from +3.5 ‰ at 500oC to ~ +1.0 ‰

at 900oC and 2.0 GPa (Wunder et al., 2006).

Application of these data to models of fluid-rock interaction in the mantle wedge reveals that

lithium is a moderately incompatible element in the mantle during mineral-fluid exchange

reactions. Because lithium is not a conservative element, it cannot be used to deconvolve the

proportions of slab-derived fluid and altered and unaltered MORB-source involved in generating

arc lavas. However, constraining how lithium behaves in the mantle provides some insight into

the lithium and lithium isotopic trends, or lack thereof, observed in arc lavas. The absence of

high Li/Y ratios in arc lavas with high B/Be, or MORB-like δ7Li in lavas with high B/Be

contents (such as the lavas from the Sunda arc, Indonesia; Tomascak et al., 2002), can be

explained by partitioning lithium into mantle minerals as fluids percolate through the mantle

wedge. In these cases, transport through the mantle wedge completely removed the lithium

signal from the slab-derived fluid. Convergent margin lavas, such as the Izu forearc lavas, with

Page 58: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

45

δ7Li values greater than the mantle values (δ7Li ~ +4 ‰) are likely the result of some component

of slab fluid-mantle interaction during percolation. It is important to note that very high fluid

fluxes are implied if a 7Li signal from slab-derived fluids is to reach the melt source by

percolation.

Low δ7Li values (< MORB; δ7Li ~ +4 ‰) that correspond with high Li/Y ratios are likely

generated near the slab and transported to the melt source with a minimum amount of interaction

with the mantle wedge; here transport through hyrdofractures is a likely mechanism. The trend of

increasing δ7Li with decreasing Li/Y, which is observed in most arc lavas (Tomascak et al.,

2002), could be viewed as a spectrum between the two scenarios. Where low Li/Y values

correspond with high δ7Li, large fluid fluxes were most likely percolating through the mantle

wedge. Where high Li/Y values correspond with low δ7Li, the fluids were likely generated at

depth and transported through the mantle through hydrofractures, having minimal interaction

with the wedge. Intermediate values could be a result of some component of both these

mechanisms.

Transport of slab-derived fluids through hydrofractures in the mantle can also explain the lack of

clear and consistent correlations between lithium and other fluid mobile elements. Fluids

transported to the melt source through hydrofractures would be subject to differing degrees of

mantle interaction (variable fluid/rock ratios and transport velocities). Lithium is moderately

compatible in the mantle and diffuses rapidly; therefore, lithium contents and isotopic

compositions will be very sensitive to variations in the types of mineral-fluid interaction.

The lithium isotopic evolution of the mantle will also be affected by these processes, as it is such

an efficient sink for lithium. Dehydration reactions in the subducting slab at depths less than

~125 km, where fluid density is relatively low, and the predicted predominance of three-fold and

four-fold coordinated lithium in the fluid will generate 7Li-rich fluids and result in localized 6Li

enrichment of the mantle. Hydrous fluids generated deeper than ~125 km are predicted to contain

lithium in coordination states greater than four-fold, and therefore likely to be enriched in 6Li, at

least initially, giving rise to a zone of 7Li-rich mantle at depth. Xenoliths with δ7Li values

greater than MORB are uncommon, but have been found in blueschists from Syros (Greece),

eclogites from Dabishan (China), Cima di Gagnone and Trescolmen (Alps) and lherzolites from

Northern Japan and SE Austrailia (Nishio et al., 2004; Marschall et al., 2007).

Page 59: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

46

Table 2.1 Composition of Starting Material

Dekalb Diopside San Carlos Olivine Crystal Bay Bytownite

Mt St. Hillaire Albite

SiO2 54.77 (0.80)1 40.95 (0.02) 49.16 (0.20) 67.88 (0.26)

Al2O3 0.66 (0.10) <0.01 32.67 (0.26) 19.61 (0.20)

FeO 0.85 (0.08) 9.31 (0.05) 0.50 (0.06) <0.03

MgO 17.31 (0.22) 49.19 (0.42) 0.13 (0.04) <0.02

CaO 25.17 (0.26) <0.02 15.09 (0.28) <0.02

Na2O 0.43 (0.08) <0.02 2.65 (0.12) 11.18 (0.46)

MnO 0.05 (0.06) 0.12 (0.02) <0.03 <0.03

NiO <0.03 0.39 (0.03) <0.03 <0.03

Total 99.29 100.86 100.27 98.75

n 11 3 8 4

Li ppm2 8.86 (0.60) 2.52 (0.46) 1.65 (0.08) <0.14

7Li (‰) +9.7 (1)3 +3.64 (0.15)4 1Numbers in parentheses represent 2 of the mean of n analyses 2Analyzed by LA-ICPMS, numbers in parentheses represent 2 of the mean of 5 analyses for diopside, 9 analyses for olivine and 2 analyses for feldspars 3Analyzed by MCICP-MS, numbers in parentheses represent 2 of the uncertainty on the measurement 4Magna et al., 2006

Page 60: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

Table 2.2 Summary of Experimental Details

#3 0.2 800 74 Au DD 14.1 3.9 - - 48.3 A cpx + quench xtlsNCDL2-2 1 800 139 Pt DD 10.66 3.9 - - 57.09 A clear cpx + orange cpxNCDL4 1 1000 67 Pt DD 9.67 3.9 - - 49.46 A clear cpx + orange cpxNCDL5 1 1100 72 Pt DD 10.27 3.9 - - 36.02 A clear cpxNCDL6 1 900 72 Pt DD 6.65 3.9 - - 36.62 A clear cpx + orange cpxNCDLR 1 800 69 Pt NCDL1 9.20 - - - 21.35 B fine grained clear cpx + ? small xtls + quench DiAb10 1 800 66 Pt DD 6.82 - 9.5 - 42.01 A clear cpx + orange cpx

NCOL1 1 1000 72 Pt SCO 5.27 3.3 - 3.2 42.04 A clear olivine + pink/red/black oxidesNCOL2 1 900 68 Pt SCO 10.87 3.3 - 3.2 50.48 A clear olivine + pink/red/black oxidesNCOL3 1 1100 70 Pt SCO 7.83 3.3 - 3.2 37.21 A clear olivineLSCO8 1 1200 72 Pt SCO 12.63 3.3 - 3.2 57.92 C clear olivineNCOLR 1 800 67 Pt NCOL2 5.03 - - - 35.48 B clear olivine + ? Mg phase + pink/red/black oxidesOlAb5 1 900 66 Pt SCO 5.12 - 4.5 - 40.42 A clear olivine

NCA1 1 1000 77 Pt CBBy 11.31 3.1 - 2.6 45.28 A melt + large clear anorthiteNCA2 1 900 69 Pt CBBy 7.95 3.1 - 2.6 45.55 A clear anorthiteNCA3 1 800 70 Pt CBBy 18.36 3.1 - 2.6 46.20 A asicular green xtls (amphibole?) + clear anorthiteNCA5 1 800 48 Pt CBBy 9.34 3.1 - 2.6 28.69 A asicular green xtls (amphibole?) + clear anorthiteNCAR 1 800 72 Pt NCA3 5.92 - - - 34.15 B zoesite

AnAb10 1 800 68 Pt CBBy 5.49 - 12.6 - 46.36 A melt/quench + clear anorthiteAnAb20 1 800 55 Pt CBBy 7.68 - 19.3 - 40.93 A melt/quench + clear anorthite

2m-hi4 1 1000 48 Pt+Re DD + SCO3 6.60 - - - 25.45 A melt/quench + clear enstatite2m-lo5 1 1000 48 Pt+Mo DD + SCO3 5.73 - - - 29.66 A blue/grey cpx and ol + black oxides + CaMo oxideYb-16 1 1000 72 Pt DD + SCO3 4.80 - - - 30.19 A clear cpx and ol + black oxides

NCDL3 1 900 71 Pt DD 9.47 3.9 - - 48.03 A clear olivine + clear cpx + orange cpx 3

capsule solid (mg)+ wt% SiO2

+ wt% Abstarting

material1fluid (mg) run productsfluid2sample P (GPa) T (oC) t (hrs)

+ wt% Al2O3

2m-1 1 900 72 Pt DD + SCO3 5.33 - - - 40.34 A clear cpx + orange cpx, ol absent2m-2 1 800 68 Pt DD + SCO3 9.04 - - - 48.01 A clear olivine + clear cpx + orange cpx 2m-3 1 1000 72 Pt DD + SCO3 8.47 - - - 47.01 A clear olivine + clear cpx 2m-R 1 800 72 Pt 2m-2 5.57 - - - 39.63 B clear ol + cpx + monticellite + black oxides

LDi-107 1 900 142 Pt DD 16.3 3.9 - - 86.7 C clear cpx + orange cpxLDi-117 1 900 72 Pt DD 5.1 3.9 - - 52.4 C clear cpx + orange cpxLDi-127 1 900 68 Pt LDi-10 4.89 13 - - 66.35 D clear cpx + orange cpxLDi-15 1 900 75 Pt DD 6.86 3.9 - - 63.63 C clear cpxLDi-17 1 1000 20 Pt DD 9.25 3.9 - - 59.82 C clear cpx + black oxidesLDi-18 1 1100 70 Pt DD 9.66 3.9 - - 31.13 C clear cpx

1 DD: Dekalb Diopside, SCO: San Carlos Olivine, CBBy: Crystal Bay Bytownite2 Fluid compositions; A: 243ppm Li, B: 96ppm Li, C: 306ppm Li, D: 180ppm Li3 82wt% DD + 18wt% SCO 4 Capsule materials result in log f O2 of -55 Capsule materials result in log f O2 of -156 0.25 mg of Yb2O3 added7 Ti outer capsule

Natalie Caciagli
Typewritten Text
Natalie Caciagli
Typewritten Text
47
Page 61: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

48

Table 2.3 Standards and Reference Material

Li (ppm)1 reference

International Standards

NBS 612 41.54 (2.87) Pearce et al. 1997

NBS 612 41.5 (2.2) this study2, LA-ICPMS

NBS 610 484.6 (21.7) Pearce et al. 1997

NBS 610 488.7 (39.6) this study3, LA-ICPMS

JG1a 79.5 (4.5) Imai et al. 1995

JG1a 92.4 (6.4) this study, LA-ICPMS

JB-2 7.78 (1.39) Imai et al. 1995

JB-2 7.9 (0.6) this study, LA-ICPMS

BCA1 13.3 Ryan and Langmuir, 1987

BCA1 12.3 (0.8) this study, LA-ICPMS

JGB-1 4.59 (.90) Imai et al. 1995

JGB-1 4.6 (0.4) this study, LA-ICPMS

In house Standards

San Carlos Olivine 2.52 (0.4) this study, LA-ICPMS

San Carlos Olivine 1.6 (0.08) Magna et al. 2006, MC-ICPMS

Dekalb Diopside 8.9 (0.6) this study, LA-ICPMS

Dekalb Diopside 7.8 (2) this study, MC-ICPMS

Kunlun Diopside 42.6 (3.) this study, LA-ICPMS

7Li(‰) reference

International Standards

IRMM016 +0.1 (1) this study, MC-ICPMS

IRMM016 -0.1 (1) Teng et al. 2004

IRMM016 +0.13 (1) Jeffcoate et a. 2004

In house Standards

UMD-1 54.8 (1) this study, MC-ICPMS

UMD-1 54.7 (1) Teng et al. 2004

San Carlos Olivine +3.64 (0.2) Magna et al. 2006, MC-ICPMS

San Carlos Olivine +1.1 (4) this study, SIMS corrected4

Dekalb Diopside +8.5 (4) this study, SIMS corrected4

Dekalb Diopside +9.7 (1) this study, MC-ICPMS

1) Numbers in parentheses represent 2 errors

2) Analysed using NIST 610 as standard unless otherwise noted

3) Analysed using NIST 612 as standard

4) Values are corrected for instrument mass fractionation (see text)

Page 62: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

Table 2.4 Run Product Major Element Composition

Sample Total n ppm Yb Di% Jd% Tsc% Fo% An%

Single Phase ExpClinopyroxene

#3 55.87 (0.12) 0.59 (0.12) 0.82 (0.11) 0.05 (0.02) 17.94 (0.20) 25.06 (0.16) 0.37 (0.07) na 100.70 16 94 3 0NCDL2-2 54.72 (0.29) 0.22 (0.10) 0.57 (0.21) 0.06 (0.03) 18.26 (0.17) 25.26 (0.26) 0.02 (0.01) nd 99.11 16 97 0 1NCDL4 54.84 (0.30) 0.23 (0.06) 0.43 (0.18) 0.05 (0.03) 17.98 (0.16) 25.82 (0.13) 0.02 (0.02) nd 99.37 14 99 0 1NCDL5 55.02 (0.30) 0.21 (0.06) 0.05 (0.09) 0.03 (0.02) 18.36 (0.16) 25.77 (0.24) 0.02 (0.01) nd 99.46 11 100 0 0NCDL6 54.77 (0.23) 0.32 (0.11) 0.78 (0.27) 0.05 (0.01) 17.87 (0.18) 25.38 (0.12) nd nd 99.17 5 95 1 1NCDLR 54.99 (0.23) 0.47 (0.13) 0.76 (0.23) 0.05 (0.03) 17.48 (0.26) 25.41 (0.24) 0.27 (0.17) 0.11 (0.16) 99.54 14 96 2 0DiAb10 54.59 (0.43) 0.32 (0.09) 1.79 (0.44) 0.04 (0.03) 17.89 (0.27) 25.28 (0.11) nd 0.04 (0.03) 99.95 8 96 0 2

OlivineNCOL1 42.51 (0.22) 0.04 (0.01) 1.12 (0.10) 0.10 (0.04) 55.59 (0.18) 0.03 (0.01) nd 0.32 (0.05) 99.72 9 99NCOL2 42.45 (0.22) 0.04 (0.01) 2.07 (0.24) 0.13 (0.03) 54.74 (0.31) 0.04 (0.01) nd 0.33 (0.11) 99.80 9 98NCOL3 43.74 (0.27) 0.07 (0.01) 0.06 (0.02) 0.05 (0.03) 56.72 (0.20) 0.06 (0.01) nd 0.03 (0.03) 100.73 9 100LSCO8 42.51 (0.14) 0.09 (0.02) 1.32 (0.63) 0.16 (0.10) 56.61 (0.64) 0.04 (0.00) nd 0.31 (0.39) 101.04 7 99NCOR 41.81 (0.42) 0.03 (0.01) 0.12 (0.03) 0.09 (0.03) 53.55 (0.70) 0.01 (0.01) nd 4.53 (0.70) 100.14 13 100OlAb5 42.62 (0.05) 0.02 (0.01) 0.83 (0.11) 0.12 (0.02) 55.73 (0.10) 0.09 (0.01) nd 0.36 (0.02) 99.77 5 99

PlagioclaseNCA1 42.99 (0.14) 37.36 (0.15) 0.03 (0.03) nd nd 19.77 (0.14) 0.15 (0.03) na 100.30 22 98NCA2 43.04 (0.13) 37.19 (0.15) 0.22 (0.03) nd nd 19.78 (0.11) 0.15 (0.04) na 100.38 14 99NCA3 44.05 (0.41) 36.32 (0.31) 0.25 (0.03) nd nd 18.81 (0.29) 0.66 (0.17) na 100.09 23 94NCA5 44.57 (0.26) 35.66 (0.22) 0.06 (0.05) 0.03 (0.02) nd 18.23 (0.22) 0.97 (0.09) 0.03 (0.02) 99.55 8 91

NCAR - zoesite 39.53 (0.44) 34.18 (0.12) 0.24 (0.13) nd 0.55 (0.10) 24.38 (0.24) nd na 98.88 20AnAb10 43.09 (0.01) 36.66 (0.10) 0.14 (0.04) 0.02 (0.01) 0.01 (0.01) 19.23 (0.16) 0.36 (0.08) nd 99.52 4 97AnAb20 43.99 (0.26) 35.87 (0.24) 0.06 (0.06) nd 0.02 (0.01) 19.15 (0.18) 0.41 (0.03) nd 99.50 10 93

NCA1 melt 43.47 (0.99) 26.14 (0.36) 0.12 (0.03) nd 1.16 (0.06) 14.78 (0.54) 0.26 (0.05) na 85.93 10NCA2 melt 60.36 (2.59) 21.64 (1.46) 0.61 (0.09) nd 0.40 (0.10) 2.28 (0.48) 0.86 (0.31) na 86.15 4

Two Phase ExpNCDL3 olivine 43.97 (0.21) 0.03 (0.01) 0.17 (0.04) 0.11 (0.02) 54.95 (0.17) 0.96 (0.25) nd nd 100.19 6 100

NCDL3 cpx 54.20 (0.38) 0.33 (0.08) 0.65 (0.13) nd 17.91 (0.13) 25.93 (0.21) nd nd 99.02 9 100 2

CaO Na2O NiOSiO2 Al2O3 FeO MnO MgO

p ( ) ( ) ( ) ( ) ( )2m-hi opx 59.23 (0.13) 0.14 (0.06) 1.92 (0.09) 0.05 (0.02) 38.52 (0.23) 0.09 (0.01) nd 0.03 (0.04) 99.99 9

2m-lo olivine 42.56 (0.41) 0.04 (0.01) 0.12 (0.10) 0.14 (0.05) 57.05 (0.60) 0.11 (0.06) nd 0.01 (0.02) 100.03 9 1002m-lo cpx 56.04 (0.30) 0.53 (0.09) nd 0.07 (0.01) 21.06 (0.49) 22.71 (0.65) nd nd 100.41 8 85 1 1

Yb-1 olivine 42.60 (0.41) 0.05 (0.04) 0.25 (0.07) 0.16 (0.04) 55.47 (1.20) 1.11 (0.28) nd 0.04 (0.02) 99.67 13 2868 100Yb-1 cpx 55.73 (0.22) 0.20 (0.05) 0.13 (0.06) 0.03 (0.02) 18.69 (0.16) 25.84 (0.11) nd 0.16 (0.13) 100.78 12 1441 98 0 12m-1 cpx 54.40 (0.34) 0.52 (0.11) 1.30 (0.37) 0.05 (0.03) 17.87 (0.22) 25.46 (0.26) nd 0.03 (0.02) 99.63 6 97

2m-2 olivine 42.67 (0.30) 0.04 (0.01) 0.02 (0.01) 0.10 (0.06) 42.67 (0.30) 0.26 (0.18) nd nd 85.77 4 1002m-2 cpx 55.30 (0.34) 0.24 (0.13) 0.36 (0.49) 0.05 (0.02) 18.40 (0.33) 25.47 (0.16) nd 0.02 (0.02) 99.84 5 97 1 0

2m-3 olivine 42.80 (0.14) nd nd 0.04 (0.02) 55.96 (0.17) 0.95 (0.04) nd nd 99.75 9 1002m-3 cpx 55.35 (0.14) 0.23 (0.04) 0.21 (0.10) nd 18.50 (0.12) 25.59 (0.10) nd nd 99.87 10 98 1 0

2m-R monticillite 38.53 (0.10) 0.03 (0.02) 0.15 (0.02) 0.12 (0.04) 26.23 (0.25) 34.66 (0.27) nd 0.07 (0.04) 99.80 8 1002m-R cpx 54.80 (0.18) 0.29 (0.04) 0.93 (0.09) 0.02 (0.02) 18.04 (0.07) 25.67 (0.13) nd 0.02 (0.02) 99.76 9 98 0 2

2m-R olivine 42.40 (0.14) 0.04 0.02 0.35 (0.11) 0.12 (0.03) 55.55 (0.41) 0.91 (0.13) nd 0.16 (0.05) 99.52 4 100

Natalie Caciagli
Typewritten Text
49
Page 63: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

50

Table 2.5 Run Product Lithium Concentration

sample Li ppm 21 n min/fluid 2 ol/cpx 2 min/melt 2

#3 13.66 3.12 2 0.07 0.01

NCDL2-2 97.87 32.12 2 0.32 0.14

NCDL4 33.20 13.96 3 0.14 0.06

NCDL5 26.64 12.04 3 0.11 0.05

NCDL6 61.96 32.29 2 0.27 0.14

NCDLR 6.67 1.66 3 0.07 0.02

DiAb10 49.92 0.61 2 0.21 0.03

NCOL1 101.36 10.18 3 0.47 0.03

NCOL2 124.3 13.32 4 0.57 0.06

NCOL3 85.6 35.00 2 0.38 0.16

LSCO8 49.20 12.22 3 0.17 0.04

NCOLR 13.37 1.84 3 0.14 0.02

OlAb5 278.90 22.66 3 1.34 0.02

NCA1 20.92 7.35 3 0.09 0.03

NCA2 20.89 2.26 2 0.09 0.02

NCA3 38.57 17.65 2 0.17 0.08

NCA5 70.21 33.57 3 0.32 0.15

NCAR* -zoisite 0.34 0.36 2

AnAb10 31.15 11.12 2 0.13 0.11

AnAb20 50.29 19.85 4 0.21 0.08

NCA1-melt 1119.93 223.99 1 0.019 0.008

NCA2-melt 1197.00 239.40 1 0.017 0.004

NCDL3 - olivine 66.27 8.10 3 0.30 0.02

NCDL3 - cpx 9.78 1.82 3 0.04 0.06 6.78 1.51

2m-1 cpx 13.56 2.83 2 0.06 0.01

2m-2 olivine 78.04 6.80 1

2m-2 cpx 50.22 4.03 2 1.55 0.18

2m-3 olivine 51.11 11.04 3

2m-3 cpx 12.79 7.07 2 4.00 2.37

2m-R olivine 31.04 2.14 1

2m-R cpx 25.84 5.16 1 1.20 0.25

2m-R monticillite 39.27 23.62 2

2m-hi enstatite 5.28 3.18 3 0.02 0.01

2m-lo olivine 29.21 8.90 2

2m-lo cpx 41.63 20.51 3 0.70 0.41

Yb-1 olivine 174.45 14.70 1

Yb-1 cpx 16.82 6.22 3 10.37 3.93

1) 2 refers to the standard deviation for n analyses and reflects the degree of sample heterogeneity

Page 64: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

51

Table 2.6 Isotopic Composition of Starting Materials and Run Products

Starting Material Products Exp T (oC) fluid mineral mineral

7Li (‰) 7Li (‰) 7Li (‰) 2(‰) 7Licpx-fluid

(‰) 7Liol-cpx

(‰)

MC-ICPMS data (UMD)

cpx-fluid experiments using 6Li doped solution

LDi-10 900 -88.4 +9.7 -90.9 1 -2.5 ±1.4

LDi-11 900 -88.4 +9.7 -91.4 1 -3.0 ±1.4

LDi-12* 900 -46.1 -90.9 -49.5 1 -3.4 ±1.4

LDi-15 900 -88.4 +9.7 -89.1 1 -0.7 ±1.4

LDi-17 1000 -88.4 +9.7 -89.5 1 -1.1 ±1.4

LDi-18 1100 -88.4 +9.7 -88.7 1 -0.3 ±1.4

cpx-fluid experiments using LSVEC solution

NCDL-6 900 0 +9.7 -2.6 1 -2.6 ±1.4

SIMS data (LLNL)

cpx-fluid experiments using LSVEC solution

NCDL-2 800 0 +9.7 -2.1 4 -2.1 ± 4.1

NCDL-4 1000 0 +9.7 -1.1 4 -1.1 ± 4.1

NCDL-5 1100 0 +9.7 -2.9 4 -2.9 ± 4.1

NCDL-6 900 0 +9.7 -3.5 4 -3.5± 4.1

Diopside #3 800/0.2GPa 0 +9.7 -5.3 4 -5.3 ± 4.1

olivine-cpx experiments using LSVEC solution

clinopyroxene

2m-2 cpx 800 0 +9.7 -4.2 4

2m-3 cpx 1000 0 +9.7 -6.5 4

2m-R cpx* 800 -2.7 -4.2 -4.4 4

olivine

2m-2 ol 800 0 +1.0 1.5 4 +5.7 ± 5.6

2m-3 ol 1000 0 +1.0 -0.7 4 +5.9 ± 5.6

2m-R ol* 800 -2.7 +1.5 -7.6 4 -3.2 ± 5.6

2m-R mtc* 800 -2.7 +1.0 -14.8 4

* denotes reversal experiment

Page 65: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

52

-8.0

-4.0

0.0

4.0

8.0

12

16

10 15 20 25 30 35 40

y = -16.287 + 0.80082x R2= 0.94585

7 Li m

easu

red

by

MC

-IC

PM

S (

o/o

o)

7Li 'uncorrected' measured by SIMS ( o/oo)

SCO

Dekalb

NCDL 6

Figure 2.1 Internal Reference Materials

Plot of δ7Li values measured by MC-ICPMS versus uncorrected δ7Li values measured by SIMS of the internal reference materials, Dekalb diopside and San Carlos olivine, as well as the run product from one experiment (NCDL6). The δ7Li values measured by MC-ICPMS for Dekalb diopside and NCDL6 are from this study and SCO is measured by MC-ICPMS from Magna et al. (2006). Error bars for ‘uncorrected’ SIMS δ7Li are 2, based on counting statistics. The error bars for δ7Li measured by MC-ICPMS are ±1 ‰ (2) or the published 2 errors (±0.3 ‰ ; Magna et al., 2006). Note that the discrepancy between values is due to instrument mass fractionation (Decitre et al. 2002). All values plot on a single line with a slope of ~1, suggesting the absence of any significant matrix effect on the lithium instrumental isotopic fractionation (see text).

