Next Article in Journal
Challenges Assessing Rock Slope Stability Using the Strength Reduction Method with the Hoek–Brown Criterion on the Example of Vals (Tyrol/Austria)
Previous Article in Journal
Changes in Unfrozen Water Contents in Warming Permafrost Soils
Previous Article in Special Issue
In-Situ Evolution of Calcite Twinning during Uniaxial Compression of Carrara Marble at Room Temperature
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Deformation of the European Plate (58-0 Ma): Evidence from Calcite Twinning Strains

1
Geology Department, Macalester College, St. Paul, MN 55105, USA
2
Institutionen för Geologiska Vetenskaper, Stockholms Universitet, 106 91 Stockholm, Sweden
3
Geologisches Institut, Universitat Bern, 3012 Bern, Switzerland
*
Author to whom correspondence should be addressed.
Submission received: 23 March 2022 / Revised: 23 May 2022 / Accepted: 7 June 2022 / Published: 20 June 2022

Abstract

:
We present a data set of calcite twinning strain results (n = 209 samples; 9919 measured calcite twins) from the internal Alpine nappes northwestward across the Alps and Alpine foreland to the older extensional margin along the Atlantic coast in Ireland. Along the coast of Northern Ireland, Cretaceous chalks and Tertiary basalts are cross-cut by calcite veins and offset by calcite-filled normal and strike-slip faults. Both Irish sample suites (n = 16 with four U-Pb vein calcite ages between 70–42 Ma) record a sub-horizontal SW-NE shortening strain with vertical extension and no strain overprint. This sub-horizontal shortening is parallel to the margin of the opening of the Atlantic Ocean (~58 Ma), and this penetrative fabric is only observed ~100 km inboard of the margin to the southeast. The younger, collisional Alpine orogen (~40 Ma) imparted a stress–strain regime dominated by SE-NW sub-horizontal shortening ~1200 km northwest from the Alps preserved in Mesozoic limestones and calcite veins (n = 32) in France, Germany and Britain. This layer-parallel shortening strain (−3.4%, 5% negative expected values) is preserved across the foreland in the plane of Alpine thrust shortening (SE-NW) along with numerous outcrop-scale contractional structures (i.e., folds, thrust faults). Calcite veins were observed in the Alpine foreland in numerous orientations and include both a SE-NW layer-parallel shortening fabric (n = 11) and a sub-vertical NE-SW vein-parallel shortening fabric (n = 4). Alpine foreland strains are compared with twinning strains from the frontal Jura Mountains (n = 9; layer-parallel shortening), the Molasse basin (n = 26; layer-parallel and layer-normal shortening), Pre-Alp nappes (n = 39; layer-parallel and layer-normal shortening), Helvetic and Penninic nappes (Penninic klippe; n = 46; layer-parallel and layer-normal shortening plus four striated U-Pb calcite vein ages ~24 Ma) and calcsilicates from the internal Tauern window (n = 4; layer-normal shortening). We provide a chronology of the stress–strain history of the European plate from 58 Ma through the Alpine orogen.

