You are on page 1of 15

Tectonophysics 457 (2008) 128–142

Contents lists available at ScienceDirect

Tectonophysics
j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / t e c t o

Grain coarsening maps for polymineralic carbonate mylonites: A calibration based on


data from different Helvetic nappes (Switzerland)
Andreas Ebert ⁎, Marco Herwegh, Alfons Berger, Adrian Pfiffner
Institute of Geological Sciences, Baltzerstrasse 1-3, CH-3012 Bern, Switzerland

A R T I C L E I N F O A B S T R A C T

Article history: Microstructures of carbonate mylonites with 0–40 vol.% second-phase particles from different Helvetic
Received 25 September 2007 nappes (Switzerland) have been analyzed. While the mean calcite grain size Dcc increases with temperature
Received in revised form 5 May 2008 (T), second phases pin calcite grain boundaries. Two types of second phases can be distinguished: small
Accepted 8 May 2008
second phases included in calcite grains and larger second phases at calcite grain boundaries, both
Available online 16 May 2008
coarsening with volume fraction and T. These trends are controlled by an interaction of different
Keywords:
processes during deformation, including diffusion, dynamic recrystallization, grain growth and pinning. In
Coupled grain growth terms of Dcc−T dependencies, the microstructures of the investigated nappes are similar. The relationship
Recrystallization between Dcc and T differs for high T and low T. These observations can be explained by a change in the
Calcite simultaneous activity of dominant mechanisms (e.g., diffusion and dislocation creep) as a function of T and
Second phases second-phase content. In contrast, the relation log(dp) − 1 / T (dp second-phase grain size) is linear with
Zener drag constant slope for all nappes indicating that growth processes of second phases are identical. Despite
Helvetic Alps differences in fluid activity and strain rate between individual nappes, the microstructural relationships of
Dcc, second phases and T remain similar. Therefore, grain coarsening maps of second phase and
recrystallization controlled calcite aggregates are a representative tool to predict deformation conditions
and mechanisms for other shear zones developed under similar physical conditions.
© 2008 Elsevier B.V. All rights reserved.

1. Introduction the grain boundary mobility and grain coarsening (e.g., Zener in
Smith, 1948; Nes et al., 1985, Olgaard and Evans, 1986; Hillert, 1988,
At mid to upper crustal levels of orogens, deformation is often Olgaard, 1990; Mas and Crowely, 1996; Evans et al., 2001, Herwegh
concentrated in high-strain shear zones consisting of carbonates (e.g., and Kunze, 2002; Herwegh and Berger, 2004; Ebert et al., 2007b). The
Schmid, 1975; Burkhard, 1990; Kennedy and Logan, 1997; Bestmann changes in microstructural processes will affect the physical and
et al., 2000; Molli et al., 2000; De Bresser et al., 2002; Ulrich et al., 2002; rheological properties of a rock (e.g., Olgaard, 1990, Doherty et al.,
Leiss and Molli, 2003; Herwegh et al., 2005b; Ebert et al., 2007a,b). In 1997). The volume fraction fp and size dp of second phases influence
these shear zones mylonitic microfabrics developed, which are the degree of pinning and/or dragging. In general, a large number of
controlled by extrinsic parameters like temperature, stress, strain rate small grains are more effective in hindering grain coarsening than a
(e.g., De Bresser et al., 2002; Herwegh et al., 2005b) and intrinsic few large grains (e.g., Alexander et al., 1994; Fan et al., 1998; Stearns
parameters like chemical impurities (e.g., Freund et al., 2001, 2004; and Harmer, 1996; Herwegh and Berger, 2004; Herwegh et al., 2005a;
Herwegh et al., 2003), second-phase particles (e.g., Olgaard and Evans, Ebert et al., 2007b). This effect can be expressed with the equation
1986; Herwegh and Berger, 2004; Ebert et al., 2007b), rates of Dmax = c ⁎ dp/fpm (Zener cited in Smith, 1948; Olgaard and Evans, 1986;
recrystallization cycles (e.g., Herwegh and Kunze, 2002), synkinematic Evans et al., 2001 and references therein) where Dmax is the matrix
fluids (e.g., Kirschner et al., 1999; Schenk et al., 2005), diffusion grain diameter, m is an exponent and c is a constant. Besides fp and dp
coefficients and grain boundary width (e.g., Paterson, 1995), and/or of second phases, their shape and dispersion are additional key
dislocation densities and distributions (e.g., Goetze and Kohlstedt, 1977; parameters (see in Ryum et al., 1983; Olgaard and Evans, 1986; Mas
Urai et al. 1986; De Bresser, 1991). and Crowely, 1996; Fan et al., 1998, Evans et al., 2001; Ebert et al.,
In nature, only few cases exist where carbonate mylonites are 2007b). The location of the second phases determines the m value.
nominally pure. The majority of carbonate mylonites contain other Second phases that are randomly distributed, located along grain
phases, here called second phases. They can pin and drag grain boundaries or at grain triple junctions, respectively, result in m values
boundaries of the matrix phase, which results in a retardation of both of 1, 1/2 and 1/3 (e.g., Olgaard and Evans, 1986).
Particularly under static conditions, both matrix and second
⁎ Corresponding author. Tel.: +41 31 631 8768; fax: +41 31 631 4843. phases can grow and consequently interact with each other during
E-mail address: ebert@geo.unibe.ch (A. Ebert). coarsening (Joesten, 1991; Hickey and Bell, 1996; Evans et al., 2001;

0040-1951/$ – see front matter © 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.tecto.2008.05.007
A. Ebert et al. / Tectonophysics 457 (2008) 128–142 129

Solomatov et al., 2002; Berger and Herwegh, 2004), referred to as carbonate rocks deformed under lower greenschist-facies conditions (de
coupled grain growth (Higgins et al., 1992; Voorhees, 1992; Alexander Bresser et al., 2002; Herwegh et al., 2005a,b). Moreover, paleopiezometers
et al., 1994; Chen and Fan, 1996; Stearns and Harmer, 1996; Fan et al., can only be applied to pure monomineralic aggregates; a restriction which
1998). In case of deformation, however, grain size reducing mechan- prohibits the use of paleopiezometer in the case of the polymineralic
isms are additionally active. Recently, Herwegh and Berger (2004), samples of this study. For this reason, we will refer in the following only to
Herwegh et al. (2005a) and Ebert et al. (2007a,b) demonstrated that temperature, although we are well aware that the coupling
the combination of both grain growth and grain size reducing of temperature, stress, and strain rate control the microfabric. (e.g., Austin
mechanisms in polymineralic systems preserve dynamic steady state and Evans, 2007; Austin et al., 2008).
microfabrics that are characteristic for the deformation conditions.
These studies, however, focused on individual large-scale shear zones 2. Geological settings
only. As a follow-up of this work, we will now compare the
microfabrics of polymineralic carbonate mylonites derived from The Helvetic nappes of Switzerland are one of the main structural units
different thrusts of the Helvetic Alps. We present the effect of content, of the Alpine orogen and represent a classical fold-and-thrust belt (e.g.,
size, type, and distribution of second phases and temperature/stress Ramsay, 1981; Pfiffner, 1993). In the western part of the Swiss Alps, the
on the calcite grain size. Based on these results, we discuss the Aiguilles Rouges massif is overlain by the Morcles nappe (Fig. 1). The
influence of different settings between the nappes and we recalibrate Morcles nappe itself is overlain by the Diablerets, Mt. Gond, and Sublage
grain coarsening maps of Herwegh et al. (2005a). Particularly the nappes (Masson et al., 1980; Ramsay, 1981; Escher et al., 1993; Pfiffner,
latter allows future application of the concept to predict deformation 1993). The Helvetic nappes of Central Switzerland are built up of a nappe
microfabrics of polymineralic carbonate rocks in mid-crustal levels. stack consisting of the Doldenhorn, Gellihorn, and Wildhorn nappes
In nature, temperature and stress go hand in hand, resulting in lower overlaying the Aar-Gastern massif (Fig. 1; Burkhard, 1988; Pfiffner et al.,
rock strengths at elevated temperature conditions (e.g., Kohlstedt et al., 1997). In eastern Switzerland, the Helvetic zone can be subdivided into
1995). While temperature can be estimated with geothermometers in a two major tectonic units, the Infrahelvetic complex in the footwall, and the
reliable manner, stress estimates are mainly based on the experimentally Glarus nappe complex in the hanging wall of the Glarus thrust (Fig. 1,
calibrated paleopiezometric relationship between recrystallized grain size Pfiffner, 1993; Schmid et al., 1996). Although all nappes were stacked and
and stress (e.g., Twiss, 1977; Rutter, 1995). Extrapolation of experimental deformed during the same time period ranging from Oligocene to
stress data towards natural conditions is particularly problematic for Miocene (Masson et al., 1980; Milnes and Pfiffner, 1980; Burkhard, 1988;

Fig. 1. Cross-sections through the western (AB), central (CD) and eastern (EF) Helvetic nappe stack of Morcles/Diablerets, Dolden/Gelli/Wildhorn, and Glarus nappe complex
(modified after Burkhard, 1988; Escher et al., 1993; Pfiffner, 1993; Herwegh and Pfiffner, 2005). Numbers represent sample locations of approximately NW–SE sample profiles
projected onto the cross-sections.
130 A. Ebert et al. / Tectonophysics 457 (2008) 128–142

Kirschner et al., 1996; Lihou, 1996; Pfiffner et al., 1997; Stampfli et al., Diablerets, Doldenhorn, Gelli-Wildhorn, and Glarus nappe) with a
2002), their internal tectonic structures vary along strike. While the focus on important influencing parameters like temperature/stress
Morcles and Doldenhorn nappe represent large-scale recumbent folds, the and second phases. Samples were collected that show strong
higher nappes are mostly thrust sheets made up of folded normal se- mylonitic fabrics, such as penetrative foliation and stretching linea-
quences. Imbrication and folding were accompanied by ductile deforma- tion. To investigate the effect of second phases, different lithologies
tion and pervasive stretching in the internal part with strain ratios of up to with varying contents and types of second phases were sampled
X/Z=400 (Ramsay, 1981; Siddans, 1983; Dietrich, 1989; Dietrich and Casey, across the shear zone at different sample locations. Although, deforma-
1989). The Glarus nappe complex differs to the western and central nappe tion continued on the retrograde P–T path resulting in further strain
stacks. It consists of one predominant major thrust, which accommodated localization within the shear zone (Ebert et al., 2007a,b), we will only
a much higher displacement of up to 40–50 km than the other nappes focus on the peak temperature structures in this study. To gain insight
(Schmid, 1975; Groshong et al., 1984; Pfiffner, 1985), while thrust related into the temperature effect, samples were collected along NW–SE
strain was localized in a narrow zone of only a few meters in width transects of each basal thrust (Fig. 1). The sample profiles were chosen in
(Schmid, 1975; Milnes and Pfiffner, 1977; Ebert et al., 2007a). such a way that they are approximately normal to the isogrades of peak
For all three mid-crustal sections, peak metamorphic temperatures metamorphic temperatures (see Herwegh and Berger, 2004; Herwegh
increase from the diagenetic-anchizone transition in frontal parts to and Pfiffner, 2005; Ebert et al., 2007a,b and references therein). Sample
lower greenschist-facies conditions in the rear (Frey et al., 1980; locations are projected onto the profile traces parallel to the axes of
Groshong et al., 1984; Burkhard and Kerrich, 1988; Frey, 1988; Dietrich large-scale folds and temperature isograds. The distances between the
and Casey, 1989; Burkhard, 1990; Burkhard et al., 1992; Rahn et al., localities along the thrusts range between 0.5 and 10 km (mean of 3–
1994, 1995, 2002; Marquer et al., 1994; Crespo-Blanc et al., 1995; 5 km), which correspond to relative differences in peak metamorphic
Kirschner et al., 1995; Burkhard and Goy-Eggenberger, 2001; Herwegh temperatures of around 15–30 °C (Fig. 2). In each nappe, the localities are
and Pfiffner, 2005; Herwegh et al., 2005a; Ebert et al., 2007a,b). numbered from north to south starting with 1 in the north (Fig. 1). The
distances between the locations in Fig. 2 are referred to distances
3. Analytical methods measured along the thrust plane, starting with 0 km at reference points
predefined in the north for each section.
In this study, we compare microstructures of polymineralic Microstructural analyses were performed on the X–Z-plane
carbonate mylonites from different large-scale shear zones (Morcles, of polished and etched rock chips of carbonate mylonites. The X-

