California State University, Northridge

Strain Localization within Arc Crust: Microstructural Investigation of the Grebe Mylonite Zone in ,

A thesis submitted in partial fulfillment of the requirements For the degree of Master of Science in Geological Sciences

By Lonnie Hufford

August 2018 The thesis of Lonnie Hufford is approved:

______Dr. Keith Klepeis Date

______Dr. Joshua Schwartz Date

______Dr. Elena Miranda, Chair Date

California State University, Northridge

ii

ACKNOWLEDGMENTS I would like to thank the following people for their help and support with this thesis: Dr. Miranda’s ability to explain microstructural concepts in an understandable way was critical to my completion of this thesis. The professionalism and desire she had for me to produce the best work I could is something I deeply appreciate and have used to grow as a scientist. Dr. Schwartz’s geochemical knowledge and patience when I was trying to interpret my data is extremely appreciated. Dr. Klepeis’ ideas prompted me to think about how the Grebe Mylonite Zone compares to other regions which helped me focus on large-scale Fiordland geology. My family for their love and support. It is a blessing to have a mother, stepfather, and sisters who are always there for you. Dr. Peng for saving the SEM, even on the weekends. Dr. Andrew Cross for his help with MTEX code. Dr. Turnbull for her help with interpreting geochemistry data. Dr. Hielscher for his help with MTEX code. All the faculty and staff at CSUN for their help and support. Caroline, Courtney, Luisa, and Solishia for being such amazing research team peers. Working through MTEX code, talking through ideas, and traveling with you was wonderful. Veselina for her willingness to help proofread and make me laugh. All the graduate students at CSUN for their support and comradery in the office.

iii

TABLE OF CONTENTS

Signature Page ii Acknowledgments iii List of Figures v List of Tables vi Abstract vii 1. Introduction 1 2. Geologic Background 6 3. Methods 8 3.1 Sample Preparation 8 3.2 EBSD 9 3.3 EBSD Data Post-Processing 10 4. Results 13 4.1 Optical Microscopy- Microstructures 13 4.2 EBSD Results 17 5. Discussion 29 5.1 Deformation and Recrystallization 29 5.2 CVA Analysis 34 5.3 Relation back to the GMZ 38 5.4 Broad Microstructural Implications 45 6. Conclusion 46 References 48

iv

LIST OF FIGURES

Fig. 1- Geologic Maps 5 Fig. 2- Photomicrograph of 17NZ95C1 14 Fig. 3- Photomicrograph of 17NZ91 16 Fig. 4- Phase map of 17NZ95C1 20 Fig. 5- AE, GOS, and Local Misorientation maps of quartz in 17NZ95C1 21 Fig. 6- AE, GOS, and Local Misorientation maps of plagioclase in 17NZ95C1 22 Fig. 7- Pole figures and CVA figures from 17NZ95C1 23 Fig. 8- Phase map of 17NZ91 26 Fig. 9- AE, GOS, and Local Misorientation maps of quartz in 17NZ91 27 Fig. 10- Pole figures and CVA figures from 17NZ91 28 Fig. 11- Stereonet with GMZ data 39

v

LIST OF TABLES

Table 1- Parameters and working conditions for SEM-EBSD data 12

vi

ABSTRACT

Strain Localization within Arc Crust: Microstructural Investigation of the Grebe Mylonite Zone in Fiordland, New Zealand

By Lonnie Justin Hufford Master of Science in Geological Sciences

We investigate strain localization and deformation within the Grebe Mylonite Zone

(GMZ), a mid-crustal, transpressional shear zone in central Fiordland, New Zealand, using electron backscatter diffraction (EBSD) analysis. The EBSD analysis includes the use of microstructural, crystallographic orientation, and Crystallographic Vorticity Axis

(CVA) data. The study focuses on two samples: a mylonite, and a monzogranite mylonite. For each sample we define the rheology-controlling mineral, deformation mechanism, style of dynamic recrystallization, and deformation geometry. We find in the diorite mylonite that even though quartz is abundant, its isolated grains allowed the interconnected and rheologically stronger plagioclase to become the dominant rheology- controlling mineral. CVA analysis yields data suggesting plagioclase and quartz. The analysis of quartz in the monzogranite mylonite yields microstructural and crystallographic preferred orientation (CPO) data consistent with two populations of grain sizes and lattice distortion. One population is consistent with Grain Boundary

Migration (GBM) dynamic recrystallization (DRX) and the other is consistent with

vii

Subgrain Rotation (SGR) DRX. The quartz data show a correlation between the wrench- dominated transpression geometry with larger, GBM DRX grains and pure shear transpression geometry with the smaller, SGR DRX grains. We present two scenarios for the Grebe Mylonite Zone. The first scenario involves plagioclase that deformed under crystal-plastic processes and quartz that experienced GBM DRX to be associated with the wrench-dominated transpressional geometry, whereas the quartz that underwent GBM

DRX have a wrench-dominated transpressional geometry and quartz grains that underwent SGR DRX have pure shear-dominated transpressional geometry that were simultaneously partitioned. The second involves a change in the shear zone conditions that may include a decrease in temperature and/or an increase in strain rate. This resulted in crystal-plastic deformation ceasing in plagioclase, and quartz GBM DRX shifting to

SGR DRX and their corresponding deformation geometries described above. This allowed shear zone conditions to be accommodated such that work hardening was counteracted by neighboring grains favorably oriented for lower temperature and/or higher strain rate deformation, and these grains preserve the transition from wrench- dominated to pure shear-dominated transpression in quartz. This study provides insight into strain localization and deformation geometry and their connected roles in plagioclase and quartz.

viii

1. INTRODUCTION

Crustal strength within shear zones is often characterized by the strength of quartz and plagioclase, the dominant rheology-controlling minerals in the middle crust, as strain is localized into these abundant phases. To understand crustal composition, rheology, and strain localization within quartz-dominated and plagioclase-dominated shear zones, work has been done through modeling, experiments, and studies on naturally deformed rocks

(Kohlstedt et al. 1995; Montési, 2007; Gerbi et al. 2010; Behr and Platt, 2011; Platt and

Behr, 2011; Gueydan et al. 2014). Research on strain localization within quartz

(commonly focused on monomineralic cases) and the parameters that affect its deformation such as temperature, strain rate, and H2O content (Hirth and Tullis, 1992;

Kruhl, 1996; Stipp et al. 2002(a); Stipp et al. 2002(b); Toy et al. 2008; Law, 2014) continue to bring better accuracy of microstructural and crystallographic preferred orientation (CPO) interpretation. This research has also been applied to plagioclase and its recrystallization microstructures and CPO (Ji and Mainprice, 1990; Kruse et al. 2001;

Stünitz et al. 2003; Hansen et al. 2013; Miranda et al. 2016). These phases are not isolated from each other within the crust (Menegon et al. 2008), and may be found deformed together (Mancktelow and Pennacchioni, 2004) in the transition from quartz- dominated to plagioclase-dominated crust. Polymineralic deformation has been studied

(Herwegh et al. 2011; Little et al. 2015, Cross et al. 2017, Mulyukova and Bercovici,

2017), but less attention has been given to understanding how strain is localized within the quartz- and plagioclase-dominated crust and how it relates to strain partitioning within the 3D kinematic deformation geometry of a shear zone. Both strain localization

1 and 3D kinematic deformation geometry can be studied using electron backscatter diffraction (EBSD).

EBSD analysis provides information on microstructures, CPO, and lattice distortion which can be used to constrain the deformation mechanisms, interpret styles of dynamic recrystallization, and estimate deformation conditions within rocks. Ductile deformation in quartz and plagioclase can be a product of dislocation creep at mid-crustal levels and is identified through the presence of CPO (Ji and Mainprice, 1990; Hirth and

Tullis, 1992; Heilbronner and Tullis, 2006; Menegon et al. 2008). This deformation may be accompanied by dynamic recrystallization (DRX) (Hirth and Tullis, 1992; Stipp et al.

2002; Halfpenny et al. 2006). DRX involves the competition of deformation and recrystallization by promoting grain boundary mobility (Ji and Mainprice, 1990; Hirth and Tullis, 1992; Halfpenny et al. 2006). Understanding what regime or regimes of DRX

(Bulging (BLG), Subgrain Rotation (SGR), or Grain Boundary Migration (GBM)) were active through interpretation of microstructures and CPO may allow researchers to interpret temperature and strain rate of the shear zone (Hirth and Tullis, 1992; Stipp et al.

2002(b); Law 2014). The deformation and DRX of both phases need to be constrained to compare the localized strain and if this affects their deformation geometries.