1. Introduction

Orogenic belts generally bury the adjacent foreland in synorogenic debris, making it difficult to study thrust-related deformation inboard of the orogen. Chinn and Konig [1] were the first to measure mechanical twins in calcite in a foreland setting and ponder the origin of this ‘far-field’ deformation ~100 km north of the Appalachian–Ouachita front in Arkansas, USA. Other finite strain observations in foreland settings [2,3,4,5] have confirmed the extent of thrust deformation inboard of the margin and thrust front (“toe”), and distal extent of this deformation was characterized by Craddock and van der Pluijm [6] ~2000 km into the Appalachian foreland and by Lacombe et al. [7] <100 km into the Burgundy region of the Alpine foreland (see also [8,9,10]). The collective effort of foreland strain studies increased the horizontal dimensions of a critically tapered sedimentary prism that evolves into a thrust belt [11,12,13,14] and also provides a pre-thrusting layer-parallel shortening (LPS) strain fabric that can be used as a passive strain marker in the allochthonous sediments of the adjacent thrust belt. In the collisional Appalachians, all the shortening axes are parallel to thrust motion (SE-NW), identical to the LPS fabric in the allochthonous foreland, so thrust motions did not involve rotation ([15,16,17], and references therein). In a tectonic setting of oblique convergence, as in the Sevier–Laramide belt, counterclockwise thrust sheet motions are documented when compared to the autochthonous E-W LPS fabric in the foreland [18,19,20]. The Alps present a collisional mountain belt that is curved (oroclinal), composed of high-grade nappes in its core (hinterland: [21,22,23,24]) that overthrust dated synorogenic sediments (molasse and flysch; [23,25,26,27,28]) and folded the foreland in amplitudes that diminish NW into the craton (Figure 1; Jura folds vs. the Paris basin; [29]). The influence of far-field stresses decreases into the foreland as a function of the plate boundary type, and it is dependent on whether the boundary is obliquely convergent or collisional and the slab dip that generated the stress field [30,31,32,33,34].
Figure 1. Topographic map of western Europe with regional geologic features related to Alps and the Alpine foreland, including extensional features of the European Central Rift System (ECRIS). Sample sites are listed (left column; red dots, for the foreland, the Alps (many colors, see Figure 2 and Figure 3) and the Irish Coast (orange dots). Locations of field photos in additional figures are indicated by white stars.
Figure 1. Topographic map of western Europe with regional geologic features related to Alps and the Alpine foreland, including extensional features of the European Central Rift System (ECRIS). Sample sites are listed (left column; red dots, for the foreland, the Alps (many colors, see Figure 2 and Figure 3) and the Irish Coast (orange dots). Locations of field photos in additional figures are indicated by white stars.
Geosciences 12 00254 g001
In this reconnaissance study (Figure 1), we sampled calcsilicates in the European continental crust in the Tauern window in the eastern Alps (Austria) and Mesozoic carbonates and calcite veins across the Alpine foreland from the Jura Mountains to southwestern Britain, as well as in Northern Ireland, where Cretaceous chalks are offset by normal faults with striated calcite and intruded or overlain by Tertiary basalts with calcite veins [35]. We analyzed the twinned calcite in this suite and compared our results with twinning strain and fabric studies obtained within the Alpine orogen: the Jura folds, [36,37], the Molasse basin, [36,38], the Pre-Alpine nappes [39], the Helvetic nappes [40,41,42] and the Helvetic–Penninic nappes [40,43]. With this robust data set, we hope to (1) interpret the variations in stress–strain response of the oroclinal Tertiary Alpine orogen across the region into the gently warped foreland (i.e., Paris basin, Purbeck thrust-monocline, London basin, etc.) toward a better understanding of thrust mechanics and (2) couple the twinning strain results associated with Pre-Alpine intra-plate deformation [44,45,46] and the opening of the North Atlantic (58 Ma), Tertiary Alpine deformation and the modern Icelandic margin [47].
Figure 2. Regional maps of the Alps with cross section line A-A’ shown in Figure 3. (A) Digital elevation model with the main drainage divides delimiting the hydrographic basins of the Rhine, Po and Danube rivers. (B) Tectonic map of the Alps and their foreland showing the main nappe systems and foreland basins.
Figure 2. Regional maps of the Alps with cross section line A-A’ shown in Figure 3. (A) Digital elevation model with the main drainage divides delimiting the hydrographic basins of the Rhine, Po and Danube rivers. (B) Tectonic map of the Alps and their foreland showing the main nappe systems and foreland basins.
Geosciences 12 00254 g002
Figure 3. Cross section (SE-NW) through the Alps from the Po Basin to the Bresse graben (see Figure 2 for the section line). PAF: Peri-Adriatic fault system.
Figure 3. Cross section (SE-NW) through the Alps from the Po Basin to the Bresse graben (see Figure 2 for the section line). PAF: Peri-Adriatic fault system.
Geosciences 12 00254 g003