Fig. 2. Temperature estimations along the analyzed sample profiles from north to south. Numbers represent the locations of analyzed microstructures of this study. Used
thermometers and references are given in the legend of each diagram.
A. Ebert et al. / Tectonophysics 457 (2008) 128–142 131

and Z-axes represent the stretching lineation and the axis oriented Doldenhorn and Glarus nappe can be found in Herwegh and Berger
normal to the foliation, respectively. Due to application of the two- (2004), Herwegh and Pfiffner (2005), Ebert (2006), and Ebert et al.
step etching procedure (see details in Herwegh, 2000; Ebert et al., (2007a,b). In case of the new temperature profile of the Diablerets
2007a) grain and interface boundaries were modeled out resulting in nappe, calcite–dolomite temperatures of this study were combined
topographic differences between calcite grains and grain boundaries, with oxygen isotope thermometry and illite crystallinity data from
as well as between calcite and second phases. Digital images of the Dietrich and Casey (1989) and Crespo-Blanc et al. (1995) (Fig. 2). All
microstructures were acquired with a Cam Scan (CS4) scanning references of temperature estimates are labelled in Fig. 2.
electron microscope. Note that two-step etching is only sensitive to The studied parts of the Helvetic nappes are characterized by peak
high angle grain boundaries and not to subgrain boundaries as metamorphic conditions that range from around 200–250 °C in frontal
confirmed by a comparison of microstructures derived from SEM and parts (N) to 350–400 °C in rear zones (S, Fig. 2). Resulting temperature
thin sections (Ebert et al., 2007c). In addition, twin boundaries also gradients along the thrust planes are all around 6 ± 1 °C per km. Least
become visible, although they can be discriminated from the high square regressions were performed to obtain temperatures for the
angle grain boundaries by their straight and parallel traces. Using the individual sample locations for data sets from each nappe. Note that
program Adobe Photoshop and the public domain software Image although the absolute error of the temperature estimates for each
SXM, grain boundaries of both calcite and second phases were hand- single thermometer is high (N30 °C, Herwegh and Pfiffner, 2005),
traced and analyzed from the obtained digital SEM images. From the average values from different geothermometers yield smaller errors
resulting grain area, the equivalent circular grain diameter was (b10–15 °C, Fig. 2). The temperature estimates given in this study are
calculated for each grain (Herwegh and Berger, 2004; Ebert et al., derived from the best-fit regression values at the specific sample
2007b). From the grain area distribution, an area-weighted average locations.
grain size Dcc was calculated (for more details see Ebert et al., 2007b),
which is close to the real 3-D grain size distribution (Berger and 4.3. Steady state microstructure of calcite
Roselle, 2001; Herwegh et al., 2005b). To quantify the second-phase
influence on the grain boundaries of calcite, the Zener parameter All analyzed calcite aggregates show steady state deformation
Z = dp/fp was determined (e.g., Zener in Smith, 1948; Herwegh and microstructures (see also Means, 1981; Herwegh and Handy, 1996).
Berger, 2004). The variable dp represents the number-weighted mean Processes like dynamic recrystallization, grain growth, mass transfer,
grain size and fp the volume fraction of the second phases. and inter- and intracrystalline deformation can stabilize the rock's
Furthermore, the types of second phases were distinguished and fabrics that are typical for specific metamorphic deformation condi-
separated into two groups, sheet silicates as predominant second tions. These fabrics are characterized by strong textures, unimodal
phase and remaining phases consisting mainly of quartz and grain size distributions that are symmetric to slightly right-sided
dolomite. This separation was made in light of determining growth skewed (Schmid et al., 1981; Herwegh and Kunze, 2002; Herwegh and
behaviors of sheet silicates and to gain insight into coupled grain Berger, 2004; Herwegh et al., 2005b; Ebert et al., 2007a,b). As
growth of the two main interacting phases calcite and sheet indicated by experiments, shear strains of γ N 5–10 are sufficient to
silicates. create a steady state microstructure (e.g., Herwegh and Handy, 1996;
Pieri et al., 2001; Rybacki et al., 2003; Barnhoorn et al., 2004).
4. Results Moreover, strains of X/Z N 100 obtained along the Helvetic nappes
(Ramsay, 1981; Siddans, 1983; Dietrich, 1989; Dietrich and Casey,
4.1. Sample description 1989), strong foliation parallel to the shear zone, and stretching
lineation confirm that the microstructures have reached high strains
All samples have been derived from deformed Mesozoic carbo- and therefore steady state.
nates. Independent of their nappe origin, they show rather similar
microfabrics for samples deformed at same temperature conditions. 4.3.1. Calcite grain size influenced by second phases
The carbonates represent strong mylonitic fabrics, which are As previously described, the effect of second phases can be
characterized by a well-developed foliation, strong textures and expressed in form of the Zener parameter Z (see Sections 1 and 3).
elongated and orientated grains. The foliation is caused by μm- to The relationship between the average grain size of calcite Dcc and Z for
mm-thick layers that differ in impurity contents. The impurities are each Helvetic nappe is given in Fig. 3. The deformation microstruc-
represented by micro-scale second-phase particles, which are sheet tures of impure and pure carbonate mylonites can be separated into
silicates, quartz, dolomite, albite, apatite, and ores. Dependent on fabrics that are controlled by second phases (Z ≪ 100 µm) and by
the stratigraphical units, their quantities vary in the range of b1 to dynamic recrystallization ( Z ≫ 100 µm), respectively. This relationship
40 vol.%. The grayish appearance of some mylonites gives evidence was already presented by Herwegh and Berger (2004) and Herwegh et
for well-dispersed organic carbon of nano-scale particle size al. (2005a) for the Doldenhorn nappe (note that some samples were
(Herwegh and Kunze, 2002). Detailed descriptions for the Morcles, recalculated due to a calibration error yielding a slight shift of some
Doldenhorn, and Glarus nappe are given in Herwegh and Berger data points in the Dcc−Z diagram of Fig. 3c). Ebert et al. (2007b)
(2004), Ebert (2006), and Ebert et al. (2007a,b). Veins of different specified this transition zone in case of samples from the Morcles
recrystallization degrees are abundant along each thrust. These were nappe. They defined the transition between both fields by second-
only incorporated into the microstructural investigations when they phase contents fp of around 2 vol.%. Depending on deformation
are totally recrystallized, as manifest by the occurrence of thin white temperatures, this boundary is oblique, i.e. with increasing tempera-
and very pure calcite bands that are parallel in orientation with ture the transition shifts to larger Z values (Fig. 3). In case of the
respect to the foliation. investigated nappes, the boundaries are situated at similar Dcc−Z
values. The pure recrystallization controlled aggregates are character-
4.2. Temperature ized by a homogeneous calcite grain size, which is independent on the
second-phase content. In contrast, second phases of quantities of 2 to
Peak metamorphic temperatures have been derived from a variety 40 vol.% pin the calcite grain boundaries and keep the grain size of
of different geothermometers including calcite–dolomite and calcite– calcite smaller for second-phase controlled fabrics. The effect of the
graphite thermometry, illite crystallinity, fluid inclusions, vitrinite impurities on the mean calcite grain size increases if larger numbers
reflectance, mineral parageneses, and/or oxygen isotope thermo- (i.e. higher volume fraction) of second phases are located on grain
meters. Details of analytics and temperature profiles for the Morcles, boundaries.
132 A. Ebert et al. / Tectonophysics 457 (2008) 128–142

Fig. 3. Relationship between mean calcite grain size Dcc, temperature and second phases for different impure carbonate mylonites from the investigated Helvetic nappes. Z = dp/fp
represents the Zener parameter (dp = grain size and fp = volume fraction of second phases). Each symbol characterizes samples from one location that is represented by the location
number in the symbol. Best fits are calculated for second phase and recrystallization controlled data, respectively, and resulting errors in grain size are given for both microstructure
types at left- and right-hand side of each trend line. In case of the Glarus nappe, only samples that were affected by cooling-induced strain localization represent enough second
phases to show a pinning effect (symbols with a cross insight of locations 3 and 4).