Page 66: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

53

1

10

100

1000

1 10 100 1000

Standard Reference Material

NBS 610NBS 612DeklabJG1aJB-2BCR1JGB-1SCO

Lith

ium

(pp

m)

Pub

lish

ed V

alue

s

Lithium (ppm)(LA-ICPMS, this study)

Figure 2.2 Standards and Reference Material

Lithium abundance of various standard reference materials (values taken from literature; see text) and internal reference material (SCO and Dekalb; measured by MC-ICPMS) plotted against lithium abundance determined in this study by LA-ICP-MS. NBS 610 was used as the standard reference material (SRM) for all analyses. The value for NBS 610 was determined using NBS 612 as the SRM. Error bars for lithium concentrations from this study are 2, based on standard deviation of replicate measurements (typically 5 or more analyses for each).

Page 67: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

54

Figure 2.3 Photomicrographs of Starting Material and Run Products

(a) Starting material San Carlos olivine mounted in oil, (b) starting material Dekalb diopside mounted in oil, (c) run product olivine from NCOL3 mounted in epoxy, (d) run product diopside from NCDL 4 mounted in epoxy and (e) NCAL 2 mounted in epoxy. LA-ICPMS pits in run product crystals are 50m in diameter. All photomicrographs are taken in plane polarized light at 100x magnification.

Page 68: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

55

100

101

102

103

104

105

106

107

0 10 20 30 40 50 60 70 80

2m-2, 800oC, 68 hrsco

unts

per

sec

ond

Time (seconds)

1.8 x105ppm 43Ca

50 ppm 7Li

laser off laser on

Figure 2.4 Time Resolved Spectra

Example of time resolved spectra for an individual clinopyroxene crystal from a cpx/fluid partitioning experiment at 800oC for 68 hrs. The first ~20 seconds of the analysis was done with the laser shutter in place for background measurements, followed by ~60 seconds of sample ablation. Note that the 7Li signal is consistent with respect to the 43Ca signal, which is an indication of homogeneity, confirming mineral-fluid equilibrium of both major and trace elements.

Page 69: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

56

-6

-5

-4

-3

-2

-1

0

1

0.5 0.6 0.7 0.8 0.9 1

Hart & DunnBrenan 1998bBrenan 1998aBlundy & DaltonBlundy 1998this study

ln D

Li

1000/T

1200 1000 80014001600ToC

0.3

0.2

5

10

14

1.5*

0.5

8

3

1.3-1.6

Figure 2.5 lnDLi cpx/fluid vs 1000/T

A plot of ln DLicpx/fluid as a function of 1000/T for cpx/fluid partitioning measured in this study, demonstrating the temperature dependence of lithium partitioning between clinopyroxene and hydrous fluids. A linear regression of the data yields: ln DLicpx/fluid = -7.38 + 7.04 * 1000/T (R2 = 0.98) where T is temperature in Kelvins. Neither NCDL1, diopside #3 (unequilibrated samples) nor NCDLR (reversal) were used in the regression. Also shown are the data from the experiments of Hart and Dunn (1993), Brenan et al. (1998a, 1998b), Blundy and Dalton (2000). The data points are labeled with the wt. % Al2O3 content of the run product clinopyroxene and suggest that lithium may be coupled with Al3+ as a substitution mechanism in clinopyroxene. Error bars for the partition coefficients measured in this study are 2, based on the standard deviation of n analyses (see Table 2.5).

Page 70: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

57

-3

-2.5

-2

-1.5

-1

-0.5

0

0.4 0.5 0.6 0.7 0.8 0.9

this studyBrenan 1998aTaura 1998Blundy & DaltonZanetti 2004

lnD

Li

1000/T (K)

1400 1200 1000 9001600

ToC

1800

1.3

19*

0.4na

98

10

Figure 2.6 lnDLi ol/fluid vs 1000/T

A plot of ln DLi ol/fluid as a function of 1000/T measured in this study, demonstrating the temperature dependence of

lithium partitioning between olivine and hydrous fluids. A weighted linear regression of the data (excluding the reversal, NCOLR) yields the relationship: ln DLi

ol/fluid = -5.93 + 6.46 * 1000/T (R2=0.82) where T is temperature in Kelvins. Also shown are the experiments of Brenan et al. (1998a), Taura et al. (1998), Blundy and Dalton (2000) and Zanetti et al. (2004). The data points are labeled with the wt. % FeO content of the run product olivine, and suggest that lithium couples with Fe3+ as an exchange mechanism in olivine. The experiment of Zanetti et al. (2004), 19*, was conducted at very reducing conditions and most likely contained less Fe3+ than the experiments in this study. Error bars for the partition coefficients measured in this study are 2, based on the standard deviation of n analyses (see Table 2.5).

Page 71: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

58

-30000

-25000

-20000

-15000

-10000

-5000

0.9 0.92 0.94 0.96 0.98 1

RT

lnD

Li

Xan

Figure 2.7 Anorthite/Fluid Lithium Partitioning

Plot of RTln DLi as a function of the mole fraction of anorthite in plagioclase (XAn). Lithium partitioning between anorthite and fluid shows a linear relationship with a negative slope over the range of XAn from 0.91 to 0.99 indicating that lithium is more compatible in albite than in anorthite. Linear regression of the six partitioning experiments yields the relationship, in Jmol-1: RTlnDLi = 162,170 – 188,820(XAn) (R

2=0.96). The primary control on lithium partitioning between plagioclase and hydrous fluids is the composition of the feldspar (see text). 2 errors for the partition coefficients measured in this study are smaller than the symbol used, and based on the standard deviation of n analyses (see Table 2.5).

Page 72: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

59

0.01

0.1

1

10

100

750 800 850 900 950 1000 1050 1100 1150

two phase exp

DiLi

ol/fluid/DLi

cpx/fluid

Li, p

pm, o

livin

e/cl

inop

yrox

ene

Temperature (oC)

reversal

Figure 2.8 Olivine/Clinopyroxene Lithium Partitioning

Dol/cpxLi as a function of temperature (oC) for two phase experiments (open symbols) and the ratio of DLi

olivine/fluid/ DLi

cpx/fluid calculated from single phase experiments (solid symbols). The constant Dol/cpxLi versus temperature from

800oC to 1000oC further suggests that olivine/clinopyroxene partitioning of lithium is independent of temperature. The partitioning of lithium between olivine and clinopyroxene calculated from single-phase fluid partitioning experiments is the same as those determined from the two-phase experiments. Error bars for the partition coefficients measured in this study are 2, based on the standard deviation of n analyses (see Table 2.5).

Page 73: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

60

0.1 1 10 100 10000.1

1

10

100

1000

10000

silicate meltcarbonatite melt

equilibrated xenolithshydrous fluidsilicate meltcarbonatite melt

hydrous fluid

low fO2

high REEreversal

Li clinopyroxene (ppm)

Li o

livin

e (p

pm)

Experiments Natural Samples

Figure 2.9 Lithium Partitioning From Mantle Xenoliths and Experimental Studies

Lithium abundances in olivine and clinopyroxene from mantle xenoliths (Seitz and Woodland 2000, Paquin and Altherr 2002, Woodland et al. 2002, Woodland et al. 2004) and experimental data (this study, Brenan et al 1998b, Blundy and Dalton 2000). Equilibrated olivine-clinopyroxene pairs tend to fall on a linear ~ 1:1 trend, those metasomatised by silicate melts, and hydrous fluids fall below the trend line, and those altered by carbonatite melt fall on and above the trend. The experimental data have higher concentrations of lithium present due to experimental and analytical requirements nevertheless they appear to follow the same trends. A silicate melt equilibrated olivine-clinopyroxene pair from Brenan et al. (1998b) plots slightly below the line projected from the equilibrated mantle xenoliths, and carbonatite melt olivine-clinopyroxene pairs from Blundy and Dalton (2000) fall on the projected line, above the silicate melt experiment. Olivine-clinopyroxene pairs equilibrated with hydrous fluids in this study fall above the equilibrated mantle xenoliths trend, not below where the hydrous fluid metasomatised samples plot (see text).

Page 74: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

61

-25

-20

-15

-10

-5

0

5

10

0.5 1 1.5 2 2.5 3 3.5 4

this study

Wunder et al 2006

reversal

Chan et al 1993Chan et al 1992

7 L

i cpx-

fluid (

o/o

o)

1000/T(K)

-15

-10

-5

0

5

10

0.6 0.8 1 1.2 1.4 1.6 1.8

Figure 2.10 Mineral/Fluid Isotopic Fractionation of Lithium

∆7Licpx-fluid (‰) as a function of 1000/T (K). Results are shown from this study, the spodumene-fluid experiments of Wunder et al. (2006), the basalt-seawater measurements of Chan et al. (1993; 350oC) and Chan et al. (1992; 2oC). Also shown is the regression line constrained by the experiments of Wunder et al. (2006). Error bars for this study are 2, based on counting statistics of the analyses.

Page 75: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

62

10 10010

100

1000

10000

100000

1000000

z (km)

year

s

Vf = 100 m/yr

Vf = 10 m/yr

Vf = 4 m/yr

Lithium

Boron

max t from U-series

max t from Ra- Th

Figure 2.11 Time for Li and B Transport to Top of Column

Plot of the time (years) required for an element front to reach the top of a chromatographic column as a function of column height (km) for a fixed column length of 100 km. Time constraints given by U-series and Ra-Th disequilibria are also shown for reference (10,000 to 40,000 years). Because lithium is more compatible in mantle minerals than boron, a lithium signal transported by fluids percolating through the mantle will lag significantly behind the boron signal. The times of transport are calculated from life of the column with respect to each element, in other words; how many times the column can be re-used before its capacity to take up more lithium or boron is reached. The lithium signal will not reach the melt source because it will preferentially partition into the mantle wedge, relative to other fluid mobile elements (such as boron).

Page 76: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

63

0

5

10

15

20

25

30

35

40

00.20.40.60.81

7 Li fl

uid

Fraction of Li in fluid remaining

0.996

0.999

0.998

Izu back arc

Izu fore arc

fraction of Li in fluid after equilibration with mantle

MORB

Figure 2.12 Evolution of the Slab Derived Fluid by due to Rayleigh Distillation

Plot of the evolution of δ7Li of a fluid percolating through the mantle as a function of fraction of lithium in the fluid remaining. The change in the isotopic composition of a slab-derived fluid with an initial δ7Li of +10 ‰ (as estimated for the Izu arc by Moriguti and Nakamura, 1998) during percolation through a mantle column. The isotopic composition of the altered fluid will increase and become heavier than the initial slab derived fluid with any amount of fluid:rock interaction. The mass fraction of Li remaining in the fluid after equilibration with the mantle column is also shown for reference. Because lithium is so readily taken up by the mantle wedge, only a small amount of lithium will remain in the fluid. This will result in extreme fractionation and lead to very 7Li-rich fluids at the top of the melt column. The isotopic composition of both the fore arc and back arc lavas of the Izu arc are shown for reference, as is the 7Li of MORB.

Page 77: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

64

0 200 400 600 800 1000 1200 14000.8

0.9

1

1.1

1.2

1.3

1.4

T (oC)

fluid

de

nsi

ty (

g/c

m3)

Benioff Zone

(3.6 GPa, 700oC)

Franciscan Subduction

(0.9 GPa, 300oC)

Alpine Subduction

(1.1 GPa, 500oC)

Li[>4.5]

Li[<4.5]

Figure 2.13 Lithium Coordination and P-T Paths

A plot of fluid density as a function of temperature, with the average Li-coordination (from Jahn and Wunder, 2009) shown as degree of shading. Also shown are the fluid densities calculated with the CORK-EOS (Holland and Powell, 1991) using Perple_X program (Connolly, 2005) for Franciscan and Alpine subduction zone P-T paths (from Ernst, 1986) and an ascent path from the slab (Benioff zone) and a fore-arc volcano, ~125 km above the slab (calculated from the thermal model of Peacock, 1993). 6Li-rich fluids will be generated when mineral fluid interaction occurs at depths greater than 125 km, assuming a mantle-slab interface temperature of 700-800oC. As the fluids percolate up through the mantle wedge and fluid density decreases and the coordination of lithium in the fluids will decrease consequently, 6Li will once again preferentially fractionate into the mineral phase and the fluids will become heavier.

Page 78: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

65

-15

-10

-5

0

5

10

15

0 0.5 1 1.5 2 2.5

Fin

al I

soto

pic

Co

mp

osi

tion

of

the

Wa

ll R

ock

Fluid/Rock

10 O/oo

0 O/oo

-10 O/oo

range of Izu arc lavas

range of most arc lavas

MORB

Figure 2.14 Evolution of 7Li of Mantle Wedge due to Hydrofractures

A plot of δ7Li of the altered wall rock as a function of fluid/rock ratio. The final isotopic shift of the wall rock depends on the extent of reaction between the mantle and the slab-derived fluids: if the fluid/rock ratio is high, the isotopic composition of the vein/hydrofracture wall rock will approach complete equilibrium with the fluid; equal to 1 ‰ less than the fluid (R-F of -1 ‰). The entire range of δ7Li values found in the Izu arc lavas (light grey shaded area) or other arc lavas (dark grey shaded are) can be achieved by metasomatising the mantle with fluids that have δ7Li between 0 and +10 ‰ and fluid/rock ratios >1.2. To generate arc lavas with δ7Li greater than MORB the isotopic composition of the metasomatising fluid must have δ7Li > +6 ‰. These values are reasonable given the isotopic composition of subducted material and typical fluid/rock ratios for hydrofractured zones.

Page 79: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

66

3 Lithium Diffusion

3.1 Introduction

To date, there have been few studies on lithium diffusion in minerals, but observations of natural

samples and limited experimental work indicates that lithium diffusion may be extraordinarily

fast. For example, Berlo et al. (2004) reported rapid mobilization of lithium in plagioclase

phenocrysts from the 1980 eruption of Mount St. Helens in Washington, USA. Plagioclase

phenocrysts erupted prior to the degassing event contained ~14 ppm lithium, whereas those

erupted immediately after contained ~5 ppm. The implication is that the magma lost a significant

amount of lithium in a seven-day period, as recorded in the lithium content of the plagioclase

phenocrysts. Similarly, Kent et al. (2007) interpreted the lithium contents of plagioclase

phenocrysts from the Mount St. Helens 2004 dome lavas as having increased due to the addition

of pre-eruptive lithium rich vapour phase within one year of the dome lava eruptions.

Recent high spatial resolution analyses of lithium isotopes have revealed significant isotopic

heterogeneity at the grain-scale that is suggestive of diffusive exchange. Both olivine and

clinopyroxene phenocrysts from Solomon Island lavas are zoned with respect to lithium and δ7Li

(Parkinson et al., 2007). The rims of the phenocrysts are enriched in lithium compared to the

cores, and the δ7Li decreases from core to rim by as much as 20 ‰ in a W-shaped profile

(Parkinson et al., 2007). A similar pattern was also observed by Jeffcoate et al. (2007) who

measured a 40 ‰ variation in a single orthropyroxene crystal from a San Carlos xenolith (Figure

3.1). The extreme grain-scale variability exhibited by lithium and lithium isotopes is not limited

to terrestrial samples. The basaltic lunar meteorite, NWA 479, examined by Barrat et al. (2005)

contains olivine and pyroxene phenocrysts that also display a wide range of δ7Li values (+2.4 to

+15.1 ‰ in olivine and -0.2 to +16.1 ‰ in pyroxene). Beck et al. (2004) examined pyroxenes in

the shergottite meteorite NWA 480, and found zoning of δ7Li from -17 ‰ in the cores to +10 ‰

in the rims but an absence of lithium compositional variation within the same crystals.

Bulk analyses of lithium isotopes in mantle xenoliths may also be reflect heterogeneity due to

kinetic effects operating on the grain scale. For example, Rudnick and Ionov (2007) reported

highly variable δ7Li in clinopyroxene and olivine grains in peridotite xenoliths from eastern

Page 80: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

67

Russia. The δ7Li values ranged from -0.8 to -14.6 ‰ for clinopyroxene and -1.7 to +11.9 ‰ for

corresponding olivine, and olivine/clinopyroxene distribution coefficients varied from 0.2 to 1.0,

which is somewhat lower than previously estimated for equilibrium partitioning.

Diffusive fractionation of lithium isotopes would explain, in some cases, the apparent

disequilibrium that exists between olivine and clinopyroxene pairs of mantle xenoliths. If lithium

diffusion is significantly faster in one mineral phase compared the other, then one phase would

be more affected by introduction or loss of lithium during transport and cooling of the xenolith.

Bulk analyses of olivine and clinopyroxene pairs from a xenolith from the Vitim volcanic field

found δ7Li to range from -17 to -18 ‰ in the pyroxenes with a δ7Li of +6 ‰ in the

corresponding olivine (Ionov and Seitz, 2008). Bulk measurements of olivine phenocrysts in

primitive magmas from a variety of localities found a relatively uniform δ7Li of +3.2 to +4.9 ‰;

however, measurements of clinopyroxene yielded highly variable δ7Li (+6.6 ‰ to -8.1 ‰;

Jeffcoate et al., 2007).

The extreme fractionation of lithium isotopes documented in these studies indicates a kinetic

mechanism rather than an equilibrium process (Lundstrom et al., 2005; Beck, 2006; Jeffcoate et

al., 2007; Parkinson et al., 2007; Rudnick and Ionov, 2007; Marchall et al., 2007). This kinetic

effect has been experimentally demonstrated by Richter et al. (2003) for diffusion of lithium in

silicate melts. Richter et al. (2003) found that 7Li could be fractionated from 6Li by tens of per

mil during diffusion between molten basalt and rhyolite or when diffusing through hydrous

fluids (Richter et al., 2006).

Although estimates of lithium diffusion coefficients have been made from gradients measured in

natural samples (Parkinson et al., 2007), to date there have been few studies to determine the

diffusion coefficients or mechanisms of lithium diffusion in common rock forming minerals. Pell

(1960) measured lithium diffusion in a p-type silicon crystal (a semicounductor material with a

deficit of electrons, therefore allowing positively charged species, to move through the material)

and investigated the effect of diffusion on the 6Li/7Li ratio. Giletti and Shanahan (1997)

measured the diffusion coefficients of various alkali elements in plagioclase feldspars at high

temperature. Coogan et al. (2005) measured the diffusion coefficient for 6Li in clinopyroxene

between 800oC and 1100oC.

Page 81: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

68

The current study presents new measurements of lithium diffusion in pyroxene and olivine

between 800oC and 1000oC, and provides the first demonstration of isotopic fractionation

induced by solid-state diffusion in a geological material. These data are critical to understanding

the origin of mineral zonation patterns and isotopic variations that have been documented in

natural samples. With this information, the times-scales of processes and events recorded by

mineral zonation patterns and isotopic variations can be extracted.

3.2 Experimental Methods

Lithium diffusion was measured parallel to the c-axis in clinopyroxene with experiments done at

atmospheric pressure, at temperatures between 800oC to 1000oC and controlled fO2. Starting

materials consisted of natural gem quality clinopyroxene (diopside) crystals from Dekalb, New

York and Kunlun, China. Table 3.1 gives the composition of the starting materials. The mineral

samples were oriented based on crystal habit and ~3 mm thick slabs were made by sectioning

perpendicular to the c-axis using a diamond saw. Slabs free of inclusions and alteration were

selected, cleaned in acetone and rinsed with ultra-pure water in an ultrasonic cleaner. The slabs

were then polished with diamond and alumina paste to 0.3 m. Slabs were sealed in silica tubes

with a solid state buffer (Ni-NiO, MnO-Mn3O4 or Fe3O4-Fe2O3) and ‘pre-annealed’ for 48 hrs at

the same fO2 and temperature conditions as subsequent diffusion experiments. This technique

was meant to homogenize lithium concentrations and equilibrate point defects in the crystals

prior to the diffusion experiments.

Examination of the annealed slabs revealed that the polished surface had roughened and that the

total concentration of lithium in the crystal had decreased uniformly, with the exception of the

outer 50 m of the crystal, which showed greater depletion than the remainder of the crystal

(Figure 3.2). The annealed slabs were then re-polished with diamond and alumina paste to 0.3

m to remove the depleted zone. The re-polished slabs were then cleaned in acetone and rinsed

with ultra-pure water in an ultrasonic cleaner. A single slab was then packed into a platinum

capsule with either a lithium-source, LiCl + 4 wt. % powdered Dekalb diopside for a diffusion-in

experiment, or a lithium-sink, NaCl + 4 wt. % powdered Dekalb diopside for diffusion-out

experiments. The platinum capsule was crimped shut, and loaded into a silica tube with a solid

state buffer. The silica tubes were gently heated in a water bath to ~100oC, and evacuated for a

minimum of 20 minutes, and sealed with a blowtorch. The ampoule containing the sample was

Page 82: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

69

equilibrated in a box-type furnace for the duration of the experiment (3 minutes to 16 days) and

quenched by removal from the furnace and air-cooling.

The Pt capsule was recovered, the crystal slab extracted and rinsed with ultra-pure water in an

ultrasonic cleaner to remove the salt rind, mounted in epoxy, ground to half-thickness parallel to

the c-axis, and polished for analysis. This method allowed for measurement of two diffusion

profiles from each slab, on either edge of the slab towards the centre of the crystal. In one case

(sample Kcpx-900.72), the slab tilted slightly while the epoxy set and grinding of this sample

truncated one side, resulting in a single diffusion profile for this run. In some cases (samples

SCO-12, Kcpx-12, Kcpx-MH), the molten salt did not completely wet the slab, again resulting in

only a single profile measurement.

A ‘zero time’ experiment was carried out to assess the effects of the sample preparation, loading

procedure, and temperature run up on the lithium profile of the slabs. A slab that had been

previously analyzed by LA-ICPMS before and after annealing (sample from Figure 3.2) was

packed in NaCl, loaded, sealed into a silica tube as described above, and placed in a box-type

furnace at 1000oC. When the internal temperature stabilized at 1000oC, after 3 minutes, the

sample was removed from the furnace and extracted from the silica tube. This slab was prepared

as above, and lithium concentration was analyzed along the midpoint of the cross-section. No

measurable change in lithium was observed (Figure 3.3), indicating that the sample pre-treatment

and initial heating did not contribute to the diffusion profiles.

3.3 Analytical Techniques

3.3.1 Major Element Analyses

Major element compositions of the starting materials were obtained using the University of

Toronto’s Cameca SX50 Electron Probe X-ray Microanalyzer (EPMA). An accelerating voltage

of 15 kV and a focused 20 nA beam was used for all samples. Diopside, basalt, anorthite and

natural and synthetic oxides were used as standards. X-ray intensities were converted to

concentrations using ZAF and Phi-Rho-Z calculations. The reported errors are the 1 variations

of the reported number of analyses (n).