2. Previous Work

Central Europe was the site of widespread Triassic rifting followed by regional Triassic–Cretaceous marine sedimentation and complex accretion of the Iberian, Carnian and Apulian terranes [48,49]. Inversion of Permo-Mesozoic basement faults has been recognized in the Alps and the Alp foreland by Kley and Voight [44] as having been active in northern Europe in the Cretaceous. The Alps formed in the mid Cretaceous (ca. 100 Ma) and progressed from the Austroalpine in the Eastern Alps (as part of the Adriatic Plate) towards the west and north; the Central Swiss Alps were tectonized from the Paleocene through the Miocene. The inversion structures in northern Europe are the result of SSW-NNE shortening that ended by the Maastrichtian (72 Ma), a result that is in contrast to the Alpine–Carpathian orogen model proposed by Zeigler et al. ([50] and references therein). Opening of the Atlantic Ocean (58 Ma) coincided with the progradation of the Alpine orogen into the Central Swiss Alps and major northwest-directed nappe imbrication [51]. The Alps deformation, as recorded by offset synorogenic flysch deposits, apatite and zircon uplift ages and radiometric dates on metamorphic minerals, reveals a progression of orogenesis from the high-grade internal nappes in the northwest to the Jura Mountain folds and foreland that youngs toward the craton (NW; [23,26,52,53,54,55,56,57]). Palinspastic restorations indicate the Alps have been shortened ~60%, including folds, in the foreland (Figure 1, Figure 2 and Figure 3), with significant transport-normal displacement on the Simplon–Rhone [58] and Insubric strike-slip systems [29,59,60], and were originally ~300–350 km wide. Alpine convergence also initiated a variety of Eocene extensional features in the foreland known as the European Cenozoic Rift System (ECRIS; [45,61,62,63]), which are unusual grabens not predicted by models of orogenic convergence where extensional features are observed within the high-grade internal portions of orogenic belts [64,65,66,67]. Evidence of extensional deformation in the Alpine nappes is poor, and the Proterozoic Keweenaw Rift–Grenville orogen provides one analog with foreland extension and rifting during a collisional orogen ([17] and references therein).
We provide a compilation of new (Irish coast, Alp foreland, Tauern window) and old (Jura, Molasse Basin, Helvetic and Penninic nappes) calcite twinning strain data across the Alps and the foreland to the northwest. Calcite, of course, is not the only means to record deformation: Siddans [68] reports a Rf/ϕ study of reduction spots in Permian mudstones from the basal Helvetic nappes along the Glarus thrust, where oblate ellipsoids (3:1.5:1 aspect ratios) have long axes oriented parallel to thrust transport and 20–30° to the thrust plane. This is a similar result to belemnite strain data in the Helvetic nappes reported by Ramsay [69] and Ramsay and Huber [70] and a fold-strain study by Pfiffner [71]. Studies of solution mass transfer (pressure solution) along the Glarus thrust reveal a volume loss of 36% of the hanging wall sediments [54,72]. Each of these strain results complements the twinning strain observations of Groshong et al. [73], namely the dominance of layer-normal shortening (LNS), in the Helvetic nappes, but they also reveal the value of using numerous techniques to measure strain [strain partitioning; [74]. Calcite and quartz tectonites have been studied in the Tauern window and the Penninic nappes [43,75,76] where calcitic pseudotachylite is found [75]. In the Alpine foreland, Lacombe et al. [7] first used a calcite twin paleostress (not strain) method to propose a complex tectonic history adjacent to the Jura Mountains, an approach similar to proxy structural methods (e.g., FAULTKIN, see [31,44,46,77,78,79,80]) and Ring and Gerdes ([45]; with U-Pb calcite ages) to interpret ~84 Ma of tectonic history in the European plate as Iberia accreted, the Atlantic Ocean opened, the Alps developed, and the ECRIS rift system evolved [81,82].

