+ All Categories
Home > Documents > Evidence for high stress in quartz from the impact site of Vredefort ...

Evidence for high stress in quartz from the impact site of Vredefort ...

Date post: 07-Jan-2017
Category:
Upload: ngonhan
View: 266 times
Download: 0 times
Share this document with a friend
10
Evidence for high stress in quartz from the impact site of Vredefort, South Africa KAI CHEN 1 ,MARTIN KUNZ 2 ,NOBUMICHI TAMURA 2 and HANS-RUDOLF WENK 1, * 1 Department of Earth and Planetary Science, University of California, Berkeley, CA 94720, USA *Corresponding author, e-mail: [email protected] 2 Advanced Light Source, Lawrence Berkeley Laboratory, 1 Cyclotron Road, Berkeley, CA 94720, USA Abstract: A microstructural investigation of stishovite-bearing quartzite from the Vredefort meteorite impact site in South Africa reveals features that are attributed to shock deformation. These include abundant mechanically induced Dauphine ´ twinning, rhombohedral deformation lamellae and associated residual stresses. Mechanical twins were identified with SEM-EBSD. Residual stresses are derived from equivalent strains that are observed on Laue diffraction images measured with microfocused synchrotron X- rays. Average lattice strains of 1400 microstrains in Vredefort quartz are much higher compared to values (500) found in quartz from granite samples subject to only regional metamorphism. While the granitic quartz shows mainly normal compressive and tensile residual stresses, in Vredefort quartz, shear stresses associated with lamellar boundaries dominate. Key-words: Vredefort impact site, quartz, Dauphine ´ twinning, residual stress. 1. Introduction Shock-induced phenomena in quartz have long been the sub- ject of detailed investigations. They include documentation of phase transformations of quartz to high-pressure polymorphs coesite and stishovite (Chao et al., 1960, 1962), deformation lamellae (Engelhardt & Bertsch, 1969; Goltrant et al., 1991, 1992) and observations of mechanical twins, both according to the Brazil law (McLaren et al., 1967) and the Dauphine ´ law (Trepmann & Spray, 2005; Trepmann, 2008). In a previous study of quartzite from the Vredefort crater in South Africa, we observed preferred orientation patterns that could be inter- preted as a result of twinning and could thus serve as a paleopiezometer (Wenk et al., 2005). No actual twins were observed. We have now returned to the same sample to use scanning electron microscopy (SEM) and electron backscat- ter diffraction (EBSD) to explore Dauphine ´ twinning and to use a newly emerging technique of synchrotron microfocus X-ray Laue diffraction to explore residual stress in shocked quartz (Tamura et al., 2002, 2003; Kunz et al., 2009a). Daly (1947) was the first to suggest that the Vredefort crater in South Africa was the result of a meteorite impact, but this hypothesis only became accepted after the discov- ery of pseudotachylites (Reimold, 1995; Gibson et al., 1997a), high-pressure phases like stishovite (Martini, 1978), and shatter-cone features (Dietz, 1961). It emerged as a classical example of a meteorite impact site (Gibson & Reimold, 2001). The impact has been dated at 2.0 Ga (Gibson et al., 1997b). Maximum shock pressures are estimated at 20–50 GPa (Leroux et al., 1994). The sample of quartzite used in this study is from Weltevrede farm, between Parys and Sasolburg, in Kromellenboogspruit, about 30 km NE of the Vredefort impact center. In this rock, Martini (1978, 1991) discovered small amounts of coesite and stishovite, as well as planar deformation fea- tures in quartz that are typical for shock deformation. The average grain size of quartz is 0.3–0.5 mm. 2. Dauphine ´ twinning: EBSD measurements In a previous study we investigated preferred orientation in a bulk sample of Vredefort quartzite with neutron diffraction and observed that, while the orientation of c- and a-axes was random, rhombohedral planes such as f20 21g and f02 21g displayed distinct and different orientation patterns (Wenk et al., 2005). It was surmised that this orientation was produced by mechanical twinning by the shock wave gen- erated during the meteor impact. It has long been established that quartz crystals subjected to high stress undergo mechan- ical Dauphine ´ twinning (Zinserling & Schubnikov, 1933; Wooster et al., 1947). During twinning, only a slight dis- placive rearrangement of atoms occurs, without any break- age of bonds or permanent macroscopic strain. However, the change in crystallographic orientation is profound. It 0935-1221/11/0023-2082 $ 4.50 DOI: 10.1127/0935-1221/2011/0023-2082 # 2011 E. Schweizerbart’sche Verlagsbuchhandlung, D-70176 Stuttgart Eur. J. Mineral. 2011, 23, 169–178 Published online January 2011
Transcript
Page 1: Evidence for high stress in quartz from the impact site of Vredefort ...

Evidence for high stress in quartz from the impact site

of Vredefort, South Africa

KAI CHEN1, MARTIN KUNZ2, NOBUMICHI TAMURA2 and HANS-RUDOLF WENK1,*

1 Department of Earth and Planetary Science, University of California, Berkeley, CA 94720, USA*Corresponding author, e-mail: [email protected]

2 Advanced Light Source, Lawrence Berkeley Laboratory, 1 Cyclotron Road, Berkeley, CA 94720, USA

Abstract: A microstructural investigation of stishovite-bearing quartzite from the Vredefort meteorite impact site in South Africareveals features that are attributed to shock deformation. These include abundant mechanically induced Dauphine twinning,rhombohedral deformation lamellae and associated residual stresses. Mechanical twins were identified with SEM-EBSD. Residualstresses are derived from equivalent strains that are observed on Laue diffraction images measured with microfocused synchrotron X-rays. Average lattice strains of 1400 microstrains in Vredefort quartz are much higher compared to values (500) found in quartz fromgranite samples subject to only regional metamorphism. While the granitic quartz shows mainly normal compressive and tensileresidual stresses, in Vredefort quartz, shear stresses associated with lamellar boundaries dominate.

Key-words: Vredefort impact site, quartz, Dauphine twinning, residual stress.