Crystallographic vorticity axis (CVA) analysis is a method involving EBSD data to interpret the geometry of deformation within individual dynamically recrystallized grains by plotting their crystallographic vorticity axes in lower-hemisphere stereonets in the kinematic reference frame (Michels et al. 2015). One method to interpret deformation geometry is to calculate the kinematic vorticity axis, Wk, of the grains within the deformed phase (Tikoff and Fossen, 1995). Wk is used to calculate finite strain (Tikoff

2 and Fossen, 1995) and can be described as relating the rate of stretching in a rock to the local and instantaneous vorticity ratios (Iacopini, et al. 2008). Michels et al. 2015 created the MTEX code to calculate dispersion axes from crystallographic orientations and suggest that this method may be used to study deformation geometry between phases and within phases within specified domains such as grain size using EBSD data. We follow their suggestion to utilize CVA analysis to study deformation geometry between quartz and plagioclase and within quartz at different grain sizes. The use of CVA analysis may allow researchers to identify the bulk vorticity axis of a rock as well as its constituent minerals while relating it to the pure-shear dominated or wrench-dominated end members of transpression deformation geometry (Michels et al. 2015; Giorgis et al. 2017).

We focus the study on strain localization in quartz and plagioclase and their relation to deformation geometry within an exhumed, mid-crustal, ductile shear zone in

Central Fiordland, New Zealand named the Grebe Mylonite Zone (GMZ) (Fig. 1A). The

GMZ is a mid-crustal shear zone within a series of transpressional, crustal-scale shear zones in Fiordland (Scott et al. 2011), and it provides a unique exposure of both plagioclase- and quartz-rich mylonites. Two samples, 17NZ95C1 and 17NZ91, with contrasting interconnected fabrics of plagioclase and quartz are used in this study. They were collected along strike of the GMZ in the South Arm region of (Fig.

1B). Our objectives were to use EBSD analysis to 1) investigate strain localization within these rocks, allowing interpretation of the deformation mechanisms of both quartz and plagioclase, and 2) conduct CVA analysis (Michels et al. 2015) to evaluate how the transpressional shear zone deformation geometry is partitioned between mineral phases and how it varies as a function of grain size. We hypothesize that the deformation

3 geometry is partitioned differently between minerals owing to rheological competence contrasts, such that strong and weak minerals record early and late-stage deformation, respectively. We also hypothesize that dynamic recrystallization results in ‘parent’ and recrystallized grain size populations that also record the transition from early to late-stage deformation geometry, respectively.

Our results show the Grebe Mylonite Zone is a shear zone that underwent crystal- plastic deformation within a transpressional setting. We interpret quartz having undergone GBM DRX and was overprinted by SGR DRX as the shear zone evolved.

Plagioclase experienced BLG DRX in rocks with an interconnected plagioclase fabric.

CVA analysis in sample 17NZ95C1 suggests wrench-dominated transpressional deformation geometry in plagioclase and quartz that underwent GBM DRX. We interpret a transition from wrench-dominated transpressional deformation geometry in GBM DRX quartz grains to pure shear-dominated transpressional deformation geometry in the SGR

DRX quartz grains in sample 17NZ91. We have used structural and EBSD analysis to determine that this was a dextral, inclined, transpressional shear zone without the use of paleoreconstruction as has been done through past research (Scott et al. 2011).

4

5

2. GEOLOGIC BACKGROUND

Subduction off the coast of Gondwana resulted in the emplacement of arc-related magmas. This subduction zone was a product of the Panthalassan oceanic lithosphere

(Marcotte et al. 2005) subducting under Gondwana’s eastern coast during the

Carboniferous, Jurassic, and Cretaceous periods (Mortimer, 2004). Arc-parallel magmatism associated with the subduction zone has resulted in the emplacement of plutons parallel to the arc axis (Allibone et al. 2009; Schwartz et al. 2017). The Median

Batholith is region of arc crust composed of Carboniferous units, the Jurassic Darran

Suite, and the Cretaceous Separation Point Suite (SPS) plutonic units (Allibone et al.

2009, Klepeis et al. 2016). Inboard arc axis migration during a contractional or transpressional regime has been suggested to be a result of a change in subduction geometry ~128 Ma (Schwartz et al. 2017). Crustal thickening of the continental arc ceased ~106 Ma and lower crustal shear zones accommodated orogenic collapse, extension, and rifting (Klepeis et al. 2016; Schwartz et al. 2016).

The Grebe Mylonite Zone is a transpressional, ductile shear zone (Allibone et al.

2009) within the Median Batholith, and is associated with a series of shear zones including the Indecision Creek Shear Zone, Escarpment Fault, and Gutter Shear Zone to the south of Lake Manapouri on Stewart Island (Allibone et al. 2009; Scott et al. 2011).

The fault ranges in width from 200-1000 m (Scott, 2013) and strikes north with its outcrop extending northward through Lake Manapouri (Allibone et al. 2009) into the

Kepler Mountains (Scott et al. 2011) and to the south towards the Princess Mountains

(Allibone et al. 2009). Brittle faults cut across and near the GMZ, while the Cenozoic

Hauroko Fault cuts the GMZ at its southern exposure where it is reported to end

6

(Allibone et al. 2009). Cross-cutting relationships involving the Fowler Pluton and West

Arm leucogranite with the GMZ have resulted in constraining a date range for deformation activity in the Cretaceous from 129 to 115 Ma (Scott and Palin, 2008;

Allibone et al. 2009; Scott et al. 2011). This age range does not constrain the activation of the GMZ to before SPS emplacement or during emplacement. Recently, LASS-ICP-MS analysis and thermochronolgy of titanite contained within the mylonitic fabric of the

GMZ within the South Arm of Manapouri has provided a 206Pb/238U age constraint of

122-119 Ma under amphibolite facies conditions during the deformation that formed the mylonitic foliation (Buritica, 2018).

Our field measurements and observations describing the Grebe Mylonite Zone originate from the eastern side of the fjord in the South Arm of Lake Manapouri (Fig. 1B) and observations demonstrate that it is a sub-vertical, ductile shear zone. The site where we collected the monzogranite mylonite sample associated with the Foulwind Suite

(17NZ91) has a mylonitic foliation of 195, 81° NW and stretching lineation of 40°/198.

We observed porphyroclasts of K-feldspar and plagioclase showing dextral and sinistral shear sense at this location. Continuing south, the site where we collected the diorite mylonite sample associated with the Ridge Suite (17NZ95C1) has a mylonitic foliation oriented at 180, 70° W. A stretching lineation measured from this site yielded an orientation of 61°/232 when measured on a rock with a mylonitic foliation of 190, 65°

NW. Porphyroclasts of plagioclase were observed with a sinistral shear sense.

7

3. METHODS

3.1 Sample Preparation

We collected 6 oriented samples from the Grebe Mylonite Zone along the South

Arm of Manapouri during the 2017 field season and use 2 of the samples within this study. The measurements and structural data were collected with southern hemisphere

Suunto compasses. Factors including fabric development (sub- or hyper-solidus), lack of or minimal mineral alteration, accessibility of sample, sample distribution within the area, and rock type were considered when choosing samples and their locations. Orientation data were collected with the structural samples describing the strike and dip of the foliation and the trend and plunge of any mineral lineations.

The oriented rock samples were processed and cut with wet saws at the GNS location in Dunedin, New Zealand. The oriented sample billets were cut parallel to lineation and perpendicular to foliation, providing a way to view any kinematic indicators within the mylonitic fabrics of the samples. The thin section billets were then sent to

Spectrum Petrographics for processing into ultra-polished thin sections. The thin sections chosen for SEM-EBSD work went through 1-2 rounds of probe polishing which entailed using a colloidal silica solution and 1 hour of machine time with 4 lbs. of force, a speed of 25 rpm, and soft stop selected. This method provides an ultra-polished thin section acceptable for EBSD data collection. Optical microscopy work was done with a Nikon eclipse 50iPOL polarized microscope.

8

3.2 EBSD

Precise preparation for the thin sections chosen for EBSD analysis was undertaken. Copper tape was strategically placed on the thin sections to allow for copper standardization. This is needed for EDS mapping and to function as a “landmark” for navigating the thin section during the imaging process. The thin sections were cleaned with isopropyl alcohol to prevent any contamination from the nitrile gloves worn during preparation. A FEI Quanta 600 W-filament Scanning Electron Microscope was used with an Oxford Instrument Nordlys Nano ultrahigh sensitivity Electron Backscatter

Diffraction detector. An angled pedestal was utilized to allow the thin section to rest at

70° facing the EBSD detector as to maximize the EBSD detector-electron pathways. The

Oxford Instruments AZtec software was used to control and set the parameters for the

SEM. Vacuum settings were pumped first to “high” (~2.0 e-3 Pa) and then to “low” (~2.0 e2 Pa). This procedure is used to prevent charging within the chamber.

The beam settings changed between sample sites so as to produce the highest quality results. The beam varied from 25-30 kV and a range of spot sizes were used. The sample magnification ranged depending on the size of the map. Working distance was normally kept between 14-18mm. A 70° pre-tilt option was selected during the imaging process along with dynamic focus to correct for stage tilt and to produce the best images possible. The working conditions are available in Table 1.

Energy Dispersive X-ray Spectroscopy (EDS) data were collected in conjunction with EBSD data on most samples to verify mineral compositions observed in thin section.