Page 83: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

70

3.3.2 Lithium Analyses

3.3.2.1 LA-ICPMS

Elemental concentrations in the run-product clinopyroxene were measured by laser ablation

inductively coupled plasma mass spectrometry (LA–ICP–MS) at the University of Toronto,

using a frequency quintupled Nd:YAG laser operating at 213 nm, coupled to a VG PQExcell

quadrupole ICP-MS. The laser was operated at 10 Hz and 3 mJ, with He flushing the ablation

cell to enhance sensitivity (Eggins et al., 1998), and produced spot sizes ~25 m in diameter and

~ 25 m deep. At the start of each session, the quadrupole lens settings were adjusted to

maximize the signal on mass 7 during ablation of NIST 610. Data were collected as time-

resolved spectra with background levels determined by counting for 20 s prior to the 60 s of

sampling by laser ablation. Analyses were collected in blocks of 20, with the first and last two

spectra acquired on standard reference materials (SRM). SRMs employed include NIST 610

silicate glass, NIST 612 silicate glass, and “in house” standards of Kunlun diopside and Dekalb

diopside. Table 3.2 lists measurements of reference materials. Data reduction was performed off-

line using the GLITTER software package. Ablation yields were corrected by referencing to the

known concentration of 43Ca that was determined previously by electron microprobe analyses.

The precision for concentration measurements is better than ±10 %. The length of the diffusion

profile was determined by measuring the distance from the edge of the slab to the edge of the

spot using the digital measurement tool included with the laser operating software. The precision

of a repeated measurement is ±5 %.

3.3.2.2 Secondary Ion Mass Spectrometry (SIMS)

In situ analyses of the isotopic composition of the run product clinopyroxene for experiment

kcpx-2 were obtained using the Cameca IMS 3f ion microprobe at Lawrence Livermore National

Laboratory, Livermore, California. Secondary ions were generated by bombardment with a 5-12

nA negatively charged 16O primary beam, accelerated through –12.5 kV and focused to ~20 μm.

The positive secondary ions were accelerated through 4.5 kV. 6Li and 7Li were measured with a

mass resolving power of 1011, and no energy offset was applied. The background (mass 5.8), 6Li, and 7Li were counted on an electron multiplier for 2 s, 10 s, and 2 s respectively over 120-

400 counting cycles, depending on count rate. The isotopic composition of the sample is

expressed as per mil values relative to the core of the slab where,

Page 84: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

71

(27) d7Li slab = [(7Li/6Li)slab - (7Li/6Li)core] /(

7Li/6Li)core x 1000

where ‘slab’ refers to the 7Li/6Li ratio from the salt-crystal interface and inwards toward the

core, and ‘core’ refers to the geometric centre of the slab, which is where the lowest 7Li/6Li were

recorded. Because the absolution value of 7Li is not needed here, this notation can used to

highlight the change in the 7Li/6Li of the salt-crystal interface (and inwards) with respect to the 7Li/6Li value of the core (which is assumed to be the original and unaltered 7Li/6Li value of the

slab). Using this notation also removes any ambiguity arising from the available standards not

matching matrix of the sample. The 2 precision of the 7Li/6Li measurements is based on

counting statistics and is approximately ±4 ‰.

3.3.3 Data Reduction

The experimental method was designed to provide a constant concentration of lithium at the

sample surface, a uniform initial concentration of lithium in the sample crystal with a fixed

diffusion boundary, and a semi-infinite diffusion medium. The solution of Fick’s Second Law

for a semi-infinite medium with a planar surface, and these boundary conditions is given by:

(28)

21,

2 Dt

xerfcCC otx

where Cx,t is the concentration of lithium at distance x (m) from the interface at time t (sec), Co is

the concentration of lithium in the crystal at x = 0, D is the diffusion coefficient (m2/s).

Following the method of Harrison and Watson (1983) the data are inverted through the error

function given:

(29)

21

,

)(21

Dt

x

C

Cinverf

o

tx

Fitting of the concentration data is accomplished by plotting x against the inverse error function

of (Cx / Co), and adjusting Co to force the intercept through the origin. This yields a straight line

with a slope of ½√(Dt), determined by least-squares regression. The modeled diffusion profile,

along with a typical concentration profile determined from LA-ICPMS, is shown in Figure 3.4a.

The resulting fit of the inverse error function is shown in Figure 3.4b.

Page 85: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

72

3.4 Results

3.4.1 Diffusion in Clinopyroxene

Table 3.3 provides a summary of the experimental conditions and the calculated diffusion

coefficients. Run product crystal slabs emerged from the Pt tubing coated with recrystallized

NaCl or LiCl. The surfaces of the slabs developed a roughened texture, but the major element

chemistry of the crystal did not appear to be changed. Addition or loss of sodium or calcium

would be the most likely exchange; however, given the low diffusion rates of those cations

(Dimanov and Sautter, 2000; Brady and McCallister, 1983), and the short duration of the

experiments, very little exchange is expected. EDS X-ray map analysis of Si, Al, Ca, Na, Mg and

Cl for the highest temperature (1000oC) diffusion-out and diffusion-in experiments confirm this

The EDS X-ray map of kcpx-R, a diffusion in experiment, is shown in Figure 3.5.

Table 3.3 lists the diffusion coefficients calculated from the concentration profiles for each

experiment. Figure 3.6 shows the diffusion coefficients calculated from a series of experiments

at 1000oC and fO2 of NNO, with durations of 2 to 12 hours. The measured values are within error

of each other, demonstrating that the diffusion coefficients are independent of time. Lithium

diffusion coefficients are also independent of the diffusion gradient, as values are the same

whether the flux of lithium is into or out of the crystal. Diffusion coefficients are plotted as a

function of inverse absolute temperature in Figure 3.7, along with values measured by Coogan et

al. (2005). A least squares regression line can be fit to the data for experiments conducted at an

fO2 of NNO and temperature ranging from 800oC to 1000oC with the Arrhenius relationship:

(30) log DLicpx (m2/s) = 5.92 (±8.51) – 2.30 (±1.06)*10,000/T R2 = 0.93

From this, a pre-exponential factor of 8.31 x 105 m2/s and activation energy of 442 ±10 kJ mol-1

is determined.

3.4.2 fO2 Series Experiments

With the exception of Kcpx-MnOMn, and Kcpx-MH, all of the experiments were done at a log

fO2 of -10.3, buffered by Ni-NiO (calibration of O’Neill and Powceby, 1993a). Kcpx-MnOMn

was conducted at a log fO2 of -6.7 (calibration of O’Neill and Powceby, 1993b), and Kcpx-MH

Page 86: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

73

was conducted at a log fO2 of -5.3, buffered by magnetite-hematite (calibration of Hemingway,

1990). The results of the experiments done at different fO2 are shown in Figure 3.8, along with

the 1000oC data of Coogan et al. (2005). In general, the diffusion coefficients are within error of

each other; however, there appears to be a slight but systematic trend of increasing values with

decreasing oxygen fugacity. A weighted least squares regression line can be fit to the data

resulting in the relationship:

(31) log DLicpx (m2/s) = -15.2 (±1.7) – 0.28 (±0.18) x log fO2 R2 = 0.75

3.4.3 Diffusion in Olivine

The low (2.5 ppm) lithium content of the San Carlos olivine starting material, relative to Kunlun

diopside (42 ppm) and Dekalb diopside (8 ppm), presented an analytical challenge in

determining diffusion coefficients by the diffusion-out method. Instead, a ‘Li-in’ method with a

Li-chloride source was used. Run durations were restricted as the high lithium vapour pressure

generated by the Li-source had a corrosive effect on the sealed silica tubes, resulting in

catastrophic failure of the ampoules after 12 hours. Nonetheless, a single measurement of lithium

diffusion into olivine was made that can be directly compared to values for clinopyroxene

(Figure 3.9). At 1000oC and fO2 of NNO the measured lithium diffusion into olivine is log D = -

14.1 (±0.12) m2/s, two orders of magnitude slower than the value for clinopyroxene at the same

conditions. The heating time of the olivine experiment was 12 hours and lithium concentration of

the crystal was unchanged beyond 100 m from the crystal-lithium source interface, whereas

given 12 hours at similar temperature and oxygen fugacity conditions, the concentration gradient

in a clinopyroxene continued to ~700 m from the crystal-source interface (Figure 3.4a).

3.4.4 Kinetic Fractionation of 7Li/6Li

An analysis of the 7Li/6Li ratio in experiment kcpx2 was carried out to investigate the

fractionation of lithium isotopes during diffusion. This sample was pre-annealed at an fO2 of

NNO at 1000oC for 48 hours, but unlike the other samples it was not re-polished to remove the

50 m zone of disturbed lithium (see above). Similar to other Li-out experiments the sample was

packed in NaCl, sealed in a Si-tube with NNO oxygen buffer and heated at 1000 oC for two

hours.

Page 87: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

74

The diffusion coefficient for 7Li, calculated from the concentration profile acquired by LA-

ICPMS, is 5.28 x 10-13 m2/s, which is within the range of other 1000oC measurements from this

study. The measured lithium concentration profile is shown in Figure 3.10a, along with the

modeled diffusion profile, and represents lithium loss during the combined ‘pre-annealing’ step

and ‘Li-out’ diffusion experiment. The modeled diffusion profile which best fits the data is that

for 2 hours of diffusion time. This is consistent with the finding that the ‘pre-annealing’ step

served to homogenize the lithium concentrations in the slab.

The Co value that was determined from the fit of the 7Li data was then used, together with the

isotopic ratio that was measured by SIMS analysis, to generate a 6Li concentration profile. This

profile was used to calculate a diffusion coefficient of 5.44 x 10-13 m2/s. The modeled isotopic

gradient is shown in Figure 3.10b, along with the isotopic gradient measured by SIMS, and

reveals a +7 ‰ change between the NaCl-crystal interface and the centre of the slab.

Although the starting material for this experiment was not characterized for pre-existing isotopic

gradients, it is likely any gradients would have been eliminated in the subsequent annealing steps

as a result of the rapid lithium diffusion documented in this study. Rather, the main uncertainty

when interpreting this result is the timescale over which the gradient was produced. The

experiment was subject to two episodes of diffusive loss: ‘controlled’ loss into the NaCl lithium-

sink and an ‘uncontrolled’ loss during the ‘pre-annealing’ phase. Despite this shortcoming, there

are two notable aspects of the measured gradient. First, the isotopic gradient appears to penetrate

farther into the crystal than the chemical gradient, ~250 m versus ~100 m. This is consistent

with mass balance considerations as the bulk of the elemental profile will be dominated by the

more abundant 7Li, which is expected to diffuse more slowly than 6Li. The second and most

significant aspect of this profile is the large isotopic difference produced between the core and

the rim, ~ +7 ‰.

3.5 Discussion

3.5.1 Effect of fO2 on lithium diffusion in clinopyroxene

As described above, a negative correlation of lithium diffusion rates with fO2 is shown in Figure

3.8, with log DLicpx decreasing from -12.4 to -13.8 as oxygen fugacity increases from log fO2 -12

to log fO2 -5.5. Previous studies have demonstrated that the magnitude of the effect of fO2 on Pb

Page 88: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

75

and Ca diffusion in clinopyroxene is about two log units in D-value over a range of ten log units

of fO2 (Cherniak, 2001; Dimanov et al., 1996). Given that the precision of this study is about 1

log unit (2), an effect of fO2 on lithium diffusion in clinopyroxene would just be discernable

from this data.

The negative dependence of lithium diffusion on fO2 is similar to that determined for Ca self-

diffusion in clinopyroxene by Dimanov et al. (1996), and in contrast to the positive fO2

dependence determined for Pb diffusion in clinopyroxene (Cherniak, 2001). Cherniak (2001)

interpreted the positive fO2 dependence of Pb diffusion in clinopyroxene as the result of the

oxidation of Fe2+ to Fe3+ creating point defects, or vacancies on the crystal lattice, which

accommodates the Pb cation as it diffuses through the crystal. According to Dimanov et al.

(1996), the negative fO2 dependence of Ca self-diffusion suggests an interstitial mechanism for

diffusion, where Ca moves through interstitial sites of the crystal lattice, and is not displacing

other cations in normal lattice sites. Tsai and Dieckmann (2002) describing the relationship

between oxygen content and point defects in olivines demonstrated how an increase of fO2 would

result in the oxidation of Fe2+ to Fe3+, thereby creating intrinsic point defects, or vacancies, in the

crystal lattice via the reaction:

(32) 6Fe2+o + 3O2 = 4Fe3+

o + 2Vo + 2FeO

where V is a vacancy in an octahedral site (o). Following the treatment of Ganguly et al. (2007),

interstitial diffusion of Li+ can be written as:

(33) Li+o + Vi ↔ Li+ i + Vo

where Li+ is the lithium cation in an octahedral site (o) or interstitial site (i), and V is either an

octahedral coordinated vacancy (o) or an interstitial vacancy (i). An increase in Vo would result

in an decrease of Li+i, and an increase in Li+

o, making interstitial diffusion less favorable.

The partitioning of lithium into olivine and clinopyroxene provides some insight into the lithium

transport process. Lithium incorporation into olivine or clinopyroxene appears to be a coupled

substitution with trivalent Fe or Al for Mg2+ or Fe2+ (see Chapter 2). Increasing the proportion of

trivalent Fe may result in a decrease in the degree of ‘misfit’ between the Li+ ion and a potential

site on the normal crystal lattice, thereby reducing the possibility of interstitial movement or

Page 89: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

76

increasing the activation energy required for a jump. Given the small size of the lithium cation,

interstitial diffusion may be plausible. However, it should be noted that the effect of decreasing

diffusion with increasing fO2 is not large, and indicates that this interstitial mechanism is

probably only a minor component of the total diffusive flux.

3.5.2 Comparison with other lithium diffusion studies

Previous work on lithium diffusion in clinopyroxene by Coogan et al. (2005) employed a natural

diopside crystal as a Li-sink, and a powdered mixture of San Carlos olivine and 6Li enriched

Li2CO3 and Li2SiO3 as the Li-source. The diopside crystals and Li-source were then packed into

Al-crucibles and heated in a gas-mixing furnace using a CO and CO2 mix to control fO2. Unlike

this study, Coogan et al. (2005) did not pre-anneal the diopside slabs to pre-equilibrate point

defects under fO2 and temperature conditions of the diffusion experiments. Also, some of the

experiments of Coogan et al. (2005) were buffered at fO2s that differed from those used in this

study by as much as four log units (e.g. at 900oC). When equivalent fO2 conditions were

employed there is good agreement between the measurements of Coogan et al. (2005) and the

results presented here (Figure 3.7).

Figure 3.11 presents the lithium diffusion measurements for clinopyroxene and olivine from this

study, along with data from previous studies of lithium diffusion in other minerals. Lithium

diffusion in olivine is 2 orders of magnitude lower than in clinopyroxene at the same

temperature. Due to the different activation energies, the data for feldspars (Giletti and

Shanahan, 1997) overlap with the clinopyroxene data only at the lowest temperature

investigated. At high temperatures, lithium diffusivities in anorthite and albite are almost 4

orders of magnitude higher than in olivine or clinopyroxene. The measurements of Pell (1960) in

p-type Si-crystal are an order of magnitude higher than those in feldspars and two orders of

magnitude higher than those in clinopyroxene. Previous researchers have also proposed an

interstitial mechanism for lithium diffusion in both Si-crystal (Pell, 1960) and feldspars (Giletti

and Shanahan, 1997) consistent with the results of this study.

3.5.3 Comparison with diffusion of other cations in clinopyroxene

Figure 3.12 compares the lithium diffusivities measured in this study, and that of Coogan et al.

(2005), with experimentally determined diffusivities for other elements in clinopyroxene.

Page 90: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

77

Lithium diffusivities measured in this study, and by Coogan et al. (2005), are more than five

orders of magnitude higher than Sr (Sneeringer et al., 1984), or Fe –Mn and Mg (Dimanov and

Sautter, 2001). Only hydrogen diffusion is more rapid (Woods, 2000). The large range of

diffusion coefficients measured in clinopyroxene most likely reflects the different mechanisms of

diffusion at work. Previous studies have suggested that the diffusion mechanism for many of the

larger, divalent cations, such as Pb, Sr, Fe, Mg, and Mn is point defect, or vacancy, controlled

(Dimanov and Wiedenbeck, 2006; Azough and Freer, 2000; Cherniak, 1998; -2001, Sneeringer

et al., 1984). In the case of REEs, U and Th an elastic diffusion model is suggested, where

movement is governed by the elastic strain of the crystal lattice, and diffusivities increase with

decreasing charge and radius (Van Orman et al., 2001). Some contribution from a point defect,

or vacancy, controlled mechanism may be generally applicable for clinopyroxene since diffusion

coefficients have been shown to increase in more Fe-rich pyroxenes (Cherniak, 2001; Woods,

2000) which have a greater ionic porosity (Cherniak, 2001 and references therein) or more‘free

space’ in their crystal lattices than Fe-poor pyroxene.

3.5.4 Geological Implications

3.5.4.1 Preservation of Lithium Signatures

Results of experimental measurements of lithium partitioning between mantle minerals and

hydrous fluids show that lithium is only slightly incompatible in olivine and clinopyroxene with

respect to hydrous fluids, and that the lithium content and isotopic signature of slab derived

fluids can be significantly modified during transport through, and interaction with the mantle

wedge (Caciagli, Chapter 2). Given the rapid diffusivities measured in this study, it is reasonable

to assume that mantle minerals in contact with lithium bearing fluids will quickly equilibrate.

This assumption can be quantitatively demonstrated following the treatment of Crank (1975).

The maximum time for centre preservation of lithium concentration in a spherical grain can be

calculated as a function of grain radius at a constant temperature. For diffusion in a sphere, the

centre will retain unaltered lithium concentrations for values of Dt/a2 ≤ 0.03, assuming that the

concentration of the sink/source remains constant at the sphere surface over the annealing

interval (Crank 1975). This relationship can be expressed as:

(34) t = 0.03 / (Da2)

Page 91: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

78

where t is time, D is the diffusion coefficient, at a given temperature, and a is grain radius.

Figure 3.13 shows a plot of maximum annealing times at 1000oC for spherical olivine and

clinopyroxene grains of a given radius (in microns) and provides constraints on the time-scale of

processes which preserve lithium zonation. An essential assumption required for modeling the

effects of interaction between a fluid flux from a subducting slab and the overlying mantle

wedge is that fluid/solid equilibrium must be maintained. This calculation demonstrates that such

an assumption appears to be valid.

The rapid exchange portrayed in Figure 3.13 is also consistent with the evidence from mantle

xenolith studies that transient late stage events, such as entrainment and transport in a magma

followed by eruption and cooling, can perturb clinopyroxene and, to a lesser extent, olivine

compositions (Aulbach et al., 2008; Parkinson et al., 2007; Rudnick and Ionov, 2007). For

example, xenoliths from the Vitim volcanic field have pyroxenes with 2-5ppm lithium and 7Li

of -17 ‰ yet coexisting olivine has 1.2 ppm lithium and 7Li of +6.3 ‰ (Ionov and Seitz, 2008).

‘Normal’ mantle is estimated to contain 1.6 ppm lithium and an average 7Li of +4 to +5 ‰

(Jagoutz et al., 1979; Tomascak, 2004). Equilibrium partitioning of lithium between olivine and

clinopyroxene results in Dol/cpx ~1 (Chapter 2, this study) so the pyroxene and olivine should

contain similar lithium concentrations. The increased lithium content of the Vitim pyroxenes is

likely the result of an influx of lithium during entrainment and cooling; however, the lithium

content and isotopic composition of the olivine is relatively unperturbed (Ionov and Seitz, 2008).

This study has demonstrated that lithium diffusion is almost two orders of magnitude faster in

clinopyroxene than in olivine, therefore the lithium content and isotopic composition of

clinopyroxene will be modified to a greater extent than the coexisting olivine during short-lived

events. The elemental and isotopic disequilibrium evident in the Vitim xenoliths allows for

maximum transport times to be determined. Estimates of alkalic magma ascent rates range from

0.2 -0.5 m/s based on hydrogen diffusion studies (Peslier and Luhr, 2006), to 1.3-2.7 m/s based

on mineral dissolution studies (Brearley and Scarfe, 1987). The estimated depths of the Vitim

xenoliths are 40-50 km (Ionov et al., 1993), therefore a magma ascent rate of 0.5-1.0 m/s will

result in a transport time of ~12-25 hrs. From Figure 3.13, at 1000oC a clinopyroxene grain with

a radius of 1000 m will be re-equilibrated in ~29 hrs, whereas an olivine grain of same size will

preserve its original concentration for almost 1000 hrs. Assuming the temperature of the

entrainment magma was 1000oC, then the time for transport, eruption and cooling of Vitim

Page 92: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

79

xenoliths was at least 29 hrs, to allow time to preturbe the lithium content of the clinopyroxene,

but less than 1000 hrs, to preserve the original olivine signature, consistent with an ascent rate of

~0.5 m/s.

3.5.4.2 Closure Temperature

Knowing the temperature for which mineral grains are ‘closed’ to lithium diffusion allows the

extent of re-equilibration during cooling to be estimated. A closure temperature for lithium

diffusion in clinopyroxene can be determined using the equation for mean closure temperature

given by Dodson (1973; eq. 23):

(35)

dtdTE

aDART

RT

E

a

oc

c

a22

ln

where Tc is the closure temperature, Ea is the activation energy, R is the gas constant, Do is the

pre-exponential factor, a is the characteristic dimension of the crystal (e.g. radius of a sphere or

cylinder or the half thickness of a plane), dT/dt is cooling rate, and A is a dimensionless

parameter relating to the weighted arithmetic mean of the closure temperature and geometry of

the grain (A = 55 for a sphere, 27 for a cylinder, and 8.7 for a plane sheet; Dodson, 1973). Figure

3.14 shows the calculated closure temperature of lithium as a function of characteristic radius for

a spherical grain for cooling rates of 1o, 10o and 100o/Myr. The closure temperatures calculated

for grains ranging from 1 to 1000 m in size is between 425oC and 550oC. This is approximately

100-200oC lower than the closure temperature for Sr in clinopyroxene for grains of equivalent

size (Figure 3.14).

3.5.4.3 The Potential for Re-Equilibration of Lithium Composition

The combination of extremely rapid lithium diffusion, and low closure temperatures results in a

significant potential for homogenization of elemental and isotopic differences between minerals.

In particular, the absence of any observed isotopic fractionation during crystallization and melt

differentiation needs to be re-assessed. Kilauea Iki lava lake, Hawaii formed in 1959 due to a

single eruption of picritic tholeiite magma; a crust formed within a few weeks and the whole lake

crystallized as a closed system afterwards. The resulting rocks span a wide compositional range,

from olivine-cumulates, and olivine tholeiites to ferrodiabase, and silicic veins (Helz et al.,

1989). The isotopic composition of a suite of samples from Kilauea Iki lava lake are plotted as a

Page 93: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

80

function of MgO content in Figure 3.15, and show no significant (> ±2 ‰) fractionation between

the more primitive, MgO-rich rocks and the more evolved, MgO-poor rocks (Tomascak et al.,

1999). However, lithium isotope fractionation is observed in granitic systems during

crystallization at 500-600oC (Tomascak, 2004). Because fractionation factors for stable isotopes

are temperature dependant (Urey, 1947), it has been concluded that crystallization of more mafic

magmas occurs at sufficiently high temperatures, 1200-1000oC, such that appreciable (>1 ‰)

equilibrium fractionation of lithium isotopes does not occur (Tomascak et al., 1999).