After Herwegh and Berger (2004) the average calcite grain size Dcc for same Z value is similar for equivalent temperatures. As a great
of a second-phase controlled microstructure follows the modified advantage of this consistency, the calcite grain size can be estimated
relationship after Zener (cited in Smith, 1948): for each second-phase content (size dp and volume fraction fp). This is
important in light of grain size and temperature dependent processes,
 mT
T dp when different nappes or samples with variable second-phase
Dcc ¼ cdZ m ¼ cd ð1Þ
fp contents have to be compared. Without correcting for similar Z
values, stress estimates and activation energies cannot be derived
where c is a temperature dependent constant, Z is the Zener parameter, from classical paleopiezometry or Arrhenius-type diagrams, respec-
dp and fp are the mean grain size and volume fraction of the second tively (see below). In this sense, Eq. (1) can be modified as recently
phases, respectively, and m⁎ is the slope of the Dcc–Z relation in Fig. 3. suggested by Herwegh et al. (2005a) into an Arrhenius-type form:
In this study, the average of the exponent m⁎ is 0.24 ± 0.06 for the
Morcles nappe, 0.33 ± 0.09 for the Diablerets nappe, 0.32 ± 0.14 for the    mT
−Qcc dp
Doldenhorn nappe, and 0.45 ± 0.02 for the Glarus nappe. Note that in Dcc ¼ c Vd exp d ð2Þ
RT fp
the latter case the value of m⁎ is only weakly defined due to the small
data set. The average value for m⁎ was calculated by weighting the where c′ is a constant, Qcc an activation energy, R the gas constant,
mean m⁎ of each data set of one nappe at constant temperature by the and T the deformation temperature in K.
number of all data points belonging to this data set. The resulting mean
m⁎ for all nappes is 0.30 ± 0.05. 4.4. Microstructure of second phases

4.3.2. Calcite grain size influenced by deformation temperature Second phases are homogeneously distributed within each ana-
Besides the second-phase influence, the calcite grain size is lyzed sample area. However, on a larger scale of some 100 μm to mm,
controlled by the deformation temperature. With increasing meta- second phases are enriched in bands that define the foliation. The
morphic conditions from anchizonal to lower greenschist-facies second phases are covering a grain size range of b1 to 10 μm (Fig. 4).
conditions the calcite grain size increases from b5 to N50 μm They can be separated into particles that are situated at grain
(Fig. 3). In the investigated Helvetic nappes, the grain size of calcite boundaries only and particles that are included in the interior of
A. Ebert et al. / Tectonophysics 457 (2008) 128–142 133

Fig. 4. The influence of volume fraction fp and temperature on the mean second phase grain size dp. Each symbol characterizes samples from one location that is represented by the
location number in the symbol. Symbols with a cross insight represent particles that are included in the calcite, while the other symbols represent the particles at the calcite grain
boundaries. For the Diablerets nappe Fig. a displays the relationship of all second phases, while in panel b the relationship of sheet silicates only is represented. The diagrams c and e
of the Morcles and Doldenhorn nappe, respectively, represent sheet silicates, while the diagrams of d and e show data points representing all second phases. The effect of retrograde
strain localization is given in panel d by samples of location 7 plotted as black squares.

calcite grains. The size dp and volume fraction fp of these included of second-phase particles, especially in case of recrystallization
second phases are always smaller than dp and fp of particles situated at controlled samples.
grain boundaries of calcite (Fig. 4). The transition between second At constant temperature conditions, the grain sizes dp of all second
phases that are dropped or situated at calcite grain boundaries is phases increase with rising volume fractions fp in all nappes (Fig. 4).
located at similar fp values in the diagrams of Fig. 4 for the Morcles and Furthermore, at constant fp, dp increases by a factor of two to three
Diablerets nappe. The variation in dp is more pronounced than with increasing deformation temperatures from around 250 to 400 °C
observed for the calcite grain size. This is due to the smaller number (Fig. 4). As in the case of the calcite, the grain sizes of the second
134 A. Ebert et al. / Tectonophysics 457 (2008) 128–142

phases are similar at identical temperatures for all four Helvetic effect of retrograde strain localization on microfabrics from the Morcles
nappes. The relationship of the second-phase grain size dp and the nappe. Again, the second phases are distinguished as either included
volume fraction fp at constant temperature can be approximated by a grains or grains at calcite grain boundaries. In localized samples, both
power law relationship in the form of: types of second phases display reduced grain sizes by ≈2 μm under
retrograde conditions independent on their volume fraction (Fig. 4c).
dp ¼ kd fpn ð3Þ
5. Calibration of grain coarsening maps from steady state
with a constant k and an exponent n. The average of n for all nappes is
deformation microstructures
around 0.15 ± 0.05. Since k is temperature dependent, Eq. (3) can be
rewritten such that it includes an activation energy term:
The fact that all phases grew with increasing temperature means that
  they will affect each other. Due to pinning, the second phases retard
−Qp
dp ¼ kVd exp d fpn ð4Þ grain growth of the matrix phase. Since the second phases grow with
RT
temperature, the matrix phase becomes affected by the second-phase
here, k′ is a constant, Qp the activation energy for second-phase growth as well as by thermally-induced grain growth itself. As
growth, R the gas constant, and T the absolute temperature. In k′ the represented before (Fig. 3), the stabilized grain size in steady state
dependency of other parameters influencing the second-phase grain mylonites depends on temperature, volume fraction and size of second-
growth can be included. These are, for example, the diffusion constant phase particles. This relationship can be calculated by combining Eqs. (2)
or the grain boundary width (e.g., see Paterson, 1995). and (4), which results in:
Additionally, second phases differ in mineralogy and shape. In this
    T
study, they were subdivided into sheet silicates, which represent the T −Qcc −Qp m mT ðn−1Þ
Dcc ¼ cVdk Vm d exp d exp d fp ð5Þ
predominant second phase, and a minor component consisting of RT RT
quartz, dolomite, apatite, ores, and other minerals. In order to evaluate
whether there exists a different growth behavior of sheet silicates in From this relation, grain coarsening maps can be obtained (see also
comparison to other second phases, the change in second-phase grain Herwegh et al., 2005a). Note that only the predominant second phase,
size as function of their volume fraction is plotted in dependence of the sheet silicates, was incorporated in the calculation because of
temperature for sheet silicates and other second phases for the presumably varying growth behaviors of different types of second
Diablerets nappe (Fig. 4a, b). Except for the samples from location 5, phases (see also Section 4.4). This equation has the potential to offer a
the small quantities of non-sheet-silicate second phases do not affect large variety of information provided that the necessary constants can
the relationship. The trends in Fig. 4a and b do not differ significantly be calibrated (see below).
given the errors. In case of location 5 (Fig. 4a), the volume fractions of
second phases other than sheet silicates are higher than for those of 5.1. Estimation of activation energies for grain coarsening maps
sheet silicates only for total second-phase contents N10 vol.%. Because
of the small differences between plots a and b in Fig. 4 between sheet 5.1.1. Calcite
silicates only and the data set including all second phases, only sheet Activation energies for the dominating processes of microstructural
silicates were considered in Fig. 4c and e in case of the Morcles and evolution can be derived from log(Dcc) vs. 1/T diagrams (Fig. 5); for
Doldenhorn nappes. Due to low second-phase contents (b5 vol.%) in details see data repository in Herwegh et al. (2005a). The grain growth
analyzed samples from the Glarus nappe (Fig. 4f), a considerable error behavior of the analyzed samples is represented in Fig. 5 for the different
in the dp versus fp relationship results. nappes. The plotted data points are derived from Fig. 3. To minimize the
Retrograde strain localization can be observed along all thrusts error, a best fit for each data set at constant temperature was
investigated (Herwegh and Pfiffner, 2005; Ebert et al., 2007a,b). The recalculated after Eq. (1) using the average m⁎ of each nappe. Note
samples of this study treated so far, all derived from the peak that different m⁎ values were obtained for second-phase and recrys-
temperature parts of the shear zones only. Fig. 4c demonstrates the tallization controlled microstructures. The calcite grain size at Z values of

Fig. 5. Arrhenius-type diagrams illustrating the change of mean calcite grain size Dcc with temperature (given in 1/T in Kelvin) for a constant Z value of (a) Z = 500 μm and (b)
Z = 30 μm, i.e. for pure (recrystallization controlled) and impure (second-phase controlled) samples, respectively. Data points of each nappe are characterized by one symbol. In panel
b the dark-grey background of the data scatter of panel a is integrated to visualize the second-phase influence on the calcite grain size. Due to the outcrop conditions at location 2 of
the Glarus nappe, only samples affected by late retrograde strain localization could be sampled and therefore had to be corrected for strain localization. Error bars represent a
maximum variation in temperature of ± 15 °C and in calcite grain size of ±10%.
A. Ebert et al. / Tectonophysics 457 (2008) 128–142 135

the data scatter increases. Although the relative variations in dp are


larger than is the case for the calcite grain size, dp shows no consistent
signs of a change in slope with increasing temperature. Therefore, we
chose a linear fit for the four nappes that is within the error (Fig. 6).
Resulting activation energies Qp were determined in the same way as
done for calcite yielding an average of Qp = 40 ± 7 kJ/mol (due to the
small data set of the Glarus nappe, these data were excluded, but if
included Qp = 45 ± 12 kJ/mol).

5.2. Discussion of resulting activation energies and their relevance on


coupled grain coarsening

The steady state microfabrics evolved in all nappes under similar


temperature/stress/strain rate conditions. Consequently, it is not
surprising to see that the relationship log(Dcc) vs. 1/T is comparable
for the different data sets (Fig. 5). Interestingly, the log(Dcc) vs. 1/T
relation is non-linear (Fig. 5). Herwegh et al. (2005b) already
discovered this behavior in a previous study and attributed it to
Fig. 6. Change of mean grain size dp of sheet silicates with temperature for a constant
volume fraction fp = 0.3 for the 4 investigated Helvetic nappes characterized by different
relative changes in grain growth and grain size reducing mechanisms.
symbols. In this light, the efficiency of grain growth increases in comparison to
grain size reduction with increasing temperature during the pre-
servation of dynamically-balanced steady state microfabrics. In order
30 and 500 μm for impure and pure samples, respectively, was used at to mathematically describe the microfabric changes of our new data
each temperature to calculate the activation energies in a log(Dcc) vs. 1/T set, which extends over several Helvetic nappes, three hypotheses can
diagram (Fig. 5). In regard of temperature uncertainties, a maximum be considered: (a) a non-linear character of the log(Dcc) vs. 1/T relation
error of ±15 °C was taken into account. Concerning variations in the that is only apparent and results solely from the data scatter, (b) a low
average grain size Dcc around the best fit in Fig. 3, a maximum error of and high temperature relationship with a transitional zone in
10% of the mean calcite grain size was assumed. These variations in Dcc between, or (c) a log(Dcc) vs. 1/T trend that changes gradually with
are presumably caused by local inhomogeneous distributions of second temperature in a non-linear manner.
phases in the calcite aggregates or local microstructural overprints by (a) Although the compilation of all nappe data results in a scatter, a
retrograde strain localization that cannot be discriminated anymore. The non-linear character of the log(Dcc) vs. 1/T relation results for each
latter point can be the case for areas located at the transition from high nappe that is clearly beyond potential data scatter. Furthermore, to
temperature to retrograde shear zones. In general, however, retrograde expand the temperature range, microstructural analyzes from other
strain localization can clearly be discriminated, even allowing recon- studies on natural carbonate mylonites were included (Fig. 7). Note
struction of high temperature microstructures, as done for one sample of that only data that were not affected by cooling-induced strain
the Glarus thrust from location 2. This step is necessary because the localization or annealing and data where the deformation tempera-
outcrop geometry of this locality only allowed the collection of samples tures were well established were used for the model. The data scatter
affected by later strain localization (Ebert et al., 2007a). The data points
of each nappe in diagrams of Fig. 5 coincide for recrystallization and
second-phase controlled samples within the error. As demonstrated
above, second-phase affected microstructures are characterized by
smaller grain sizes than is the case for pure samples (Fig. 3), an effect that
is also abundant in Arrhenius-type diagrams (Fig. 5b). In the log(Dcc) vs.
1/T space, both recrystallization and second-phase controlled data of
each nappe indicate a non-linear relationship. The best fits above 320–
330 °C represent a steeper slope than is the case for the low temperature
data. The question, whether this trend can be defined by a linear fit with
a constant slope, by a gradual change in slope, or rather requires a
combination of a high and low temperature trend characterized by two
different slopes, will be discussed below. Consequently, activation
energies vary depending on the best fit used. Using the slope of the
whole data set yields activation energies of Qcc = 45± 12 kJ/mol and 56±
12 kJ/mol for impure and pure samples, respectively. Otherwise, Qcc
determined from the steep slope of the high temperature part only
is 69 ± 10 kJ/mol (second-phase controlled) and 84 ± 10 kJ/mol
(recrystallization controlled) and for the low temperature part with
a reduced slope, Qcc is 25± 4 kJ/mol (impure) and 33 ± 4 kJ/mol (pure).
Fig. 7. Change of mean calcite grain size Dcc with temperature for pure microfabrics for
this study (black dots) and other field studies derived from literature (gray dots)
5.1.2. Second phases
(Be = Bestmann et al., 2000, Bu = Burkhard, 1990, Co = Covey-Crump and Rutter, 1989 and
As for calcite, activation energies for second-phase grain growth Urai et al., 1990, De = De Bresser, 1991, Fe = Fernández et al., 2004, He = Heitzmann, 1987,
can be derived from the slope in log(dp) vs. 1/T diagrams (Herwegh Hw = Herwegh & Pfiffner, 2005, Ke = Kennedy and Logan, 1997 and Kennedy and White,
et al., 2005a). To compare similar second-phase microstructures, 2001, Ma = Mas and Crowley, 1996 & Urai et al., 1990, Mo = Molli et al., 2000, Sc = Schenk
second-phase grain sizes dp for varying temperatures at a constant et al., 2005, Ul = Ulrich et al., 2002). Note that number-weighted grain sizes were
recalculated into area-weighted grain sizes. Only totally recrystallized mylonites that
volume fraction of fp = 0.3 were selected. This fp value was chosen to were not affected by retrograde strain localization, annealing and/or second-phase
minimize the error, because there the data density of analyzed influence were incorporated into this study. Moreover, only data with well-estimated
samples is the highest. Note, for other fp, the relationship persists but temperatures were included.
136 A. Ebert et al. / Tectonophysics 457 (2008) 128–142