O, Na, Mg, Al, Si, P, S, K, Ca, Ti, Fe, and Zr were primarily chosen for detection based on the ability of the detector and the compositions of the rocks being analyzed. Using the

9 data that the EDS maps provided in conjunction with the “optimize pattern” and search for mineral phase operations in the AZtec software, the best fit minerals were selected based on the lowest Mean Angular Deviation (MAD) numbers and Kikuchi pattern matches. The software used a maximum of 12 bands to compare Kikuchi patterns produced from the sample to patterns from multiple mineral databases in conjunction with EDS data. The study focuses primarily on quartz and plagioclase deformation and these were deemed critical for correct indexing. If accessory minerals were seen in thin section and detected through EDS they were included in the indexing process.

3.3 EBSD Data Post-Processing

Post-processing was completed using HKL Channel 5 software with Windows 7 per the methods described in Stewart and Miranda, (2017). We used this software to complete several functions including map stitching, cleaning and refining data, map creation, pole and inverse pole figure creation, and misorientation analysis. Grain boundaries are defined as having ≥ 10° of mis-orientation and misorientations from 2-10° are considered subgrain boundaries as specified in other EBSD studies (Trimby et al.

1998; Halfpenny et al., 2012; Stewart and Miranda, 2017). Data were cleaned and refined by extrapolating wild spikes and extrapolating zero solutions up to 7 neighboring pixels.

Removing systematic misindexing was completed when necessary on plagioclase to remove or lessen indexing errors. This function was completed by rotating the systematic spikes incrementally from “low” to “high”.

EBSD maps are used to interpret microstructures, lattice distortion, and where lattice distortion occurred. A series of EBSD maps were created with the Channel 5 software to analyze the phases present, the orientation of the grains, the accommodated

10 strain, and the location of low-angle boundaries (2-<10°). Phase maps define where grains in the sample area, grain boundaries (both high and low-angle boundaries), and twinning exist. All Euler (AE) maps are utilized to compare the orientations of grains of a single phase. Grain Orientation Spread (GOS) maps measure the average curvature of a grain and relate this to the amount of lattice distortion. Local Misorientation maps calculate where low-angle boundaries (<10°) within the grains exist. These maps used to display the intensity and location of subgrain wall development within grains.

We used MTEX (Hielscher and Schaeben, 2008) to plot crystallographic orientation and crystallographic vorticity axis data. To plot crystallographic orientation, we first calculated the orientation distribution function (ODF) from the EBSD data, then constructed lower-hemisphere, equal area pole figures in the kinematic reference frame

(parallel to lineation and perpendicular to foliation) to interpret CPO. We plotted crystallographic vorticity axis (CVA) data in lower-hemisphere, equal area stereonets, where the data are shown as contoured plots of the mean vorticity axis relative to foliation and lineation. MTEX was also used to rotate the quartz data in 17NZ91 to align the foliation with the kinematic reference frame. We use the methods and code described in Michels et al. (2015) to plot the CVA data.

Subsets of grains were created in both 17NZ91 and 17NZ95C1 based on grain size and/or lattice distortion. In 17NZ95C1, separate subsets of biotite, plagioclase and quartz were distinguished. The quartz was further subset into a band of interconnected grains (n=120) to separate them from the isolated population. In 17NZ91, 4 population subsets were distinguished. The first population consists of all the quartz grains (n=387) within the interconnected anastomosing band. The second population consists of grains

11

(n=10) relatively large with lobate grain boundaries and high lattice distortion. The third population of quartz consists of grains (n=86) between 50-100 µm and have relatively low lattice distortion. The fourth population of quartz consists of grains (n=163) 50 µm to smallest grain size and these have relatively low lattice distortion.

12

4. RESULTS

4.1 Optical Microscopy- Microstructures

17NZ95C1- Diorite Mylonite

17NZ95C1 (Fig. 2) consists of quartz, biotite, and plagioclase with trace amounts of apatite, muscovite, epidote, and zircon. Biotite and elongate plagioclase (~0.16 mm) define the foliation. The majority of quartz grains have rounded grain boundaries (~0.07 mm) and exist between plagioclase and biotite in isolated pockets. An interconnected quartz-rich band is located parallel to the mylonitic foliation. This interconnected quartz displays both lobate and straight grain boundaries and undulose extinction within several grains with grains up to 0.65 mm. Some albite twinning is preserved in plagioclase, with a few coarse-grained (~1 mm) sigma porphyroclasts present. Both sinistral and dextral porphyroclasts were observed.

13

14

17NZ91-Monzogranite Mylonite

17NZ91 (Fig. 3) is dominantly composed of quartz, plagioclase, and potassium feldspar with minor amounts of muscovite, biotite, garnet, epidote, and rutile. Quartz is present in anastomosing, interconnected bands with few preserved porphyroclasts. Finer grained quartz (~0.06 mm) appear next to coarser grains (~0.8 mm) in regions of the rock, such as next to feldspar porphyroclasts (~1.5 mm). Many of the coarser quartz grains have lobate grain boundaries, while finer grained quartz grains display rounded and straightened boundaries. Sigma porphyroclasts composed of both plagioclase and K- feldspar are common throughout the thin section. These sigma porphyroclasts generally display sinistral shear. Outside of the anastomosing quartz bands, potassium feldspar and quartz are observed interconnected with myrmekite.

15

16

4.2 EBSD Results

17NZ95C1- EBSD MAPS

QUARTZ

Two populations of quartz are distinguished within this site with one composed of isolated grains between plagioclase and biotite and the second forming an interconnected band (Fig. 4). Many of the interconnected quartz grains in the All Euler (AE) map display a gradation of similar orientations amongst adjacent grains (Fig. 5A). The coarser grains within the quartz-rich band display moderate levels of lattice distortion, while the smaller grains outside of the interconnected band have comparatively lower lattice distortion in the Grain Orientation Spread (GOS) map (Fig. 5B). The Local Misorientation map displays low-angle boundaries (<1°) developed parallel to and crosscutting Dauphiné twin boundaries (Fig. 5C).

PLAGIOCLASE

Plagioclase forms an interconnected fabric with distinct microstructures and deformation features. The majority of grain boundaries are elongate with straight to slightly rounded boundaries. Several of the coarser grains appear rounded or lobate (Fig.

4). A few grain bulges are observed, but are sparse amongst the population (Fig. 4). The

AE map displays a gradation of colors with many adjacent grains exhibiting similar orientations. The pink to light brown grains (Fig. 6A) are an example of these similar orientations separated by the quartz-rich band. The GOS map displays plagioclase grains with relatively low lattice distortion, though some grains with higher lattice distortion are in close proximity to the quartz-rich band (Fig. 6B). Lower lattice distortion grains are

17 observed bulging into higher lattice distortion grains (Fig. 6B). These bulges were measured in size from ~6 µm to 30 µm. The Local Misorientation map displays a lack of subgrain wall development within the majority of plagioclase grains (Fig. 6C). The areas where minor subgrain wall development occurs is in close proximity to other phases. The plagioclase grains share similar orientations with neighbor grains, exhibit low lattice distortion, and have low subgrain wall development.

BIOTITE

Biotite is present within the samples and helps define the foliation within the mylonite. The grain boundaries appear to have no deformation microstructures present such as mica fish. Biotite’s grain boundaries are straight and have slight curvatures in several grains (Fig. 4).

17NZ95C1- CPO AND CVA

QUARTZ

Quartz within the interconnected band exhibits CPO presented in the (0001) c- axis pole figure plot and the CVA plot has diffuse maxima (Fig. 7A). A strong, slightly diffuse maximum in the middle of the pole figure and weak maxima on the primitive

~10° clockwise from the pole to foliation is present in the c-axis pole figure (Fig. 7A).

Between the primitive and center of the pole figure, a weak and tightly clustered maximum is observed in the (0001) pole figure (Fig. 7A). The CVA plot shows strong, but diffuse maxima near the center of the stereonet, with weaker maxima oriented 20° from the lineation direction not extending to the primitive circle (Fig. 7A).

18

PLAGIOCLASE

Plagioclase exhibits CPO within pole figures and maxima in several orientations in the CVA plot. Plagioclase pole figure data exhibit strong CPO with slightly diffuse maxima plotting perpendicular to foliation on the (001) pole figure plot and parallel to lineation on the [100] plot (Fig. 7B). The CVA plot of plagioclase displays a diffuse, off- center maximum close to the primitive and weaker maxima on the primitive parallel to lineation (Fig. 7B).

BIOTITE

Biotite contributes in defining the foliation within the mylonite and displays CPO and strong maxima in the CVA plot. Strong CPO is present with maxima perpendicular to foliation in the (001) pole figure and weaker, diffuse girdles parallel to lineation in the

[100] and [110] plots (Fig. 7C). The CVA plot displays strong, slightly diffuse maxima on the primitive (Fig. 7C).

19

20

21

22

23

17NZ91- EBSD MAPS

Quartz exists in interconnected, anastomosing bands forming the majority of the mylonite’s matrix (Fig. 8). The All Euler (AE) map for quartz displays populations of adjacent grains with similar orientations (yellow grains) present with interstitial grains of populations with different orientations (green and blue grains) (Fig. 9A). Finer grains located at the boundaries of the coarser, lobate quartz grains display differences in orientation in comparison to the grains they rim. Grains displaying similar orientations and similar Dauphiné twins are separated by a feldspar porphyroclast (Fig. 9A).