1. Introduction

Shock-induced phenomena in quartz have long been the sub-ject of detailed investigations. They include documentation ofphase transformations of quartz to high-pressure polymorphscoesite and stishovite (Chao et al., 1960, 1962), deformationlamellae (Engelhardt & Bertsch, 1969; Goltrant et al., 1991,1992) and observations of mechanical twins, both accordingto the Brazil law (McLaren et al., 1967) and the Dauphine law(Trepmann & Spray, 2005; Trepmann, 2008). In a previousstudy of quartzite from the Vredefort crater in South Africa,we observed preferred orientation patterns that could be inter-preted as a result of twinning and could thus serve as apaleopiezometer (Wenk et al., 2005). No actual twins wereobserved. We have now returned to the same sample to usescanning electron microscopy (SEM) and electron backscat-ter diffraction (EBSD) to explore Dauphine twinning and touse a newly emerging technique of synchrotron microfocusX-ray Laue diffraction to explore residual stress in shockedquartz (Tamura et al., 2002, 2003; Kunz et al., 2009a).

Daly (1947) was the first to suggest that the Vredefortcrater in South Africa was the result of a meteorite impact,but this hypothesis only became accepted after the discov-ery of pseudotachylites (Reimold, 1995; Gibson et al.,1997a), high-pressure phases like stishovite (Martini,1978), and shatter-cone features (Dietz, 1961). It emergedas a classical example of a meteorite impact site (Gibson &Reimold, 2001). The impact has been dated at 2.0 Ga

(Gibson et al., 1997b). Maximum shock pressures areestimated at 20–50 GPa (Leroux et al., 1994). The sampleof quartzite used in this study is from Weltevrede farm,between Parys and Sasolburg, in Kromellenboogspruit,about 30 km NE of the Vredefort impact center. In thisrock, Martini (1978, 1991) discovered small amounts ofcoesite and stishovite, as well as planar deformation fea-tures in quartz that are typical for shock deformation. Theaverage grain size of quartz is 0.3–0.5 mm.

2. Dauphine twinning: EBSD measurements

In a previous study we investigated preferred orientation in abulk sample of Vredefort quartzite with neutron diffractionand observed that, while the orientation of c- and a-axes wasrandom, rhombohedral planes such as f2021g and f0221gdisplayed distinct and different orientation patterns (Wenket al., 2005). It was surmised that this orientation wasproduced by mechanical twinning by the shock wave gen-erated during the meteor impact. It has long been establishedthat quartz crystals subjected to high stress undergo mechan-ical Dauphine twinning (Zinserling & Schubnikov, 1933;Wooster et al., 1947). During twinning, only a slight dis-placive rearrangement of atoms occurs, without any break-age of bonds or permanent macroscopic strain. However,the change in crystallographic orientation is profound. It

0935-1221/11/0023-2082 $ 4.50DOI: 10.1127/0935-1221/2011/0023-2082 # 2011 E. Schweizerbart’sche Verlagsbuchhandlung, D-70176 Stuttgart

Eur. J. Mineral.

2011, 23, 169–178

Published online January 2011

Page 2: Evidence for high stress in quartz from the impact site of Vredefort ...

corresponds to a twofold rotation about the c-axis andreverses positive and negative rhombs. In terms of elasticproperties, the pole to the positive rhomb f2021g is close tothe softest direction of quartz, and that of the negative rhombf0221g close to the stiffest direction. In experiments it wasobserved that at differential compressive stresses above 100MPa, crystals with negative rhombs perpendicular to thecompression direction become reoriented by Dauphine twin-ning in a way that the positive unit rhomb f2021g alignsperpendicular to compression (Tullis & Tullis, 1972; Wenket al., 2007a). While the preferred orientation pattern sug-gests twinning, actual twins were never observed.

Here we returned to the same sample studied by neutrondiffraction to investigate the surface structure of the quart-zite with the SEM. It is apparent from the contrast in back-scatter electron images that quartz grains are subdivided intodomains with boundaries of 120� angles (Fig. 1). DetailedEBSD scans over individual grains confirm that the bound-aries are Dauphine twin boundaries, i.e. both domains sharec-axes and a-axes but are related by a 180� rotation about thec-axis. In terms of Bunge Euler angles that define the crystalorientation relative to the sample f1 and � specify theorientation of the c-axis. They are insensitive to Dauphinetwins. On the other hand, f2 is a measure of the crystalrotation about the c-axis, and maps of f2 display Dauphinetwins (Fig. 2a, b, different gray shades). Misorientations of56�–64� between domains, i.e. twin boundaries, are dis-played as red lines. The images clearly show the presenceof twin domains in quartz crystals.

3. Residual stress: synchrotron microbeamanalyses

3.1. Experimental method

The Vredefort quartzite sample was glued on a glass slidewith CrystalbondTM, then cut by a diamond saw and ground,

with SiC and Al2O3 abrasives, to a 30mm thick thin section. Inthe thin section grains were identified which display planardeformation features with optical microscopy (Fig. 3). Thesegrains have c-axes at low angles to the sample surface (‘‘flashfigures’’). The thin-sectioned sample was then removed fromthe glass slide and mounted on an Al holder with a 1 cm holeacross the region of interest of the thin section.

Several grains were analyzed and results were similar. Weonly report here data for a single grain which shows planardeformation features (PDFs) in good contrast (Fig. 3). ThePDFs in the grain were almost perpendicular to the samplesurface. The synchrotron X-ray Laue microdiffractionexperiment was carried out at Beamline 12.3.2 of theAdvanced Light Source (ALS) at Lawrence BerkeleyNational Laboratory (Kunz et al., 2009b). A polychromaticX-ray beam was focused to � 1 � 1 mm2 by a pair of

Fig. 2. Orientation maps of quartz grains based on EBSD measure-ments. Euler angle f2 is used, since it is sensitive to Dauphinetwinning. Red lines are boundaries of orientations related by a180� rotation about the c-axis. (a) Detail of the quartz grain with adistinct boundary. Inserted is a forward scatter SEM image of thesame area. EBSD measurements with the HKL-Oxford system. (b)Larger scan over a different grain divided into twin domains mea-sured with the TSL-EDAX system.

Fig. 3. Optical micrograph, plane polarized light, of a grain in the thinsection which shows planar deformation features. The black squareindicates the raster scanned area by microfocused X-ray beam.

Fig. 1. Backscatter electron image of quartz grain divided intodomains with different contrast. White bar corresponds to 100 mm.

170 K. Chen, M. Kunz, N. Tamura, H.-R. Wenk

Page 3: Evidence for high stress in quartz from the impact site of Vredefort ...