Recent experiments have found lithium isotopic fractionation between clinopyroxene and fluids

to persist to high temperatures (Chapter 2, Wunder et al., 2006). Furthermore, the case can be

made that the degree of lithium isotopic fractionation between minerals and fluids should be

comparable to the degree of fractionation that occurs between minerals and melts. Lithium

isotopic fractionation between minerals and fluids depends on the difference in the zero point

potential energy (ZPE) between the phases of interest. 7Li is heavier and has a lower vibrational

frequency, and therefore a lower ZPE than 6Li (Chacko et al., 2001). The phase that will undergo

the greatest reduction in ZPE will preferentially take 7Li over 6Li (Chacko et al., 2001). This has

been demonstrated by Ab initio calculations, which have predicted that during mineral-solution

reactions 6Li, is preferentially incorporated into octahedrally coordinated sites in the solid, and 7Li is preferentially incorporated into the dominantly tetrahedrally coordinated sites in the fluid

(Yamaji et al., 2001). Experimental measurements of lithium isotopic fractionation between

clinopyroxene and fluids (Chapter 2) and spodumene and fluids (Wunder et al., 2006) also

confirm this behavior. Given that silicate melts also have a tetrahedral structure, with lithium in

tetrahedrally coordinated sites (Cormier et al., 1998; Majérus et al., 2003) then a degree of

isotopic fractionation, similar to that which occurs between minerals and fluids, is expected

between minerals and silicate melts. This study found the isotopic fractionation between

clinopyroxene and fluids at 1000oC to be approximately -1‰, which is in agreement with the

empirically determined regression of Wunder et al. (2006), which predicts min-fluid

fractionation of -0.65 ‰ to -1.1 ‰ at 1200oC to 1000oC. Although this fractionation is small,

the overall magnitude of this effect can be significant, resulting in gradients of up to 6-7 ‰

across the Kilauea differentiation sequence, (Chapter 2, Figure 2.12) if closed system Rayleigh

distillation continues to completion.

Page 94: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

81

Surprisingly, significant (> ±2 ‰) lithium isotopic fractionation is not evident in the Kilauea Iki

lava lake rocks. Magmatic differentiation of Kilauea Iki lava lake took place between 1200-1100 oC, with the last veins forming at ~1000oC (Helz et al., 1989). The eruption occurred in 1959,

and samples from the upper portion of the lake were sub-solidus by mid-1969, which gives a

maximum cooling rate of 25 o/yr for the rocks in the upper zone. Because closure temperatures

are dependant on cooling rate (see Equation 35), the closure temperature for grains ranging from

1-1000 m would increase to 600-800oC respectively. Even considering the rapid cooling rate

and increased closure temperatures, the timescales required for re-equilibration of lithium

isotopes by diffusion (weeks) are much less than the timescales required for cooling and

crystallization (years). Thus, it seems plausible that the overall homogeneity of lithium isotopes

between samples from the Kilauea Iki lava lake implies that isotopic re-equilibration occurred

beyond the grain-scale with an external reservoir.

In support of this notion is the existence of a convecting geothermal system within the upper

portion of the lava lake (Hardee et al., 1981). The geothermal system did not extend to the rocks

of the lower part of the lake; however, interaction with a volatile phase is still possible in this

zone via the fractures and pores in these rocks (Hardee et al., 1981). A convecting volatile phase

such as the one that developed in the upper zone of the lava lake, or volatiles moving through

fractures and pores in the rocks of the lower zone, can efficiently transport lithium. Evidence for

lithium transport in a volatile phase is provided by several studies that report changes in the

lithium content of phenocrysts phases from other volcanic conduit systems. For example,

plagioclase phenocrysts erupted prior to the 1980 Plinian eruption of Mout St. Helens contained

~14 ppm lithium, whereas those erupted seven days later contained ~5 ppm. Similarly, Kent et

al. (2007) found that the lithium contents of plagioclase phenocrysts from the Mount St. Helens

2004 dome lavas had increased due to the addition of pre-eruptive lithium rich vapor phase

within one year of the dome lava eruptions. The implication is that the Mount St. Helens magmas

can gain or lose a significant amount of lithium over very short timescales, as recorded in the

lithium content of the plagioclase phenocrysts. Likewise, a study of lithium the Tin Mountain

pegmatite and host rocks found gradients in both lithium concentration and isotopic composition

to persist much farther (>30 m) into the country rock than other alkali elements such as Rb and

Cs, which were limited to <2 m from the contact (Teng et al., 2006). These findings demonstrate

that even in a crystalline rock, when a volatile phase is present, lithium will be extremely mobile;

Page 95: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

82

the estimated lithium diffusivity in the host rock amphibolite interacting with an interconnected

fluid or volatile network along grain edges in the Tin Mountain pegmatite is approximately 2 x

10-16 m2/s at 350oC (Teng et al., 2006). Given the rapid diffusivity of lithium and the efficiency

of lithium transport in a volatile phase, a setting such as the Kilauea Iki lava lake would not be

expected to preserve signatures of isotopic fractionation that might arise from crystallization and

magma differentiation. Isotopic fractionation due to crystallization would be most reliably

investigated by direct experimental measurements.

The potential for lithium re-equilibration during emplacement and cooling of rocks exists in

many different settings. The lithium contents of rocks exhumed at convergent margins,

especially eclogites thought to represent remnants of subducted oceanic crust, have been the

focus of much study with the aim to understand the potential for lithium recycling between the

crust and mantle (Zack et al., 2003; Marschall et al., 2007). The P-T path of exhumed eclogites

must be taken into account when examining the lithium concentrations of these rocks. A

Franciscan type P-T path, where the retrograde and prograde paths are similar, requires that the

rocks be cooled as they are exhumed. In this setting exhumation, and consequently, cooling, is

relatively slow (Ernst, 1988). It is unlikely that these rocks will preserve the lithium signatures

that developed at peak metamorphic conditions, unless the peak metamorphic temperatures are <

500oC, which is the case for blueschist facies rocks. Alpine metamorphic paths are characterized

by nearly isothermal decompression due to very rapid exhumation (Ernst, 1988). In this case, the

lithium signatures displayed in exhumed Alpine eclogites may be representative of those

achieved at peak metamorphic conditions, assuming interaction with retrograde fluids is at T <

500oC. Given the relatively low closure temperature, any post emplacement heating or secondary

metamorphic events must be carefully considered when interpreting lithium concentrations.

3.5.4.4 Diffusion-Induced Isotopic Fractionation

Richter et al. (2003) showed that slightly faster transport of 6Li than 7Li results in diffusive

fractionation in silicate melts. Diffusive fractionation has been suggested as the mechanism for

generating the 20-40 ‰ gradients observed in some phenocrysts and xenoliths. This study has

documented isotopic fractionation due to diffusion in a solid medium. A 7 ‰ gradient was

generated 300 m into the crystal during the course of the 2-hour diffusion experiment (Figure

3.10b). Although preliminary, this result is a demonstration of the magnitude of the isotopic

Page 96: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

83

gradient that can be produced by diffusion in a very short time-scale. This fractionation occurs

because lighter isotopes diffuse faster than heavy ones; the relative diffusivity between isotopes

is given by:

(36) D1/D2 = (m2/m1)

where D1/D2 is the ratio of the diffusivity of isotopes of mass m1 and m2, and the exponent is

empirically determined (Richter et al., 1999; Richter et al., 2003; Richter et al., 2009). A of ½

is characteristic of self-diffusion in a gas, for diffusion in liquids or melts ranges from 0.025 to

0.215, depending on the isotopes (Ge isotopes have < 0.025; Ca isotopes, of 0.05 to 0.1; Li

isotopes, ~0.215; Richter et al., 1999; Richter et al., 2003).

A -value was estimated from the calculated diffusion coefficients of 7Li and 6Li. A diffusion

coefficient for 7Li was calculated from the measured concentration profile following the method

described in section 3.3.3. The modeled diffusion profile for 7Li was then used, together with the

isotopic ratio that was measured by SIMS analysis, to generate a 6Li concentration profile. This

profile was used to calculate a diffusion coefficient of 5.44 x 10-13 m2/s. A value of 0.2 was

determined from the ratio of diffusivities calculated from the concentration profile and isotopic

profile of experiment kcpx-2. This is similar to the of 0.215 determined for lithium isotopes

diffusing in melt by Richter et al. (1999).

This study has shown that isotopic fractionation can occur due to diffusion of lithium in a

mineral. Diffusive isotopic fractionation has been proposed as the means to produce the extreme

isotopic compositions and complex isotopic profiles observed in xenoliths and phenocrysts

(Parkinson et al., 2007; Jeffcoate et al., 2007). In Figure 3.16 one half of the W-shaped profile

observed in an orthropyroxene crystal from a San Carlos xenolith (Figure 3.1; Jeffcoate et al.,

2007) is modeled using the diffusivities and the (0.2) determined from this study. The lithium

isotopic gradient is calculated for 10 hrs, 1000oC and with a surface concentration/initial mineral

concentration (Cs/Co) of 3.25. Because the lithium concentration at the surface of the grain is

higher than the interior of the grain, the faster diffusion of 6Li into the grain results in a 7Li

profile with a ‘trough’ or one-half of a W-profile. The calculated model is a very good fit for the

orthopyroxene data and demonstrates that lithium diffusion in a solid is a viable mechanism for

producing extreme and complex isotopic gradients.

Page 97: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

84

3.6 Conclusions

Measured lithium diffusion in a natural clinopyroxene at 800-1000oC ranges from 6.5 x 10-16 to 1

x 10-12 m2/s and is 3.5 x 10-15 m2/s in San Carlos olivine at 1000oC. The lithium diffusion

coefficient is independent of the diffusion gradient as values are the same if the flux of lithium is

into or out of the crystal. These rapid diffusivities can be used to determine the timescales for

retention of lithium signatures in mantle minerals, which can in turn be used constrain transport

times for xenoliths.

Closure temperatures have also been calculated for clinopyroxene and have been found to be

low, ranging from 400-600oC, depending on cooling rates and grain size. The low closure

temperatures and rapid diffusivities indicate that there is great potential for lithium re-

equilibration during emplacement and cooling of rocks in many geological scenarios.

Furthermore, this implies that the absence of discernable isotopic gradients in high temperature

differentiation sequences is not necessarily evidence for a lack of isotopic fractionation during

crystallization. Direct experimental measurements would be the most reliable way to determine

the degree of isotopic fractionation during crystallization; however, to date this information is

lacking.

Lithium and lithium isotopes are frequently used as tracers of surface inputs into the mantle;

however, great care must be taken when interpreting the lithium contents of minerals. Given the

rapid diffusion of lithium and its low closure temperature, lithium contents in minerals can be

significantly modified due to diffusion over very short timescales. The results of this study show

that lithium is most suitable when interpreting short duration events, such as volcanic eruptions

or degassing events (Berlo et al. 2004, Kent et al 2007) or cooling rates (Coogan et al., 2005).

Furthermore, due to the high mass ratio (1.166) between 7Li and 6Li, lithium isotopes are subject

to diffusive fractionation, even during diffusion in solids, and the isotopic composition of lithium

in minerals can be modified over very short timescales.

Page 98: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

85

Table 3.1 Composition of Starting Material

Kunlun Diopside Dekalb Diopside San Carlos

Olivine

SiO2 55.39 (0.23)1 54.77 (0.8) 40.95 (0.02)

Al2O3 0.90 (0.04) 0.66 (0.10) <0.01

FeO 0.74 (0.05) 0.85 (0.08) 9.31 (0.05)

MgO 17.83 (0.13) 17.31 (0.22) 49.19 (0.42)

CaO 24.82 (0.1) 25.17 (0.26) <0.02

Na2O 0.59 (0.03) 0.43 (0.08) <0.02

MnO 0.07 (0.03) 0.05 (0.06) 0.12 (0.02)

NiO nd nd 0.39 (0.03)

Total 100.34 99.29 100.86

n 25 11 13

Li ppm2 42.6 (2.5) 8.9 (0.6) 2.5 (0.5)

7Li (‰) +13.0 (1.0)3 +9.7 (1.0) +3.64 (0.15)4 1Numbers in parentheses represent 2 of the mean of n analyses 2Analyzed by LA-ICPMS, numbers in parentheses represent 2 of the mean of 10 analyses 3Analyzed by MCICP-MS, numbers in parentheses represent 2 of the uncertainity on the measurement 4Magna et al., 2006

Page 99: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

86

Table 3.2 Measurements of Standards

LA-ICP-MS1 Li (ppm) reference

International Standards

NBS 612 41.5 41.54 (2.87) Pearce et al. 1997

NBS 610 488.7 484.6 (21.7) Pearce et al. 1997

JG1a 92.4 79.5 (4.5) Imai et al. 1995

JB-2 7.9 7.78 (1.39) Imai et al. 1995

BCA1 12.3 13.3 Ryan and Langmuir, 1987

JGB-1 4.6 4.59 (.90) Imai et al. 1995

In house Standards

Kunlun Diopside 42.6 this study

1) 2 errors for LA-ICPMS abundance analyses are better than 10% based on counting statistics

Page 100: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

87

Table 3.3 Summary of Experiments

Experiment T

(oC) log fO2 t (min) sample

Li source/sink

1D (m2/s) Co log D (m2/s)

zerotime 1000 -10.3 5 kunlun NaCl nd

kcpx3 1000 -10.3 180 kunlun NaCl 4.06E-13 (0.73)2 4.25 (0.42)2 -12.39 (1.12)2

kcpx6 1000 -10.3 360 kunlun NaCl 1.06E-12 (0.17) 2.83 (0.51) -11.97 (0.93)

kcpx12 1000 -10.3 720 kunlun NaCl 1.01E-12 (0.11) 13.08 (2.35) -11.99 (0.65)

kcpx2 1000 -10.3 120 kunlun NaCl 5.28E-13 (0.47) 6.28 (0.31) -12.27 (0.76)

kcpxMH 1000 -5.3 345 kunlun NaCl 1.71E-14 (0.21) 4.19 (0.29) -13.77 (0.86)

kcpxMnOMn 1000 -6.7 364 kunlun NaCl 1.10E-13 (0.21) 6.82 (0.55) -12.96 (1.24)

kcpx900-72 900 -10.3 4485 kunlun NaCl 4.47E-14 (0.71) 3.27 (0.29) -13.35 (1.05)

kcpx900-36 900 -10.3 2150 kunlun NaCl 6.23E-15 (0.88) 2.06 (0.21) -14.21 (1.)

kcpx800-16d 800 -10.3 23100 kunlun NaCl 6.47E-16 (2.43) 6.56 (0.52) -15.19 (2.86)

kcpxR 1000 -10.3 362 kunlun LiCl 1.18E-13 (0.17) 1.77 (0.12) -12.93 (0.95)

sco12 1000 -10.3 735 sco LiCl 3.44E-15 (0.91) 2.64 (0.32) -14.46 (1.92)

1D shown is the average of Ds calculated from two traverses, with the exception of SCO12, Kcpx-12, Kcpx-MH, and Kcpx-900.72, see text for details 2number in parenthesis represents 2 error

Page 101: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

88

-50

-40

-30

-20

-10

0

0

0.5

1

1.5

2

2.5

3

0 500 1000 1500 2000 2500

7 Li

Li pp

m

microns

Figure 3.1 Li Elemental and Isotopic Gradients in San Carlos Opx

7Li (‰) and lithium concentration (ppm) plotted against distance from crystal edge (m) for cross sections of a orthopyroxene xenolith from San Carlos, Arizona. The isotopic gradient shows a 35 ‰ decrease from rim to core, in a W-shaped profile. Data from Jeffcoate et al. (2007).

Page 102: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

89

0

10

20

30

40

50

60

70

0 200 400 600 800 1000

DekalbKunlun

Li (

ppm

)

X(m)

starting concentration

Figure 3.2 Effect of fO2 Anneal

Lithium concentration (ppm) plotted against distance from the salt-crystal interface (m) for cross sections of a Dekalb diopside slab (circles) and a Kunlun diopside slab (squares) after the fO2 anneal. The total concentration of lithium in the crystal is uniformly decreased with respect to pre-treated material (solid lines) with the exception of the outer 50 m of the crystal, which is slightly more depleted. The annealed slabs were then re-polished with diamond and aluminum paste to 0.3 m to remove the depleted zone. Error bars are 2, based on analytical uncertainty of lithium concentration.

Page 103: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

90

10

20

30

40

50

60

0 50 100 150 200 250

lith

ium

(p

pm)

x (m)

Figure 3.3 Zero time Experiment

Lithium concentration (ppm) plotted against distance from the salt-crystal interface (m) for a cross section of the ‘zero time’ experiment. No measurable change in lithium was observed, indicating that the sample preparation and sample loading had no effect on the lithium concentration. Error bars are 2, based on analytical uncertainty of lithium concentration.

Page 104: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

91

-14.0

-12.0

-10.0

-8.0

-6.0

-4.0

-2.0

0.0

2.0

0 200 400 600 800 1000

bac

kgro

un

d c

orr

ecte

d li

thiu

m (

pp

m)

microns

0.0

0.2

0.4

0.6

0.8

1.0

1.2

0 100 200 300 400

kcpx12

inv

erf

(C

/Co

)

microns

Figure 3.4 Results for Experiment Kcpx-12

(a) Background corrected lithium concentration measured by LA-ICPMS (circles), plotted against distance from salt-crystal interface for a cross section of Kcpx12. Also plotted is the model diffusion profile (curve) calculated from the concentration gradient. (b) Plot of distance from the salt-crystal interface, x (m), versus the inverse of the error function of (Cx / Co) and the resulting fit of least-squares regression. Error bars are 2, based on analytical uncertainty of lithium concentration.

a)

b)

Page 105: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

92

Figure 3.5 X-Ray Maps of Run Product

SEI and EDS x-ray maps of Kcpx-R (1000oC, 6 hour heating in LiCl). No detectable gradients were noted in major element concentration (Si, Al, Ca, Na, Mg) or Cl.

SEI

Page 106: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

93

-18.0

-16.0

-14.0

-12.0

-10.0

-8.0

0 5 10 15

diffusion outdiffusion in

log

D (

m2 /s

)

t (hrs)

Figure 3.6 Time Series

Diffusion coefficients measured from a series of diffusion-out experiments (solid circles) at 1000oC, fO2 of NNO and times ranging from 2 to 12 hours. The measured diffusion coefficients are within error of each other (2 based on analytical uncertainty of lithium concentration) and are independent of experiment duration. Also within error of the other experiments is Kcpx-R, a diffusion-in experiment (open circle), which demonstrates that the lithium diffusion coefficient is independent of the diffusion gradient as D-values are the same if the flux of lithium is into or out of the crystal. Error bars are 2, based on analytical uncertainty of lithium concentration.

Page 107: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

94

-22.0

-20.0

-18.0

-16.0

-14.0

-12.0

-10.0

-8.0

7 7.5 8 8.5 9 9.5

this studyCoogan et al., 2005olivine

log

D (

D in

m2 /s

)

10000/T(K)

1100oC 1000oC 900oC 800oC

Figure 3.7 Measured Lithium Diffusion Coefficients

Log DLi values measured in this study (solid circles), and Coogan et al. 2005 (open circles), plotted against 10,000/T (K). The diffusion coefficient for lithium in clinopyroxene is temperature dependent from 800 to 1000 oC with an Arrhenius relationship of log DLi

cpx = 5.92 (±8.51) – 192 (±10)/RT (R2=0.87) Also shown is the datumn from the olivine experiment. Error bars are 2, based on analytical uncertainty of lithium concentration. Error bars are 2, based on analytical uncertainty of lithium concentration.

Page 108: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

95

-18.0

-17.0

-16.0

-15.0

-14.0

-13.0

-12.0

-11.0

-10.0

-16 -14 -12 -10 -8 -6 -4

Coogan et al., 2005

log

D (

D in

m2 /s

)

log fO2

Figure 3.8 fO2 Experiment Series

Log DLicpx measured in this study (solid circles), and Coogan et al. 2005 (open circles), plotted as a function of log fO2.

As log fO2 increases from -12 to -5.3 (Kcpx-MH; log fO2 buffered by magnetite-hematite) log DLicpx decreases from -

12.4 (from the study of Coogan et al., 2005) to -13.8 (Kcpx-MH). A weighted least squares regression line yields the equation: log DLi

cpx = -15.2(±1.7) – 0.28 (±0.18) x log fO2 (R2=0.75). The least squares regression line does not

include the data of Coogan et al., (2005). Error bars are 2, based on analytical uncertainty of lithium concentration. Error bars are 2, based on analytical uncertainty of lithium concentration.

Page 109: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

96

0

1

2

3

4

0 50 100 150 200

bac

kgro

un

d c

orr

ecte

d li

thiu

m (

pp

m)

microns

Figure 3.9 Lithium Diffusion Profile in San Carlos Olivine

The background corrected lithium measured by LA-ICPMS (circles) plotted against distance from the salt-crystal interface and model diffusion profile (curve) from a cross section of the 12-hour SCO12 experiment. At 1000oC and fO2 of NNO the measured lithium diffusion into olivine is log D = -14.1 (±0.12) m2/s, which is two orders of magnitude slower than the measured lithium diffusion into clinopyroxene at the same conditions, see text. Error bars are 2, based on analytical uncertainty of lithium concentration.

Page 110: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

97

-12

-10

-8

-6

-4

-2

0

2

4

0 100 200 300 400 500 600

Kcpx2, Li out

1000oC, 2 hoursLi corrmodel

back

gro

und

cor

rect

ed

Li (

ppm

)

x (m)

-5.0

0.0

5.0

10

15

0 100 200 300 400 500 600 700

Kcpx2, Li out

1000oC, 2 hours

(7 Li/6 Li

- 7 Li

/6 Li c

ore

)/7 Li

/6 Li c

ore

* 1

000

m from rim to core

Figure 3.10 Lithium Diffusion and Isotopic Fractionation in Kcpx-2

(a) The background-corrected lithium (ppm) measured by LA-ICPMS (circles) plotted against distance from the salt-crystal interface and model diffusion profile (curve) from a cross section of Kcpx-2. (b) A plot of the7Li/6Li ratio (normalized by 7Li/6Li ratio of the core of the slab) measured by SIMS plotted against distance from the salt-crystal interface (m) and a model isotopic gradient (dashed curve). The solid line is the assumed core value (the original and unaltered 7Li/6Li ratio of the slab). Error bars are 2, in (a) based on analytical uncertainty of lithium concentration and in (b) based on counting statistics from SIMS analysis.

a)

b)

Page 111: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

98

-22.0

-20.0

-18.0

-16.0

-14.0

-12.0

-10.0

-8.0

-6.0

4 8 12 16 20

log

DLi (

m2 /s

)

10,000/T (K)

20040060080010001400T (oC)

Si-crystal

albite & anorthite

cpx, this study

cpx, Coogan

ol, this study

Figure 3.11 Comparison of Lithium Diffusion Coefficients

A plot of log DLi as a function of10,000/T (K) for lithium diffusion in geologically significant minerals. Lithium diffusion, as calculated from the Do and Ea of the following studies: in a p-type Si-crystal (short-dash line) from Pell (1960), feldspar data (anorthite and albite; dashed-dotted line) from Giletti and Shanahan (1997), clinopyroxene (long-dash line) from Coogan et al. (2005) and this study (solid line). Also shown is the measurement of lithium diffusion in olivine (solid square) from this study.

Page 112: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

99

-30.0

-25.0

-20.0

-15.0

-10.0

5 6 7 8 9 10 11

log

D (

m2 /s

)

10,000/T (K)

HLi, Coogan et al

Li, this study

Fe, Mn-Mg

Sr

Ca-Mg

Ca

Fe

Pb

Si

YbDyNdCeLu

U

Th

O

1400 1200 1000 800

T(oC)

Figure 3.12 Comparison of Diffusivities Measured in Pyroxene

A plot of log DLi vs. 10,000/T (K) for the lithium diffusivities measured in this study (solid circles) and that of Coogan et al. (2005, open circles) with experimentally determined diffusivities for other elements in clinopyroxene as calculated from the Do and Ea of those studies. Hydrogen data (H) are from Woods (2000), strontium (Sr) from Sneeringer et al. (1984), Fe, Mn-Mg interdiffusion from Dimanov and Sautter (2000), Ca-Mg interdiffusion from Brady and McCallister (1983), lead (Pb) from Cherniak (2001), Ca self diffusion (Ca) from Dimanov et al. (1996), iron (Fe) from Azough and Freer (2000), oxygen (O) from Ryerson and McKeegan (1994), uranium (U) and thorium (Th) from Van Orman (1998), REE (Yb, Dy, Nd, Ce, Lu) from Van Orman (2001) and silicon (Si) from Béjina and Jaoul (1996).

Page 113: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

100

-3

-2

-1

0

1

2

3

4

1 10 100 1000

log

tim

e (

ho

urs

)

grain size (microns)

olivine

clinopyroxene

100 hrs

10 hrs

1 hr

10 min

30 sec

10 sec

1000 hrs

1 year

Figure 3.13 Retention of Lithium Composition

Centre retention time (hours) at 1000oC as a function of grain size (m) for spherical grains from 1 to 1000 m. Curves are calculated for Dta2 = 0.03 and represent the maximum time an olivine or clinopyroxene grain can remain at 1000 oC and retain its original core concentration, unaltered by diffusion. See text for details.