Quarts Grain Orientation Spread (GOS) and Local Misorientation maps show a difference in the amount of lattice distortion with grain size. In the GOS map, coarser grains with lobate grain boundaries display hotter colors indicating higher lattice distortion, while smaller grains with cooler colors display dominantly low amounts of lattice distortion (Fig 9B). Several of the coarsest grains have extensive Dauphiné twinning and high amounts lattice distortion, while finer grains are located along the larger grains’ boundaries or within junctions of larger grains and have lower lattice distortion while the majority lack extensive Dauphiné twinning. The higher lattice distortion, coarse grains display significant subgrain development with trends occurring normal to foliation and sparse trends sub-parallel to foliation in the Local Misorientation map (Fig. 9C). Dauphiné twin boundaries intersect with these subgrain walls throughout the map with subgrains truncating at the Dauphiné twin boundaries. Subgrain wall development is weaker and less abundant within the finer grains. The subgrains observed in the grains with higher lattice distortion are of similar size to the smaller, lower lattice distorted grains with straightened to rounded grain boundaries.

24

17NZ91- CPO AND CVA

We observe differences in CPO and CVA maxima with changes in quartz grain size. The pole figures of the entire quartz population display CPO within the (0001) c- axis pole figure plot with 3 maxima offset ~30° clockwise from the pole to foliation (Fig.

10A). None of these maxima extend to the primitive of the pole figure. Several of the lobate, coarse grains displaying high strain are plotted as a subset and display strong CPO with a tight maximum near the center of the pole figure (Fig. 10B). The subset consisting of quartz grains from 50 to 100 µm exhibit CPO as a diffuse, broken girdle which does not reach the primitive, oriented 40° counterclockwise from the pole to foliation (Fig.

10C). In the smallest grains (~8 µm) to 50 µm subset, CPO is present within the (0001) plot as a weak, broken cross-girdle with a weak, isolated maximum on the primitive and weak to moderate pattern forming the girdle (Fig. 10D). The weak maxima on the primitive are oriented ~15° clockwise from the pole to foliation. The CVA plot for the entire quartz population displays a strong maximum in the center of the stereonet and diffuse, weaker maxima on the primitive oriented 10° counter clockwise from the lineation direction (Figure 10A). The lobate, high strain, coarse grain population displays a strong maximum in the center of the stereonet with a weak, diffuse girdle reaching the primitive oriented 15° counterclockwise from the lineation direction (Fig. 10B). The 50 to 100 µm grain population displays a strong, tight maximum in the center of the stereonet and strong, tight maxima on the primitive oriented 15° counterclockwise from the lineation direction (Fig. 10C). The grain population ranging from the smallest grains

(~8 µm) to 50 µm displays a diffuse, moderate maximum in the center of the stereonet, a

25 strong maximum rotated 20° clockwise from the pole to foliation which does not reach the primitive, and strong, diffuse maxima on and near the primitive (Fig 10D).

26

27

28

5. DISCUSSION

Our data is suggestive that the Grebe Mylonite Zone is a dextral, inclined, transpressional shear zone containing quartz and plagioclase-dominated rocks which have undergone dynamic recrystallization. We find changes between quartz and plagioclase with regards to transpressional end members of deformation geometry, wrench- and pure shear-dominated transpressional geometry, and regime of DRX correlated with grain size.

Wrench-dominated transpressional deformation geometry is recognized in GBM DRX quartz grains and pure shear-dominated transpressional deformation geometry is recognized in SGR DRX quartz grains. The data confirms that strain localized differently between plagioclase and quartz, and within quartz at different grain sizes and dynamic recrystallization regimes. Differences in deformation geometry between quartz and plagioclase and within quartz may be attributed to simultaneous strain partitioning between wrench-dominated and pure shear-dominated transpressional deformation or through progressive deformation as the shear zone evolved.

5.1 Deformation and Recrystallization

17NZ95C1

QUARTZ

The interconnected band of quartz displays microstructures consistent with DRX and CPO as a result from dislocation creep. The quartz grains have slightly lobate microstructures, though many display fairly straight grain boundaries (Fig. 4). Similar orientations of adjacent grains within the AE map suggest grain boundary mobility. The

GOS map displays lattice distortion within the quartz, while the Local Misorientation

29 map displays low-angle boundaries which support crystal-plastic deformation (Fig. 5).

The quartz (0001) pole figure displays CPO consistent with prism slip, rhomb and basal slip (Stipp et al. 2002(b); Toy et al. 2008; Morales et al. 2014). The CPO data paired with the slight lobate grain boundaries suggest that the interconnected quartz band underwent GBM DRX (Hirth and Tullis, 1992; Stipp et al. 2002(b)) and then partial static recrystallization.

PLAGIOCLASE

Limited preserved microstructures in conjunction with CPO is consistent with plagioclase having undergone DRX. Plagioclase is elongated and interconnected throughout the rock with sporadic porphyroclasts observed. It is the dominant, interconnected mineral composing the fabric. Plagioclase grains bulging into other plagioclase grains are observed. Similar orientations of adjacent grains are seen in the AE map and most grains exhibit low lattice distortion in the GOS map which suggest crystal- plastic deformation, but with low accommodated strain in the grains. There is limited low-angle boundary development within the plagioclase observed in the Local

Misorientation map. Plagioclase pole figures display CPO consistent with the known slip system (001)[100] (Kruse et al. 2001), and is consistent with deformation by dislocation creep. The few preserved plagioclase grains bulging into other plagioclase grains suggest

BLG DRX (Ji and Mainprice, 1990; Passchier and Trouw, 2005; Halfpenny et al. 2006;

Platt and Behr, 2011), but the prevalence and grain scale of the straightened grain boundaries are indicative of partial static recrystallization.

30

17NZ95C1 CONDITIONS

Partial static recrystallization has resulted in the overprinting of some DRX microstructures in quartz and plagioclase. The straightened grain boundaries, grain size, and low lattice distortion within plagioclase is suggestive of partial static recrystallization

(Bestmann et al 2005; Otani and Wallis, 2006). The interconnected quartz band’s lobate boundaries and higher lattice distortion within the grains coupled with the fairly straightened boundaries several of the grains suggest partial static recrystallization but not to the degree that plagioclase underwent. Plagioclase appears to have a higher degree of static recrystallization than quartz because of the larger disparity of DRX microstructures and less lattice distortion than quartz. This may suggest that the quartz band was able to deform for an extended period of time after plagioclase crystal-plastic deformation ceased.

As both plagioclase and quartz were crystal-plastically deformed, plagioclase displays characteristics of having been the dominant rheology-controlling mineral and may be used to help constrain deformation temperature. The deformed interconnected quartz band constitutes a small percentage of the thin section, thus the interconnected fabric of the plagioclase allows it to be the rheology-controlling mineral. We interpret plagioclase as having undergone BLG DRX and quartz as GBM DRX. A temperature range of 500-700° C (Stipp et al. 2002(b); Passchier and Trouw, 2005) is associated with these DRX regimes observed in quartz and plagioclase.

31

17NZ91

QUARTZ

The quartz grains within 17NZ91 display DRX microstructures and CPO consistent with dislocation creep. Many of the larger quartz grains have lobate grain boundaries, and the smaller grains (<100 µm) have grain boundaries that are straight to round. When all of the grains are plotted in a pole figure, the CPO is consistent with activation of the prism and rhomb slip systems (Stipp et al. 2002(b); Toy et al.

2008; Morales et al. 2014). The microstructures and CPO of all the quartz grains observed together are consistent with dislocation creep as the dominant deformation mechanism.

The data suggest that increasing quartz grain size correlates with a change in microstructures consistent with SGR to GBM DRX and a change in CPO patterns displaying basal, rhomb and prism slip to dominantly prism slip. We find that a change in grain size is correlated with a change in microstructures and CPO and thus, different regimes of DRX. The few large grains with high amounts of lattice distortion such as the red grain (Fig. 9B) have lobate grain boundaries consistent with GBM DRX, distinctly different than the smaller grain sizes described below. These larger grains exhibit CPO consistent with prism slip. We compared the microstructures and CPO of the interstitial quartz grains that are 50 to 100 µm to the entire quartz population within the band of quartz. The majority of these grains have straightened or rounded boundaries and are mostly the interstitial grains with low lattice distortion between larger grains with higher lattice distortion (Fig. 9B). The CPO forms a broken girdle consistent with activation of the basal, rhomb, and prism slip systems. These grains have

32 microstructures consistent with SGR DRX. The quartz grain size population composed of

8 to 50 µm grains with low amounts of lattice distortion display CPO which form a weak cross girdle consistent with the activation of the basal, rhomb, and prism slip systems (Fig. 10D). The smaller grains have round to straight grain boundaries and are located along the rim of larger grains or within them (Fig. 9A). These grains have microstructures that are consistent with SGR DRX.