3. Methods

3.1. Calcite Twin Analysis

The calcite strain-gage technique (CSGT) of Groshong [83] uses intracrystalline twinning of rock-forming calcite grains to derive a three-dimensional orientation of the stress and strain ellipsoids during twinning. Although the result is a strain tensor, a similar orientation of the stress tensor is calculated in the case of coaxial deformation [84,85] he CSGT has been used to constrain three-dimensional stress and strain tensor directions in calcite veins [6,16,86], limestone [6,39,87,88,89,90,91,92,93,94,95,96], marble [97], amygdaloidal basalt [47,98,99] calcite ocelli in lamprophyres [100] and fault gouge where calcite is dated with U–Pb methods [101,102,103]. Lacombe et al. [104]; this volume have reviewed the history and methods of calcite twin analysis.
Under temperatures of ca. 200 °C, intracrystalline deformation of calcite results in the formation of e-twins. The formation of calcite e-twins requires a shear stress exceeding ca. 10 MPa [105,106,107,108]. Calcite offers three glide systems for e-twinning. From U-stage measurements of width, frequency and orientation of twins, along with the crystallographic orientation of the host crystals, a strain tensor can be calculated using a least-squares technique [83]. To remove “noise” from the dataset, a refinement of the calculated strain tensor can be achieved by stripping 20% of the twins with the highest deviations [42]. This procedure has been used if the number of measured grains was large (n > 20). In cases where the data appear to be inhomogeneous, the separation of incompatible twins (“NEV” = negative expected values) from compatible twins (“PEV” = positive expected values) of the initial dataset allows for separate calculations of two or more least-squares deviatoric strain tensors. Thus, the CSGT can be used to obtain information on superimposed deformations [83,109] and differential stress magnitudes [110]. The validity of this stripping procedure was demonstrated in experimental tests where the reliability depends on the overall complexity of deformation and the number of grains with twins [109,111]. The stripping procedure was used in cases of high proportions of NEVs and many measured grains. An experimental re-evaluation of the CSGT has shown that measurements of about 50 grains on one thin section or 25 grains on two mutually perpendicular thin sections yield the best results [42,112,113] The chance to extract the records of more than two deformations from one dataset is limited when dealing with natural rocks [106]. Individual analyses of veins, matrix, nodules, etc., allow the acquisition of several strain tensors without applying statistical data stripping. Research on the complexity of rotational strains in fault zones is limited to the efforts of Gray et al. [114] and Craddock et al. [115], although the technique is more robust now that calcite can be dated using U-Pb methods [101]. Application of the CSGT requires the following assumptions to be valid: (1) low temperatures (dominance of Type I and Type II twins), (2) random c-axis orientations of calcite, (3) homogenous strain, (4) coaxial deformation, (5) volume constancy, (6) low-porosity materials and (7) low bulk strain (<15%). If these conditions are not fully met, calculated strain tensor datasets could be biased, modified or random. Strain tensors were calculated from calcite e-twin datasets using the software package of Evans and Groshong [112]. Fabric interpretations are based on the orientation of the shortening axis (e1), which usually plots near the contoured maxima of the Turner [84] compression axes, and if e1 is ~20° from the sample bedding great circle, layer-parallel shortening (LPS) is a valid fabric interpretation. Layer-normal shortening (LNS) is the fabric where e1 is at a high angle (>45°) to the bedding great circle; vein-normal (VNS) or vein-parallel (VPS) shortening are additional potential fabric interpretations (Table 2).

3.2. Calcite U-Pb Geochronology with LA-ICPMS-MC

Samples from Northern Ireland were examined under plane, polarized and cathodoluminescence (CL) microscopy to identify distinct precipitation phases yielding ages for veins PUE-17, 21, 23 and 26 (Table 1). U-Pb analyses of calcite were conducted at the University of California, Santa Barbara, following the protocol of Nuriel et al. [101], and are found in Appendix B.