Page 114: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

101

400

500

600

700

800

900

1 10 100 1000

T (

oC

)

characteristic radius (microns)

100o/Myr

10o/Myr

1o/Myr

dT/dt = 100o/MyrLiSr

Figure 3.14 Comparison of Closure Temperature of Li and Sr in Clinopyroxene

A plot of calculated closure temperatures (oC) versus characteristic grain size (m) for a spherical grain with cooling rates of 1o, 10o and 100 o/Myr. The solid curves are values calculated for lithium using the diffusion parameters measured in this study. The dashed curve is for Sr calculated using the diffusion parameters of Sneeringer et al. (1994). Note the large difference in the closure temperature for the two elements. Whereas the value for Sr (and other more highly charged cations) is more than 1000 oC higher for a given grain size and cooling rate.

Page 115: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

102

0

1

2

3

4

5

6

7

8

0 5 10 15 20 25 30

7 Li

MgO (wt%)

Figure 3.15 Lithium Isotopic Compositions of Kilauea Iki Lava Lake Rocks

A plot of measured lithium isotopic composition of Kilauea Iki samples as a function of MgO content, which represents degree of magmatic differentiation. There is no significant fractionation observed in these samples; however, the expected isotopic fractionation due to Rayleigh distillation in a closed system is approximately 6-7 ‰, assuming ~0.999 (consistent with predicted from min-fluid fractionation at 1100oC). Kilauea Iki lava lake data from Tomascak et al. (1999). Error bars are based on the external reproducibility (±1.1 ‰)

Page 116: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

103

-50

-40

-30

-20

-10

0

0 200 400 600 800 1000 1200

7 Li

microns from crystal edge

Figure 3.16 Li Isotopic Gradient in San Carlos Opx and Modeled Profile

7Li (‰) and lithium concentration (ppm) plotted against distance from crystal edge (m) for cross sections of a orthopyroxene xenolith from San Carlos, Arizona (closed symbols) and the modeled isotopic gradient (curve). The right half of the W-shaped profile observed in an orthropyroxene crystal from a San Carlos xenolith is modeled using the diffusivities and the (determined from this study. The lithium isotopic gradient shown here is calculated for 10 hrs, 1000oC and with a surface concentration/initial mineral concentration (Cs/Co) of 3.25.

Page 117: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

104

4 Technique Development to Study Muscovite-Fluid Partitioning of Nitrogen

The following represents work that was undertaken in collaboration with Gray Bebout at Lehigh

University, Bethlehem, Pennsylvania. The results of this study demonstrated the feasibility of the

technique and provided the basis for a successful NSF application awarded to Bebout et al. in

2007.

4.1 Introduction

Nitrogen isotopes are often employed to unravel the mechanisms involved in a wide variety of

systems such as: N-cycling at convergent margins and recycling of surface material to the mantle

(Sadofsky and Bebout, 2004; Bebout and Fogel, 1992), the origins of fluids involved in orogenic

Au-deposits (Jia and Kerrich, 1999), the origin of the Earth’s atmosphere and hydrosphere and

their chemical evolution (Jia and Kerrich, 2004; Pinti et al., 2001; Sephton et al., 2002) and the

production and migration of hydrocarbons due to metamorphism of organic matter (Williams et

al., 1995).

Nitrogen has two stable isotopes, 14N and 15N, whose abundances are ~99.64 % and ~0.36 %,

respectively. Enrichments in nitrogen isotopes are described as:

(37)

10001NN/

NN/N

std1415

smp1415

15

where ‘smp’ refers to the sample and ‘std’ refers to the standard, typically atmospheric N2.

Nitrogen is a crucial nutrient in biological systems and its global cycle is thought of as primarily

a biological one. The present day nitrogen cycle is a result of the rise of oxygenic photosynthesis

and aerobic respiration (Falkowski and Godfrey, 2008). Biological cycling of nitrogen involves

the reduction of N2 to NH4+, for the synthesis of protein, as it is transferred from the atmosphere

and hydrosphere to the biosphere (Falkowski and Godfrey, 2008). A large portion (7.5 x 1020 g)

of the Earth’s nitrogen content is also stored in sedimentary rocks (Holloway and Dalhgren, 2002

and references therein). The transfer of nitrogen from surface reservoirs to the mantle and then

Page 118: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

105

back to the atmosphere is an important component of the global cycle (Berner, 2006). Burial of

organic matter and NH4+ in sedimentary rocks and the subsequent weathering of these rocks will

cycle some nitrogen from the biosphere to the atmosphere (Holloway and Dahlgren, 2002).

Subduction of sedimentary rocks and transfer of nitrogen to the mantle from these rocks will also

result in the cycling of nitrogen back to the atmosphere via volcanic and metamorphic degassing

(Berner, 2006; Bebout and Fogel, 1992).

Nitrogen in the atmosphere occurs primarily as N2 with a 15N = 0 ‰ (Faure, 1986). Nitrogen in

the hydrosphere and in soils can occur as nitrate (NO3-), nitrite (NO2

-) ammonium (NH4+),

ammonia (NH3), oxides (NO, NO2 and N2O) and amino acids. Pelagic sediments are enriched in

nitrogen, ranging from 100 ppm N to as much as 2000 ppm N and have 15N ranging from +2 ‰

to +10 ‰ (Holloway and Dalhgren, 2002; Tolstikin, 1998), relative to mid-ocean ridge basalts

(MORB) which have 0.1 to 0.3 ppm N and 15N -5‰ (Marty and Humbert, 1997; Busigny et al.,

2005). Therefore, nitrogen isotopes can serve as an ideal tracer of surface-mantle interaction, as

well as a tracer of organic and sedimentary sources of fluids and melts in the mantle (Figure 4.1;

Hallam and Eugster, 1976; Sadofsky and Bebout, 2000; Bebout, 1997).

In silicate rocks, nitrogen occurs primarily as ammonium (NH4+), which replaces K+ (Hallam and

Eugster, 1976); however, 10-20 % of the ammonium in igneous rocks is extractable with weak

KCl solutions, suggesting that some of the ammonium may occur as soluble salts or introduced

by biological activity along grain boundaries (Faure, 1986). The primary mineral hosts for

nitrogen in igneous and metamorphic rocks are sheet silicates, and especially white micas

(Bebout et al., 1999; Bebout, 1997; Sadofsky and Bebout, 2000).

Various researchers have documented a correlation between nitrogen content, nitrogen isotopic

composition, and metamorphic grade of sedimentary rocks. Häendel et al. (1986) found that the

nitrogen content of both regional-metamorphic and contact-metamorphic rocks from Erzgebirge,

Germany, decreased from ~500 to ~20 ppm towards the contact zone, whereas the δ15N of the

rocks increased from ~ +5 ‰ up to +15 ‰ towards the contact. This correlation of nitrogen,

δ15N, and metamorphic grade was also reported in many other settings including: the contact

aureole of the Skiddaw Granite, Lake District, England (Bebout et al., 1999), the contact aureole

of the Cooma metamorphic complex, southeastern Australia (Jia, 2006), and the Catalina schist

subduction-zone metamorphic complex, California, USA (Bebout and Fogel, 1992). The

Page 119: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

106

increasing δ15N and decreasing nitrogen content with increasing temperature (and therefore,

increasing metamorphic grade) has been interpreted as progressive release of nitrogen enriched

in 14N, leaving the residual rocks enriched in 15N (Häendel et al., 1986; Bebout et al., 1999; Jia,

2006).

To date, a systematic investigation of the degree of isotopic fractionation and the extent of

partitioning that occurs during dehydration reactions has been lacking. Both the isotopic

fractionation, 15Nsolid-fluid, and the nitrogen partitioning between mica and fluids needs to be

known if nitrogen contents are to be used to determine the extent of slab dehydration . This study

presents a technique to measure nitrogen partitioning and isotopic fractionation between fluids

and mica. With this information, more accurate models can be developed to constrain the extent

of slab dehydration.

4.2 Theoretical Considerations

4.2.1 N-speciation and Isotopic Fractionation

In crustal metamorphic and magmatic settings the most likely nitrogen species are N2 (Bebout,

1997) or NH3 (Equation 4, see below). Based on analyses of mineral separates from

metamorphosed rocks, many workers have postulated that NH4+ - N2 or NH4

+ - NH3 exchange

during devolatilization increases the δ15N in the residual metamorphic rocks (Bebout, 1997;

Sadofsky and Bebout, 2000; Bebout et al, 1999). That is:

(38) 14NH4+

musc + 15N14N,aq = 15NH4+

musc + 14N2,aq

(39) 14NH4+

musc + 15NH3,aq = 15NH4+

musc + 14NH3,aq

Using MINDO/3, a semi-empirical molecular orbital calculation method, Hanschmann (1981)

predicted a large difference in the isotopic fractionation between N-species in fluids (as N2 or

NH3) and NH4+ in solids (Figure 4.2). All previous studies have based their calculations on the

work of Hanschmann (1981), who calculated fractionation factors between N2, NH3, and NH4+

molecules. At any given temperature, the fractionation that occurs during NH4+ - NH3 is larger

than would occur at that temperature during NH4+ - N2 exchange. For example, at 500 oC NH4

+ -

NH3 exchange results in a 7 ‰ fractionation, whereas NH4+ - N2 results in only a 3 ‰

fractionation. To date, direct experimental measurements of isotopic fractionation during NH4+ -

NH3 or NH4+ - N2 exchange are lacking.

Page 120: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

107

Therefore, knowledge and control of the fN2 and fNH3 of the experiments is necessary to

determine N speciation and understand the isotopic fractionation. Estimation of N- species is

possible given knowledge of the activity product ratios (K) for the speciation of water and N2 in

the experiment:

(40) H2 + ½O2 H2O K1 = fH2O / (fH2 * fO2 ½)

(41) ½ N2 + 3/2H2 NH3 K2 = fNH3 / (fN2 ½ * fH2

3/2)

(42) PT = Pgas = (fH2O/H2O) + (fH2/H2) + (fO2/O2) + (fN2/N2) + (fNH3/NH3)

fNH3 can be calculated from K2 (from JANF tables) if fH2 or fN2 are known.

One possible fN2-buffer is the Cr-CrN buffer (CCN) used by Hallam and Eugster (1976);

(43) CrN Cr + ½N2 K4 = fN2 ½ log fN2 = -2GoCrN/2.303RT

Assuming the fH2 external to capsule (the intrinsic fH2 of the reaction vessel), is equal to the

internal fH2, then both fH2 and fN2 are known. The CCN buffer can be used together with the

Inconel pressure vessels (which intrinsically generate an fO2 of approximately one log unit below

Ni-NiO; Matthews et al., 2003). The two buffers, NNO-1 and CCN, result in an fH2 such that fN2

<< fNH3 (Figure 4.3). In this manner, control of fH2 can function to buffer fN2 and fNH3. The

appropriate fH2 buffers needed such that fN2 >> fNH3 have yet to be determined; however, there

are a variety of solid materials (e.g. Cu metal, graphite) that can be used as spacers inside

pressure-vessels during experiments that can be employed to buffer redox conditions inside the

pressure vessel (Matthews et al., 2003).

Many studies have also found that the degree of fractionation is consistent with isotopic

exchange shifting from N2-dominated release to NH3-dominated release (Häendel et al., 1986;

Bebout and Fogel, 1992; Jia, 2006). For example, the data for greenschist and amphibolite (300-

600 oC) facies rocks of the Cooma metasedimentary complex have isotopic compositions that are

shifted only +1 ‰ from lowest grade metasediments, which suggest N2-NH4+ dominates the

isotopic exchange. Whereas the upper amphibolite facies (> 600oC) rocks are shifted almost +7

‰, which implies NH3-NH4+ fractionation (Jia, 2006). However, this process is not well

understood (Häendel et al., 1986; Bebout and Fogel, 1992; Jia, 2006). To date there is very little

Page 121: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

108

information on the degree of isotopic fractionation that may occur during fluid-mineral

partitioning of nitrogen, and no experimental determinations have been made. These data are

essential for developing quantitative models of N-isotope systematics and N exchange during

fluid-mineral partitioning and constrain devolatilization processes.

4.2.2 Buffering pH

The experimental method employed needs to ensure mica stability. Previous experimental work

on micas has used KCl solutions to buffer pH and stabilize mica (Lynton et al., 2005):

(44) 1.5 k-spar + H+ 0.5 mus + 3 qtz + K+

However, this reaction generates potassium feldspar (k-spar), which also takes up NH4+. This

introduces some ambiguity to the mass balance assumptions necessary for quantitative

determination of partitioning and fractionation; therefore, keeping the proportion of k-spar to a

minimum is necessary. NH4Cl is ideal because it serves both as a nitrogen source and potentially

provides a way to buffer the pH, which is necessary for subsequent fluid speciation calculations.

(45) NH4+ ↔ NH3 + H+

By adjusting the concentration of the NH4Cl solution, the pH of the system can be controlled.

For example, a 1 M NH4Cl solution has a pH of 4.6, and a 0.1 M solution has a pH of 5.1. As

Equation 45 shows, a fixed pH would also constrain the nitrogen speciation.

Another method to stabilize mica is to use mineral mixtures with a very high ratio of finely

powdered muscovite to quartz and k-spar (50:1:1; Table 4.1). The powdered mica would

promote fluid-mineral exchange, probably by some combination of dissolution/re-precipitation

and diffusive exchange. The dissolution of some of the mica would also serve to buffer the pH of

the solution. To determine the contribution of the k-spar to the nitrogen content of the solid

residue, experiments using variable ratios of k-spar to quartz + muscovite can be conducted.

Another method to assess the contribution from potassium feldspar would use larger mica

fragments. These larger fragments would undergo diffusive exchange with the fluid, and then be

mechanically separated from the residual solid after the experiment.

Page 122: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

109

4.3 Experimental Methodology

Experiments to measure N isotopic fractionation involved equilibrating mica, of known N-

isotopic composition, with a solution of known N-isotopic composition using the cold-seal

pressure vessels at the University of Toronto's High Pressure Lab. All experiments utilized a

natural muscovite from Ontario, Canada (40 ppm N, 15N = +6.2 ‰), and synthetic quartz and

orthoclase that were ground and screened to 200-300 mesh, as a starting material. Isotopically

characterized NH4Cl (15N = -4.1 ‰) was combined with ultra-pure water to make a 1M

solution. Table 4.1 lists the details of the preliminary experiments and the initial results.

Experiments using mineral mixtures with a very high ratio of muscovite to quartz and k-spar

(50:1:1; Table 4.1), were conducted. For these experiments, N-02, N-07, N-08, and N-09 the

muscovite was ground and screened to 100-200 mesh.

Another strategy to ensure mica stability and limit interference from potassium feldspar used

larger mica fragments. These fragments, which would undergo diffusive exchange with the fluid,

could be mechanically separated from the residual solid after the experiment. Experiments N-04

and N-05 utilized 3 x 10 mm flakes of muscovite.

The sample mixtures and the 1M NH4Cl solution were loaded into Pt-capsules (5 mm OD x 1.5

cm height) and sealed. The whole assembly was weighed, placed in a drying oven for several

hours, and weighed again to determine if the welding had resulted in an airtight seal. The

capsules were loaded into Inconel pressure vessels, pressurized to 15,000 psi (~1 kbar), and

heated while the pressure was adjusted accordingly to maintain 15000 psi. Experiments were

quenched from maximum temperature to room temperature in 4 min; this was accomplished by

cutting power to the heaters and alternately misting the pressure vessels with water and air-

cooling them with pressurized air. The capsules were recovered; the samples extracted and rinsed

with ultra pure water to remove NH4Cl precipitate and quench products.

4.4 Analytical Methods

All samples were analyzed at Lehigh University on a Finnigan MAT 252 isotope ratio mass

spectrometer, using the new Gas Bench II metal-high vacuum extraction line (Bebout et al.,

2007). Powdered samples were loaded, along with varying amounts of the Cu/CuOx reagent, into

Page 123: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

110

6 mm (o.d.) quartz tubes that were previously combusted in atmosphere at 550°C for 2 h to

remove organic and other contaminants. Samples were evacuated overnight, heated with heating

tape to 100°C, and then sealed under vacuum. The tubes were then heated to 950–1050°C in a

programmable furnace, held at these peak temperatures for at least 3 h, then cooled slowly,

particularly through the 700–500°C temperature range, to ensure proper speciation of gases as

H2O, CO2, and N2 (Bebout and Fogel, 1992). After loading the sealed tube onto the tube cracker,

and at least 1 h of evacuation, the tube was cracked and the released gas expanded into a series of

traps to remove any condensable gases, presumably mostly H2O and CO2 but also some Ar and

other minor contaminants. After passing through the last trap, the gas was expanded and isotopic

analyses were undertaken using the GBII system. Nitrogen concentrations are obtained by

measurement of voltage on the m/z 28 peak for calibrated volumes in the mass spectrometer

(either a variable-volume/bellows or micro-volume inlet); voltage vs. moles N2 is calibrated by

extractions and analyses of non-silicate standards with known N contents. A 1 of 0.14 ‰ and

0.12 nmol for analyses of 12-225 nmol of N2 is reported by Bebout et al. (2007) for this system.

4.5 Results

4.5.1 Nitrogen Partitioning

The addition of K was found to promote potassium feldspar growths (k-spar) as minute

intergrowths of ms + k-spar on the surface of the mica, which could not be mechanically

separated. The potassium feldspar forms clusters of idiomorphic, monoclinic crystals 2-5 m in

length near the edges of the muscovite sheet, as well as where the surface of the muscovite has

formed ‘steppes’ either due to dissolution or re-precipitation (Figure 4.4).

Another method employed to ensure mica stability and limit interference from potassium

feldspar used larger mica fragments. These fragments diffusively exchanged K+ and NH4+ with

the fluid, and were mechanically separated from the residual solid after the experiment.

However, experiments that used this technique (N-04, N-05; Table 4.1) produced unsatisfactory

results. The mica fragments were found to have lower nitrogen contents than experiments

utilizing finely powdered mica. The samples that utilized muscovite pieces as a starting material

contained ~250 ppm N, whereas the experiments that had muscovite powder as a starting

material contained ≥1000 ppm N (Figure 4.5).

Page 124: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

111

The experiments using mineral mixtures with a very high ratio of muscovite to quartz and k-spar

(50:1:1; Table 4.1), resulted in some loss of sample (presumably dissolved in solution during the

experiment); however, they resulted in a remarkably reproducible nitrogen content (1250 ±182

ppm). This suggests that using finely powdered mica promotes fluid-mineral exchange, probably

by some combination of dissolution/re-precipitation and diffusive exchange.

4.5.2 Nitrogen Isotopic Fractionation

Experiments, in which muscovite (δ15N = +6.2 ‰) was equilibrated with a 1 M NH4Cl solution

(δ15N = -4.1 ‰) resulted in solid residues that showed an overall decrease in δ15N (δ15N = +0.35

‰, +1 ‰, +1.5 ‰; Table 4.1, Figure 4.6). Consistent with previous work, which has shown that 15N is preferentially concentrated in the solid relative to the fluid, the final solid in the

experiments has a more positive δ15N than the NH4Cl of the fluid. The apparent mica-fluid

fractionation is ~4.5 ‰ to 5.6 ‰, assuming that the fluid composition is constant and equal to

the NH4Cl. The work of Hanschmann (1981) predicts a fractionation of 2.5 ‰ between N2,aq-

NH4+ and ~8 ‰ between NH3,aq-NH4

+ at 500oC.

An average mica-fluid = +4.8 (+0.6) ‰ is calculated from this preliminary data (Figure 4.7). The

magnitude of 15Nmica-fluid measured in this study falls between the NH4+-N2 and NH4

+-NH3

curves determined by Haschmann (1981). Because the speciation of N was not controlled in this

study, it is uncertain if the intermediate 15Nmica-fluid measured here is a result of isotopic fraction

between some mix of N2 and NH3 species or isotopic fractionation due to quench and sample

retrieval procedures. This result is very encouraging as it is consistent with the shift observed in

the δ15N of mica samples from the Catalina Schist, where the residual solids are enriched in 15N

with respect to the fluids released by dehydration (Bebout, 1997). This is also consistent with the

behaviour of N-isotopes in metamorphic rocks in the contact aureole of the Skiddaw Granite,

Lake District, England (Bebout et al., 1999), the contact aureole of the Cooma metamorphic

complex, southeastern Australia (Jia, 2006), and the regional-metamorphic and contact-

metamorphic rocks from Erzgebirge, Germany (Häendel et al., 1986). The magnitude of the

isotopic shifts measured in these rocks range from ~1 ‰ to 7 ‰, which is consistent with

fractionations occurring between the NH4+-N2 and NH4

+-NH3 curves (as determined by

Hanschmann, 1981). It is interesting to note that natural samples have calculated 15Nmica-fluid

values that also fall between the two curves as well, this phenomenon has been noted by several

Page 125: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

112

researchers and has been attributed to a shift in the speciation of the fluids; however, this process

is not well understood (Häendel et al., 1986; Bebout and Fogel, 1992; Jia, 2006).

4.6 Discussion

4.6.1 Utility of NH4Cl as Nitrogen Source

The use of NH4Cl seemed ideal because it would serve both as a nitrogen source and provide a

way to buffer the pH, which is necessary for subsequent fluid speciation calculations. However,

a nitrogen source that does not cause precipitation of N-rich solids would be preferable because

it would allow for ‘on-line’ (see below) analysis of vapor, liquid and solid components of the

experiment. Silver azide, AgN3, has been employed in previous studies as a N-source (Keppler,

1989), and has the benefit of producing a large amount of N, and a relatively inert residue (Ag

metal). However, at unconfined conditions, such as in an improperly sealed capsule, AgN3

releases N explosively making it rather hazardous in the laboratory, both for laboratory personnel

and equipment. Anovitz et al. (1998) have evaluated a variety of solid-N sources for

experiments. They concluded that Cu3N is an ideal N source; the release of N is much slower

than AgN3 and it is inexpensive and commercially available. They also examined CrN and

concluded that it did not dissociate readily enough at metamorphic temperatures to generate a

sufficient amount of nitrogen; however, the benefit of using CrN is that the CrN-N buffer fixes

the relative fugacities of N2 and NH3 in the fluid (see section 4.2.2). It is important to note that

using CrN, AgN3 or Cu3N as a nitrogen source would require some strategy to buffer pH and

stabilize muscovite. Because a satisfactory alternative to using NH4Cl as a pH buffer has not yet

been determined, a 1 M NH4Cl solution was used as the nitrogen source for these experiments.

Considering the difficulties that were encountered with the NH4Cl residue during analyses (see

below) the experimental method will ultimately need to make use of one of the other nitrogen

sources discussed.

4.6.2 Analytical Considerations

Originally, it was planned to pierce the capsule on-line, step-heat, and measure the isotopic

composition of the gas, liquid and solid components as they were released to determine the

fractionation by measuring each phase. This method would forgo the rinsing, powdering and

combusting of the sample, thereby reducing change for contamination and having the ability to

Page 126: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

113

analyze all run-products, not just the solid. To this end a device was constructed that would

enable the capsules to be punctured and heated on the extraction line (Figure 4.8); however,

difficulties were encountered with the discovery of a NH4Cl residue that formed on the inner

surface of the puncturing tool and inside the capsule. The use of NH4Cl as a N-source and a pH

buffer, while desirable from an experimental point of view, resulted in complications during the

analysis. A natural consequence of taking a solution from high pressure and temperature to room

temperature and pressure is that the solubility of various components of the solution (such as

dissolved gases and salts, as well as hydrous silica and alumina complexes) decreases

dramatically. This change in solubility is accommodated by the exsolution of the dissolved gases

and the precipitation of salts and amorphous silica and alumina complexes – described as quench

products. Typically, the run products of interest are the solid phases and these are recovered by

puncturing the capsule, evaporating off the solution (which results in further precipitation of the

remaining solute load) and rinsing away the evaporite and quench products.