Kirkpatrick-Baez (KB) mirrors. The sample was mounted ona high-resolution x–y scan stage and positioned at the focalpoint of the X-rays by assistance of a laser triangulation unit(Keyence LK-G152). The energy of the X-ray ranged from 4to 22 keV. The diffraction experiment was conducted intransmission mode. In this mode a two-dimensional (2D)X-ray MARCCD detector was mounted with its centre at 2yof 55� as shown in Fig. 4a. The CCD has a 133 mm diameterand 1024 � 1024 pixel resolution. The diffraction geome-tries, including the distance from the CCD (�78 cm) to thesample, the centre position of the CCD, and the CCD tiltangles, were carefully calibrated by indexing a diffractionpattern of a strain-free Si sample. A 20 mm� 20 mm area onthe selected grain was raster scanned using 0.5 mm step size.The scanned area is indicated by the black square in Fig. 3.Since the crystal grain size in our sample was much biggerthan the X-ray beam size, the white X-ray beam generated asingle-crystal Laue diffraction pattern at each position wherethe X-ray interacted with the sample. An example for a Lauediffraction pattern in transmission mode from the Vredefortquartz is shown in Fig. 4b. Miller indices are shown for someof the reflections. Conventionally the CCD detector is posi-tioned at a high 2y angle and samples are analyzed in reflec-tion mode to get better strain resolution. Transmission modewas employed in this study rather than reflection modebecause it is easier to identify the PDFs when they areperpendicular to the sample surface while parallel to theincoming X-ray beam. In reflection mode the incident X-ray beam penetrates more than one lamella at each scanningspot, so that any features created by lamellae are smearedout. This also leads to many difficulties and failures inindexing the diffraction patterns.

3.2. Data analysis

All 1600 diffraction patterns were indexed by a customwritten software package (XMAS, Tamura et al., 2003),which allowed mapping the grain orientation, strain/stresstensor, and plastic deformation of the scanned area.

Diffraction peaks are detected based on a user-definedpeak to background threshold. The positions on the CCD(x-pixel and y-pixel) as well as width of each individualdiffraction peak were determined by fitting intensity (I)using a 2D Lorentzian function (Equation (6.6), p. 168,Noyan & Cohen, 1987). Note that the quantitative inter-pretation of these intensities, affected by numerous arte-facts specific to white-beam experiments, is notstraightforward and no attempt was made in this study toinclude any intensity information in the analysis. In someof the diffraction patterns, splitting of the diffraction peakswas observed, indicating that sub-grains with slightly dif-ferent orientations, separated by low-angle grain bound-aries, were contained in the diffraction volume (Fig. 4b). Insuch cases, the strongest set of diffraction peaks, corre-sponding to the sub-grain with largest volume, was ana-lyzed, as indicated by the square in the insertion of Fig. 4b.Grain orientation was indexed by fitting the position of theLaue diffraction peaks based on the lattice parameters ofunstrained quartz and trigonal symmetry (a¼ 4.921 A, c¼5.4163 A, space group P3121; Glinnemann et al., 1992).We realize that reported lattice parameters of naturalquartz crystals vary sufficiently to influence deduced strainvalues. In the present case, the observed strain values,however, are much larger than the effect from variablereference parameters. In order for a straightforward repre-sentation of crystal orientations, a set of laboratory coordi-nates is defined in the following way: Z-axis is normal tothe sample plane; the sample surface plane is thus the X–Yplane perpendicular to the X-ray beam. X is horizontal andY is vertical. The X–Y–Z coordinates in transmissionmode are shown in Fig. 4a. Grain orientation at each spotis expressed by a 3 � 3 matrixax ay azbx by bzcx cy cz

24

35, where ui (u¼ a, b, c and i ¼ x, y, z) refers

to the component of vector u in i-direction.It is worth emphasizing that the indexing algorithm does

not take diffraction intensities into account and thus treatstrigonal quartz as hexagonal, so that positive rhombs fh0hlgcannot be distinguished from negative rhombs f0hhlg, i.e. a60� orientation ambiguity exists. It is therefore not possibleto use this method to identify Dauphine twins.

By comparing the angles between experimentally mea-sured diffraction peak positions with the theoretically cal-culated ones derived from unstrained hexagonal latticeparameters, orientation, as well as lattice parameters for adistorted triclinic cell, were refined by using least squaresfitting, and a strain tensor was obtained for each positionwhere a Laue pattern was acquired by comparing theorientation matrices before and after strain refinement(Pavese, 2005). The algorithm described by Pavese wasfurther simplified, especially for non-orthogonal crystalstructures, compared to the one used in old version ofXMAS (Chung & Ice, 1999). Since only polychromaticLaue diffraction patterns were collected, no informationabout lattice volume and thus hydrostatic strain could beobtained in this study. Based on previous optimizations of

Fig. 4. (a) Scheme of the X-ray microdiffraction experimental con-figuration in transmission mode and (b) one of the diffraction pat-terns taken in transmission mode with the Miller indices for some ofthe diffraction peaks.

Evidence for high stress in Vredefort quartz 171

Page 4: Evidence for high stress in quartz from the impact site of Vredefort ...

experimental conditions, we can safely assume that if thecalculated positions of more than 20 diffraction peaks onthe detector were less than 0.2 pixel from the measuredpeak positions, the pattern was indexed successfully;otherwise this pattern was rejected.

The strain tensor obtained using this method is a devia-toric strain. Negative values refer to compression, positivevalues to tension. The magnitude of the deviatoric strainwas estimated by ‘‘equivalent strain’’ (Liu, 2005), which isdescribed as Equation (1):

eeq ¼ 23

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffie011�e022ð Þ2þ e022�e033ð Þ2þ e033�e011ð Þ2þ6 e212þe213þe223ð Þ

2

r(1)

where eij (i, j¼ 1–3) are the components of a strain tensor.Note that it depends on all components of the deviatoricstrain tensor.