Correlating grain size with different regimes of DRX, we find that quartz grains consistent with SGR DRX, 100 µm and below, are distinct from the larger grain sizes with respect to subgrain size, lattice distortion, and Dauphiné twins. The size range of this grain population is also similar to the size of the subgrains developed within the larger, lobate grains (Fig. 9C), suggestive of subgrain rotation. The 100 µm and below grain sizes also have little to no Dauphiné twinning, whereas many of the larger grains do. The microstructures, amount of lattice distortion, and difference in the amount of twinning in comparison to the larger grains suggest a distinction. The grain population of

100 µm and below is a product of SGR DRX, while the larger grains underwent GBM

DRX.

17NZ91 CONDITIONS AND TIMING

The GBM DRX grains may signify the highest temperature deformation conditions recorded in quartz, whereas the SGR DRX grains may signify a lower temperature deformation. Workers have studied the effect of water content and strain rate

(Stipp et al. 2002(a); Toy et al. 2008; Law, 2014) on quartz CPO development. If CPO is affected by these factors in this sample, the data may only be able to offer a range of temperature estimates. We did not complete piezometry on quartz, as grains having

33 undergone GBM DRX have been reported to only display minimum stress values (Stipp et al. 2010). We did not measure water content in the mylonite and have interpreted temperatures from CPO and microstructures. We estimate temperatures of up to 700°C for the GBM grains, and temperatures of ~550° C for the SGR grains (Stipp et al.

2002(b); Passchier and Trouw, 2005).

Different regimes of DRX have been reported to occur through progressive, down temperature deformation resulting in overprinting microstructures (Stipp et al. 2002(a)).

We interpret the low amount of lattice distortion and lack of subgrains and Dauphiné twins within the smaller grains as not having as much accommodated strain as the larger, lobate grains. An increase in strain rate can also lead to weakening of the grains (Platt and Behr 2011), which may have promoted grain-size reduction of the lobate grains. If

SGR DRX occurred at the same time with GBM DRX or after GBM DRX is difficult to definitively determine without more data. We interpret that SGR DRX occurred simultaneously with GBM DRX or overprinted the GBM DRX microstructures, which resulted in the smaller recrystallized grains with low amounts of lattice distortion.

5.2 CVA Analysis

17NZ95C1

The CVA analysis conducted on 17NZ95C1 yields data suggestive of a monoclinic shear zone with quartz and plagioclase displaying wrench-dominated transpression deformation geometry and biotite displaying data consistent with pure shear-dominated transpression deformation geometry. Monoclinic transpressional shear zones may show pure shear-dominated or wrench-dominated end members of

34 transpression, with the vorticity axis either parallel to lineation or with the vorticity axis perpendicular to the lineation, respectively (Michels et al. 2015). Triclinic shear zones are described as the vorticity axis oriented oblique to lineation when plotted in the kinematic reference frame (Michels et al. 2015). The plagioclase CVA plot dominantly corresponds to wrench-dominated transpressional geometry and does have some maxima oblique to the lineation consistent with a trinclinic shear zone (Fig. 7B and Fig. 7D) (Michels et al.

2015). The quartz grains within the interconnected band display maxima that may be suggestive of wrench-dominated transpression and a triclinic shear zone geometry (Fig.

7A and Fig. 7D) (Michels et al. 2015). The biotite CVA figure is consistent with pure shear transpression (Fig. 7C and Fig. 7D). .

The CVA analysis conducted on 17NZ95C1 yields data suggesting quartz, plagioclase, and biotite can have differing deformation geometries between them, which correspond to various periods of shear zone activity and/or deformation partitioning. The presence of wrench-dominated transpression and triclinic shear zone maxima in the quartz and plagioclase CVA plots is reconciled in several ways. First, it may be interpreted as a shift from one deformation geometry to another in quartz and plagioclase.

A second alternative interpretation is that plagioclase exerted a geometrical control on quartz crystal-plastic deformation either while both were deforming or as quartz continued to deform after plagioclase ceased deforming. Michels et al. 2015 suggest three options for individual grains resulting in differences from the bulk of the CVA data stemming from interactions of neighboring grains, multiple slip system activation, or differing amounts of strain between grains with temporal or strain rate variation. We interpret the bulk of the quartz and plagioclase CVA data to be consistent with wrench-

35 dominated transpression. With biotite acting as one of the weakest phases during deformation (Kronenburg, 1991), we assume that it will record the latest style of deformation and deformation geometry. Biotite displays strong maxima consistent with pure shear deformation geometry in the CVA plot that may be indicative of the last style of dominant deformation in the shear zone (Fig. 7C and Fig. 7D), however the lack of microstructural evidence of DRX within the biotite must be considered. The biotite may have undergone partial static recrystallization, but we cannot confirm it underwent DRX.

The rock is partially statically recrystallized and we make assumptions on the microstructures and the ability to perform CVA analysis on this rock. We observe DRX microstructures such as bulges in the plagioclase and lobate grain boundaries in the interconnected quartz band that are suggestive of DRX as well as minor amounts of lattice distortion in the plagioclase and moderate amounts of lattice distortion in the quartz. We make the assumption that because we find evidence of DRX, these grains will still yield pertinent data in regards to CVA analysis i.e., crystallographic distortions used to determine the vorticity axis.

17NZ91

Analyzing the quartz grains within the populations described above, we find the large grains with high amounts of lattice distortion are associated with wrench-dominated transpression and starting at 100 µm and below, the grains with low amounts of lattice distortion display a transition to pure shear-dominated transpression geometry. When analyzing all of the bulk vorticity CVA of the quartz grains, a strong maximum consistent with wrench-dominated transpression is present along with diffuse, weaker maxima consistent with pure shear-dominated transpression (Fig. 10A). The largest grains with

36 higher amounts of lattice distortion exhibit a maximum consistent with wrench- dominated transpression (Fig. 10B). The intermediate-sized grains (50 to 100 µm) with low amounts of lattice distortion display maxima consistent with both wrench-dominated and pure shear-dominated deformation geometries (Fig. 10C). The smallest grains (8 µm to 50 µm) with low amounts of lattice distortion display maxima consistent with wrench- dominated and pure shear-dominated transpressional geometries and some maxima oblique to lineation consistent with triclinic shear zone geometry (Fig. 10D). Similar to the discussion on multiple deformation geometries within the CVA plots in 17NZ95C1, several ways may be considered to reconcile the data. We interpret this CVA plot to have a majority of the data parallel to lineation or near-parallel to lineation which is consistent with pure shear-dominant transpression geometry.

We find the larger quartz grains with higher lattice distortion associated with wrench-dominated transpressional geometry and smaller grains with lower lattice distortion associated with pure shear-dominated transpressional geometry also correspond to the GBM and SGR DRX interpreted above. The larger grains with lobate grain boundaries and higher amounts of lattice distortion are associated with GBM DRX and have CVA plots consistent with wrench-dominated transpressional geometry. The interstitial quartz grains 50 to 100 µm with low amounts of lattice distortion associated with SGR DRX have a CVA plot consistent with pure and wrench-dominated transpressional deformation geometries. This grain size may correlate to a shift in deformation geometry styles as it records a difference in DRX regime and association of deformation geometry compared to the larger, lobate grains having higher amounts of lattice distortion. The quartz grains 8 to 50 µm with low amounts lattice distortion

37 associated with SGR DRX display a CVA plot that we interpret as pure shear-dominant transpressional geometry.

5.3 Relation to the GMZ

GMZ Orientation

The GMZ has been reported to be a sinistral, transpressional shear zone through the use of paleoreconstruction predicated on angular unconformities to the east (Fet al.

2011). This paleoreconstruction involves Tertiary angular unconformities east of Lake

Manapouri dipping 30° to the east (Scott et al. 2011), but we propose a different method to characterize the GMZ. The current orientation of the GMZ at the South Arm of Lake

Manapouri is consistent with a steeply dipping fault (Fig. 11). A rose diagram plotted from the poles to mylonitic foliation displays a maximum shortening direction of ESE to

WNW when compared to the brittle Grebe Fault orientation south of Lake Manapouri near Lake Hauroko (Fig. 11) which would suggest a dextral transpressional shear zone.

With the maximum shortening direction being arc-normal, this is consistent with the

Indecision Creek Shear Zone to the north of the GMZ and shear zones to the south such as the Gutter Shear Zone (Klepeis and Clark, 2004; Marcotte et al. 2005; Allibone and

Tulloch, 2008). The stretching lineations measured in the South Arm of Lake Manapouri are steeply plunging (Fig. 11) and consistent with an inclined transpressional shear zone

(Jones et al. 2004).