of high temperature data from literature in Fig. 7 can be explained by consumption cycles of second phases described above, dropped
different strain rates or by an effect of annealing. Latter ones can be second phases continuously become released from the calcite interior,
excluded as Urai et al. (1990) redefined the microfabrics of Covey- mainly by grain boundary migration. Therefore, small-sized second
Crump and Rutter (1989) and Mas and Crowely (1996) into phases are always abundant along the calcite grain boundaries.
deformation fabrics and as the fabrics of Ulrich et al. (2002) and However, with increasing grain boundary mobility, calcite grains are
Molli et al. (2000) clearly indicate deformation microstructures. more readily able to overgrow a larger number of second-phase
Additionally, strain rates are similar to those in the Helvetic realm. particles. Moreover, the lower the volume fraction of second phases,
Heitzmann (1987) and Ulrich et al. (2002) estimated strain rates of the larger the inter-particle distances; a fact already manifest by the
10− 12–10− 13 s− 1, Kennedy and Logan (1997) 10− 11 s− 1, and shear zone Zener parameter Z. Due to larger transport distances, mass transfer is
descriptions of Bestmann et al. (2000) and Molli et al. (2000) should reduced resulting in slower growth rates of second phases and
give similar strain rates to those of the Helvetic nappes. Despite therefore in smaller grain sizes of the second phases (see also Fig. 4).
potential differences in some physical parameters (e.g., strain rate) This in turn yields lower pinning forces that act on calcite grain
between these data and the Helvetic data, the coincidence of the data boundaries. Therefore, a higher percentage of second-phase particles
points is appealing and the existence of a non-linear trend is further become included within calcite grains if their volume fractions are low
supported. The power law fit derived from averaging over the entire (Fig. 8). This implies that a considerable amount of second phases are
Helvetic data would fit also with the high temperature data from disconnected from the nutrients required to grow, which, in
literature but yields too large grain sizes for temperatures below combination with larger mass transport distances, further reduces
250 °C (Fig. 7). For these reasons, the assumption of a single linear fit the average growth rate of second phases.
to the log(Dcc) vs. 1/T data can be excluded. The influence of the growing second phases on simultaneously
(b) Alternatively, a low and high temperature relationship can be coarsening calcite grains can be seen in Fig. 9 where, for constant
assumed consisting of two different slopes and a rather abrupt second-phase volume fractions fp, the grain growth trends of second
transition between 300 and 350 °C (Fig. 7). With integrated data from phases and calcite were calculated. It is evident that second phases
literature, above 300–350 °C the activation energies Qcc yield 60 ± 8 grow with increasing temperature (Fig. 6). Hence, for a constant fp, the
and 80 ± 8 kJ/mol for impure and pure microfabrics, respectively. At number of particles must decrease and the inter-particle distances
lower temperatures, Qcc reveal values of 32 ± 3 (impure) and 40 ± 3 kJ/ between second phases must increase simultaneously. This in turn
mol (pure). Therefore, different processes in high and low tempera- allows calcite grains to grow further until their grain boundaries
ture as well as in pure and impure samples, respectively, may have become pinned again by the now larger second phases.
taken place, i.e. were rate limiting. Thus, temperature/stress and In case of grain coarsening maps of Fig. 9, averaged values derived
second-phase pinning are the most relevant parameters to affect from the whole data set of all nappes were used (see Section 4) to
microstructure modifying processes. These processes are stored determine (i) the temperature effect (dashed lines), (ii) the grain
deformation energy driven grain boundary migration and subgrain coarsening trends at constant fp (oblique vertical lines), and (iii) the
rotation recrystallization as well as surface energy driven grain boundary between second-phase and recrystallization controlled field.
boundary migration. While dynamic recovery of the microfabric is The effect of different calcite activation energies Qcc becomes visible by
provoked by the first two processes and results in a reduction of differences between Fig. 9a–d. Note that the shift and/or change in slope
dislocations, the latter process is responsible for grain growth. In light of temperature and fp trends due to variations in m, n and Qp (see error
of the non-liner log(Dcc) vs. 1/T relation, microstructures below 300– bars in Fig. 9a) are distinctly lower than caused by different Qcc.
350 °C controlled by processes with lower activation energies reflect Therefore, we focus in the following on the effect of Qcc on coupled grain
microfabrics that are rate limited by grain growth, while above 300– coarsening. Four grain coarsening maps were constructed for the
350 °C the grain growth component is steadily increasing, forcing following conditions: (a and b) Qcc derived from the high and low
grain size reduction to be rate limiting. temperature branch of Fig. 5, respectively, (c) average Qcc obtained by a
(c) Rather than only one rate limiting active process at high and linear regression over the entire data set, and (d) a combination of low
low temperature, respectively, simultaneous activity of both processes and high temperature Qcc for the corresponding temperature ranges
over the whole temperature range can be considered as well. In this below and above 325 °C from Fig. 7. (a) In Fig. 9a, the calculated grain
case, the presumed transition in the log(Dcc) vs. 1/T diagram can be coarsening map yields too small grain sizes Dcc for 300 °C and slightly
interpreted as a gradual change in predominance between these two
rate limiting processes.
Second-phase pinning also influences the deformation and growth
processes at constant temperature. While in pure samples the
microstructure is affected by the competition between grain growth
and grain size reducing processes only, dragging of grain boundaries
impedes an additional influence in case of impure microfabrics.
Nonetheless, also in second-phase controlled microstructures the
non-linear change in the log(Dcc)log(Dcc) vs. 1/T relation occurs (Fig. 5).
The estimated activation energies from impure microstructures are
lower than in pure ones. To which processes these activation energies
can be related is yet unclear, because the microstructural changes with
temperature in Fig. 5 do not indicate, whether pinning, growth of
calcite or second phases, diffusion or mass transfer along grain
boundaries, grain size reduction or other processes are rate limiting.
Since both the matrix grains and second phases change their
microstructures with increasing temperature, grain growth of both
phases has to occur simultaneously, which requires a close interaction Fig. 8. The relation between included (ndropped) and all (nall) second phases is given in
of the processes involved in this coupled grain coarsening behavior dependence of volume fraction of second phases fp and temperature (different
symbols); the Diablerets nappe as an example. Symbols represent microstructures of
(Figs. 8 and 9). With increasing temperature, the grain boundary locations 2–5 (samples from location 1 are characterized by too less included second
mobility of calcite increases (e.g., Evans et al., 2001) and so does the phases). The relative number of included second phases increases with lower volume
ability to drop a second-phase particle. Due to the growth and fractions and larger calcite grain sizes, i.e. with rising temperature.
A. Ebert et al. / Tectonophysics 457 (2008) 128–142 137

Fig. 9. Grain coarsening maps of calcite in second-phase (light grey) and recrystallization (white area) controlled microstructures derived from Eq. (5) with mean values of activation
energies of second phases Qp, exponents m and n, and constants c′ and k′ derived from averaged values from all 4 Helvetic nappes. Variations in activation energies of calcite Qcc
result in different grain coarsening maps (panels a–d), see discussion. The letters for Qcc in grey and white above each map represent Qcc values for second-phase and recrystallization
controlled microstructures, respectively. The change of Dcc at constant volume fractions fp of second phases is represented by oblique lines. In case of diagram (d) trends ≥350 °C and
b 350 °C are calculated with Qcc determined from high and low temperature data of Fig. 7, respectively. Dark-grey fields reflect areas with fpN0.5, for which equation 5 is not valid
anymore.