The six tensor components defining shape and orienta-tion of a strain tensor can be visualized as a strain ellipsoidrelative to an arbitrary sample coordinate system. A devia-toric strain tensor can be represented in a polar coordinatesystem with a spherical projection (Fig. 5a), in which ezz isat the centre, exx along the horizontal axis, eyy along thevertical axis, and the gray scale indicates the magnitude ofthe strain value in certain direction. For example, thedeviatoric strain tensor at the position where the arrow ispointing in Fig. 7 (exx) is expressed by a symmetric 3 � 3matrix in laboratory coordinates (Equation (2))

e ¼�0:28 �0:15 �0:17�0:15 0:12 0:60�0:17 0:60 0:17

24

35� 10�3 (2)

By plotting it in the polar coordinate system (Fig. 5a) withsubroutine VELO in BEARTEX (Wenk et al., 1998), wefound the highest tensile strain,þ0.79� 10�3, is applied inthe direction 102�, 47�, 45� with respect to x-, y-, andz-axis, respectively. Perpendicular to this direction we findalmost uniform compressive deviatoric strain. The strongestvalue (�0.46� 10�3) is found roughly within the y–z plane,45� oblique to z-direction towards �y-axis. The principalstrains of this tensor are �0.33 � 10�3, þ0.79 � 10�3, and�0.46 � 10�3, and the principal vectors in laboratory

coordinates are [�0.97, �0.19, �0.10], [�0.21, 0.68, 0.71],and [�0.07, 0.71, �0.70], respectively. Thus, if this tensoracted on a sphere, it resulted in a biaxial ellipsoid with onelong axis and an almost circular shortened equatorial plane.The representation illustrates that it is difficult to visualize thestrain ellipsoid intuitively from individual tensor componentsand, for a better understanding, a three-dimensional picture isrequired. The equivalent strain value of this distortion is 0.79� 10�3 (Equation (1)).

In order to study the resolution limit of the technique, a scanwas made on a sample of a synthetic quartz single crystal. Weassume this hydrothermally grown crystal to be completelystrain free, so that the sample itself was taken as a standard torefine experimental parameters. Another advantage of usingthe synthetic quartz as a calibration standard is that the error ofstrain measurement due to the uncertainty of reference latticeparameters is effectively reduced. The equivalent strainobserved in the synthetic quartz is ,0.1� 10�3.

Ideally the stress tensor can be calculated, when knowingthe strain tensor in lattice coordinates and the respectivestiffness constants, by applying Hooke’s law sij ¼ cijkl � ekl.However, for this case the strain tensor in lattice coordinatescould not be determined unequivocally because of the 60�

ambiguity of orientation indexing by using the hexagonalsetting of the trigonal quartz lattice. Furthermore, becausethe stiffness constant of quartz differs significantly between<2021> and <0221> directions, the stress tensor could notbe derived by applying Hooke’s law. Therefore, in this report,residual stress was estimated by multiplying equivalent strainby the isotropic Young’s modulus.

The peak width is determined by instrumental broadening,X-ray beam penetration, grain size, geometrically necessarydislocations (GND) and geometrically necessary boundary(GNB) densities (Barabash et al., 2002, 2004; Valek et al.,2003). It has an influence on the precision of the peak positionfitting, and thus orientation indexing and strain measurement.For this specific study of a thin-sectioned sample, peak broad-ening is controlled by the GNDs and GNBs, i.e. plasticdeformation of the sample. The spatial resolution of thistechnique is given by the scan step size (0.5 mm). The orien-tation- and strain resolutions were about 0.01� and 10�4 s.u.,respectively (Tamura et al., 2009). Peaks are fitted withLorentzian functions in 2D and the two main axes areextracted. For the average peak width we use an average oflarge axes over all indexed reflections.

3.3. Results

The orientation map of the scanned area of the Vredefortsample is shown in Fig. 6a. The white squares in the mapsare caused by the failure of indexing the correspondingdiffraction pattern. Figure 6a shows the angle between thecrystal c-axis and the z-axis of the laboratory coordinates(i.e., the normal to the sample plane). Values close to 90�

confirm that the c-axis is roughly within the sample plane.Figure 6b displays the angle between the crystal a-axis andthe y-axis of the laboratory coordinates. More detailedinformation demonstrates that the projection of the c-axis

Fig. 5. (a, b) Deviatoric strain value distributions at the positionindicated by the arrows in Fig. 7 and 11, respectively, in a polarcoordinate system. (a) is Vredefort quartzite and (b) is a quartz grainin a moderately deformed granite. Equal area projection. For (b) y isperpendicular to the foliation and x parallel to the lineation.

172 K. Chen, M. Kunz, N. Tamura, H.-R. Wenk

Page 5: Evidence for high stress in quartz from the impact site of Vredefort ...

onto the x–y plane lies within 5� off the x-axis, while the<10�10> or <01�10> direction is less than 10� oblique to thez-axis and the in-plane orientation distribution is about 5�

within the scanned region.These orientation maps allow identifying deformation

features as oblique lamellae (dashed lines in Fig. 6b). Mostof the unindexed spots appear close to the lamellar bound-aries, i.e. regions where domains are plastically deformed.This causes the diffraction peaks to be highly streakedwhich in turn makes them inaccessible to indexing.

The planar deformation features (dashed lines in Fig. 6b)are inclined relative to the y-axis by about 25�. Assuming thePDFs to be perpendicular to the sample surface, i.e. the z-axislies more or less in the PDF, the normal to the plane can bewritten in x–y–z coordinates as a vector p¼ [�1, tan 25�, 0]T,and T stands for transpose here. The Miller indices (hkil) ofthe planar features can be calculated by applying the equation:

hPDF ¼Mp (3)

where M is the 3� 3 orientation matrix of the quartz latticeas described in the previous section, p is the normal to thePDFs in x–y–z coordinates, and vector hPDF represents theMiller indices for the PDFs (hPDF¼ [h, k, l]T) in reciprocalcrystal lattice units.

As expected for a single crystal, the crystal orientation inthe raster scanned region remained almost constant. Thedeformation lamellae can thus be indexed approximatelyparallel to f10�13g or f01�13g when applying an averageorientation matrix to Equation (3) (positive and negativerhombs can not be distinguished with this method, sincediffraction peak intensities are not taken into account). Thisresult agrees with previous observations in shocked quartz(e.g., Vernooij & Langenhorst, 2005; Trepmann & Spray,2006). Vernooij & Langenhorst reported the observation ofthree different PDF features, including rhombohedral planes,f1013g andf1013g, and prismatic planesf1010g. Two typesof PDFs were described in the Charlevoix impact structure byTrepmann & Spray (2006). Type one consisted mainly ofrhombohedral planes, including both f1013g and f0113gplanes, while type two was detected as basal PDFs (0001).