38

Having defined the orientation of the GMZ, we can interpret shear sense and its relation to transport direction. Past workers have reported opposing shear sense indicators located in the GMZ (Scott et al. 2011). Quartz and feldspar porphyroclasts in hand sample reveal both sinistral and dextral shear sense, but the dextral shear sense clasts appear dominant. This is consistent with a top to the NE, reverse fault sense. The

17N95C1 and 17NZ91 thin sections display sinistral shear porphyroclasts, consistent with the dominant dextral shear porphyroclasts described in the hand samples because of how the thin sections were cut relative to the hand samples. These sinistral porphyroclasts in the thin sections are supportive of a top-to-the-NE, reverse fault. The CPO data also display opposing shear sense, with 17NZ95C1 displaying top-to-the-SW, dextral sense of shear sense for quartz and plagioclase. 17NZ91 displays top-to-the-SW, dextral shear

39 sense of shear for all of the quartz grains, but top-to-the-NE, sinistral shear sense for the

100 µm and below quartz grains. Overall, this supports a top-to-the-NE, dextral transpressional shear zone with thermochronological data supporting our interpretations on the style of deformation and recrystallization.

GMZ Thermochronology Comparison There are thermochronologic data that we can use to compare our interpretations of microstructures and quartz CPO. Recent thermochronologic data provided from titanite contained within the mylonitic foliation from sample 17NZ96 (Fig. 1B), yielded a temperature of 658° C (Buritica, 2018). This temperature is consistent with temperatures associated with GBM microstructures and the onset of rhomb and prism slip systems (Stipp et al. 2002(b); Passchier and Trouw, 2005; Toy et al. 2008; Morales et al.

2014), similar to what we document in quartz within 17NZ95C1 and 17NZ91. They are also consistent with the plagioclase bulges (Ji and Mainprice, 1990; Stünitz et al. 2003;

Paschier and Trouw, 2005) observed in 17NZ95C1. The SGR DRX microstructures observed in the grains 100 µm and below with low amounts of lattice distortion that appear along boundaries of larger grains or as interstitials between them within 17NZ91 could be associated with lower temperatures or strain rates (Stipp et al. 2002(a); Stipp et al. 2002(b); Stipp et al. 2010) compared to the quartz grains displaying lobate grain boundaries in 17NZ95C1 and 17NZ91. This marks a potential change in deformation conditions from the GBM to the SGR DRX grains. A change in temperature, strain rate, or crystallographic orientation favorable for slip may have occurred.

40

GMZ Evolution

A change in subduction geometry has the potential to be linked to changes in deformation conditions and deformation geometry between and within minerals. The subduction zone off the coast of Gondwana has been described as a contractional to transpressional system which may have undergone a change in subduction geometry around 128 Ma (Schwartz et al. 2017). The Grebe Mylonite Zone has been reported as a transpressional shear zone (Scott et al. 2011) which is consistent with our interpretations.

Our CVA plots display a change in the deformation geometry between phases and at the grain scale. In 17NZ95C1, the plagioclase and quartz CVA plots are consistent with a wrench-dominated transpressional deformation geometry, though variation in the spread of where vorticity axes lie exists. In 17NZ91, the CVA plot for the lobate quartz grains is consistent with wrench-dominated transpression, however as grain size decreases, the wrench-dominated transpression geometry interpretation weakens and the pure shear- dominated transpression interpretation strengthens. The titanite age dates from sample

17NZ96 cannot support the subduction geometry change scenario as there is a ~10 million year discrepancy between the subduction geometry change and the GMZ titanite age. As transpressional deformation geometry can be complex, we suggest other scenarios for the potential change in deformation conditions and deformation geometry differences between phases and grain sizes.

Shifts in transpressional deformation geometry can be a function of several different factors which can manifest as changes in finite strain markers. Changes in transpression geometry have been reported to be a function of inclination change (within an inclined shear zone), obliquity of shear zone strike, and dip of the shear zone, which

41 affect the orientation of the finite strain ellipsoid and the lineation orientations (Jones et al. 2004). We use the orientation of the crystallographic vorticity axis (infinitesimal strain marker) with respect to the mineral stretching lineation (finite strain marker) to interpret the transpressional deformation geometry (Michels et al. 2015; Giorgis et al. 2016).

Keeping with the complexity of an inclined shear zone in mind, we present two scenarios consistent with how inclined transpressional shear zones function and our interpretations:

1) Wrench-dominated and pure shear-dominated deformation geometry occurred simultaneously while being partitioned into plagioclase and quartz grains larger than 100

µm and quartz grains 100 µm and smaller, respectively or 2) deformation geometry shifted from wrench-dominated to pure shear-dominated transpression during the period of shear zone activity.

The first scenario involves phases preferentially partitioning wrench-dominated transpression to the plagioclase and quartz grains (larger than 100 µm) and pure shear- dominated transpression to the quartz grains (100 µm and smaller) and biotite simultaneously. This scenario would entail GBM and SGR DRX having been active at the same time within the quartz in 17NZ91. A majority of the smaller grains associated with SGR DRX appear in the junctions of lobate grains of different orientations, not surrounded and encompassed by grains having underwent GBM DRX. This would suggest that as the grains underwent GBM, some grains were not oriented favorably for high-temperature slip and underwent subgrain development leading to SGR DRX. SGR

DRX microstructures are not observed in the interconnected quartz band of 17NZ95C1.

The rheological strength and interconnectedness of plagioclase within 17NZ95C1 in comparison to 17NZ91 may account for this. If the plagioclase experienced limited

42 amounts of crystal-plastic deformation in comparison to quartz, it could control the orientations and geometry that quartz could deform under and in.

Strain partitioning of deformation geometry has been reported with different phases and grain sizes (Jones et al. 2004). This implies heterogeneous strain partitioning within a shear zone. Pure shear-dominated transpression and wrench-dominated transpression deformational components are not partitioned equally in shear zones (Jones and Tanner, 1995; Tikoff and Fossen, 1995). The unequal partitioning results in preferential accommodation of pure shear-dominated transpressional components over the wrench-dominated transpressional components in shear zones (Tikoff and Fossen,

1995). This may be manifested in 17NZ95C1 and 17NZ91 as the wrench-dominated transpressional component partitioned into the rheologically stronger plagioclase and the larger quartz grains with higher amounts of lattice distortion and the pure shear- dominated transpressional component partitioned into the biotite and smaller quartz grains with lower amounts of lattice distortion. This interpretation suggests grains recrystallized through SGR DRX partition pure shear-dominated transpressional components more efficiently than wrench-dominated transpressional components. The timing of when the grains were dynamically recrystallized and partitioned the wrench- dominated and pure shear-dominated components needs to be established to test this scenario further.

The use of quartz thermometry, such as TitaniQ (Kohn and Northrup, 2009), may prove beneficial to exploring this scenario further. Researchers have studied differing microstructures and have linked different DRX regimes to different Ti-in-quartz concentrations, which could signify different events in the period of shear zone activity

43

(Ashley et al. 2013; Nachlas et al. 2014; Haertel et al. 2015). Obtaining and comparing temperatures from the larger and smaller grains could provide insight as to whether these

DRX regimes acted simultaneously and recorded differing deformation geometries. This argument would be predicated on the idea that the SGR and GBM DRX regimes operated under different temperatures and that the grain populations displaying these regimes are associated with different temperatures and different periods of shear zone activity.

The second scenario involves a transition from a wrench-dominated to pure shear- dominated transpressional geometry throughout the period of shear zone activity. The difference in shear zone deformation geometry may be a result of strain partitioning between and within mineral phases over time. In the rocks used in this study, it entails the plagioclase having underwent BLG DRX and quartz underwent GBM DRX. If shear zone conditions changed such as a shift in temperature or strain rate, plagioclase was not able to easily undergo crystal-plastic deformation and higher amounts of strain were localized into the quartz. The quartz then began to undergo SGR DRX and provided a down temperature and/or higher strain overprint on the GBM DRX quartz microstructures. The quartz populations in 17NZ91 display GBM DRX with the larger grains (above 100 µm) with higher amounts of lattice distortion and SGR DRX with the smaller grains (100 µm and below) with lower amounts of lattice distortion. The smaller grains that underwent SGR DRX are of similar size to the subgrains within the larger, lobate grains. This would suggest that the smaller grains are produced through SGR within the larger lobate grains (Halfpenny et al. 2006) and occur after GBM DRX resulting in an overprinting relationship. The SGR quartz grains have lower amounts of lattice distortion in comparison to the GBM quartz grains suggesting that the GBM grains

44 were work-hardened and the SGR grains were recrystallized after the GBM grains. GBM is the DRX regime in the interconnected band of quartz recorded in 17NZ95C1 which was partially statically recrystallized. As plagioclase was the dominant rheology- controlling mineral in 17NZ95C1, this may have not allowed SGR DRX to occur as quartz had limited interconnectedness. This scenario would have produced an overprinting, progressive, down-temperature and/or higher strain rate microstructures.

5.4 Broad Microstructural Analysis Implications

This study has important implications on the relation between strain localization and deformation geometry. Phase topology is observed to be a critical factor in determining the rheology-controlling mineral rather than just rheological strength and the deformation geometry between dynamically recrystallized phases. CVA analysis allows for interpretation of relative deformation timing and deformation geometry evolution in partially statically recrystallized rocks. The comparison of CVA between quartz and plagioclase and within different grain sizes and dynamic recrystallization regimes of quartz suggest that phases can have differing deformation geometries between them and within them, varying with grain size and DRX regime. It holds importance as dynamic recrystallized quartz grains of specific regimes such as SGR DRX may preferentially accommodate pure shear-dominated transpression over wrench-dominated transpression.