In puncturing and extracting the run products ‘on-line’, all the run products (vapor phase, liquid

phase and solid phase), are of interest. However, it is necessary to ascertain what proportion of

the total N component this NH4Cl residue represents. That is, what proportion of this solid was

dissolved in the solution at the conditions of the experiment and how much precipitated during

quenching? Was some of the nitrogen dissolved in the vapor phase at the conditions of the

experiment? It is uncertain whether the NH4Cl residue was the result of precipitation during

quenching of the experiment or evaporation due to heating during analysis, (it is most likely a

combination of both). This also complicated the N analysis of vapor phase, as there is likely

isotopic fractionation occurring between the NH4Cl dissolved in the solution and the coexisting

liquid/vapor phase during the exsolution of the gas phase at quench. Due to the ambiguities

regarding the nature of the vapour and liquid phases, partitioning was determined from the

composition of the solid phase (with any residual NH4Cl rinsed away).

4.6.3 Experimental Considerations

The mica fragments were found to have lower nitrogen contents than experiments utilizing finely

powdered mica. The samples that utilized muscovite pieces as a starting material contained ~250

ppm N, whereas the experiments that had muscovite powder as a starting material contained ≥

1000 ppm N. The discrepancy between the results from muscovite powder experiments and the

Page 127: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

114

muscovite flake experiments may be a result of nitrogen exchanging faster in the smaller,

powdered muscovite grains. This suggests that equilibration had not been reached in the larger

grain sizes by 6 days at 550oC, which is consistent with the sluggish exchange kinetics

documented in past studies (Lynton et al., 2005; Bos et al., 1988). Another possibility is that the

orthoclase is contributing nitrogen to the analysis. The ambiguity of these results highlights the

importance of quantifying the amount of nitrogen incorporated into orthoclase and its effect on

the mass balance of the experiment. Even if complete separation of orthoclase and muscovite is

not possible, a control on the amount of orthoclase would at least allow an evaluation of the

contribution of this mineral to the overall nitrogen content of the run product.

4.6.4 Isotopic Fractionation Experiments and Atmospheric Contamination

A complicating factor in calculating mica-fluid fractionation by mass balance is the presence of

trapped atmospheric nitrogen (δ15N = 0 ‰) in the capsule. This amount of added nitrogen would

be variable and difficult to control. In the above experiments, trapped nitrogen would result in a

positive shift of the overall composition of the fluid. The magnitude of the fractionation

calculated neglecting the additional nitrogen-source would represent a maximum value. It is

possible to calculate the percentage of nitrogen contribution from trapped air in the capsule using

estimated capsule volumes and the known atmospheric nitrogen concentration. Estimates of the

contribution from atmospheric nitrogen to the mass balance are provided in Table 4.2. The above

calculations also show that an increase in the concentration of the NH4Cl solution of up to 3 M

would produce enough nitrogen to buffer the solution against the atmospheric contribution, and

provide an ‘infinite reservoir’ of constant N-isotopic composition. For the mass of solution used

and the estimated capsule volume, the composition of the solution is estimated to have shifted a

maximum of +0.2 ‰ from the value for the pure NH4Cl solution; this would result in a

maximum mica-fluid of +5.4 (+0.6) ‰.

4.7 Suggestions for Future Work

This study has produced promising new results for nitrogen partitioning and isotopic

fractionation between aqueous fluids and muscovite. This work also highlights the challenges in

making these measurements and their interpretations. From an experimental design point of

view, minimizing the contribution from orthoclase to the nitrogen content during analysis,

Page 128: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

115

ensuring mica stability and equilibrium between fluids and mica is established, could be

circumvented by growing mica in a nitrogen-rich environment. This technique has been

successfully utilized by Pöter et al. (2004) and has not been investigated here. From an analytical

point of view, the choice of a N-source needs to be carefully considered so as not to add

ambiguity during the analysis. A solid source such as Cu3N (Anovitz et al., 1998) is attractive

because the residue (Cu) is relatively inert and would not result in further reactions/fractionation

during quench or analysis.

The experiments conducted in this study have demonstrated that measurable fractionation occurs

between fluids and micas equilibrated at metamorphic conditions. The magnitude of

fractionation is consistent with estimates from analysis of metasedimentary rocks (Häendel et al.,

1986; Bebout and Fogel, 1992; Bebout, 1997; Jia, 2006) and theoretical estimates based on N2-,

NH3-NH4+ calculations (Hanschmann, 1981). The above techniques can be utilized to better

constrain the N-isotopic fractionation that occurs during fluid-rock interaction. With better

constraints, the relationship between N and δ15N might be used to estimate metamorphic grade

for rocks with no indicator mineralogy, or provide insight to the degree of fluid-rock interaction

or retrograde metamorphism. This data would allow the δ15N content of metasedimentary

convergent margin rocks to be used to better quantify the composition of material re-cycled deep

into the mantle, pinpoint the origins of fluids involved in orogenic Au-deposits or even further

our understanding of the evolution of the Earth’s atmosphere and oceans.

Page 129: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

116

Table 4.1 Experiments and Results

Final

Name T (oC) t (days)

Composition (wt. %)

Mica (mg)

sol'n (mg)

prep ppm N 15N M-F

N-02 550 5 1 mus pdr : 1 qtz : 4 orth 46 50 rinsed 900

N-04 550 6 1 mus pcs : 1 qtz : 3 orth 72 rinsed - qtz/orth removed 270

N-05 650 6 1 mus pcs : 1 qtz : 4 orth 70 rinsed - qtz/orth removed 230

N-07 500 6 50 mus pdr : 1 qtz : 1 orth 47 60 30.1 mg recovered/rinsed 1302 +0.35 4.5

N-08 500 16 50 mus pdr : 1 qtz : 1 orth 51 57 25.7 mg recovered/rinsed 1407 +1 5.1

N-09 500 10 50 mus pdr : 1 qtz : 1 orth 61 58 31.5 mg recovered/rinsed 1052 +1.5 5.6

Table 4.2 Percentage of Nitrogen Contribution from Air

% atmospheric N2 in capsule

solution concentration

composition of atmospheric N2 and

1MNH4Cl solution mix in capsule

Solution added to capsule (mg)

1 M 2 M 3 M 15N

50 11 % 6 % 4 % -3.9

75 7 % 4 % 3 % -4.0

100 6 % 3 % 2 % -4.0

Page 130: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

117

0.1 1 10 100 1000 10000

Marine Seds

AOC

MORB

Seawater

Lakes, Rivers

N content (ppm)

a)

-10 -5 0 5 10 15 20 25 30

Marine Seds

AOC

MORB

Seawater

Lakes, Rivers

15N

b)

Figure 4.1 Summary of N Concentration and Isotopic Composition

The nitrogen content (ppm) a) and isotopic composition b) of terrestrial reservoirs. Data for seawater, lakes, rivers from Faure (1986), nitrogen content of marine sediments from Holloway and Dahlgren (2002), data for MORB from Busigny et al. (2005), and isotopic composition of marine sediments from Tolstikin (1998).

Page 131: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

118

2 4 6 8 10 12300

400

500

600

700

800

15N

T (

oC

)

NH4

+-NH3

NH4

+-N2

Figure 4.2 Calculated N2-, and NH3-NH4+ Fractionation

Plot of the isotopic fractionation as calculated from quantum theory that occurs during NH4+ - N2 and NH4

+- NH3 exchange as a function of temperature. Data from Hanschmann (1981) is based on isotopic fractionation predicted to occur between N2, NH3 and NH4

+ molecules.

Page 132: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

119

(i)

(ii)

Figure 4.3 Relationship of fH2, fN2, and fNH3

(i) Plot of resulting log fN2/fNH3 versus temperature when using CCN buffer Inconel pressure vessel to buffer fH2

(ii) (ii) Plot of fH2 versus temperature showing lines of fNH3/ fN2 ratios

Page 133: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

120

Figure 4.4 Scanning Electron Micrograph of Muscovite Texture

Scanning electron micrograph of Ontario Muscovite sample heated to 540oC and 1.5 kbar with 1 M KCl solution for seven days. K-feldspar forms clumps of 2-5 m long monoclinic crystals on the surface of the mica sheet.

Mus

Mus + k-spar

Page 134: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

121

0

200

400

600

800

1000

1200

1400

1600

4 6 8 10 12 14 16 18

500oC powder

550oC powder

550oC piece

650oC piece

Nit

rog

en (

pp

m)

time (days)

Figure 4.5 Nitrogen contents of run products

Plot of measured nitrogen content (ppm) of run product micas versus experiment duration (days). Error bars are 2 and represent the error of the analysis (10 %). The experiments using powdered muscovite (circles) contained the highest nitrogen contents suggesting that experiments using larger pieces of mica (squares) have not equilibrated with respect to nitrogen content of the solution. The experiments using powdered mica as a starting material have a constant N content in experiments conducted for times of 6 to 16 days, suggesting that these samples have reached equilibrium values.

Page 135: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

122

-5

-2.5

0

2.5

5

7.5

10

4 6 8 10 12 14 16 18

15N

(o

/oo

)

time (days)

NH4Cl

ont mus

Figure 4.6 Nitrogen isotopic compositions of run products

Plot of the measured nitrogen isotopic composition of experiments (15N in ‰) versus experiment duration (in days). Also shown for reference is the isotopic composition of the starting material, Ontario Muscovite and the NH4Cl used for the solution. Consistent with the N content measurements, isotopic equilibrium appears to have been reached, as the isotopic composition does not change with experiment duration from 6 to 16 days. 2 errors (approximately symbol size) are ±0.2 ‰ and represent the error of the analysis.

Page 136: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

123

2 4 6 8 10 12300

400

500

600

700

800

15N

T (

oC

)

N2-NH

4

+

NH3-NH

4

+

Figure 4.7 Isotopic shifts of run products

Plot of temperature versus 15N between NH4+, N2 and NH3 molecules. Curves constructed with data from

Hanschmann (1981). Open circles are measured 15N from 500oC experiments. 2 errors (approximately symbol size) are ±0.6 ‰ and represent the error of the analysis propagated through the calculation of 15N (15Nmin-

15Nfluid).

Page 137: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

124

Figure 4.8 Puncturing Device

Puncturing device constructed at the University of Toronto Physics Machine shop based on plans generously provided by Ethan Baxter. The sample is placed in the sample well; the device is sealed and then attached to the nitrogen extraction line. The sample is punctured by turning the screw, which lowers the puncturing tool, then the sample is heated (by means of an electrical heating coil or heating tape) to extract the nitrogen.

Page 138: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

125

5 Summary of Results and Conclusions

This study has found lithium to be moderately incompatible in the mantle during mineral – fluid

exchange reactions. The measured DLi ranges from 1.34 – 0.17 in olivine, to 0.32 – 0.09 in

plagioclase and, 0.32 – 0.04 in clinopyroxene. Lithium partitioning between clinopyroxene and

hydrous fluids is a function of temperature, decreasing with increasing temperature from 800oC

to 1100oC at 1 GPa, and appears to be a function of clinopyroxene Al2O3 content. Lithium

partitioning between olivine and fluid is not strongly a function of temperature, but appears to be

sensitive to FeO content. Lithium partitioning in anorthite is a function of feldspar composition,

similar to the partitioning of other cations in the feldspar-fluid system. Lithium partitioning

between olivine and clinopyroxene is independent of temperature; however, preliminary

experiments examining the effect of REE content and fO2 suggest that DLiol/cpx may be a function

of crystal chemistry.

Additionally, isotopic fractionation between clinopyroxene + fluid and olivine + clinopyroxene

has been measured. The isotopic fractionation between clinopyroxene and fluid at 900oC is ~ +1

‰ (±2 ‰) and the measured isotopic exchange between olivine and clinopyroxene is ~ +5 ‰

(±4 ‰). Isotopic fractionation between clinopyroxene and fluids is a function of temperature and

consistent with what has been observed in the spodumene – fluid system. The fractionation

between spodumene and hydrous fluids results in an enrichment of 7Li in the fluid from 3.5 ‰ at

500oC to ~1.0 ‰ at 900oC and 2.0 GPa (Wunder et al., 2006).

Given the incompatibility in mantle minerals, and the slight fractionation that occurs during

clinopyroxene-olivine-fluid interaction, the lithium content and isotopic signature of slab-derived

fluids can be significantly modified during transport through and interaction with the mantle

wedge. Because diffusion is so rapid, complete equilibration of the fluid with the mantle wedge

can be assumed if fluid transport from the slab to the melt source occurs by percolation;

therefore, characteristics of the lithium signal, such as the Li/Y ratio and the isotopic

composition, can provide some insight to the mechanism of transport. For example, the absence

of high Li/Y ratios in arc lavas with high B/Be, or MORB-like δ7Li in lavas with high B/Be

contents (such as the lavas from the Sunda arc, Indonesia; Tomascak et al., 2002), can be

Page 139: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

126

explained by partitioning lithium into mantle minerals as fluids percolate through the mantle

wedge. In these cases, transport through the mantle wedge completely removes the lithium signal

from the slab-derived fluid. In cases where island arc lavas have δ7Li values greater than the

mantle values (δ7Li ~ +4 ‰), such as the Izu fore arc lavas, significant slab fluid-mantle

interaction has likely occurred, as would be the case if transportation was by percolation. It is

important to note that very high fluid fluxes are implied if a lithium signal from slab-derived

fluids is to reach the melt source by percolation. When low δ7Li values (< MORB; δ7Li ~ +4 ‰)

correspond with high Li/Y ratios the fluids transported to the melt source with a minimum

amount of interaction with the mantle wedge, and transport through hyrdofractures is a likely

mechanism.

The trend of increasing δ7Li with decreasing Li/Y, which is observed in most arc lavas

(Tomascak et al., 2002), could be viewed as a spectrum between the two scenarios. Where low

Li/Y values correspond with high δ7Li, large fluid fluxes were most likely percolating through

the mantle wedge. Where high Li/Y values correspond with low δ7Li, the fluids were likely

generated at depth and transported through the mantle through hydrofractures, having minimal

interaction with the wedge. Intermediate values could be a result of some component of both

these mechanisms.

Transport of slab-derived fluids through hydrofractures in the mantle can also explain the lack of

clear and consistent correlations between lithium and other fluid mobile elements. Fluids

transported to the melt source through hydrofractures would be subject to differing degrees of

mantle interaction (variable fluid/rock ratios and transport velocities). Lithium is moderately

compatible in the mantle and diffuses rapidly; therefore, lithium contents and isotopic

compositions will be very sensitive to variations in mineral-fluid interaction.

The diffusion coefficient of lithium in clinopyroxene measured in this study is temperature

dependent, increasing from -15.19 ± 2.86 m2/s at 800oC to -11.97 ± 0.86 m2/s at 1000oC. These

diffusion coefficients are consistent with those determined by Coogan et al. (2005) for 6Li in

clinopyroxene. Lithium diffusion is independent of the diffusion gradient; values are the same if

the flux of lithium is into the crystal or out of the crystal. The lithium diffusion coefficient in

clinopyroxene has a slight negative dependence with fO2, which suggests a component of

interstitial diffusion. A single measurement of lithium diffusion in olivine was made at 1000oC

Page 140: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

127

and an fO2 of NNO, the measured lithium diffusion into olivine is log D = -14.1 (m2/s), two

orders of magnitude slower than the measured lithium diffusion into clinopyroxene at similar

conditions.

More significantly, isotopic fractionation of lithium isotopes can occur as a result of diffusion in

a mineral. This means that even the isotopic composition of lithium grains can be modified over

very short timescales. If a component of interstitial diffusion exists, as this study suggests, then

the isotopic diffusivity ratio, D6/D7 may increase with increasing temperature.

Closure temperatures calculated for lithium diffusion in clinopyroxene range from ~400 to

~500oC. Considering the rapid cooling rate and increased closure temperatures, the timescales

required for re-equilibration of lithium isotopes (weeks) are much less than the timescales

required for cooling and crystallization of magmatic rocks (kyr-Ma). These results suggest that

isotopic fractionation during crystallization and magmatic differentiation is unlikely to be

preserved due to rapid diffusion and re-homogenization of lithium isotopic compositions. The P-

T history of the samples must be evaluated before lithium signatures are interpreted. It is likely

that the lithium content of minerals can only reliably represent chemical exchange in the very

latest stages of the sample’s history, or if there is no inter-granular reservoir for lithium

exchange.

This study has developed new and promising techniques to measure isotopic fractionation of

nitrogen between muscovite and fluids. These experiments have demonstrated that measurable

fractionation occurs between fluids and micas equilibrated at metamorphic conditions. These

results are very encouraging as they are consistent with the shift observed in the δ15N of

metasedimentary rocks (Häendel et al., 1986; Bebout and Fogel, 1992; Bebout, 1997; Jia, 2006)

and theoretical estimates based on N2-, NH3-NH4+ calculations (Hanschmann, 1981). The above

techniques can be utilized to better constrain the N-isotopic fractionation that occurs during

fluid-rock interaction. This data would allow the δ15N content of metasedimentary convergent

margin rocks to be used to better quantify dehydration processes during subduction, pinpoint the

origins of fluids involved in orogenic Au-deposits or even further our understanding of the

evolution of the Earth’s atmosphere and oceans.

Page 141: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

128

References

Abart, R. (1995). Phase equilibrium and stable isotope constrains in the formation of metasomatic garnet-vesuvianite veins (SW Adamello, N Italy). Contributions to Mineralogy and Petrology, 122: 116-133.

Angel, R. J., R. M. Hazen, T. C. McCormick, C. T. Prewitt, and J. R. Smyth. (1988). Comparative compressibility of end-member feldspars. Physics and Chemistry of Minerals, 15 (4): 313-318.

Anovitz, L. M., J. G. Blencoe, D. B. Joyce, and J. Horita. (1998). Precise measurement of the activity/composition relations of H2O-N2 and H2O-CO2 fluids at 500oC, 500 bars. Geochimica et Cosmochimica Acta, 62 (5): 815-829.

Aulbach, S., R. Rudnick, and W. F. McDonough. (2008.) Li-Sr-Nd isotope signatures of the plume and cratonic lithospheric mantle beneath the margin of the rifted Tanzanian craton (Labait). Contributions to Mineralogy and Petrology, 155 (1): 79-92.

Ayers, J. C. (1998). Trace element modeling of aqueous fluid; peridotite interaction in the mantle wedge of subduction zones. Contributions to Mineralogy and Petrology, 132 (4): 390-404.

Ayers, J.C., J. B. Brenan, E. B. Watson, D. A. Wark, and W. G. Minarik. (1992). A new capsule technique for hydrothermal experiments using the piston-cylinder apparatus. American Mineralogist, 77 (9-10): 1080-1086.

Azough, F., and R. Freer. (2000). Iron diffusion in single-crystal diopside. Physics and Chemistry of Minerals, 27 (10): 732-740.

Barrat, J. A., M. Chaussidon, M. Bohn, Ph. Gillet, C. Göpel and M. Lesourd. (2005). Lithium behavior during cooling of a dry basalt; an ion microprobe study of the lunar meteorite northwest Africa 479 (NWA 479). Geochimica et Cosmochimica Acta, 69 (23): 5597-5609.

Bebout, G. E. (1997). Nitrogen isotope tracers of high-temperature fluid-rock interactions; case study of the Catalina Schist, California. Earth and Planetary Science Letters, 151 (1-2): 77-90.

Bebout, Gray E., and M. L. Fogel. (1992). Nitrogen-isotope compositions of metasedimentary rocks in the Catalina schist, California; implications for metamorphic devolatilization history. Geochimica et Cosmochimica Acta, 56 (7): 2839-2849.

Bebout, G. E., J. G. Ryan, W. P. Leeman, and A. E. Bebout . (1999). Fractionation of trace elements by subduction-zone metamorphism; effect of convergent-margin thermal evolution. Earth and Planetary Science Letters, 171 (1): 63-81.

Bebout, G. E., B. D. Idelman, L. Li, and A. Hilkert. (2007). Isotope-ratio-monitoring gas chromatography methods for high-precision isotopic analysis of nanomole quantities of silicate nitrogen. Chemical Geology, 240 (1-2): 1-20.

Beck, P., J. A. Barrat, M. Chaussidon, Ph. Gillet, and M. Bohn. (2004). Li isotopic variations in single pyroxenes from the northwest Africa 480 shergottite (NWA 480); a record of degassing of Martian magmas? Geochimica et Cosmochimica Acta, 68 (13): 2925-2933.

Page 142: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

129

Béjina, F., and O. Jaoul. (1996). Silicon self-diffusion in quartz and diopside measured by nuclear micro-analysis methods. Physics of the Earth and Planetary Interiors, 97: 145-162.

Bennett, S. L., J. Blundy and T. Elliott. (2004). The effect of sodium and titanium on crystal-melt partitioning of trace elements. Geochimica et Cosmochimica Acta, 68 (10): 2335-2347.

Benton, L.D., J. G. Ryan, and I. P. Savov. (2004). Lithium abundance and isotope systematics of forearc serpentinites, conical seamount, Mariana forearc; insights into the mechanics of slab-mantle exchange during subduction. Geochemistry, Geophysics, Geosystems – G3, 5 (8).

Benton, L. D., J. G. Ryan, and F. Tera. (2001). Boron isotope systematics of slab fluids as inferred from a serpentine seamount, Mariana forearc. Earth and Planetary Science Letters, 187 (3-4): 273-282.

Berlo, K., J. Blundy, S. Turner, K. Cashman, C. Hawesworth, and S. Black. (2004). Geochemical precursors to volcanic activity at Mount St. Helens, USA. Science, 306 (5699): 1167-1169.

Blundy, J. D., and J. Dalton. (2000). Experimental comparison of trace element partitioning between clinopyroxene and melt in carbonate and silicate systems, and implications for mantle metasomatism. Contributions to Mineralogy and Petrology, 139 (3): 356-371.

Blundy, J. D., and B. J. Wood. (1991). Crystal-chemical controls on the partitioning of Sr and Ba between plagioclase feldspar, silicate melts, and hydrothermal solutions. Geochimica et Cosmochimica Acta, 55 (1): 193-209.

Blundy, J. D., J.A. C. Robinson, and B. J. Wood. (1998). Heavy REE are compatible in clinopyroxene on the spinel lherzolite solidus. Earth and Planetary Science Letters, 160 (3-4): 493-504.

Bos, A., W. Duit, Ad. M. J. van der Eerden, J. Jansen, H. Ben. (1988). Nitrogen storage in biotite – An experimental study of the ammonium and potassium partitioning between 1M-phlogopite and vapor at 2 kb. Geochimica et Cosmochimica Acta, 52 (5): 1275-1283.

Boyd, F. R., and J.L. England. (1960). Apparatus for phase-equilibrium measurements at pressures up to 50 kilobars and temperatures up to 1750oC. Journal of Geophysical Research, 65 (2): 741-748.

Brady, J. B., and R. H. McCallister. (1983). Diffusion data for clinopyroxenes from homogenization and self-diffusion experiments. American Mineralogist, 68 (1-2): 95-105.

Brearley, M. and C. M. Scarfe. (1986). Dissolution rates of upper mantle minerals in an alkali basalt melt at high pressure: an experimental study and implications for ultramafic xenolith survival. Journal of Petrology, 27 (5): 1157-1182.

Brenan, J.M., E. Neroda, C. C. Lundstrom, H. F. Shaw, F. J. Ryerson, and D. L. Phinney. (1998a). Behaviour of boron, beryllium, and lithium during melting and crystallization constraints from mineral-melt partitioning experiments. Geochimica et Cosmochimica Acta, 62 (12): 2129-2141.

Page 143: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

130

Brenan, J.M., F. J. Ryerson, and H. F. Shaw. (1998b). The role of aqueous fluids in the slab-to-mantle transfer of boron, beryllium, and lithium during subduction experiments and models. Geochimica et Cosmochimica Acta, 62 (19-20): 3337-3347.

Brooker, R. A., R. H. James, and J. D. Blundy. (2004). Trace elements and Li isotope systematics in Zabargad peridotites; evidence of ancient subduction processes in the Red Sea mantle; lithium isotope geochemistry. Chemical Geology, 212 (1-2): 179-204.

Busigny, V., C. Laverne, and M. Bonifacie. (2005). Nitrogen content and isotopic composition of oceanic crust at super fast spreading ridge: A profile in altered basalts from ODP Site 1256, Leg 206. Geochemistry, Geophysics, Geosystems G3, 6 (12).