too high Dcc at 400 °C. (b) In case of low temperature Qcc (Fig. 9b), the low feldspar occur at temperatures above 700–800 °C (Zhu et al., 1994;
temperature trends fit well with the natural data set, but the grain sizes Mazzucato and Gualtieri, 2000; Cultrone et al., 2001).
at higher temperatures are much too small (compare with Fig. 3). Both,
low and high temperature Qcc, give good agreements with the nature 5.3. Comparison with experiments
only for either low or high temperatures, respectively. (c and d)
Coarsening maps for both scenarios fit very well with the natural data Note that the activation energies Q estimated in the aforemen-
set of the Helvetic nappes. However, the use of different Qcc for low and tioned manner are solely based on the quantified microstructural
high temperatures, respectively, yields slightly better fits for tempera- changes. Therefore, these values represent estimates of the activation
tures of 400 and 250 °C than is the case for the linear regression over the processes responsible for the development of deformation micro-
entire data set. Thus, the calculated grain coarsening maps further structures. This approach is in contrast to grain growth and rock
support the hypothesis of at least two processes with different activation deformation experiments, where conventionally activation energies
energies that were actively involved in the preservation of steady state for processes affecting the mechanical response of a rock are
microfabrics. Moreover, Qcc must differ between pure and impure determined. For that reason, absolute values of activation energies
samples, otherwise the boundary between second-phase and recrys- derived from these two different approaches cannot be directly
tallization controlled fields would be vertical at a constant Z value. In compared. However, we can (1) compare relative changes between
contrast to Herwegh et al. (2005a), grain growth trends at constant fp in experimental and natural data sets or (2) recalculate microstructure
the maps of Fig. 9c and d do not cross the boundary between the second- based activation energies for the experimentally obtained micro-
phase and recrystallization controlled field, indicating that the boundary structures in the same way as done for the ones of this study.
between the two fields is controlled only by second-phase volume
fraction as small as 2 vol.%. (1) Generally, experimental studies yield larger Q estimates than
The coupled grain coarsening trends in our maps are only valid if result from natural studies. In case of static grain growth in
no mineral reactions occur. In case of calcite and second phases like experiments, values tend to be between 100 and 230 kJ/mol (Evans
sheet silicates and quartz, reactions to wollastonite, gehlenite and et al., 2001 and references therein). Q estimations for diffusion and
138 A. Ebert et al. / Tectonophysics 457 (2008) 128–142

nappes were observed, although the fluid history might have been
different (see above). Variations in Q between other field observations
and our study may be explained either by effects like second-phase
pinning, retrograde strain localization, or annealing. Furthermore,
compared to strain rates in experiments, which are several orders of
magnitude faster and shift the relationship log(Dcc) vs. 1/T to the left
(Fig. 10), strain rate variations between 10− 11 to 10− 13 s− 1 between
different shear zones of this study have a subordinate influence only
and range in the same error as the data scatter itself. However, the
shift of experimental data in Fig. 10 clearly indicate that strain rate has
an influence on the microstructure but its variations between natural
shear zones are too small to result in distinct different microstruc-
tures. Thus, the effect of strain rate has to be considered if lab data are
extrapolated to nature.

6. General discussion

Fig. 10. Change of mean calcite grain size Dcc with temperature for pure microfabrics of
natural samples (right side) and experiments (left side). Grain sizes of experiments
6.1. The Zener parameter revisited
were recalculated into area-weighted grain sizes. Symbols and references of natural
microstructures see Fig. 7. References for data points of experimental studies are given In the past couple of years, various studies revealed that the Zener
in the diagram. Note that an increase in strain rate induces a shift of the data scatter to relation represents a useful way to describe microfabrics of poly-
the left. Activation energies of low and high temperature microstructures are similar in
mineralic rocks that consist of a predominant matrix phase (e.g., Mas
experiments and nature, respectively.
and Crowely, 1996; Krabbendam et al., 2003; Herwegh and Berger,
2004; Herwegh et al., 2005a,b). The original version of the Zener
parameter Z was in the form of Z = dp/fp (Zener in Smith, 1948). Zener
dislocation creep represent larger values of 200–300 kJ/mol and
and others (see references in Olgaard and Evans, 1986) presumed
200–500 kJ/mol, respectively (De Bresser et al., 2002 references
random dispersions of second phases in the poly-phase aggregates. If
therein; Herwegh et al., 2003; Freund et al., 2004). For natural
second phases are dispersed along grain boundaries and grain triple
samples, Q values of 30–50 kJ/mol were modeled for static grain
junctions, the stabilized matrix grain size is proportional to dp/fpm,
growth (Herwegh and Berger, 2003), while 78–125 kJ/mol were
where m = 1/2 or 1/3, respectively (e.g., Olgaard and Evans, 1986; Fan
obtained for deformation processes in natural mylonites (De
et al., 1998). In our study a similar relation to Mas and Crowely (1996)
Bresser et al., 2002). Despite these differences in the absolute
with Dcc/dp = 1.5 ⁎ f−p 1/3 (Fig. 11) was obtained for second-phase
estimates between nature and experiment, the general tendency
controlled microstructures suggesting that the second phases are
of larger activation energies for deformation than for static
predominantly distributed along grain triple junctions. In our case,
processes seems to hold in both cases.
however, the observed microstructures indicate rather non-systema-
(2) It is appealing to see that deformation fabrics of Carrara marbles
tic distributions of second phases because they are situated within
deformed at different temperatures but similar strain rates (Pieri
calcite grains, along grain boundaries, at four-grain junctions, and in
et al., 2001; Barnhoorn et al., 2004) indicate a variation in
case of large second-phase grains they even occur along grain
microstructure based activation energies between high
boundaries of several calcite grains (Herwegh and Berger, 2004,
(75 kJ/mol) and low temperature (44 kJ/mol) fabrics yielding
Ebert et al., 2007b). A discrimination into second phases located at
similar values as obtained for the natural samples (Fig.10). In terms
specific locations in the microfabric becomes therefore rather subjective.
of rheology, such dependence between temperature/stress and Q
Following Zener's original definition, we therefore defined m to be
was already suggested by Renner et al. (2002). These authors
1, i.e. we excluded a specification of second-phase locations in the
attributed changing Q values to variations in the resistance of a
crystal to deformation due to the so-called back stress. However, it
remained unclear how exactly Q was changing, an information
which might now be obtained from the microstructure. In terms of
composite flow (see above, Herwegh et al., 2005b), the tempera-
ture/stress dependent changes in the simultaneously active
dislocation and diffusion creep deformation mechanisms would
also result in a change in ‘bulk’ activation energy. To be conclusive
with this respect, however, more high-strain experiments per-
formed over a wide range of strain rate and temperature
conditions are required. Nonetheless, there still remain differences
caused by a variety of different settings in experiments and nature
probably affecting the competition of different deformation
processes (e.g., higher temperatures, faster strain rates, different
impurities and/or deviations in fluid contents).
The effects of second phases, chemical impurities, and dry or wet
deformation conditions affect microstructures in experiments as well
as in nature (e.g., De Bresser et al., 2002; Herwegh et al., 2003;
Herwegh and Berger, 2004; Freund et al., 2004). De Bresser et al.
(2005) suggest that the effect of water in experiments seems not to Fig. 11. Change of the relationship mean calcite grain size Dcc / second phase grain size
dp versus volume fraction of second phases fp for different temperature conditions
have a large influence at low Q values similar to this study, even if the (represented by the symbols and the location numbers in the symbols) of the Morcles
occurrence of fluids would enhance diffusion processes. This agrees nappe for second-phase controlled microfabrics ( fp N 1%). The best fits for all locations
with our study, where no microstructural differences between the are given as well as the mean fit derived from all best fits.
A. Ebert et al. / Tectonophysics 457 (2008) 128–142 139

microfabrics. This has the advantage that in the Z versus Dcc space, fluid flows and strain rates have caused drastic changes in grain
microstructural data are spread over wider ranges of Z values, while they size can be excluded. Despite differences in the fluid flux, this
collapse to a narrow Z range for m≪1. As a consequence, a clear either suggests that (1) in an incremental point of view, always
discrimination between second-phase and recrystallization controlled a similar amount of fluid was present in all nappes, (2) the time
microstructures becomes possible allowing to detect the critical volume integrated fluid flux documented by stable isotopes is not
fraction of the second phases required to dominate the calcite microfabric. related to the high temperature history of the shear zones but
Furthermore, temperature induced variations are more clearly visible. reflects a late deformation episode related to strain localization,
or (3) the amount of fluid was not important for the formation
6.2. The role of annealing, retrograde strain localization, fluids and strain of the steady state microfabrics. At the present stage it is not
rate on calcite microstructures conclusive if the fluid flow changed with time or if the amount
of fluids was different between the nappes. However, mass
Care has to be taken in respect to cooling-induced retrograde strain transfer via grain boundary migration, diffusion or dissolution–
localization as well as annealing because both settings will change the precipitation is affected by fluids along grain boundaries and
microstructure. A considerable effect of annealing on the mean calcite therefore fluids should influence the recrystallization behavior
grain size can be neglected as the mylonites of this study reflect typical and consequently the microstructure of a rock. Concerning the
deformation microfabrics with symmetric to right-sided skewed grain tectonites investigated in this study, fact is that grain
size distributions as well as by elongated grain shapes. These boundaries were wet, perhaps even saturated and that the
characteristics are contrasting to results derived from studies on microstructures resemble between the different shear zones.
experimentally annealed microfabrics showing polygonal, isometri- If fluid saturation was different, the observation suggests that
cally shaped grains and evolutions towards left skewed grain size as long as enough fluid is present, i.e. the grain boundaries are
distributions (Heilbronner and Tullis, 2002; Barnhoorn et al., 2005). wet, the microfabric is not affected by the amount of
Furthermore, the absence of bimodal grain size distributions, higher fluid passing the rock. This is in agreement with Schenk
twin densities and smaller grain sizes allows excluding a late cooling- et al. (2005), who did not observe any microstructural
induced strain localization (Ebert et al., 2007a,b). Exceptions are few variations due to different fluid activities in natural large-
samples from the Morcles nappe that have specifically been collected scale shear zones.
to study the effect of localization on the microfabric (Fig. 4d). These
show similar relationships in the grain coarsening maps but at lower 6.2.1. Processes related to steady state grain sizes of second phases
temperatures and therefore at lower grain sizes as obtained for peak In regard of second phases, similar and temperature invariant slopes
metamorphic conditions. Consequently, all other microstructures of for all nappes suggest that only one rate limiting mechanism (Fig. 6) is
this study reflect dynamically stabilized fabrics that are controlled by controlling the second-phase microstructure over the temperature
peak metamorphic conditions and second phases (a). In addition, range investigated. The second-phase grain size dp increases linearly
(b) different fluid activities and strain rates between the Helvetic nappes with higher volume fractions fp at constant temperature, which indicates
have also to be considered to potentially affect the microfabrics as a transport controlled growth (Fig. 4). The lower fp the larger are the
proposed by different experimental studies (e.g., Rutter, 1974; Schmid inter-particle distances and the slower is the growth rate caused by a
et al., 1977, 1987; Walker et al., 1990; De Bresser et al., 2005). reduced mass transfer rate along the calcite grain boundaries between
the second-phase particles. At this point, it is important to note that the
(a) The influence of temperature and second phases on the calcite grain coarsening of second phases observed in this study cannot solely
microfabric was already discussed in detail by Herwegh and be related to a simple Ostwald ripening and/or a coalescence of particles
Kunze (2002), Herwegh and Berger (2004) and Ebert et al. as proposed by various authors (e.g., Eberl et al. 1990; Alexander et al.,
(2007b). Their assumptions and conclusions can be transferred 1994; Chen and Fan, 1996; Fan and Chen, 1997). Grain size distributions,
to all analyzed microstructures of this study (Fig. 3). The calcite grain shapes, and grain orientations rather reflect a cyclical behavior of
microstructures can be subdivided into second-phase and nucleation, grain growth, deformation and dissolution, where nowadays
recrystallization controlled fabrics (Fig. 3). While in pure microfabrics just reflect a frozen stage of a rather dynamic behavior
samples, cycles of recrystallization lead to a constant steady (discussed in detail in Herwegh and Jenni, 2001 and Ebert et al., 2007b).
state grain size, in second-phase rich samples, the calcite grain This is also nicely reflected in samples undergoing retrograde strain
growth is reduced due to pinning. An increase in temperature localization. In Fig. 4d, both, included second phases and those at calcite
results in an enhanced growth rate promoting larger steady grain boundaries have reduced their grain sizes due to ongoing
state calcite grain sizes in both types of microstructures. Within deformation under retrograde conditions. The localization effect on
the errors, all data show the same log(Dcc) vs. 1/T relation for the microfabrics is further supported by the occurrence of bimodal grain
both recrystallization and second-phase controlled microfab- size distributions of second phases and calcite, as well as by higher twin
rics (Fig. 5a,b). densities with low temperature type II twins (after Burkhard, 1993),
(b) Variations in stable isotope compositions of calcite across and strongly sutured calcite grain boundaries and relicts of former larger
along individual thrusts as well as between the different calcite grains (Ebert et al., 2007b). Particularly, the observation that
Helvetic nappes indicate differences in fluid flow (Burkhard and included second phases also have reduced their grain sizes implies that
Kerrich, 1988; Marquer et al., 1994; Crespo-Blanc et al., 1995; calcite grain boundaries must have moved to release the formerly
Kirschner et al., 1999; Badertscher et al., 2002). These included particles, which in turn readapted their grain sizes to the new
inferences are further supported by a varying intensity in the extrinsic conditions of deformation. If such a retrograde adaptation
occurrence of synkinematic veins. Strain rate estimates suggest would not have occurred, at least the dropped second phases should still
variations in strain rate from 10− 11 s− 1 for the Doldenhorn represent grain sizes similar to those of non-localized samples from the
nappe (Herwegh et al., 2005b) and the high temperature part of same outcrop (location 7). Moreover, Fig. 4 indicates that annealing, i.e.
the Glarus thrust (Ebert et al., 2007a) to 10− 12 s− 1 in the Morcles second-phase grain growth after deformation can be excluded because
nappe (Ebert et al., 2007b). The errors of these strain rate otherwise a grain size jump between non and included second-phase
calculations can be as high as one order of magnitude. As grain sizes should occur. This can be explained by the fact that once
mentioned above, there exist no differences in grain size of second phases are dropped, they cannot grow anymore after deforma-
calcite and second phases at similar temperatures between the tion has stopped, which is in contrast to non-included ones. Latter ones
nappes (Figs. 5 and 6). Therefore, the assumption that different should then represent larger dp values or if Ostwald ripening has
140 A. Ebert et al. / Tectonophysics 457 (2008) 128–142