This provides a potential link between dynamic recrystallization and strain partitioning should be studied further.

45

6. CONCLUSIONS

We find that studying strain localization of quartz and plagioclase within this mid- crustal shear zone brings insight regarding their deformation geometries, styles of recrystallization, and grain sizes. Rather than just rheological strength, the interconnectedness of a phase is a critical factor in determining the dominant rheology- controlling mineral. Differences in deformation geometry is observed to have occurred between phases within a partially statically recrystallized rock. Grain size is observed correlated to styles of DRX and different deformation geometries within quartz.

The Grebe Mylonite Zone is a mid-crustal, dextral, transpressional shear zone that records crystal-plastic deformation, partial static recrystallization, and dynamic recrystallization. Our interpretations are consistent with two scenarios. The first involves simultaneous strain partitioning between wrench-dominated deformation and pure shear deformation between phases and grain sizes. The second scenario involves the production of recrystallized, low lattice distortion grains through SGR DRX after GBM DRX in quartz grains which allowed the shear zone to deform as the conditions changed. The shift in deformation conditions invoked a change in deformation geometry and what grains were optimally suited to accommodate strain.

Our findings using EBSD to study the GMZ have resulted in the following statements:

 The Grebe Mylonite Zone is an inclined, dextral, transpressional shear

zone.

46

 CVA analysis may provide transpression history on partially statically

recrystallized rocks.

 Different phases may record differing deformation geometries dependent

on when they were undergoing crystal-plastic deformation.

 Interconnected fabric plays a critical role in determining the rheology-

controlling mineral and may dictate how or if weaker, more easily

deformable minerals plastically deform and recrystallize.

 Single, interconnected phases can have differing populations of grain sizes

and lattice distortion. These different populations can record different

regimes of DRX and deformation geometry.

 The smaller, low lattice distortion recrystallized grains may be critical for

the development of long-lived shear zones as progressive, lower

temperature deformation continues.

 Grains chosen for quartz thermometry (TitaniQ) must be carefully chosen

as they may record differing events and deformation geometries of the

shear zone, even within the same interconnected band.

47

REFERENCES

Allibone, A.H., Jongens, R., Scott, J.M., Tulloch, A.J., Turnbull, I.M., Cooper, A.F., Powell, N.G., Ladley, E.B., King, R.P., and Rattenbury, M.S., 2009, Plutonic rocks of the Median Batholith in eastern and central Fiordland, New Zealand: field relations, geochemistry, correlation, and nomenclature: New Zealand Journal of Geology and Geophysics, v. 52, p. 101-148, doi: 10.1080/00288300909509882 Allibone, A.H. and Tulloch, A.J., 2008, Early Cretaceous dextral transpressional deformation within the Median Batholith, Stewart Island, New Zealand: New Zealand Journal of Geology and Geophysics, v. 51, p. 115-134, doi: 10.1080/00288300809509854 Ashley, K.T., Webb, L.E., Spear, F.S., and Jay, B. Thomas, 2013, P-T-D histories from quartz: A case study of the application of the TitaniQ Thermobarometer to progressive fabric development in metapelites: Geochemistry, Geophysics, Geosystems, v. 12, n. 9, p. 3821-3843, doi: 10.1002/ggge.20237 Behr, W.M. and Platt, J.P., 2011, A naturally constrained stress profile through the middle crust in an extensional terrane: Earth and Planetary Science Letters, v. 303, i. 3-4, p. 181-192, doi: 10.1016/j.epsl.2010.11.044 Bestmann, M., Piazolo, S., Spiers, C.J., and Prior, D.J., 2005, Microstrucutral evolution during initial stages of static recovery and recrystallization: new insights from in-situ heating experiments combined with electron backscatter diffraction analysis: Journal of Structural Geology, v. 27, p. 447-457, doi: 10.1016/j.jsg.2004.10.006 Buritica, L., 2018, Temporal and Spatial Variations in Magmatism and Transpressional Deformation in the Middle to Lower-Crust of a Cretaceous Arc, Median Batholith, Fiordland, New Zealand, [M.S. Thesis]: Northridge, California State University. Cross, A.J., Hirth, G., and Prior, D.J., 2017, Effects of secondary phases on crystallographic preferred orientations in mylonites: Geology, v. 45, n. 10, p. 955-958, doi: 10.13/G38936.1 Gerbi, C., Culshaw, N., and Marsh, J., 2010, Magnitude of weakening during crystal-scale shear zone development: Journal of Structural Geology, V. 1, I, 1, p. 107-117, doi: 10.1016/j.jsg.2209.10.002 Giorgis, S., Michels, Z., Dair, L., Braudy, N., and Tikoff, B., 2016, Kinematic and Vorticity analyses of the western Idaho shear zone, USA, Lithosphere, doi: 10.1130/L518.1 Gueydan, F. Précigout, Montési, L.G.J., 2014, Strain weakening enables continental plate tectonics: Tectonophysics, v. 631, p. 189-196, doi: 10.1016/j.tecto.2014.02.005

48

Haertel, M. Herwegh, M., and Pettke, T., 2013, Titanium-in-quartz thermometry in synkinematic quartz veins in a synkinematic quartz veins in a retrograde crustal-scale normal fault: Tectonophysics, v. 608, p. 468-481 Halfpenny, A., Prior, D.J., and Wheeler, J., 2006, Analysis of dynamic recrystallization and nucleation in a quartzite mylonite: Tectonophysics, v. 427, p. 3-14, doi: 10.1016/j.tecto.2006.05.016 Halfpenny, A., Prior, D.J., and Wheeler, J., 2012, Electron backscatter diffraction analysis to determine the mechanisms that operated during dynamic recrystallization of quartz-rich rocks: Journal of Structural Geology, v. 36, p. 2-15, doi: 10.1016/j.jsg.2012.01.001 Hansen, L.N., Cheadle, M.J., John, B.E., Swapp, S.M., Dick, H.J.B., Tucholke, B.E., and Tivey, M.A., 2013, Mylonitic deformation at the Kane oceanic core complex: Implications for the rheological behavior of oceanic detachment faults: Geochem, Geophysics, and Geosystems, v. 14, p. 3085- 3108, doi: 10.1002/ggge.20184 Heilbronner, R. and Tullis, J., 2006, Evolution of C axis pole figures and grain size during dynamic recrystallization: Results from experimentally sheared quartzite: Journal of Geophysical Research, v. 111, B10202, doi: 10.1029/2005JB004194 Herwegh, M., Linckens, J., Ebert, A., Berger, A., and Brodhag, S.H., 2011, The role of second phases for controlling microstructural evolution in polymineralic rocks: A review: Journal of Structural Geology, v. 33, p. 1728- 1750, doi: 10.1016/j.jsg.2011.08.011 Hielscher, R. and Schaeben, H., 2008, A novel pole figure inversion method: specification of the MTEX algorithm: Journal of Applied Crystallography, v. 41, p. 1024-1037, doi: 10.1107/S0021889808030112 Hirth, G. and Tullis, J., 1992, Dislocation creep regimes in quartz aggregates: Journal of Structural Geology, v. 14, n. 2, p. 145-159 Iacopini, D., Carosi, R., Montomoli, C., and Passchier, C.W., 2008, Strain analysis and vorticity of flow in the northern Sardinian Variscan Belt: Recognition of a partitioned oblique deformation event: Tectonophysics, v. 446, p. 77-96, doi: 10.1016/j.tecto.2007.10.002 Ji, S. and Mainprice, D., 1990, Recrystallization and Fabric Development in Plagioclase: The Journal of Geology, v. 98, n. 1, p.65-79 Jones, R.R., Holdsworth, R.E., Clegg, P., McCaffrey, K., and Tavarnelli, E., 2004, Inclined Transpression: Journal of Structural Geology, v. 26, p. 1531- 1548, doi: 10.1016/j.jsg.2004.01.004 Jones, R.R., and Tanner, P.W.G., 1995, Strain partitioning in transpression zones: Journal of Structural Geology, v. 17, n. 6, p. 793-802