The indexing of the PDF is confirmed by plotting theð0113Þ onto a spherical projection. As shown in Fig. 6c, thepole of ð0113Þ of this orientation is within the x–y planeand about 24.5� with respect to x-axis, which matches the

observation that the planar features are normal to thesample plane and about 25� oblique to the y-axis.

Figure 7 shows maps for the six individual strain com-ponents of the Vredefort sample. Again, white spots in themaps indicate that spots failed to be indexed correctly. Weassign negative and positive signs to represent compressiveand tensile strains, respectively. In general, the residualstrains are in the order of several thousand microstrains.High shear strain is observed close to the lamellar bound-aries, which we attribute to the mechanism that formed thelamellae, even though the strain component magnitudestrongly depends on selection of coordinates.

Figure 8a shows the equivalent strain distribution in theVredefort sample. Based on the definition, equivalent strainis always positive. Generally the equivalent strain level inthis sample is about 1.5 � 10�3. High equivalent strainvalues are observed at the lamellar boundaries (light) com-pared to the equivalent strain inside the lamellae (dark). Thehistogram of equivalent strain, with the most frequent valueat 1400 microstrain, is displayed in Fig. 8b. By applyingHooke’s law, s ¼ C � e, where C is Young’s modulus and eis equivalent strain, residual stress s is estimated. As C, wetook the average value, 59.4 GPa (Pincus, 1996) in ourcalculation. The histogram of residual stress distribution isdisplayed in Fig. 8c. The distribution maximum is at 90 MPaand FWHM is about 100 MPa.

Evidence for elastic strain is documented in lattice distor-tion as described above. There is also plastic strain and ameasure of plastic deformation is the diffraction peak widthand peak shape. Peak width has been employed for plasticitystudies for decades (Morosin & Graham, 1984, 1985; Ungaret al., 1984, 1998, 2001). The distribution map of peak width,as described in Section 3.2, is plotted in Fig. 9a. For eachdiffraction pattern, the peak width is defined to be the averagewidth of all the peaks that are indexed. The average peakwidth at the bottom centre of the scanned area, correspondingto high equivalent strain, is broadened, indicating highlyplastic deformation in this region, while to the right side ofthe bottom of the map, the peaks are much sharper, showingthat the center of the deformation lamella is much lessdeformed. The results are also summarized in a histogram(Fig. 9b). Diffraction peaks from the Vredefort sample aregenerally broad and peak widths show large variations, indi-cating the GND densities in the Vredefort sample are high

Fig. 6. Orientation maps of the Vredefort sample showing the angles (a) between crystal c-axis and sample Z-axis, (b) between crystal a-axisand sample Y-axis, and (c) stereographic projection showing a ð01�1�3Þ pole. See text for explanations.

Evidence for high stress in Vredefort quartz 173

Page 6: Evidence for high stress in quartz from the impact site of Vredefort ...

and spatially non-uniform. More detailed observation showsthat in some patterns the diffraction peaks are not onlyelongated but also split, indicating that subgrain boundariesare formed because of the realignment of the dislocations thathave the same sign (Fig. 4b).

An example of a typical pattern in the vicinity of thelamellar boundary is shown in Fig. 10a. A detail of a singlepeak, shown in the insertion of Fig. 10a, indicates that thediffraction peaks are not only streaked but also split. Thescheme in Fig. 10b shows how grain bending and subgrainare formed due to formation and alignment of GNDs. Bycomparing the peak streaking directions with the simulatedresults by using XMAS as described previously by Chen et al.(2008), the dislocation line direction is out of the (0001)plane, in agreement with the model proposed by Vernooij& Langenhorst (2005). Assuming a Burgers vector1=3<1120> (Vernooij & Langenhorst, 2005) and a grain

size of 0.4 mm, the dislocation density can be calculated bymeasuring the peak streaking angle and applying the Cahn-Nye relation r ¼ ð1=RbÞ (Cahn, 1949; Nye, 1953), where ris dislocation density, R is the bending radius of a crystalgrain, b is the length of the Burgers vector. In our example,the grain is bent by about 5� around c-axis in a 20 mm rangeand thus the bending radius is estimated to be about 200 mm.The Burgers vector is calculated to be 2.8 A, resulting in adislocation density of 2� 109/cm2, which is the same as thatreported by Vernooij & Langenhorst (2005), based on directTEM observations.

4. Discussion

The results on Vredefort quartz confirm the presence ofmicrostructural features indicative of strong deformation.

Fig. 8. (a) Map of equivalent strain in Vredefort quartz, (b) equivalent strain, and (c) residual stress distribution histogram.

Fig. 7. Map of strain component exx, exy, exz, eyy, eyz, ezz in Vredefort quartz. Colours indicate magnitude of deviatoric strain, white indicatesthat spot could not be indexed.

174 K. Chen, M. Kunz, N. Tamura, H.-R. Wenk

Page 7: Evidence for high stress in quartz from the impact site of Vredefort ...

EBSD measurements prove the presence of mechanicalDauphine twins that have been suggested, based on bulkpreferred orientation (Wenk et al., 2005) and documentedfrom other impact sites (Trepmann & Spray, 2005;Trepmann, 2008). In metamorphic quartz rocks such twinboundaries are rare (e.g., McLaren & Hobbs, 1972; Liddellet al., 1976; Knipe, 1990; Wang et al., 1993; Wenk et al.,2007b) and crystals have either never twinned, or twinninghas gone to completion, i.e. the crystal is completely re-oriented.

Laue microbeam diffraction suggests preservation ofsignificant residual stress of 40–140 MPa. This is

remarkable, considering the Precambrian age of theVredefort impact. But we should keep in mind that alsohighly metastable phases like coesite and stishovite sur-vived in this same sample. This can be compared withresidual stress observed in other materials. In copper stress,50 MPa was reported by Poulsen et al. (2001) with thetechnique of three-dimensional X-ray diffraction. Muchhigher stress (about 3 GPa) was detected in metal nitridecoatings with high energy X-ray diffraction (Almer et al.,2003). Daymond et al. (2000) measured high macroscopicstress in rolled austenitic steel (up to 1400 MPa) usingneutron diffraction. Laue patterns are indicative of residualelastic strain as well as plastic strain. The direction andmagnitude of distorted peaks allows deducing the disloca-tion line direction and density. We find a value of 2� 109/cm2, assuming a Burgers vector of 1=3<1120>.