occurred a shift of grain size distributions to larger grain sizes (e.g., Eberl datascatter observed in the samples investigated. Since tectonic nappe
et al., 1990). settings are similar in thrusts all over in the mid to upper earth crust, as
indicated by similar strain rates of high temperature shear zones in the
6.3. Rheology range of 10− 11–10− 13 s− 1 (e.g., Heitzmann, 1987; Kennedy and Logan,
1997; Ulrich et al., 2002; Herwegh et al., 2005b; Ebert et al., 2007a,b), it is
The previously described temperature/stress/strain rate induced very likely that microstructures of mylonites from other similar shear
variations in microfabrics might also affect rheology. The microfabric zones show similar microstructures as shown in this study. Similarly,
features mentioned above and the occurrence of dynamic recovery fluids are mostly abundant and as concluded in this study, the amount of
processes indicate dislocation creep as dominant deformation fluids is unimportant in light of dynamic microstructural modifications
mechanism. As suggested by Ter Heege et al. (2002, 2004) and Bai as long as it is above a specific threshold. Moreover, in case of carbonate
and Raj (2005), a combination of dislocation and diffusion creep might mylonites, chemical variations of calcite as well as types of second
occur in microfabrics, where large and small grains in the same phases do not differ much in the earth crust at mid-crustal levels.
aggregate deform by these two mechanisms, respectively. As Therefore, the coarsening maps derived from 5 Helvetic nappes of this
suggested recently by numerical models of Herwegh et al. (2005b), study can be expected to be valid for other settings in the earth crust at
the contribution of the two deformation mechanisms during the similar temperature ranges located between diagenesis to greenschist-
preservation of a bulk steady state fabric may change with facies metamorphic conditions.
temperature. Compared to dislocation creep, the component of However, the settings and resulting fabrics are not directly
diffusion creep may increase with rising temperature conditions. transferable to experimental studies. Apparent different microstruc-
However, in the temperature range investigated in this study, the tures between nature and experiments are likely caused by large
contribution of diffusion creep has to remain small because of strong strain rate differences of 5–10 orders of magnitude as Fig. 10 reflects.
textures and the occurrence of dislocations in both second-phase and In spite of this, experiments also indicate that activation energies and
recrystallization controlled microfabrics (Herwegh and Berger, 2004; therefore bulk processes change with temperature/stress (Fig. 10).
Ebert et al., 2007a,b). Note that the assumption of composite flow is The grain coarsening maps provide an important base for all future
only valid, if the grains nucleated in the aforementioned cycles are investigations, where predictions about changes in polymineralic
small enough to be deformed under diffusion creep conditions. The fabrics with increasing depth are required. Potential field for the
validity of this assumption has to be tested in future approaches. application of grain coarsening maps will therefore be: numerical
Second phases did not significantly influence the rock's rheology as modeling, rock mechanics, petrology, geodynamics and geophysics.
already discussed in Ebert et al. (2007b). Impure layers seem to be
characterized by similar strength as pure ones as can be inferred from Acknowledgments
parallel bands of pure and impure layers and lacking evidences of
boudinage or fracturing of the layers with progressive strain. The We acknowledge the anonymous reviewers for their helpful and
deformation behavior of the second phases by granular flow and their constructive reviews. This manuscript also benefited from stimulating
recrystallization are represented and discussed in detail for carbonate discussions with Nick Austin and Brian Evans. We thank the Swiss
mylonites from the Doldenhorn thrust in Herwegh and Jenni (2001) National Sciences Foundation for the financial support of this project
and the Morcles thrust in Ebert et al. (2007b). by grants 21-66889.01 and 200020-103720.

7. Summary and conclusions


References
Under deformational conditions, the whole system consisting of Alexander, K.B., Becher, P.F., Waters, S.B., Bleier, A., 1994. Grain growth kinetics in alumina–
matrix mineral and second phases is dynamically active. Microstruc- zirconia (CeZTA) composites. Journal of the American Ceramic Society 77 (4), 939–946.
tures are controlled by temperature, stress, strain rate and intrinsic Austin, N., Evans, B., 2007. Paleowattmeters: a scaling relation for dynamically
recrystallized grain size. Geology 35 (4), 343–346.
parameters like the occurrence of second phases. Dependent on these Austin, N., Evans, B., Herwegh, M., Ebert, A., 2008. Strain localization in the Morcles
parameters, stabilized systems evolve, where the average grain sizes of Nappe (Helvetic Alps, Switzerland). Swiss Journal of Geosciences. doi:10.1007/
matrix mineral and second phases are coupled with each other in s00015-008-1264-2.
Badertscher, N.P., Abart, R., Burkhard, M., McCaig, A., 2002. Fluid flow pathways along
dependence on the volume fraction of the phases. As Figs. 9 and 10
the Glarus overthrust derived from stable and Sr-isotope patterns. American
display, each microstructure represents specific deformation conditions Journal of Science 302 (June), 517–547.
and can therefore be used in future studies to determine physical Bai, J., Raj, R., 2005. Influence of grain size variability on the strain rate dependence of
parameters like temperature or microstructural parameters describing the stress exponent in mixed-mode power law and diffusional creep. Metallurgical
and Materials Transactions A 36A (11), 2913–2919.
coupled grain coarsening. In this sense, the advantage of dynamic steady Barnhoorn, A., Bystricky, M., Burlini, L., Kunze, K., 2004. The role of recrystallisation on
state microfabrics is their time and strain invariance. Due to the fact that the deformation behaviour of calcite rocks: large strain torsion experiments on
steady state microstructures can already be reached at low strains of Carrara marble. Journal of Structural Geology 26 (5), 885–903.
Barnhoorn, A., Bystricky, M., Burlini, L., Kunze, K., 2005. Post-deformational annealing of
γ = 5–10, microstructures of high-strain shear zones will reflect specific calcite rocks. Tectonophysics 403 (1–4), 167–191.
extrinsic and intrinsic conditions under which the system was evolved. Berger, A., Herwegh, M., 2004. Grain coarsening in contact metamorphic carbonates:
The large data set of this study combined with data from other tectonic effects of second-phase particles, fluid flow and thermal perturbations. Journal of
Metamorphic Geology 22 (5), 459–474.
settings show that it is possible to predict grain coarsening in stabilized Berger, A., Roselle, G., 2001. Crystallization processes in migmatites. American
deformation microfabrics of polymineralic carbonate systems. There- Mineralogist 86 (3), 215–224.
fore, our grain coarsening map is a useful tool for field geologists, Bestmann, M., Kunze, K., Matthews, A., 2000. Evolution of a calcite marble shear zone
complex on Thassos Island, Greece: microstructural and textural fabrics and their
modelers, and material scientists (ceramists and metallurgists), since
kinematic significance. Journal of Structural Geology 22 (11–12), 1789–1807.
microstructures behave similar in solid materials during hot working Burkhard, M., Goy-Eggenberger, D., 2001. Near vertical iso-illite-crystallinity surfaces
and high temperature annealing. Although differences persist in cross-cut the recumbent fold structures of the Morcles nappe, Swiss Alps. Clay
Minerals 36 (2), 159–170.
synkinematic fluid flow, strain rate, mineral compositions, shear zone
Burkhard, M., Kerrich, R., 1988. Fluid regimes in the deformation of the Helvetic nappes,
geometry and large-scale tectonic structures between different nappes Switzerland, as inferred from stable isotope data. Contributions to Mineralogy and
studied in this project, relationships between matrix grain size of calcite, Petrology 99 (4), 416–429.
second-phase influence and temperature/stress are similar. Thus, the Burkhard, M., 1988. L'Helvétique de la bordure occidentale du massif de l'Aar (évolution
tectonique et métamorphique). Eclogae Geologicae Helvetiae 81 (1), 63–114.
variations in fluid flow, strain rate and rock types between shear zones Burkhard, M., 1990. Ductile deformation in micritic limestones naturally deformed at
at the best induce variations in microstructures responsible for the low temperatures (150–350 °C). In: Knipe, R.J., Rutter, E.H. (Eds.), Deformation
A. Ebert et al. / Tectonophysics 457 (2008) 128–142 141