Chacko, T., D. R. Cole and J. Horita. (2001). Equilibrium oxygen, hydrogen and carbon isotope fractionation factors applicable to geologic systems. Reviews in Mineralogy and Geochemistry, 43 (1): 1-81.

Chan, L. H., J.M. Edmond, G. Thompson, and K. Gillis. (1992). Lithium isotopic composition of submarine basalts; implications for the lithium cycle in the oceans. Earth and Planetary Science Letters, 108 (1-3): 151-160.

Chan, L. H., J.M. Edmond, and G. Thompson. (1993). A lithium isotope study of hot springs and metabasalts from mid-ocean ridge hydrothermal systems; special section on Atlantic hydrothermal activity. Journal of Geophysical Research, 98 (B6): 9653-9659.

Chan, L. H., and M. Kastner. (2000). Lithium isotopic compositions of pore fluids and sediments in the Costa Rica subduction zone; implications for fluid processes and sediment contribution to the arc volcanoes. Earth and Planetary Science Letters, 183 (1-2): 275-290.

Chan, L. H., W. P. Leeman, and C. F. You. (2002). Lithium isotopic composition of Central American volcanic arc lavas; implications for modification of subarc mantle by slab-derived fluids; correction. Chemical Geology, 182 (2-4): 293-300.

Cherniak, D. J. 2001. Pb diffusion in Cr-diopside, augite, and enstatite, and consideration of the dependence of cation diffusion in pyroxene on oxygen fugacity. Chemical Geology, 177 (3-4): 381-397.

Cherniak, D. J. 1998. Pb diffusion in clinopyroxene. Chemical Geology, 150 (1-2): 105-117.

Connolly, J. A. D. (2005). Computation of phase equilibria by linear programming: A tool for geodynamic modeling and its application to subduction zone decarbonation. Earth and Planetary Science Letters, 236: 524-541.

Coogan, L. A., S. A. Kasemann, and S. Chakraborty. (2005). Rates of hydrothermal cooling of new oceanic upper crust derived from lithium-geospeedometry. Earth and Planetary Science Letters, 240 (2): 415-424.

Coplen, T. B. (1996). Atomic weights of the elements 1995. Pure and Applied Chemistry, 68 (12): 2339-2359.

Cormier, L. P. H. Gaskell, G. Calas, J. Zhao, and A. K. Soper. (1998). Environment around Li in the LiAlSiO4 ionic conductor glass: A neutron-scattering and reverse Monte Carlo study. Physical Review B, 57 (14): R8067-R8070.

Crank, J., (1975). The Mathematics of Diffusion. 2nd edn. Oxford: Oxford University Press.

Page 144: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

131

Davidson, J. P. (1996). Deciphering mantle and crustal signatures in subduction zone magmatism. Subduction top to bottom, Geophysical Monograph 96: 251-262.

De Grave, E. and S. G. Eeckhout. (2003). 57Fe moessbauer-effect studies of Ca-rich, Fe-bearing clinopyroxenes; part III, diopside. American Mineralogist, 88 (7): 1145-1152.

Decitre, S., E. Deloule, L. Reisberg, R. James, P. Agrinier, and C. Mével. (2002). Behaviour of Li and its isotopes during serpentinization of oceanic peridoites. Geochemistry, Geophysics, Geosystems - G3, 3 (1)

Dimanov, A., O. Jaoul, and V. Sautter. (1996). Calcium self-diffusion in natural diopside single crystals. Geochimica et Cosmochimica Acta, 60 (21): 4095-4106.

Dimanov, A., and V. Sautter. (2000). "Average" interdiffusion of (Fe,Mn)-Mg in natural diopside. European Journal of Mineralogy, 12 (4): 749-760.

Dimanov, A., and M. Wiedenbeck. (2006). (Fe, Mn)-Mg interdiffusion in natural diopside; effect of pO2. European Journal of Mineralogy, 18 (6): 705-718.

Dodson, M. H. (1973). Closure temperature in cooling geochronological and petrological systems. Contributions to Mineralogy and Petrology, 40 (3): 259-274.

Eggins, S. M., R. L. Rudnick, and W. F. McDonough. (1998). The composition of peridotites and their minerals; a laser-ablation ICP-MS study. Earth and Planetary Science Letters, 154 (1-4): 53-71.

Elliott T., T. Plank, A. Zindler, W. White, and B. Bourdon. (1997). Element transport from slab to volcanic front at the Mariana Arc. Journal of Geophysical Research, 102 (B7): 24,991-15,019.

Elliott T., A. Jeffcoate, and C. Bouman. (2004). The terrestrial Li isotope cycle: light-weight constraints on mantle convection Earth and Planetary Science Letters, 220: 233-245

Ernst, W. G. (1988). Tectonic history of subduction zones inferred from retrograde blueschist P-T paths. Geology, 16 (12): 1081-1084.

Faure, G. (1986). Principles of isotope geology. 2nd edn. New York: John Wiley & Sons.

Falkowski, P. G., and L. V. Godfrey (2008). Electrons, life and the evolution of the Earth’s oxygen cycle. Philosophical Transactions of the Royal Society B, 363 (1504): 2705-2716.

Foustoukos, D. I., R. H. James, M. E. Berndt, and W. E. Seyfried. (2004). Lithium isotopic systematics of hydrothermal vent fluids at the main Endeavour field, northern Juan de Fuca ridge. Chemical Geology, 212 (1-2): 17-26.

Ganguly, J., M. Ito and X. Zhang. (2007). Cr diffusion in orthopyroxene; experimental determination, 53Mn-53Cr thermochronology, and planetary applications. Geochimica et Cosmochimica Acta, 71 (15): 3915-3925.

Giletti, B. J., and T. M. Shanahan. (1997). Alkali diffusion in plagioclase feldspar. Chemical Geology, 139 (1-4): 3-20.

Grove, T. L. C. B. Till, E. Lev, N. Chatterjee, and E. Médard. (2009). Kinematic variables and water transport control the formation and location of arc volcanoes. Nature, 459: 694-697.

Page 145: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

132

Gurenko, A. A., and H. U.Schmincke. (2002). Orthopyroxene-bearing tholeiites of the Iblean plateau (Sicily); constraints on magma origin and evolution from glass inclusions in olivine and orthopyroxene; melt inclusions at the millennium; toward a deeper understanding of magmatic processes. Chemical Geology, 183 (1-4): 305-331.

Häendel, D., K. K. Mühle, H. M. Nitzsche, G. Stiehl, and U. Wand. (1986). Isotopic variations of the fixed nitrogen in metamorphic rocks. Geochimica et Cosmochimica Acta, 50: 749–758.

Halama, R., W. F. McDonough, R. L. Rudnick, and K. Bell. (2008). Tracking the lithium isotopic evolution of the mantle using carbonatites. Earth and Planetary Science Letters, 265 (3-4): 726-42.

Hallam, M. and H. P. Eugster (1976). Ammonium silicate stability relations. Contributions to Mineralogy and Petrology, 57(3): 227-244.

Hanschmann, G. (1981). Berechnung von isotopieeffekten auf quantenchemischer grundlage am beispiel stickstoffhaltiger molecule. ZfI-Mitteilungen Akademie der Wissenschaften der DDR : Berichte des Zentralinstituts für Isotopen- und Strahlenforschung, 41: 19–31.

Hardee, H. C. J. C. Dunn, J. D. Ward, and R. D. Hills. (1981). Probing the melt zone of Kilauea Iki Lava Lake, Kilauea Volcano, Hawaii. Geophysical Research Letters, 8 (12): 1211-1214.

Harrison, T. M. and E. B. Watson. (1983). Kinetics of zircon dissolution and zirconium diffusion in granitic melts of variable water content. Contributions to Mineralogy and Petrology, 84 (1): 66-72.

Hart, S. R., and T. Dunn. (1993). Experimental cpx/melt partitioning of 24 trace elements. Contributions to Mineralogy and Petrology, 113 (1): 1-8.

Hemingway, B. S. (1990). Thermodynamic properties for bunsenite, NiO, magnetite, Fe3O4, and hematite, Fe2O3, with comments on selected oxygen buffer reactions. American Mineralogist 75, (7-8): 781-790.

Hill, E., B. J. Wood, and J. D. Blundy. (2000). The effect of Ca-tschermaks component on trace element partitioning between clinopyroxene and silicate melt; element partitioning in geochemistry and petrology. Lithos, 53 (3-4): 203-215.

Holland, T. and R. Powell. (1991). A Compensated-Redlich-Kwong (CORK) equation for volumes and fugacities of CO2 and H2O in the range 1 bar to 50 kbar and 100-1600oC. Contributions to Mineralogy and Petrology, 109: 265-273.

Holloway, J. R. (1971). Composition of fluid phase solutes in a basalt-H2O-CO2 system. Geological Society of America Bulletin, 82 (1): 233-237.

Holloway, J.M., and Dahlgren, R.A., (2002). Nitrogen in rock: Occurrences and biogeochemical implications. Global Biogeochemical Cycles, 16: 1118.

Huh, Y., L. H. Chan, and J. M . Edmund. (2001). Lithium isotopes as a probe of weathering processes; Orinoco river. Earth and Planetary Science Letters, 194 (1-2) 189-199.

Huh, Y., L. H. Chan, L. Zhang, and J. M. Edmund. (1998). Lithium and its isotopes in major world rivers; implications for weathering and the oceanic budget. Geochimica et Cosmochimica Acta, 62 (12): 2039-2051.

Page 146: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

133

Imai, N., S. Terashima, S. Itoh and A. Ando. (1995). 1994 Compilation of analytical data for minor and trace elements in seventeen GJG geochemical reference samples, “Igneous Rock Series”. Geostandards Newsletter, 19 (2): 135-213.

Ionov, D. A., I.V. Ashchepkov, H. G. Stosch and G. Witt. (1993). Garnet peridotite xenoliths from the Vitim volcanic field, Baikal region: the nature of the garnet-spinel peridotite transition zone in the continental mantle. Journal of Petrology, 34 (6): 1141-1175.

Ionov, D. A., and H. M. Seitz. (2008). Lithium abundances and isotopic compositions in mantle xenoliths from subduction and intra-plate settings; mantle sources vs. eruption histories. Earth and Planetary Science Letters, 266 (3-4): 316-331.

Ishikawa, T., and E. Nakamura. (1994). Origin of the slab component in arc lavas from across-arc variation of B and Pb isotopes. Nature, 370 (6486): 205-208.

Ishikawa, T., and F. Tera. (1999). Two isotopically distinct fluid components involved in the Mariana arc; evidence from Nb/B ratios and B, Sr, Nd, and Pb isotope systematics. Geology, 27 (1): 83-86.

Ishikawa, T., F. Tera, and T. Nakazawa. (2001). Boron isotope and trace element systematics of the three volcanic zones in the Kamchatka arc. Geochimica et Cosmochimica Acta, 65 (24): 4523-4537.

Jahn, S. and B. Wunder. (2009). Lithium speciation in aqueous fluid at high P and T studied by ab initio molecular dynamics and consequences for Li-isotope fractionation between minerals and fluids. Geochimica et Cosmochimica Acta, 73 (18): 5428-5434.

Jagoutz, E., H. Palme, H. Baddenhausen, K. Blum, M. Cendales, G. Dreibus, B. Spettel, V. Lornez and H. Wänke. (1979). The abundances of major, minor and trace elements in the earth's mantle as derived from primitive ultramafic nodules; early solar system and lunar regolith. Proceedings of the Lunar and Planetary Science Conference, 10 (2): 2031-2050.

James, R. H., D. E. Allen, and W. E. Seyfried. (2003). An experimental study of alteration of oceanic crust and terrigenous sediments at moderate temperatures (51 to 350oC); insights as to chemical processes in near-shore ridge-flank hydrothermal systems. Geochimica et Cosmochimica Acta, 67 (4): 681-691.

James, R. H., and M. R. Palmer. (2000). The lithium isotope composition of international rock standards. Chemical Geology, 166 (3-4): 319-326.

Jeffcoate, A. B., T. Elliott, S.A. Kasemann, D. Ionov, K. Cooper, R. Brooker. (2007). Li isotope fractionation in peridotites and mafic melts. Geochimica et Cosmochimica Acta, 71 (1): 202-218.

Jia, Y., and R. Kerrich. (1999). Nitrogen isotope systematics of mesothermal lode gold deposits: metamorphic, granitic, meteoric water, or mantle origin? Geology, 27: 1051-1054.

Jia, Y., and R. Kerrich. (2004). Secular evolution of 15N in crustal fluids; implications for the origin of earth's early atmosphere and hydrosphere; proceedings of the eleventh international symposium on water-rock interaction. Proceedings - International Symposium on Water-Rock Interaction, 11: 945-949.

Jia, Y. (2006). Nitrogen isotope fractionations during progressive metamorphism: A case study from the Paleozoic Cooma metasedimentary complex, southeastern Australia. Geochimica et Cosmochimica Acta, 70: 5201-5214.

Page 147: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

134

Kent, A. J., J. Blundy, K. V. Cashman, K. M. Cooper, C. Donnelly, J. S.Pallister, M. Reagan, M. C. Rowe, and C. R. Thornber. (2007). Vapor transfer prior to the October 2004 eruption of Mount St. Helens, Washington. Geology, 35 (3): 231-4.

Keppler, H., (1989). A new method for the generation of N2-containing fluids in high-pressure experiments. European Journal of Mineralogy, 1: 135-137.

Kobayashi, K., R. Tanaka, T. Moriguti, K. Shimizu, and E. Nakamura. (2004). Lithium, boron, and lead isotope systematics of glass inclusions in olivines from Hawaiian lavas; evidence for recycled components in the Hawaiian plume; lithium isotope geochemistry. Chemical Geology, 212 (1-2): 143-161.

Lagache, M., and S. C. Dujon. (1987). Distribution of strontium between plagioclases and 1 molar aqueous chloride solutions at 600oC, 1.5 kbar and 750oC, 2 kbar. Bulletin De Mineralogie, 110 (5): 551-561.

Leeman, W. P. (1996). Boron and other fluid-mobile elements in volcanic arc lavas; implications for subduction processes. Subduction top to bottom, Geophysical Monograph 96: 269-276.

Leeman, W. P., S. Tonarini, L. H. Chan, and L. E. Borg. (2004). Boron and lithium isotopic variations in a hot subduction zone; the southern Washington Cascades; lithium isotope geochemistry. Chemical Geology, 212 (1-2): 101-124.

Lundstrom, C. C., M. Chaussidon, A. T. Hsui, P. Kelemen, and M. Zimmerman. (2005). Observations of Li isotopic variations in the Trinity ophiolite; evidence for isotopic fractionation by diffusion during mantle melting. Geochimica et Cosmochimica Acta, 69 (3): 735-751.

Luth, R.W., and D. Canil. (1993). Ferric iron in mantle-derived pyroxenes and a new oxybarometer for the mantle. Contributions to Mineralogy and Petrology, 113 (2): 236-248.

Lynton, S. J., R. J. Walker, and P. A. Candela. (2005) Lithium isotopes in the system Qz-Ms-fluid: An experimental study. Geochimica et Cosmochimica Acta, 69 (13): 3337-3347.

Magna, T., U. H. Wiechert, and A. N. Halliday. (2004). Low-blank separation and isotope ratio measurement of small samples of lithium using multiple-collector ICPMS. International Journal of Mass Spectrometry, 239 (1): 67–76.

Magna, T., U. Wiechert, and A. N. Halliday. (2006). New constraints on the lithium isotopic compositions of the Moon and terrestrial planets. Earth and Planetary Science Letters, 243 (3-4): 336-353.

Majérus, O., L. Cormier, G. Calas and B. Beuneu. (2003). Structural modifications between lithium-diborate glasses and melts: implications for transport properties and melt fragility. Journal of Physical Chemistry B, 107: 13044-13050.

Marschall, H. R., R. Altherr, T. Ludwig, A. Kalt, K. Gméling, and Z. Kasztovszky. (2006). Partitioning and budget of Li, Be, and B in high-pressure metamorphic rocks. Geochimica et Cosmochimica Acta, 70: 4750-4769.

Marschall, H. R., P. A. E. Pogge von Strandmann, H. M. Seitz, T. Elliott and Y. Niu. (2007). The lithium isotopic composition of orogenic eclogites and deep subducted slabs. Earth and Planetary Science Letters, 262 (3-4): 563-580.

Page 148: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

135

Marty, B. and F. Humbert (1997). Nitrogen and argon isotopes in oceanic basalts. Earth and Planetary Science Letters, 152 (1-4): 101-112.

Matthews, W., R. L. Linnen, and Q. Guo. (2003). A filler-rod technique for controlling redox conditions in cold-seal pressure vessels. American Mineralogist, 88 (4): 701-707.

Millot, R., C. Guerrot, and N. Vigier. (2004). Accurate and high-precision measurement of lithium isotopes in two reference materials by MC-ICP-MS. Geostandards and Geoanalytical Research, 28 (1): 153-159.

Moriguti, T., and E. Nakamura. (1998). Across-arc variation of Li isotopes in lavas and implications for crust/mantle recycling at subduction zones. Earth and Planetary Science Letters, 163 (1-4): 167-174.

Moriguti, T., T. Shibata, and E. Nakamura. (2004). Lithium, boron and lead isotope and trace element systematics of Quaternary basaltic volcanic rocks in northeastern Japan: mineralogical controls on slab-derived fluid composition. Chemical Geology, 212: 81-100.

Navon, O., and E. Stolper. (1987). Geochemical consequences of melt percolation; the upper mantle as a chromatographic column. Journal of Geology, 95 (3): 285-307.

Nishio, Y., S. Nakai, J. Yamamoto, H. Sumino, T. Matsumoto, V. Prikhod’ko, and S. Arai. (2004). Lithium isotopic systematics of the mantle-derived ultramafic xenoliths; implications for EM1 origin. Earth and Planetary Science Letters, 217 (3-4): 245-261.

O'Neill, H. St C., and M. I. Pownceby. (1993). Thermodynamic data from redox reactions at high temperatures; I, an experimental and theoretical assessment of the electrochemical method using stabilized zirconia electrolytes, with revised values for the Fe- "FeO", Co-CoO, Ni-NiO and Cu-Cu2O oxygen buffers, and new data for the W-WO2 buffer. Contributions to Mineralogy and Petrology, 114 (3): 296-314.

O'Neill, H. St C., and M. I. Pownceby. (1993). Thermodynamic data from redox reactions at high temperatures; II, the MnO-Mn3O4 oxygen buffer, and implications for the thermodynamic properties of MnO and Mn3O4. Contributions to Mineralogy and Petrology, 114 (3): 315-320.

Paquin, J., and R. Altherr. (2002). Subduction-related lithium metasomatism during exhumation of the Alpe Arami ultrahigh-pressure garnet peridotite (Central Alps, Switzerland). Contributions to Mineralogy and Petrology, 143 (5): 623-640.

Parkinson, I. J., S. J. Hammond, R. H. James, and N. W. Rogers. (2007). High-temperature lithium isotope fractionation; insights from lithium isotope diffusion in magmatic systems. Earth and Planetary Science Letters, 257 (3-4): 609-621.

Peacock, S. M. (1993). The importance of blueschist → eclogite dehydration reactions in subducting oceanic crust. Geological Society of America Bulletin, 105 (5): 684-694.

Pearce, N. J. G., et al., (1997). A compilation of new and published major and trace element data for NIST SRM 610 and NIST SRM 612 glass reference materials. Geostandards Newsletter, 21(1):115-144

Pell, E.M., (1960). Diffusion of Li in Si at high T and the isotope effect. Physical Review, 119 (3): 1014-1021

Page 149: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

136

Peslier, A. H., and J. F. Luhr. (2006). Hydrogen loss from olivines in mantle xenoliths from Simcoe (USA) and Mexico: Mafic alkalic magma ascent rates and water budget of the sub continental lithosphere. Earth and Planetary Science Letters, 242: 302-319.

Pinti, D., K. Hashizume, and J. Matsuda. (2001). Nitrogen and argon signatures in 3.8 to 2.8 Ga metasediments; clues on the chemical state of the Archean ocean and the deep biosphere. Geochimica et Cosmochimica Acta, 65 (14): 2301-2315.

Pistiner, J. S., and G. M. Henderson. (2003). Lithium-isotope fractionation during continental weathering processes. Earth and Planetary Science Letters, 214 (1-2): 327-339.

Plank, T., L. B. Cooper, and C. E. Manning. (2009). Emerging geothermometers for estimating slab surface temperatures. Nature Geoscience Progress Article, NGEO0614. Retrieved from www.nature.com/naturegeoscience

Pöter, B., M. Gottschalk, and W. Heinrich. (2004). Experimental determination of the ammonium partitioning among muscovite, K-feldspar, and aqueous chloride solutions. Lithos, 74: 67-90.

Purton, J. A., N. L.Allan, and J. D. Blundy. (1997). Calculated solution energies of heterovalent cations in forsterite and diopside; implications for trace element partitioning. Geochimica et Cosmochimica Acta, 61 (18): 3927-3936.

Richter, F. M., A. M. Davis, D. J. DePaolo, and E. B. Watson. (2003). Isotope fractionation by chemical diffusion between molten basalt and rhyolite. Geochimica et Cosmochimica Acta, 67 (20): 3905-3923.

Richter, F. M., N. Dauphaus, and F. Z. Teng. (2009). Non-traditional fractionation of non-traditional isotopes: Evaporation, chemical diffusion and Soret diffusion. Chemical Geology, 258: 92-103.

Rudnick, R. L., and D. A. Ionov. (2007). Lithium elemental and isotopic disequilibrium in minerals from peridotite xenoliths from Far East Russia: Product of recent melt/fluid-rock reaction. Earth and Planetary Science Letters, 256 (1-2): 278-293.

Ryan, J. G., and P. R. Kyle. (2004). Lithium abundance and lithium isotope variations in mantle sources; insights from intraplate volcanic rocks from Ross island and Marie Byrd Land (Antarctica) and other oceanic islands; lithium isotope geochemistry. Chemical Geology, 212 (1-2): 125-142.

Ryan, J. G., and C. H. Langmuir. (1987). The systematics of lithium abundances in young volcanic rocks. Geochimica et Cosmochimica Acta, 51 (6): 1727-1741.

Ryerson, F. J., and K. D. McKeegan. (1994). Determination of oxygen self-diffuion in åkermanite, anorthite, diopside, and spinel: Implications for oxygen isotopic anomalies and the thermal histories of Ca-Al-rich inclusions. Geochimica et Cosmochimica Acta, 58 (17): 3713-3734.

Sadofsky, S. J., and G. E. Bebout (2000). Ammonium partitioning and nitrogen-isotope fractionation among coexisting micas during high-temperature fluid-rock interactions; examples from the New England Appalachians. Geochimica et Cosmochimica Acta, 64 (16): 2835-2849.

Page 150: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

137

Sadofsky, S. J., and G. E. Bebout (2004). Nitrogen geochemistry of subducting sediments; new results from the Izu-Bonin-Mariana margin and insights regarding global nitrogen subduction. Geochemistry, Geophysics, Geosystems – G3, 5.

Seitz, H. M., G. P. Brey, Y. Lahaye, S. Durali, and S. Weyer. (2004). Lithium isotopic signatures of peridotite xenoliths and isotopic fractionation at high temperature between olivine and pyroxenes; lithium isotope geochemistry. Chemical Geology, 212, (1-2): 163-177.

Seitz, H. M., and A. B. Woodland. (2000). The distribution of lithium in peridotitic and pyroxenitic mantle lithologies; an indicator of magmatic and metasomatic processes. Chemical Geology, 166 (1-2): 47-64.

Sephton, M. A., K. Amor, I. A. Franchi, P. B. Wignall, R. Newton and J. P. Zonneveld. (2009). Carbon and nitrogen isotopic disturbances and an end-Norian (Late Triassic) extinction event. Geology, 30: 1119-1122.

Seyfried, W. E., X. Chen, and L. H. Chan. (1998). Trace element mobility and lithium exchange during hydrothermal alteration of seafloor weathered basalt; an experimental study at 350oC, 500 bars. Geochimica et Cosmochimica Acta, 62 (6): 949-960.