4. Results

The opening of the Atlantic Ocean between Europe and Greenland (58 Ma; magnetochron 24) preceded the collision of Apulia into southern Europe that formed the Alps ~30 Ma, so our results are presented from oldest to youngest (Figure 1, Table 1 and Table 2). Alpine geology is presented in Figure 2 and Figure 3 (cross section). Field photos can be found in Figure 4, Figure 5 and Figure 6. Stereographic plots of twinning strain data can be found in Figure 4 (N. Ireland), Figure 7 and Figure 8 (Alps and foreland), in Figure 9 and Figure 10 (regional plots) and in Figure 11 (anomalous results). These results are compared with additional strain studies from the Alps (Table 2, see Appendix A for field locations).

4.1. Irish Tertiary Province

In central Northern Ireland, near Moneymore, a number of quarries expose interlayered Cretaceous chalks and limestones and Tertiary (70–42 Ma) basalt sills, flows and crosscutting dikes. Calcite veins are present throughout, and normal faults with calcite fillings (various strikes, steep northerly dips and striated calcite steps: top-down-to-the-north) are common in the limestones. Four calcite veins have U-Pb ages between 70–42 Ma, similar in age to calcite vein fillings along the Atlantic margin in the Faroe Islands (Figure 4; [116]). Along the northern coast, between Port Rush and Ballycastle, similar basalt–limestone contacts are exposed, but more pervasive structures are present; at Ballintoy Harbor, orthogonal (N-S, 90° and E-W, 90°) faults are present, and both preserve sinistral offsets (Figure 4). Calcite veins are also thicker and more pervasive here, and a normal fault hanging wall anticline is present with calcite fillings associated with the normal fault offset. All thirteen samples (n = 202 grains) from 3 different locations preserve the same calcite strain fabric: SE-NW sub-horizontal shortening and vertical extension that is parallel to the Atlantic margin opened by SE-NW extension. The calcite preserves modest strains (average shortening −3.1%), with no suggestion of any strain overprint (average NEVs 8%) and a modest differential stress associated with the opening of the Atlantic Ocean (average −387 bars, Table 1 and Table 2). The highest strain magnitude (−6.3%) is at the coast, and the lowest (−1.7%) is inland near Moneymore, a distance of ~50 km. Fabric interpretation is based on the orientation of the maximum shortening axis (e1) relative to bedding, and all 16 samples preserve a layer-parallel shortening (LPS) strain where e1 is +/− 20° from bedding. Conversely, as all the samples are veins or fault fillings, the orientation of e1 relative to the vein-filling provides an additional fabric insight: 4 VPS (vein-parallel shortening) and 12 VNS (vein-normal shortening) fabrics are preserved. Interestingly, samples 7, 8, and 11–14, which have sinistral and normal kinematics, respectively, all preserve the same NE-SW margin-parallel, sub-horizontal shortening.