Mechanisms, Rheology and Tectonics, vol. 54. Geological Society, London, Special axial compression and shear. In: de Meer, S., Drury Martyn, R., de Bresser, J.H.P.,
Publications, pp. 241–257. Pennock Gill, M. (Eds.), Deformation Mechanisms, Rheology and Tectonics; Current
Burkhard, M., 1993. Calcite twins, their geometry, appearance and significance as stress- Status and Future Perspectives, vol. 200. Geological Society, London, Special
strain markers and indicators of tectonic regime: a review. Journal of Structural Publications, pp. 191–218.
Geology 15 (3–5), 351–368. Heitzmann, P.,1987. Calcite mylonites in the Central Alpine “zone”. Tectonophysics 135 (1–3),
Burkhard, M., Kerrich, R., Maas, R., Fyfe, W.S., 1992. Stable and Sr-isotope evidence for 207–215.
fluid advection during thrusting of the Glarus nappe (Swiss Alps). Contributions to Herwegh, M., Berger, A., 2003. Differences in grain growth of calcite: a field-based
Mineralogy and Petrology 112 (2 – 3), 293–311. modeling approach. Contributions to Mineralogy and Petrology, 145, 600–611.
Chen, L.-Q., Fan, D., 1996. Computer simulation model for coupled grain growth and Herwegh, M., Berger, A., 2004. Deformation mechanisms in second-phase affected
Ostwald ripening—application to Al2O3–ZrO2 two-phase systems. Journal of the microstructures and their energy balance. Journal of Structural Geology 26 (8),
American Ceramic Society 79 (5), 1163–1168. 1483–1498.
Covey-Crump, S.J., Rutter, E.H., 1989. Thermally-induced grain growth of calcite marbles Herwegh, M., Handy, M.R., 1996. The evolution of high-temperature mylonitic
on Naxos Island, Greece. Contributions to Mineralogy and Petrology (Historical microfabrics: Evidence from simple shearing of a quartz analogue (norcamphor).
Archive) 101 (1), 69–86. Journal of Structural Geology 18 (5), 689–710.
Crespo-Blanc, A., Masson, H., Sharp, Z., Cosca, M., Hunziker, J., 1995. A stable and 40Ar/39Ar Herwegh, M., Jenni, A., 2001. Granular flow in polymineralic rocks bearing sheet
isotope study of a major thrust in the Helvetic nappes (Swiss Alps): evidence for fluid flow silicates: new evidence from natural examples. Tectonophysics 332 (3), 309–320.
and constraints on nappe kinematics. Geological Society of America Bulletin 107 (10), Herwegh, M., Kunze, K., 2002. The influence of nano-scale second-phase particles on
1129–1144. deformation of fine grained calcite mylonites. Journal of Structural Geology 24 (9),
Cultrone, G., Rodriguez-Navarro, C., Sebastian, E., Cazalla, O., De La Torre, M.J., 2001. 1463–1478.
Carbonate and silicate phase reactions during ceramic firing. European Journal of Herwegh, M., Pfiffner, O.A., 2005. Tectono-metamorphic evolution of a nappe stack: a
Mineralogy 13 (3), 624–634. case study of the Swiss Alps. Tectonophysics 404 (1–2), 55–76.
De Bresser, J.H.P., 1991. Intracrystalline deformation of calcite. Geologica Ultraiectina. Herwegh, M., 2000. A new technique to automatically quantify microstructures of fine
PhD Thesis, Utrecht University, The Netherlands. grained carbonate mylonites: two-step etching combined with SEM imaging and
De Bresser, J.H.P., Evans, B., Renner, J., 2002. Predicting the Strength of Calcite Rocks image analysis. Journal of Structural Geology 22 (4), 391–400.
under Natural Conditions, Deformation Mechanisms, Rheology and Tectonics: Herwegh, M., Berger, A., Ebert, A., 2005a. Grain coarsening maps: a new tool to predict
Current Status and Future Perspectives, vol. 200. Geological Society, London, Special microfabric evolution of polymineralic rocks. Geology 33 (10), 801–804.
Publications, pp. 309–329. Herwegh, M., de Bresser, J.H.P., ter Heege, J.H., 2005b. Combining natural micro-
De Bresser, J.H.P., Urai, J.L., Olgaard, D.L., 2005. Effect of water on the strength and structures with composite flow laws: an improved approach for the extrapolation
microstructure of Carrara marble axially compressed at high temperature. Journal of lab data to nature. Journal of Structural Geology 27 (3), 503–521.
of Structural Geology 27 (2), 265–281. Herwegh, M., Xiao, X.H., Evans, B., 2003. The effect of dissolved magnesium on diffusion
Dietrich, D., Casey, M., 1989. A new tectonic model for the Helvetic nappes. In: Coward, creep in calcite. Earth and Planetary Science Letters 212 (3–4), 457–470.
M.P., Dietrich, D., Park, R.G. (Eds.), Conference on Alpine Tectonics. Geological Hickey, K.A., Bell, T.H., 1996. Syn-deformational grain growth: matrix coarsening during
Society Special Publications. Geological Society of London, London, United King- foliation development and regional metamorphism rather than by static annealing.
dom, pp. 47–63. European Journal of Mineralogy 8, 1351–1373.
Dietrich, D., 1989. Fold-axis parallel extension in an arcuate fold- and thrust belt: the Higgins, G.T., Wiryolukito, S., Nash, P., 1992. The kinetics of coupled phase coarsening in
case of the Helvetic nappes. Tectonophysics 170 (3–4), 183–212. two-phase structures. Materials Science Forum 94–95, 671–676.
Doherty, R.D., Hughes, D.A., Humphreys, F.J., Jonas, J.J., Jensen, D., Juul Kassner, M.E., Hillert, M., 1988. Inhibition of grain growth by second-phase particles. Acta Metallurgica
King, W.E., McNelley, T.R., McQueen, H.J., Rollett, A.D., 1997. Current issues in 36 (12), 3177–3181.
recrystallization: a review. Materials Science and Engineering A 238 (2), 219–274. Hui Zhu, Newton, R.C., Kleppa, O.J., 1994. Enthalpy of formation of wollastonite (CaSiO3)
Eberl, D.D., Srodon, J., Kralik, M., Taylor, B.E., Peterman, Z.E., 1990. Ostwald ripening of and anorthite (CaAl2Si2O8) by experimental phase equilibrium measurements and
clays and metamorphic minerals. Science, New Series 248 (4954), 474–477. high temperature solution calorimetry. American Mineralogist 79, 134–144.
Ebert, A., 2006. Microfabric evolution in pure and impure carbonate mylonites and their Joesten, R.L., 1991. Kinetics of coarsening and diffusion-controlled mineral growth.
role for strain localization in large-scale shear zones. PhD thesis, University of Bern, Reviews in Mineralogy and Geochemistry 26 (1), 507–582.
Switzerland. Kennedy, L.A., Logan, J.M., 1997. The role of veining and dissolution in the evolution of
Ebert, A., Herwegh, M., Evans, B., Pfiffner, A., Austin, N., Vennemann, T., 2007b. fine-grained mylonites: the McConnell thrust, Alberta. Journal of Structural
Microfabrics in carbonate mylonites along a large-scale shear zone (Helvetic Alps). Geology 19 (6), 785–797.
Tectonophysics 444, 1–26. Kennedy, L.A., White, J.C., 2001. Low-temperature recrystallization in calcite: mechan-
Ebert, A., Herwegh, M., Pfiffner, A., 2007a. Cooling induced strain localization in isms and consequences. Geology 29 (11), 1027–1030.
carbonate mylonites within a large-scale shear zone (Glarus thrust, Switzerland). Kirschner, D.L., Cosca, M.A., Masson, H., Hunziker, J.C., 1996. Staircase 40Ar/39Ar spectra
Journal of Structural Geology 29, 1164–1184. of fine-grained white mica: timing and duration of deformation and empirical
Ebert, A., Rieke-Zapp, D., Herwegh, M., Ramseyer, K., Gnos, E., Decrouez, D., 2007c. constraints on argon diffusion. Geology 24 (8), 747–750.
Quantitative analysis of microstructures of coarse-grained marbles: an approach Kirschner, D.L., Masson, H., Sharp, Z.D., 1999. Fluid migration through thrust faults in the
using bireflectance of calcite to visualize the microstructure. Rendiconti della Helvetic nappes (Western Swiss Alps). Contributions to Mineralogy and Petrology
Società Geological Italiana, 5, Nuova Serie, pp. 81–82. 136, 169–183.
Escher, A., Masson, H., Steck, A., 1993. Nappe geometry in the Western Swiss Alps. Kirschner, D.L., Sharp, Z.D., Masson, H., 1995. Oxygen isotopic thermometry of quartz-calcite
Journal of Structural Geology 15 (3–5), 501–509. veins: unraveling the thermal-tectonic history of the subgreenschist facies Morcles
Evans, B., Renner, J., Hirth, G., 2001. A few remarks on the kinetics of static grain growth nappe (Swiss Alps). Geological Society of America Bulletin 107 (10), 1145–1156.
in rocks. International Journal of Earth Sciences 90 (1), 88–103. Kohlstedt, D.L., Evans, B., Mackwell, S.J., 1995. Strength of lithosphere: constraints
Fan, D., Chen, L.-Q., 1997. Computer simulation of grain growth and Ostwald ripening in imposed by laboratory experiments. Journal of Geophysical Research, B, Solid Earth
alumina–zirconia two-phase composites. Journal of the American Ceramic Society and Planets 100 (B9), 17587–17602.
80 (7), 1773–1780. Krabbendam, M., Urai, J.L., van Vliet, L.J., 2003. Grain size stabilisation by dispersed
Fan, D., Chen, L.-Q., Chen, S.-P.P., 1998. Numerical simulation of Zener pinning with graphite in a high-grade quartz mylonite: an example from Naxos (Greece). Journal
growing second-phase particles. Journal of the American Ceramic Society 81 (3), of Structural Geology 25 (6), 855–866.
526–532. Leiss, B., Molli, G., 2003. ‘High-temperature’ texture in naturally deformed Carrara
Fernández, F.J., Brown, D., Álvarez-Marrón, J., Prior, D.J., Pérez-Estaún, A., 2004. marble from the Alpi Apuane, Italy. Journal of Structural Geology 25 (4),
Microstructure and lattice preferred orientation of calcite mylonites at the base of 649–658.
the southern Urals accretionary prism. Journal of the Geological Society 161 (1), 67–79. Lihou, J.C., 1996. Structure and deformational history of the Infrahelvetic flysch units,
Freund, D., Rybacki, E., Dresen, G., 2001. Effect of impurities on grain growth in synthetic Glarus Alps, eastern Switzerland. Eclogae Geologicae Helvetiae 89 (1), 439–460.
calcite aggregates. Physics and Chemistry of Minerals 28 (10), 737–745. Marquer, D., Petrucci, E., Iacumin, P., 1994. Fluid advection in shear zones: evidence from
Freund, D., Wang, Z., Rybacki, E., Dresen, G., 2004. High-temperature creep of synthetic geological and geochemical relationships in the Aiguilles Rouges Massif (Western
calcite aggregates: influence of Mn-content. Earth and Planetary Science Letters Alps, Switzerland). Schweizerische Mineralogische und Petrographische Mitteilun-
226 (3–4), 433–448. gen 74, 137–148.
Frey, M., Teichmüller, M., Teichmüller, R., Mullis, J., Künzli, B., Breitschmid, A., Gruner, U., Mas, D.L., Crowely, P.D., 1996. The effect of second-phase particles on stable grain size in
Schwizer, B., 1980. Very low-grade metamorphism in external parts of the Central regionally metamorphosed polyphase calcite marbles. Journal of Metamorphic
Alps: illite crystallinity, coal rank and fluid inclusion data. Eclogae geologicae Geology 14 (2), 155–162.
Helvetiae 73 (1), 173–203. Masson, H., Herb, R., Steck, A., 1980. Helvetic Alps of Western Switzerland, excursion
Frey, M., 1988. Discontinuous inverse metamorphic zonation, Glarus Alps, Switzerland: no. 1. In: Trümpy, R. (Ed.), Geology of Switzerland part II. Wepf, Basel, Switzerland,
evidence from illite crystallinity data. Schweizerische Mineralogische und pp. 109–153.
Petrographische Mitteilungen 68, 171–184. Mazzucato, E., Gualtieri, A.F., 2000. Wollastonite polytypes in the CaO–SiO2 system.:
Goetze, C., Kohlstedt, D.L., 1977. The dislocation structure of experimentally deformed part I. Crystallisation kinetics. Physics and Chemistry of Minerals 27 (8), 565–574.
marble. Contributions to Mineralogy and Petrology, 59, 293–306. Means, W.D., 1981. The concept of steady-state foliation. Tectonophysics 78 (1–4),
Groshong, R.H., Pfiffner, O.A., Pringle, L.R., 1984. Strain partitioning in the Helvetic thrust 179–199.
belt of eastern Switzerland from the leading edge to the internal zone. Journal of Milnes, A.G., Pfiffner, O.A., 1977. Structural development of the Infrahelvetic Complex,
Structural Geology 6 (1–2), 5–18. eastern Switzerland. Eclogae Geologicae Helvetiae 70 (1), 83–95.
Heilbronner, R., Tullis, J., 2002. The effect of static annealing on microstructures and Milnes, A.G., Pfiffner, O.A., 1980. Tectonic evolution of the Central Alps in the cross
crystallographic preferred orientations of quartzites experimentally deformed in section St. Gallen-Como. Eclogae Geologicae Helvetiae 73 (2), 619–633.
142 A. Ebert et al. / Tectonophysics 457 (2008) 128–142