Shannon, R. D., and C. T. Prewitt (1969). Effective ionic radii in oxides and fluorides. Acta Crystallographica Section B, 25 (5): 925-946.

Sneeringer, M., S. R. Hart, and N. Shimizu. (1984). Strontium and samarium diffusion in diopside. Geochimica et Cosmochimica Acta, 48 (8): 1589-1608.

Stolper, E., and S. Newman. (1994). The role of water in the petrogenesis of Mariana Trough magmas. Earth and Planetary Science Letters, 121 (3-4): 293-325.

Suzuki, T., and M. Akaogi. (1995). Element partitioning between olivine and silicate melt under high pressure. Physics and Chemistry of Minerals, 22 (7): 411-418.

Taura, H., H. Yurimoto, K. Kurita, and S. Sueno. (1998). Pressure dependence on partition coefficients for trace elements between olivine and the coexisting melts. Physics and Chemistry of Minerals, 25 (7): 469-484.

Taylor, H. P. Jr. (1977). Water/rock interactions and the origin of H2O in granitic batholiths. Journal of the Geological Society of London, 133 (6): 509-558.

Teng, F. Z., W. F. McDonough, R. L. Rudnick, C. Dalpé, P. B. Tomascak, B. W. Chappell and S. Gao. (2004). Lithium isotopic composition and concentration of the upper continental crust. Geochimica et Cosmochimica Acta, 68 (20): 4167-4178.

Teng, F. Z., W. F. McDonough, R. L. Rudnick, and R. J. Walker. (2006). Diffusion-driven extreme lithium isotopic fractionation in country rocks of the Tin Mountain pegmatite. Earth and Planetary Science Letters, 243 (3-4): 701-710.

Tomascak, P. B. (2004). Developments in the understanding and application of lithium isotopes in the earth and planetary sciences; geochemistry of non-traditional stable isotopes. Reviews in Mineralogy and Geochemistry, 55: 153-195.

Tomascak, P. B., C. H. Langmuir, P. J. le Roux, and S. B. Shirey. (2008). Lithium isotopes in global mid-ocean ridge basalts. Geochimica et Cosmochimica Acta, 72 (6): 1626-1637.

Page 151: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

138

Tomascak, P. B., E. Widom, L. D. Benton, S. L. Goldstein, and J. G. Ryan. (2002). The control of lithium budgets in island arcs. Earth and Planetary Science Letters, 196 (3-4): 227-238.

Tomascak, P. B., J. G. Ryan, and M. Defant. (2000). Lithium isotopic evidence for light element decoupling in the Panama subarc mantle. Geology, 28 (6): 507-510.

Tomascak, P. B., F. Tera, R. T. Helz, and R. J. Walker. (1999). The absence of lithium isotope fractionation during basalt differentiation: New measurements by multi-collector ICP-MS. Geochimica et Cosmochimica Acta, 63: 901-910.

Tsai, T. L., and R. Dieckmann. (2002). Variation of the oxygen content and point defects in olivines, (FexMg1-x)2SiO4, 0.2 ≤ x ≤ 1.0. Physics and Chemistry of Minerals, 29: 680-694.

Ulmer, P. (2001). Partial melting in the mantle wedge – the role of H2O in the genesis of mantle-derived ‘arc-related’ magmas. Physics of the Earth and Planetary Interiors, 127 (1-4): 215-232.

Urey, H. C. (1947). The thermodynamic properties of isotopic substances. Journal of the Chemical Society, 1947: 562-581.

Van Orman, J. A., T. L .Grove, and N. Shimizu. (2001). Rare earth element diffusion in diopside; influence of temperature, pressure, and ionic radius, and an elastic model for diffusion in silicates. Contributions to Mineralogy and Petrology, 141 (6): 687-703.

Van Orman, J. A., T. L .Grove, and N. Shimizu. (1998). Uranium and thorium diffusion in diopside. Earth and Planetary Science Letters, 160 (3-4): 505-519.

Wenger, M., and T. Armbruster. (1991) Crystal chemistry of lithium; oxygen coordination and bonding. European Journal of Mineralogy, 3 (2): 387-99.

Williams, L.B., R. E. Ferrell Jr., I. Hutcheon and A. J. Bakel, M. M. Walsh and H. R. Krouse. (1995). Nitrogen isotope geochemistry of organic matter and minerals during diagenesis and hydrocarbon migration. Geochimica et Cosmochimica Acta, 54: 765–779.

Woodland, A. B., H. M. Seitz, and G. M. Yaxley. (2004). Varying behaviour of Li in metasomatised spinel peridotite xenoliths from western Victoria, Australia; trace element fingerprinting; laboratory studies and petrogenetic processes. Lithos, 75 (1-2): 55-66.

Woodland, A. B., H. M.Seitz, and R. Altherr, H. Marschall, B. Olker, and T. Ludwig. (2002). Li abundances in eclogite minerals; a clue to crustal or mantle origin? Contributions to Mineralogy and Petrology, 143 (5): 587-601.

Woods, S., S. Mackwell, and D. Dyar. (2000). Hydrogen in diopside; diffusion profiles. American Mineralogist, 85 (3-4): 480-487.

Wunder, B. A. Meixner, R. L. Romer, and W. Heinrich. (2006). Temperature-dependent isotopic fractionation of lithium between clinopyroxene and high pressure hydrous fluids. Contributions to Mineralogy and Petrology, 151 (1): 112-120.

Wunder, B. A. Meixner, R. L. Romer, A. Feenstra, G. Schettler, W. Heinrich. (2007). Lithium isotope fractionation between Li-bearing staurolite, Li-mica and aqueous fluids: An experimental study. Chemical Geology, 238 (3-4): 277-290.

Yamaji, K., Y. Makita, H. Watanabe, A. Sonoda, H., Kanoh, T. Hirotsu, and K. Ooi. (2001) Theoretical estimation of lithium isotopic reduced partition function ratio for lithium ions in aqueous solution. Journal of Physical Chemistry A, 105 (3): 602-613

Page 152: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

139

Zack, T., P. B. Tomascak, R. L. Rudnick, C. F. Dalpe, and W. F. McDonough. (2003). Extremely light Li in orogenic eclogites; the role of isotope fractionation during dehydration in subducted oceanic crust. Earth and Planetary Science Letters, 208 (3-4): 279-290.

Zanetti, A., M. Tiepolo, R. Oberti, and R. Vannucci. (2004). Trace-element partitioning in olivine; modelling of a complete data set from a synthetic hydrous basanite melt; trace element fingerprinting; laboratory studies and petrogenetic processes. Lithos 75, (1-2): 39-54.

Zhang, L., L. H. Chan, and J. M. Gieskes. (1998). Lithium isotope geochemistry of pore waters from ocean drilling program sites 918 and 919, Irminger Basin. Geochimica et Cosmochimica Acta, 62 (14): 2437-2450.

Page 153: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

140

Appendix 1

6 Summary of Boron Work

Boron is commonly used as a tracer of mineral-fluid reactions; however, to date there is very

little information on the degree of isotopic fractionation that may occur during fluid-mineral

partitioning. Without this information, the mixing models developed by many workers to

describe such systems as mantle metasomatism or arc magmatism are qualitative at best.

6.1 11B notation

Boron has two stable isotopes, 11B and 10B, whose abundances are ~80 % and ~20 %,

respectively. Most commonly, enrichments in boron isotopes are described as:

(1)

10001BB/

BB/B

std1011

smp1011

11

where smp refers to the sample and std refers to the standard, typically NBS SRM 951 for silicate

materials.

6.2 Evidence of Boron Mobility from Arc Lavas

Boron and boron isotopes are often employed to unravel the mechanisms involved during the

slab to mantle transfer of material in subduction zones (Ishikawa et al., 2001; Bebout et al., 1999;

Sano et al., 2001; Ishikawa and Tera, 1999; Benton et al., 2001). Due to boron’s affinity for

phyllosilicates, both altered oceanic crust (AOC) and pelagic sediments are enriched in these

elements relative to both mid-ocean ridge basalts (MORB) and oceanic island basalts (OIB)

(Table A1; Leeman, 1996).

Page 154: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

141

Table A1: Summary of B concentration and isotopic composition

B (ppm) 11B (‰)

Marine Sediments 15-160a -6.6 to +4.8a

AOC 3-63b +1 to +10b

OIB < 2c -15 to +1c

MORB < 0.1d -6 to -1d

Seawater 24e +39e

aIshikawa and Nakamura, 1993, bIshikawa and Nakamura, 1992 cPeacock and Hervig ,1999; dIshikawa and Tera, 1999; eVengosh et al.,

1995;

Because pelagic sediments and AOC are isotopically distinct from MORB and OIB, it is thought

that this slab input is reflected in the boron isotopic composition of arc lavas, which often differs

from the mantle (Ishikawa and Nakamura, 1992; --, 1993; Smith et al., 1995).

The extent to which these elements are lost by either dehydration or metasomatism to the

overlying mantle wedge or are incorporated into the mantle is not well known. Fluid and

sediment collected in fore arc environments provide evidence for mobilization and isotopic

fractionation of boron as a result of dehydration reactions in the subducting slab (Benton et al.,

2001). Clasts and muds from a serpentinite seamount in the Mariana fore arc were found to have

higher concentrations of boron and are enriched in 11B compared to seafloor sediments of the

area.

Several island arcs display higher ratios of boron to relatively immobile elements (such as Zr and

Nb in the front-arc regions) which systematically decrease towards the back-arc regions. For

example the Kamchatka arc lavas have the greatest enrichments in boron to Nb or Zr at the arc

Page 155: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

142

front; however, these enrichments decrease to MORB values with increasing slab depth

(Ishikawa et al., 2001; Wunder et al., 2005).

Figure A1 11B as a function of B/Nb in Kamchatka Arc Lavas

Plot of 11B versus B/Nb in arc lavas. Because the solid/melt partition coefficients for B and Nb are indistinguishable from one another, this trend cannot be the result of igneous processes. Rather this is suggestive of continuing mobilization of boron into the arc source region by B and 11B enriched fluids derived from dehydration reactions in the down going slab (data from Ishikawa et al., 2001).

Arc lavas display boron isotopic compositions that differ from mantle and MORB (Sadofsky and

Bebout, 2000). In both the Kamchatka arc and the Mariana arc, δ11B is most enriched at the arc

front where the highest B/Nb or B/Zr ratios occur (Ishikawa et al., 2001; Ishikawa and Tera,

1999). The δ11B in Kamtchatka lavas ranges from -4 to +6 ‰, and in Mariana lavas δ11B ranges

from +2.9 to +6.2 ‰ (Ishikawa et al., 2001; Ishikawa and Tera, 1999). In many cases researchers

point to the 11B enrichments in arc lavas as tracers of input from the subducted slab, either as

AOC or sediments, and propose various mixing models for these volcanic arcs; however, the

extent of boron isotope fractionation between fluid and residual solid are unknown and

quantitative modeling of this process is not possible.

6.3 Evidence of Boron Mobility from Eclogites

Only a few studies have examined the effects of progressive dehydration on isotopic

fractionation of boron isotopes. The δ11B values obtained from subduction zone metamorphic

rocks range from –11 to –3 ‰, which is generally lower than the δ11B values for seafloor

sediments and AOC (Peacock and Hervig, 1999).

Page 156: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

143

Figure A2 Evidence of Boron Mobility from Eclogites

Plot of 11B in eclogites versus estimated peak metamorphic temperature. Progressive dehydration results in lighter boron isotopic fraction in metamorphic rocks (from Peacock and Hervig, 1999)

These light stable isotopes are fractionated between phases, namely the exsolved fluids and the

host minerals, and unless the extent of this shift is known these isotopes cannot be fully utilized

as tracers.

6.4 Summary of Experimental Methodology

Experiments were planned to equilibrate muscovite with a fluid of known isotopic composition

at high pressures and temperatures using the cold-seal pressure vessels and piston-cylinder

apparatus at the University of Toronto's High Pressure Lab. Natural muscovite crystals from an

unknown locality in Ontario, Canada were to be used for most, if not all, of the boron

experiments.

6.5 Details of Boron Study

In silicate minerals boron is typically bonded to O; however, in silicate minerals the boron occurs

in tetrahedral coordination and is known to substitute for Al3+ or Si4+ (Schreyer et al., 2000). In

fluids, the isotopic enrichment depends on coordination of the species. Tetrahedrally coordinated

boron is 10B enriched, and trigonally coordinated boron is 11B enriched; therefore, the isotopic

fractionation depends almost entirely on the relative partitioning of B(OH)4- and B(OH)3

o

(Palmer, 1992), which in turn will depend on the speciation of boron in aqueous fluids. The boric

acid – borate equilibria can be written as:

(1) H OHB(OH)B(OH) 2-4

o3

Page 157: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

144

And the equilibrium constant K3 is expressed as:

(2) o3B(OH)

H4B(OH)3 a

aaK

Where a is the activity of the subscripted species.

The equilibrium constant (K) for reaction (2) can be derived from previous work of Mesmer et

al. (1972) calculated with SUPCRT92 (Helgeson et al., 1978; Johnson et al., 1992; Shock et al.,

1989). The Helgeson-Kirkham-Flowers equation of state for aqueous species limits the

applicability of SUPCRT92 to ≤5 kbar, and so extrapolation is necessary. It has been shown that

the logarithm of the equilibrium constant for many mineral hydrolysis reactions is linear with

logH 2O at constant T (Eugster and Baumgartner, 1987; Mesmer et al., 1988; Anderson et al.,

1991; Manning, 1994).

Figure A05 log K3 vs log density of water at 500°C, 2 - 4.5 kbar,

Calculated at 0.5 kbar increments using SUPCRT92. Error bars correspond to propagated uncertainties in thermodynamic data from Shock et al. (1989). These values can be reasonably fit by a straight line. This implies that a limiting slope method can be used to extrapolate K3 to P>5 kbar.

Using extrapolated values of K3, pK3 values for up to 30kbar have been calculated from:

(3) o3B(OH)4B(OH) logloglog 3 aa pHK

Using these values, the boric acid – borate equilibria was calculated for 2 – 20 kbar and 400-800 oC and are shown in Figure A06. At low P and T these values are consistent with those

Page 158: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

145

determined by Mesmer et al. (1972). Also shown for reference are neutral pH, and the pH that

would be buffered by the assemblage muscovite + potassium feldspar + quartz + KCl. In the case

of a dehydrating slab at the blueschist – eclogite transition, pH would be buffered at ~ 6

(Manning, 1998). As Figure A6 shows, a fluid with pH = 6 would have B(OH)4- as the dominant

species at 20 kbar but end up with B(OH)3o by 5 kbar. The role that boric acid – borate equilibria

during dehydration reactions plays in the isotopic fractionation has not been previously

addressed.

Palmer et al. (1992) conducted a study on boron-isotope fractionation with synthetic tourmaline,

and proposed that B(OH)3o initially is adsorbed onto the mineral surface. This initial adsorption

is suggested to control the isotopic fractionation because the boron symmetry changes to

‘psuedo-tetrahedral’, becomes enriched in 10B, and is incorporated into the structure without

further fractionation. Adsorption of dissolved B(OH)4- is not favoured because it involves the

breaking of a B-O bond. In their study, Palmer et al. (1992) concluded isotopic fractionation

between tourmaline and aqueous fluids decreased with increasing pressure. Palmer et al. (1992)

further state that B(OH)3o was the only B-species in the experiments because the formation of

trigonally coordinated species is favored at high pressures. This is contrary to recent studies that

have shown an increase in polymerization of hydrated species with increasing pressure (Zotov

and Keppler, 2002) and with predictions based on thermodynamic data.

Page 159: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

146

400

450

500

550

600

650

700

750

800

0 2 4 6 8 10 12 14

2 kbar

Te

mpe

ratu

re (

o C)

pH

B(OH)3 B(OH)

4

-

mu

s+

ksp

ar+

qtz

+K

Cl

neu

tral

pH

blu

esc

his

t -

ecl

og

ite

400

450

500

550

600

650

700

750

800

0 2 4 6 8 10 12 14

5 kbar

Te

mpe

ratu

re (

o C)

pH

B(OH)3

B(OH)4

-

neu

tral

pH

blu

esc

his

t -

ecl

og

ite

mu

s+

ks

pa

r+q

tz+

KC

l

400

450

500

550

600

650

700

750

800

0 2 4 6 8 10 12 14

10kbar

Te

mpe

ratu

re (

o C)

pH

B(OH)3 B(OH)

4

-

neu

tral

pH

mu

s+

ksp

ar+q

tz+

KC

l

blu

esc

his

t -

ecl

og

ite

400

450

500

550

600

650

700

750

800

0 2 4 6 8 10 12 14

20kbar

Te

mpe

ratu

re (

o C)

pH

B(OH)3

B(OH)4

-

neu

tral

pH

eclo

git

e -

blu

esch

ist

mu

s+

ksp

ar+

qtz

+K

Cl

Figure A6 calculated boric acid – borate equilibria for 2 – 20 kbar and 400-800 oC

Also shown for reference is neutral pH and pH as would be buffered by the assemblage muscovite + potassium feldspar + quartz + KCl. Note the effect of pH on speciation of boron in fluids, a fluid with pH = 6 would start off with B(OH)4

- as the dominant species at 20 kbar but end up with B(OH)3o by 5 kbar.

6.6 Boron Analyses

Initially it was planned to analyze the boron isotope composition of the run product mica using

the CAMECA ims1270 high-resolution secondary ion mass spectrometer at the Department of

Page 160: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

147

Terrestrial Magmatism under the supervision of Erik Hauri. The main concern going into the

study was to produce a large enough mica grain suitable for in situ analysis. Reconnaissance

analyses at UCLA revealed that there are significant challenges to boron analyses in micas. High

amounts of sample charging occurred during analyses (resulting in an apparent 20 ‰

fractionation from one side of the mount to the other, reproducible when sample mount was

rotated 90o). Switching to mono-collector mode appeared to resolve the initial charging issue;

however, this resulted in very long analyses. Compounded by the fact that the ion yield for micas

is very low so collection times are, by necessity, long; the amount of charging on the sample

increased. Increasing the concentration of boron in the micas would improve the counting

statistics and facilitate the analyses somewhat. One possibility would be to begin with boron rich

mica and attempt to measure diffusion out.

Page 161: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

148

6.7 References for Boron Study Anderson, G. M., S. Castet, et al. (1991). "The density model for estimation of thermodynamic

parameters of reactions at high temperatures and pressures." Geochimica et Cosmochimica Acta 55(7): 1769-1779.

Benton, L. D., et al. (2001). "Boron isotope systematics of slab fluids as inferred from a serpentine seamount, Mariana Forearc." Earth and Planetary Science Letters 187(3-4): 273-282.

Eugster, H. P. and L. Baumgartner (1987). "Mineral solubilities and speciation in supercritical metamorphic fluids

Faure, G. (1986). "[Monograph] Principles of isotope geology."

Helgeson, H. C., J. M. Delany, et al. (1978). "[Monograph] Summary and critique of the thermodynamic properties of rock-forming minerals."

Ishikawa, T. and E. Nakamura (1992). "Boron isotope geochemistry of the oceanic crust from DSDP/ODP Hole 504B." Geochimica et Cosmochimica Acta 56(4): 1633-1639.

Ishikawa, T. and E. Nakamura (1993). "Boron isotope systematics of marine sediments." Earth and Planetary Science Letters 117(3-4): 567-580.

Ishikawa, T. and F. Tera (1999). "Two isotopically distinct fluid components involved in the Mariana Arc; evidence from Nb/B ratios and B, Sr, Nd, and Pb isotope systematics." Geology 27(1): 83-86.

Ishikawa, T., et al. (2001). "Boron isotope and trace element systematics of the three volcanic zones in the Kamchatka Arc." Geochimica et Cosmochimica Acta 65(24): 4523-4537.

Johnson, J. W., E. H. Oelkers, et al. (1992). "SUPCRT92: A software package for calculating standard molal thermodynamic properties of minerals, gases, aqueous species, and reactions from 1 to 5000bar and 0 to 1000oC." Comp. Geosci 18: 899-947.

Leeman, W. P. (1996). "Boron and other fluid-mobile elements in volcanic arc lavas; implications for subduction processes

Manning, C. E. (1994). "The solubility of quartz in H2O in the lower crust and upper mantle." Geochimica et Cosmochimica acta 58: 4831-4839.

Manning, C. E. (1998). "Fluid composition at the blueschist-eclogite transition in the model system Na2O-MgO-Al2O3-SiO2-H2O-HCl." Swiss bulletin of Mineralogy and Petrology 78(2): 225-242.

Manning, C. E. and S. L. Boettcher (1994). "Rapid-quench hydrothermal experiments at mantle pressures and temperatures." American Mineralogist 79(11-12): 1153-1158.

Mesmer, R. E., C. F. Baes, et al. (1972). "Acidity Measurements at Elevated-Temperatures .6. Boric-Acid Equilibria." Inorganic Chemistry 11(3): 537-&.

Mesmer, R. E., W. L. Marshall, et al. (1988). "Thermodynamics of Aqueous Association and Ionization Reactions at High-Temperatures and Pressures." Journal of Solution Chemistry 17(8): 699-718.

Page 162: Experimental Constraints on Lithium Exchange between … · Natalie Caciagli Warman Doctor of Philosophy Department of Geology University of Toronto 2009 Abstract Clinopyroxene, olivine,

149

Palmer, M. R., D. London, et al. (1992). Experimental determination of fractionation of 11B/10B between tourmaline and aqueous vapor; a temperature- and pressure-dependent isotopic system. Frontiers in isotope geosciences. R. S. Harmon and R. W. Hinton. Amsterdam, Elsevier. 101: 123-129.

Paquin, J. and R. Altherr (2001). "New constraints on the P-T evolution of the Alpe Arami garnet peridotite body (Central Alps, Switzerland)." Journal of Petrology 42(6): 1119-1140.

Peacock, S. M. and R. L. Hervig (1999). Boron isotopic composition of subduction-zone metamorphic rocks. Interactions between slab and sub-arc mantle; dehydration, melting and element transport in subduction zones. D. S. Draper, A. D. Brandon and H. Becker. Amsterdam, Elsevier. 160: 281-290.

Sano, T., T. Hasenaka, et al. (2001). "Boron contents of Japan Trench sediments and Iwate basaltic lavas, Northeast Japan arc; estimation of sediment-derived fluid contribution in mantle wedge." Earth and Planetary Science Letters 186(2): 187-198.

Schreyer, W., U. Wodara, et al. (2000). "Synthetic tourmaline (olenite) with excess boron replacing silicon in the tetrahedral site; I, Synthesis conditions, chemical and spectroscopic evidence." European Journal of Mineralogy 12(3): 529-541.

Shock, E. L., Helgeson, H.C., and Sverjensky, D.A. (1989). " Calculation of the thermodynamic and transport properties of aqueous species at high pressures and temperatures: Standard partial molal properties of inorganic neutral species." Geochimica Cosmochimica Acta 53: 2157-2183.

Smith, H. J., A. J. Spivack, et al. (1995). The boron isotopic composition of altered oceanic crust. The mantle-ocean connection. H. Staudigel, F. Albarede, D. Hilton and T. Elliott. Amsterdam, Elsevier. 126: 119-135.

Tolstikhin, I. N. and B. Marty (1998). The evolution of terrestrial volatiles; a view from helium, neon, argon and nitrogen isotope modelling. The degassing of the Earth [modified]. M. R. Carroll, S. C. Kohn and B. J. Wood. Amsterdam, Elsevier. 147: 27-52.

Tomascak, P. B., et al. (2000). "Lithium isotope evidence for light element decoupling in the Panama subarc mantle." Geology 28(6): 507-510.

Vengosh, A., , et al. (1995). "Chemical and boron isotope compositions of non-marine brines from the Qaidam Basin, Qinghai, China." Chemical Geology 120(1-2): 135-154.

Wunder, B., et al. (2005). “The geochemical cycle of boron; constraints from boron isotope partitioning experiments between mica and fluid”. Lithos 84 (3-4): 206-16.

Zotov, N. and H. Keppler (2002). "Silica speciation in aqueous fluids at high pressures and high temperatures." Chemical Geology 184(1-2): 71-82.