Our data can be compared with an experiment on a quartzgrain in weakly deformed granite (Kunz et al., 2009a), whichhas been investigated using the same technique but in reflec-tion mode. In the previous study only equivalent strains weremapped, rather than strain tensor components and this provedto be somewhat misleading because in reflection geometrythe strain component eyz is very sensitive to alignment errors.Originally the diffraction geometry was calibrated using athin Si crystal. Recalibration using quartz itself as a calibrantreduced the magnitude of eyz significantly and thus loweredthe values for equivalent strain, compared to the valuesreported by Kunz et al. (2009a). The deviatoric lattice straincomponents in the granite sample shown in Fig. 11a–f areabout one half of the values of the Vredefort sample. The out-of-plane normal strain (ezz) of the deformed granite sample isfound to be significantly smaller compared to the in-planenormal strains (exx and eyy) due to the fact that the smallelastic strain was greatly relaxed by cutting the free samplesurface. exx is negatively related to eyy because only thedeviatoric strain tensor is obtained from the Laue diffractiontechnique and lattice volume is assumed to be kept constant.Looking at the 3D strain tensor in granite for a typical spot(arrow) we find that indeed the strain ellipsoid is more or lessaligned with the sample axes, which were chosen according

Fig. 10. (a) Diffraction pattern showing streaked diffraction peaks.An enlarged image of reflection 110 is shown. Note that streaking is

Fig. 9. (a) Average peak width distribution map and (b) average peak width histogram in Vredefort quartz.

Evidence for high stress in Vredefort quartz 175

Page 8: Evidence for high stress in quartz from the impact site of Vredefort ...

to macroscopic structural elements (Fig. 5b). Component exx,indicative of extension, is parallel to the lineation in thissample, whereas component eyy, indicative of compression,is perpendicular to the foliation. All other components arerelatively small. This pattern is quite different from Vredefort(Fig. 7) where diagonal as well as non-diagonal componentsdisplay large variations and suggest that shear parallel tolamellar boundaries played a significant role. However,keep in mind that the coordinate system xyz, on which thetensor components are based, is arbitrary. In this paper we usethe sample coordinate system related to surface and transla-tions. It is important to look at all components of the straintensor, i.e. normal and ‘‘shear’’ components to assess the truestrain (Fig. 5). A bulk measure for the magnitude of the strainis the ‘‘equivalent strain’’ Equation (1).

The equivalent strains of the synthetic quartz, granite, andVredefort samples are summarized in the histogram inFig. 12a. The equivalent strain in the deformed granite has adistribution with a maximum at 0.5� 10�3 and the distribu-tion width is about 0.3� 10�3, which is much narrower thanthe Vredefort sample. It corresponds to an equivalent stress of18 MPa (note this is lower than previously reported by Kunzet al., 2009a due to a different calibration and the sensitivity ofthe diagonal component eyz on the calibration). The averageequivalent strain in the Vredefort sample was about 1.5 �10�3; the distribution peak is at 1.3� 10�3 and the distribu-tion width is about 0.9� 10�3 FWHM, which is significantlyhigher and more scattered compared to the granite sample.The correlation of the peak width (and thus plastic strain) withequivalent strain is not very strong. We attribute this to alarger scatter in the determination of the peak positions (andthus strain) in cases where the plastic strain is large and the

Fig. 11. Map of strain components in quartz from deformed granite: exx, exy, exz, eyy, eyz, and ezz. This sample was measured in reflection modeand diffraction geometry was calibrated internally.

Fig. 12. Distributions of (a) equivalent strain and (b) diffraction peakwidths on synthetic quartz single crystal, quartz in granite, and quartzin the Vredefort sample.

176 K. Chen, M. Kunz, N. Tamura, H.-R. Wenk

Page 9: Evidence for high stress in quartz from the impact site of Vredefort ...

peaks streaked. Furthermore, plastic deformation asexpressed in peak streaking, partly releases elastic strain.

5. Conclusions

In our study we have documented mechanical Dauphinetwins as well as high residual deviatoric strains, especiallyclose to the lamellar boundaries, in quartz from theVredefort impact site. Using synchrotron X-ray Lauemicrodiffraction, high orientation and strain resolutionsare obtained with high spatial resolution. The equivalentstrain in the Vredefort sample is about three times higherthan in a deformed granite sample. The residual stress of100 MPa in the Vredefort sample is calculated by applyingHooke’s law on equivalent strain and Young’s modulus.The method has promise in recording residual stress innaturally deformed rocks that can then be used for quanti-fying the deformation history.

Acknowledgments: We acknowledge support from DOE-BES (DE-FG02-05ER15637) and NSF (EAR-0836402)and access to ALS beamline 12.3.2. ALS is supported bythe Director, Office of Science, Office of Basic EnergySciences, Materials Science Division, of the USDepartment of Energy under Contract No. DE-AC02-05CH11231. The micro-diffraction program at the ALSbeamline 12.3.2 was made possible by NSF grant #0416243. We acknowledge help by Scott Sitzman(Oxford Instruments) with some of the EBSD measure-ments and thank Uwe Reimold for providing the sample.We are appreciative for comments by reviewers A. Paveseand R.J. Angel, which helped to improve the manuscript.

References

Almer, J., Lienert, U., Peng, R.L., Schlauer, C., Oden, M. (2003):

Strain and texture analysis of coatings using high-energy X-rays.

J. Appl. Phys., 94, 697.

Barabash, R.I., Ice, G.E., Larson, B.C., Yang, W. (2002):

Application of white X-ray microbeams for the analysis of

dislocation structures. Rev. Sci. Instrum., 73, 1652.

Barabash, R.I., Ice, G.E., Tamura, N., Valek, B.C., Bravman, J.C.,

Spolenak, R., Patel, J.R. (2004): Quantitative characterization of

electromigration-induced plastic deformation in Al(0.5wt%Cu)

interconnect. Microelectron. Eng., 75, 24–30.

Cahn, R.W. (1949): Recrystallization of single crystals after plastic

bending. J. Inst. Met., 76, 121.