Molli, G., Conti, P., Giorgetti, G., Meccheri, M., Oesterling, N., 2000. Microfabric study on Schmid, S.M., Boland, J.N., Paterson, M.S., 1977. Superplastic flow in finegrained
the deformational and thermal history of the Alpi Apuane marbles (Carrara limestone. Tectonophysics 43 (3–4), 257–291.
marbles), Italy. Journal of Structural Geology 22 (11–12), 1809–1825. Schmid, S.M., Panozzo, R., Bauer, S., 1987. Simple shear experiments on calcite rocks:
Nes, E., Ryum, N., Hunderi, O., 1985. On the Zener drag. Acta Metallurgica 33 (1), 11–22. rheology and microfabric. Journal of Structural Geology 9 (5–6), 747–778.
Olgaard, D.L., Evans, B., 1986. Effect of 2nd-phase particles on grain-growth in calcite. Schmid, S.M., Pfiffner, O.A., Froitzheim, N., Schönborn, G., Kissling, E., 1996.
Journal of the American Ceramic Society 69 (11), C272–C277. Geophysical–geological transect and tectonic evolution of the Swiss-Italian Alps.
Olgaard, D.L., 1990. The role of second phase in localizing deformation. In: Knipe, R.J., Tectonics 15 (5), 1036–1064.
Rutter, E.H. (Eds.), Deformation Mechanisms, Rheology and Tectonics. Special Siddans, A.W.B., 1983. Finite strain patterns in some Alpine nappes. Journal of Structural
Publications, vol. 54. Geological Society, London, pp. 175–181. Geology 5 (3–4), 441–448.
Paterson, M.S., 1995. A theory for granular flow accommodated by material transfer via Smith, C.S., 1948. Grains, phases, and interphases: an interpretation of microstructure.
an intergranular fluid. Tectonophysics 245 (3–4), 135–151. Transactions of the American Institute of Mining and Metallurgical Engineers 175,
Pfiffner, O.A., 1985. Displacements along thrust faults. Eclogae Geologicae Helvetiae 78 (2), 15–51.
313–333. Solomatov, V.S., El-Khozondar, R., Tikare, V., 2002. Grain size in the lower mantle:
Pfiffner, O.A., 1993. The structure of the Helvetic nappes and its relation to the constraints from numerical modeling of grain growth in two-phase systems.
mechanical stratigraphy. Journal of Structural Geology 15 (3–5), 511–521. Physics of The Earth and Planetary Interiors 129 (3–4), 265–282.
Pfiffner, O.A., Lehner, P., Heitzman, P., Mueller, S., Steck, A., 1997. Deep Structure of the Stampfli, G.M., Borel, G.D., Marchant, R., Mosar, J., 2002. Western Alps geological
Swiss Alps: Results from NFP 20. Birkhäuser, Basel. 380 pp. constraints on western Tethyan reconstructions. In: Rosenbaum, G., Lister, G.S.
Pieri, M., Burlini, L., Kunze, K., Stretton, I., Olgaard, D.L., 2001. Rheological and microstructural (Eds.), Reconstruction of the Evolution of the Alpine-Himalayan Orogen. Journal of
evolution of Carrara marble with high shear strain: results from high temperature the Virtual Explorer, vol. 8, pp. 77–106.
torsion experiments. Journal of Structural Geology 23 (9), 1393–1413. Stearns, L.C., Harmer, M.P., 1996. Particle-inhibited grain growth in AI2O3-SiC: I,
Rahn, M., Mullis, J., Erdelbrock, K., Frey, M., 1994. Very low-grade metamorphism of the experimental results. Journal of the American Ceramic Society 79 (12), 3013–3019.
Taveyanne greywacke, Glarus Alps, Switzerland. Journal of Metamorphic Geology Ter Heege, J.H., de Bresser, J.H.P., Spiers, C.J., 2002. The influence of dynamic
12, 625–641. recrystallization on the grain size distribution and rheological behavior of Carrara
Rahn, M., Mullis, J., Erdelbrock, K., Frey, M., 1995. Alpine metamorphism in the North marble deformed in axial compression. In: de Meer, S., Drury, M.R., de Bresser, J.H.P.,
Helvetic Flysch of the Glarus Alps, Switzerland. Eclogae Geologicae Helvetiae 88 (1), Pennock, G.M. (Eds.), Deformation Mechanisms, Rheology and Tectonics: Current
157–178. Status and Future Perspectives, vol. 200. Geological Society, London Special
Rahn, M., Steinmann, M., Frey, M., 2002. Chloritoid composition and formation in the Publications, pp. 331–353.
eastern Central Alps: a comparison between Penninic and Helvetic occurrences. Ter Heege, J.H., De Bresser, J.H.P., Spiers, C.J., 2004. Composite flow laws for crystalline
Schweizerische Mineralogische und Petrographische Mitteilungen 82, 409–426. materials with log-normally distributed grain size: theory and application to
Ramsay, J.G., 1981. Tectonics of the Helvetic Nappes. In: McClay, K.R., Price, N.J. (Eds.), olivine. Journal of Structural Geology 26 (9), 1693–1705.
Thrust and Nappe Tectonics. Geological Society, London, Special Publications, vol. 9, Twiss, R.J., 1977. Theory and applicability of a recrystallized grain size paleopiezometer.
293–309. Pure and Applied Geophysics 115 (1–2), 227–244.
Renner, J., Evans, B., Siddiqi, G., 2002. Dislocation creep of calcite. Journal of Geophysical Ulrich, S., Schulmann, K., Casey, M., 2002. Microstructural evolution and rheological
Research 107 (B12), 2364. doi:10.1029/2001JB001680. behaviour of marbles deformed at different crustal levels. Journal of Structural
Rutter, E.H., 1974. The influence of temperature, strain rate and interstitial water in the Geology 24 (5), 979–995.
experimental deformation of calcite rocks. Tectonophysics 22 (3–4), 311–334. Urai, J.L., Means, W.D., Lister, G.S., 1986. Dynamic recrystallization of minerals. In:
Rutter, E.H., 1995. Experimental study of the influence of stress, temperature, and strain Hobbs, B.E., Heard, H.C. (Eds.), Mineral and Rock Deformation: Laboratory Studies.
on the dynamic recrystallization of Carrara marble. Journal of Geophysical The Paterson Volume. Geophysical Monograph, vol. 36. American Geophysical
Research, B, Solid Earth and Planets 100 (B12), 24,651–24,664. Union, Washington, D.C., pp. 161–199.
Rybacki, E., Paterson, M.S., Wirth, R., Dresen, G., 2003. Rheology of calcite-quartz Urai, J.L., Schuiling, R.D., Jansen, B.H., 1990. Alpine deformation on Naxos (Greece). In:
aggregates deformed to large strain in torsion. Journal of Geophysical Research 108 Knipe, R.J., Rutter, E.H. (Eds.), Deformation Mechanisms, Rheology and Tectonics,
(B2). doi:10.1029/2002JB001833. vol. 54. Geological Society, London, Special Publications, pp. 509–522.
Ryum, N., Hunderi, O., Nes, E., 1983. On grain boundary drag from second phase Voorhees, P.W., 1992. Ostwald ripening of two-phase mixtures. Annual Review of
particles. Scripta Metallurgica 17 (11), 1281–1283. Materials Science 22 (1), 197–215.
Schenk, O., Urai, J.L., Evans, B., 2005. The effect of water on recrystallization behavior Walker, A.N., Rutter, E.H., Brodie, K.H., 1990. Experimental study of grain-size sensitive
and grain boundary morphology in calcite-observations of natural marble flow of synthetic, hot-pressed calcite rocks. In: Knipe, R.J., Rutter, E.H. (Eds.),
mylonites. Journal of Structural Geology 27 (10), 1856–1872. Deformation Mechanisms, Rheology and Tectonics. Special Publications, vol. 54.
Schmid, S.M., Casey, M., Starkey, J., 1981. The microfabric of calcite tectonites from the Geological Society of London, pp. 259–284.
Helvetic Nappes (Swiss Alps). In: Price, N.J., McClay, K.R. (Eds.), Thrust and Nappe
Tectonics, vol. 9. Geological Society, London Special Publication, pp. 151–158.
Schmid, S.M., 1975. The Glarus overthrust: field evidence and mechanical model.
Eclogae Geologicae Helvetiae 68 (2), 247–280.

You might also like