49

Klepeis, K.A. and Clark, G.L., 2004, The evolution of an exposed mid-lower crustal attachment zone in Fiordland, New Zealand: In: Grocott, J., McCaffrey, K.J.W., Taylor, G., and Tikoff, B., Vertical Coupling and Decoupling in the Lithosphere: Geological Society, London, Special Publications, v. 227, p. 197-229 Klepeis, K.A., Schwartz, J., Stowell, H., and Tulloch, A., 2016, Gneiss domes, vertical and horizontal mass transfer, and the initiation of extension in the hot lower-crustal root of a continental arc, Fiordland, New Zealand: Lithosphere, v. 8, n. 2, p. 116-140, doi: 10.1130/L490.1 Kohlstedt, D.L., Evans, B., and Mackwell, S.J., 1995, Strength of the Lithosphere: Constraints imposed by laboratory experiments: Journal of Geophysical Research, v. 100, n. B9, p. 17587-17607 Kohn, M.J. and Northrup, C.J., 2009, Taking mylonites’ temperatures: Geology, v. 37, n. 1, p. 47-50, doi: 10.1130/G25081A.1 Kronenberg, A.K., Kirby, S.T., and Pinkston, J., 1990, Basal Slip and Mechanical Anisotropy of Biotite: Journal of Geophysical Research, v. 95, n. B12, p. 19,257-19,278 Kruhl, J.H., 1996, Prism- and basal-plane parallel subgrain boundaries in quartz: a microstructural geothermobarometer: Journal of Metamorphic Geology, v. 14, p. 581-589 Kruse, R., Stünitz, and Kunze, K., 2001, Dynamic recrystallization processes in plagioclase porphyroclasts: Journal of Strucutral Geology, v. 23, p. 1781- 1802 Law, J.D., 2014, Deformation thermometry based on quartz c-axis fabrics and recrystallization microstructures: A review: Journal of Structural Geology, v. 66, p. 129-161, doi: 10.1016/j.jsg.2014.05.023 Little, T.A., Prior, D.J., Toy, V.G. and Lindroos, Z.R., 2015, The link between strength of lattice preferred orientation, second phase content and grain boundary migration: A case study from the Alpine Fault zone, New Zealand: Journal of Structural Geology, v. 81, p. 59-77, doi: 10.1016/j.jsg.2015.09.004 Mancktelow, N.S. and Pennacchioni, G., 2004, The influence of grain boundary fluids on the microstructure of quartz-feldspar mylonites: Journal of Structural Geology, v. 26, i. 1, p. 47-69. Doi: 10.1016/S0191-8141(03)00081- 6 Marcotte, S.B., Klepeis, K.A., Clarke, G.L., Gehrels, G., and Hollis, J.A, 2005, Intra-arc transpression in the lower crust and its relationship to magmatism in a Mesozoic magmatic arc: Tectonophysics, v. 407, p. 135-163, doi: 10.1016/j.tecto.2005.07.007

50

Menegon, L., Pennacchioni, Heilbronner, R., and Pittarello, L., 2008, Evolution of quartz microstructure and c-axis crystallographic preferred orientation within ductiley deformed granitoids (Arolla unit, Western Alps): Journal of Structural Geology, v. 30, p. 1332-1347, doi:10.1016/j.jsg.2008.07.007 Michels, Z.D., Kruckenberg, S.C., Davis, J.R., and Tikoff, B., 2015, Determining vorticity axes from grain-scale dispersion of crystallographic orientations: Geology, doi: 10.1130/G36868.1 Miranda, E.A., Hirth, G., and John, B.E., 2016, Microstructural evidence for the transition from dislocation creep to dislocation-accommodated grain boundary sliding in naturally deformed plagioclase: Journal of Structural Geology, v. 92, p. 30-45, doi: 10.1016/j.jsg.2016.09.002 Montési, L.G.J., 2007, A constitutive model for layer development in shear zones near the brittle-ductile transition: Geophysical Research Letters, v. 34, doi: 10.1029/2007GL029250 Morales, L.F.G, Lloyd, G.E, and Mainprice, D., 2014, Fabric transitions in quartz via viscoplastic self-consistent modeling part I: Axial compression and simple shear under constant strain: Tectonophysics, v. 636, p. 52-69, doi: 10.1016/j.tecto.2014.08.011 Mortimer, N., 2004, New Zealand’s Geological Foundations: Gondwana Research, v. 7, n. 1, p. 261-272 Mulyukova, E. and Bercovici, D., 2017, Formation of lithospheric shear zones: Effect of temperature on two-phase grain damge: Physics of the Earth and Planetary Interiors, v. 270, p. 195-212, doi: 10.1016/j.pepi.2017.07.011 Nachlas, W.O., Whitney, G.L., Teyssier, C., Bagley, B., and Mulch, A., 2014, Titanium concentration in quartz as a record of multiple deformation mechanisms in an extensional shear zone: Geochemistry, Geophysics, Geosystems, v. 15, p. 1374-1397, doi: 10.1002/2013GC005200 Otani, M. and Wallis, S., 2006, Quartz lattice preferred orientation patterns and static recrystallization: Natural examples from the Ryoke belt, Japan: Geology, v. 34, n. 7, p. 561-564, doi: 10.1130/G22430.1 Passchier and Trouw, 2005, Microtectonics: Springer, 2nd edition Platt, J.P. and Behr, W.M., 2011, Grainsize evolution in ductile shear zones: Implications for strain localization and the strength of the lithosphere: Journal of Structural Geology, v. 33, p. 537-550, doi: 10.1016/j.jsg.2011.01.018 Ramezani, J. and Tulloch, A.J., 2009, TIMS U-Pb geochronology of dioritic- granitic rocks Fiordland: http://data.gns.cri.nz/paperdata/list.jsp, doi: 10.21420/G2CC7Z Schwartz, J.J., Klepeis, K.A., Sadorski, J.F., Stowell, H.H., Tulloch, A.J., and Coble, M.A., 2017, The tempo of continental arc construction in the Mesozoic

51

Median Batholith, Fiordland, New Zealand: Lithosphere, v. 9, no. 3, p. 343- 365, doi: 10.1130/L610.1 Schwartz, J.J., Stowell, H.H., Klepeis, K.A., Tulloch, A.J., Kylander-Clark, A.R.C., Hacker, B.R., and Coble, M.A., 2016, Thermochronology of extensional orogenic collapse in the deep crust of Zealandia: Geosphere, v. 12, n. 3, doi: 10.1130/GES01232.1 Scott, J.M., 2013, A review of the location and significance of the boundary between the Western Province and Eastern Province, New Zealand: New Zealand Journal of geology and Geophysics, v. 56, no. 4, p. 476-293, doi: 10.1080/00288306.2013.812971 Scott, J.M., Cooper, A.F., Tulloch, A.J., and Spell, T.L., 2011, Crustal Thickening of the Early Cretaceous paleo-Pacific Gondwana margin: Gondwana Research, v. 20, p. 380-394, doi: 10.1016/j.gr.2010.10.008 Scott, J.M. and Palin, J.M., 2008, LA-ICP-MS U-Pb zircon ages from Mesozoic plutonic rocks in eastern Fiordland, New Zealand: New Zealand Journal of Geology and Geophysics, v. 51, p. 105-113, doi: 10.1080/00288300809509853 Stipp, M., Stünitz, H., Heilbronner, R., and Schmid, S.M, 2002, Dynamic recrystallization of quartz: Correlation between natural and experimental conditions: In: DeMeer, S., Drury, M.R., De Bresser, J.H.P., Pennock, G.M., Deformation Mechanisms, Rheology and Tectonics: Current Status and Future Perspectives, Geological Society London Special Publications, v. 200, p. 171- 190, doi: 10.1144/GSL.SP.2001.200.01.11 Stipp, M., Stünitz, H., Heilbronner, R., and Schmid, S.M., 2002, The eastern Tonale fault zone: a ‘natural laboratory’ for crystal plastic deformation of quartz over a temperature range from 250 to 700 °C: Journal of Structural Geology, v. 24, i. 12, p. 1861-1884, doi: 10.1016/S0191-8141(02)00035-4 Stipp, M., Tullis, J., Scherwath, M., and Behrmann, J.H., 2010, A new perspective on paleopiezometry: Dynamically recrystallized grain size distributions indicate mechanism changes: Geology, v. 38, n. 8, p. 759-762, doi: 10.1130/G31162.1 Stewart, C.A. and Miranda, E.A, 2017, The Rheological Evolution of Brittle- Ductile Transition Rocks During the Earthquake Cycle: Evidence for a Ductile Precursor to Pseudotachylyte in an Extensional Fault System, South Mountains, Arizona: Journal o Geophysical Research Solid Earth, v. 122, i. 12, p. 10,643-10,665, doi: 10.1002/2017JB014680 Stünitz, H., Gerald, J.D.F, and Tullis, J., 2003, Dislocation generation, slip systems, and dynamic recrystallization in experimentally deformed plagioclase single crystals: Tectonophysics, v. 372, p. 215-233, doi: 10.1016/S0040-1951(03)00241-5

52

Tikoff, B. and Fossen, H., 1995, The limitations of three-dimensional kinematic vorticity analysis: Journal of Structural Geology, v. 17, n. 12, p. 1771-1784 Toy, V.G., Prior, D.J., and Norris, R.J., 2008, Quartz fabrics in the Alpine Fault mylonites: Influence of pre-existing preferred orientations on fabric development during progressive uplift, v. 30, p. 602-621, doi:10.1016/j.jsg.2008.01.001 Trimby, P.W., Prior, D.J., and Wheeler, J., 1998, Grain boundary hierarchy development in a quartz mylonite: Journal of Structural Geology, v. 20, n. 7, p. 917-935 Turnbull, I.M., Allibone, A.H., and Jongens, R., 2010, Geology of the Fiordland Area: Lower Hutt, Institute of Geological and Nuclear Sciences

53