Chao, E.C.T., Shoemaker, E.M., Madsen, B.M. (1960): First natural

occurrence of coesite. Science, 132, 220–222.

Chao, E.C.T., Fahey, J.J., Littleb, J., Milton, D.L. (1962): Stishovite,

SiO2, a very high pressure new mineral from Meteor Crater, AZ.

J. Geophys. Res., 67, 419–421.

Chen, K., Tamura, N., Valek, B.C., Tu, K.-N. (2008): Plastic defor-

mation in Al (Cu) interconnects stressed by electromigration and

studied by synchrotron polychromatic X-ray microdiffraction. J.

Appl. Phys., 104, 013513.

Chung, J.S. & Ice, G.E. (1999): Automated indexing for texture and

strain measurement with broad-bandpass X-ray microbeams. J.

Appl. Phys., 86, 5249–5255.

Daly, R.A. (1947): The Vredefort ring structure of South Africa. J.

Geol., 55, 125–145.

Daymond, M.R., Tome, C.N., Bourke, M.A.M. (2000): Measured

and predicted intergranular strains in textured austenite steel.

Acta Mater., 48, 553–564.

Dietz, R.S. (1961): Vredefort ring structure: meteorite impact scar. J.

Geol., 69, 499–516.

Engelhardt, W. & Bertsch, W. (1969): Shock-induced planar defor-

mation structures in quartz from Ries crater, Germany. Contrib.

Mineral. Petrol., 20, 204–234.

Gibson, R.L., Armstrong, R.A., Reimold, W.U. (1997a): The age and

thermal evolution of the Vredefort impact structure: a single-grain

U–Pb zircon study. Geochim. Cosmochim. Acta, 61, 1531–1540.

Gibson, R.L. & Reimold, W.U. (2001): The Vredefort impact struc-

ture, South Africa (the scientific evidence and a two-day excur-

sion guide). Council Geosci. Pretoria Mem. 92, 111 p.

Gibson, R.L., Reimold, W.U., Wallmach, T. (1997b): Origin of

pseudotachylite in the lower Witwatersrand Supergroup,

Vredefort Dome (South Africa): constraints from metamorphic

studies. Tectonophysics, 283, 241–262.

Glinnemann, J., King, H.E., Schulz, H., Hahn, T., La Placa, S.J.,

Dacol, F., (1992): Crystal structures of the low-temperature

quartz-type phases of SiO2 and GeO2 at elevated pressure. Z.

Kristallogr., 198, 177–212.

Goltrant, O., Cordier, P., Doukhan, J.-C. (1991): Planar deformation

features in shocked quartz; a transmission electron microscopy

investigation. Earth Planet. Sci. Lett., 106, 103–115.

Goltrant, O., Leroux, H., Doukhan, J.-C., Cordier, P. (1992):

Formation mechanisms of planar deformation features in natu-

rally shocked quartz. Phys. Earth Planet. Inter., 74, 219–240.

Knipe, R.J. (1990): Microstructural analysis and tectonic evolution

in thrust systems: examples from the Assynt region of the Moine

Thrust Zone, Scotland. in ‘‘Deformation Processes in Minerals,

Ceramics and Rocks’’, D.J. Barber & P.G. Meredith, eds. Unwin

Hyman, London, 228–261.

Kunz, M., Chen, K., Tamura, N., Wenk, H.-R. (2009a): Residual

stress and domain orientation in natural quartz. Am. Mineral.,

94, 1059–1062.

Kunz, M., Tamura, N., Chen, K., MacDowell, A.A., Celestre, R.S.,

Church, M.M., Fakra, S., Domning, E.E., Glossinger, J.M., Plate,

D.W., Smith B.V., Warwick, T., Padmore, H.A., Ustundag, E.

(2009b): A dedicated superbend X-ray microdiffraction beamline

for materials, geo-, and environmental sciences at the advanced

light source. Rev. Sci. Instrum., 80, 035108.

Leroux, H., Reimold, W.U., Doukhan, J.-C. (1994): A TEM inves-

tigation of shock metamorphism in quartz from the Vredefort

Dome, South Africa. Tectonophysics, 230, 223–239.

Liddell, N.A., Phakey, P.P., Wenk, H.-R. (1976): The microstructure

of some naturally deformed quartzites. in ‘‘Electron Microscopy

in Mineralogy’’, H.-R. Wenk, ed. Springer, Berlin, Heidelberg,

419–427.

Liu, A.F. (2005): Mechanics and mechanism of fracture. ASM

International, Schaumburg, Illinois, 654 p.

Martini, J.E.J. (1978): Coesite and stishovite in the Vredefort Dome,

South Africa. Nature, 272, 715–717.

Evidence for high stress in Vredefort quartz 177

Page 10: Evidence for high stress in quartz from the impact site of Vredefort ...

Martini, J.E.J. (1991): The nature, distribution and genesis of coesite

and stishovite associated with pseudotachylite of the Vredefort

Dome, South Africa. Earth Planet. Sci. Lett., 103, 285–300.

McLaren, A.C. & Hobbs, B.E. (1972): Transmission electron micro-

scope investigation of some naturally deformed quartzites. in

‘‘Flow and Fracture of Rocks’’, H.C. Heard, I.Y. Borg, N.C.

Carter, C.B. Raleigh, eds. Geophys. Monogr. Am. Geophys.

Union, Washington, DC, 16, 55–66.

McLaren, A.C., Retchford, J.A., Griggs, D.T., Christie, J.M. (1967):

Transmission electron microscope study of Brazil twins and

dislocations experimentally produced in natural quartz.

Physica Status Solids, 19, 631–644.

Morosin, B. & Graham, R.A. (1984): X-ray diffraction line-broad-

ening studies on shock-modified rutile and alumina. Mater. Sci.

Eng., 66, 73–87.

—, — (1985): X-ray diffraction line broadening of shock-modified

titanium carbide. Mater. Lett., 3, 119–123.

Noyan, I.C. & Cohen, J.B. (1987): Residual Stress – Measurement

by Diffraction and Interpretation. Springer, Heidelberg, 276 p.

Nye, J.F. (1953): Some geometrical relations in dislocated crystals.

Acta Metallurgica, 1, 153.

Pavese, A. (2005): About the relations between finite strain in non-

cubic crystals and the related phenomenological P–V equation

of state. Phys. Chem. Minerals, 32, 269–276.

Pincus, H.J. (1996): Interlaboratory testing program for rock properties

– round two – confined compression: Young’s modulus, Poisson’s

ratio, and ultimate strength. Geotech. Test. J., 16, 138–163.

Poulsen, H.F., Nielsen, S.F., Lauridsen, E.M., Schmidt, S., Suter,

R.M., Lienert, U., Margulies, L., Lorentzen, T., Jensen, D.J.

(2001): Three-dimensional maps of grain boundaries and the

stress state of individual grains in polycrystals and powders. J.

Appl. Cryst., 34, 751–756.

Reimold, W.U. (1995): Pseudotachylites in impact structures – gen-

eration by frictional melting and shock brecciation? A review

and discussion. Earth Sci. Rev., 39, 247–264.

Tamura, N., MacDowell, A.A., Celestre, R.S., Padmore, H.A.,

Valek, B., Bravman, J.C., Spolenak, R., Brown, W.L., Marieb,

T., Fujimoto, H., Batterman, B.W., Patel, J.R. (2002): High

spatial resolution grain orientation and strain mapping in thin

films using polychromatic submicron X-ray diffraction. Appl.

Phys. Lett., 80, 3724–3727.

Tamura, N., MacDowell, A.A., Spolenak, R., Valek, B.C., Bravman,

J.C., Brown, W.L., Celestre, R.S., Padmore, H.A., Batterman,

B.W., Patel, J.R. (2003): Scanning X-ray microdiffraction with

submicrometer white beam for strain/stress and orientation map-

ping in thin films. J. Synchrotron Radiat., 10, 137–143.

Tamura, N., M., Chen, K., Celestre, R.S., MacDowell, A.A.,

Warwick, T. (2009): A superbend X-ray microdiffraction beam-

line at the advanced light source. Mater. Sci. Eng. A, 524, 28–32.

Trepmann, C.A. (2008): Shock effects in quartz: compression versus

shear deformation – an example from the Rochechouart impact

structure, France. Earth Planet. Sci. Lett., 267, 322–332.

Trepmann, C.A. & Spray, J.G. (2005): Planar microstructures and

Dauphine twins in shocked quartz from the Charlevoix impact

structure, Canada. Geol. Soc. Am. Spec. Pap. 384, 315–328.

—, — (2006): Shock-induced crystal-plastic deformation and post-

shock annealing of quartz from crystalline target rocks of the

Charlevoix Structure, Canada. Eur. J. Mineral., 18, 161–173.

Tullis, J. & Tullis, T. (1972): Preferred orientation of quartz pro-

duced by mechanical Dauphine twinning: thermodynamics and

axial experiments. in ‘‘Flow and Fracture of Rocks’’, H.C.

Heard, I.Y. Borg, N.C. Carter, C.B. Raleigh, eds. Geophys.

Monogr. Am. Geophys. Union, Washington, DC, 16, 67–82.

Ungar, T., Mughrabi, H., Ronnpagel, D., Wilkens, M. (1984): X-ray

line-broadening study of the dislocation cell structure in

deformed [001]-oriented copper single crystals. Acta

Metallurgica, 32, 333–342.

Ungar, T., Ott, S., Sanders, P.G., Borbely, A., Weertman, J. R.

(1998): Dislocations, grain size and planar faults in nanostruc-

tured copper determined by high resolution X-ray diffraction

and a new procedure of peak profile analysis. Acta Mater., 46,

3693–3699.

Ungar, T., Gubicza, J., Ribarik, G., Borbely, A. (2001): Crystallite

size distribution and dislocation structure determined by diffrac-

tion profile analysis: principles and practical application to cubic

and hexagonal crystals. J. Appl. Cryst., 34, 298–310.

Valek, B.C., Tamura, N., Spolenak, R., Caldwell, W.A., MacDowell,

A.A., Celestre, R.S., Padmore, H.A., Bravman, J.C., Batterman,

B.W., Nix, W.D., Patel, J.R. (2003): Early stage of plastic

deformation in thin films undergoing electromigration. J. Appl.

Phys., 94, 3757.

Vernooij, M.G.C. & Langenhorst, F. (2005): Experimental repro-

duction of tectonic deformation lamellae in quartz and compar-

ison to shock-induced planar deformation features. Meteorit.

Planet. Sci., 40, 1353–1361.

Wang, J.N., Boland, J.N., Ord, A., Hobbs, B.E. (1993):

Microstructural and defect development in heat treated

Heavytree quartzite. in J.N. Boland & J.D. Fitzgerald, eds.

‘‘Defects and processes in the solid state: geoscience applica-

tions (The McLaren Volume)’’, Elsevier, Amsterdam, 359–382.

Wenk, H.-R., Matthies, S., Donovan, J., Chateigner, D. (1998):

BEARTEX: a Windows-based program system for quantitative

texture analysis. J. Appl. Crystallogr., 31, 262–269.

Wenk, H.-R., Lonardelli, I., Vogel, S.C., Tullis, J. (2005): Dauphine

twinning as evidence for an impact origin of preferred orienta-

tion in quartzite: an example from Vredefort, South Africa.

Geology, 33, 273–276.

Wenk, H.-R., Bortolotti, M., Barton, N., Oliver, E., Brown, D.

(2007a): Dauphine twinning and texture memory in polycrystal-

line quartz. Part 2: in situ neutron diffraction compression

experiments. Phys. Chem. Minerals, 34, 599–607.

Wenk, H.-R., Monteiro, P., Shomglin, K. (2007b): Relationship

between aggregate microstructure and concrete expansion. A

case study of deformed granitic rocks from the Santa Rosa

Mylonite Zone. J. Mater. Sci., 43, 1278–1285.

Wooster, W.A., Wooster, N., Rycroft, J.L., Thomas, L.A. (1947):

The control and elimination of electrical (Dauphine) twinning in

quartz. J. Inst. Electr. Eng., 94, 927–938.

Zinserling, K. & Schubnikov, A. (1933): Uber die Plastizitat des

Quarzes. Z. Kristallogr., 85, 454–461.

Received 22 April 2010

Modified version received 1 October 2010

Accepted 8 November 2010

178 K. Chen, M. Kunz, N. Tamura, H.-R. Wenk


Recommended