4.2. Internal Alps

Alpine exhumation has exposed the Austroalpine Variscan continental crust to the south of the Tauern Window, the Austrian Alps (Figure 1, Figure 2 and Figure 3). Variscan ductile deformational is characterized by garnet–staurolite and kyanite-bearing biotite–plagioclase paragneisses that are mainly the result of non-coaxial progressive shearing D1-D2 with isoclinal folds and a subsequent folding (D3) around the lineation L2 [117]. Both deformation events are older than mica cooling age of 300 Ma [118,119]. The paragneisses include dm-scaled sheath-like folded calcsilicate-gneiss bodies, which are elongated parallel to lineation L2 (Figure 6). Quartz c-axis fabrics signal deformation partitioning with non-coaxial shearing parallel to lineation and elongation axis along the margin of the body and coaxial deformation (flattening and constriction) in the center. Porphyroblastic garnets and local C-S offsets in the paragneiss host rock indicate the same sense of shear. Calcite is intergrown with all the main metamorphic mineral phases, including as tail growths on garnet porphyroblasts, and mechanical twins are thus interpreted as primary recorders of the Variscan stress–strain history. Twelve thin sections were analyzed (n = 769), and all had a high percentage of the secondary negative expected values (30–35%), such that two strain events were separated (PEV and NEV; Erickson et al. 2006; Table 1 and Table 2). The two events are shortening parallel to the axis of the sheath fold (~N-S; PEV) and vertical shortening (NEV), presumably related to Alpine exhumation and nappe stacking. The NEV twinning strain overprint records extension parallel to the short axes of the bodies, corroborating the quartz fabric data. The average shortening strain was −4.62%, and the average differential stress associated with twinning was −339 bars.
Figure 4. Tertiary basalts (A) from the northern coast of Ireland with calcite veins and horizontal striations on Cretaceous limestones (C). Calcite twinning strains for all data (Table 1 and Table 2, Appendix A; solid circles = shortening axes, open circles = extension axes in (B). Tera–Wasserburg plots of U-Pb ages on vein calcite ((DG); Appendix B).
Figure 4. Tertiary basalts (A) from the northern coast of Ireland with calcite veins and horizontal striations on Cretaceous limestones (C). Calcite twinning strains for all data (Table 1 and Table 2, Appendix A; solid circles = shortening axes, open circles = extension axes in (B). Tera–Wasserburg plots of U-Pb ages on vein calcite ((DG); Appendix B).
Geosciences 12 00254 g004
Figure 5. Field photos from the distal Alpine foreland. (A) Ramp anticline in Jurassic limestones and evaporites with a south-dipping thrust, Markt Nordheim, Germany. (B) Vertical Cretaceous limestones, Isle of Wight, UK, with inset of oblique-slip striations. (C,D) Thrust contraction structures in Jurassic limestones, Lulworth Cove, UK.
Figure 5. Field photos from the distal Alpine foreland. (A) Ramp anticline in Jurassic limestones and evaporites with a south-dipping thrust, Markt Nordheim, Germany. (B) Vertical Cretaceous limestones, Isle of Wight, UK, with inset of oblique-slip striations. (C,D) Thrust contraction structures in Jurassic limestones, Lulworth Cove, UK.
Geosciences 12 00254 g005
Figure 6. (A) Calcsilicate sheath folds in the Tauern window near Tyrol, Austria, including a section cut perpendicular to the fold axis (inset; penny for scale) and (B) photomicrograph of garnet (PL) surrounded by twinned calcite. Strain data in Table 1 and Table 2.
Figure 6. (A) Calcsilicate sheath folds in the Tauern window near Tyrol, Austria, including a section cut perpendicular to the fold axis (inset; penny for scale) and (B) photomicrograph of garnet (PL) surrounded by twinned calcite. Strain data in Table 1 and Table 2.
Geosciences 12 00254 g006

4.3. Alpine Nappes and Molasse

Calcite strain data from the Alps are available from: the oldest, internal Helvetic–Penninic nappes ([40]; 29 samples), the Helvetic nappes ([73]; 14 samples plus 4 new strain results), the Helvetic flysch-molasse [41]; 20 samples], the Pre-Alps ([39]; 39 samples), the Molasse basin ([36,38]; 26 samples), and the youngest, frontal Jura folds ([36]; 9 samples, see also 37; Table 1 and Table 2; Figure 7 and Figure 8). Limestones in the Penninic nappes preserve a mix of LPS and LNS fabrics with no particular structural pattern. A comparison of e1 strain magnitudes reveals that the LNS twinning shortenings strains are higher (−7.5%) than the LPS shortening strains (−5.4%); these are the highest strain values due to the presence of thick twins [106]. The Helvetic nappes preserve a combination of layer-parallel and layer-normal shortening (sub-vertical) that is generally in the plane of tectonic transport, where the vertical shortening is associated with nappe stacking. We also analyzed four samples of fault-striated calcite from the Helvetic nappes, samples analyzed by Ring and Gerdes (2016) for U-Pb ages and fault kinematics; we find a horizontal shortening strain in the plane of nappe transport for calcite with U-Pb ages between 21.8 and 25.3 Ma. Shortening strains for the Helvetic LPS (−3.8%) and LPS (−3.98%) fabrics are comparable. The Pre-Alps (Penninic klippe) include 26 limestone samples and 13 calcite veins and a split of 20 LNS and 19 LPS fabrics with shortening and extension axes in no clear pattern. A comparison of e1 strain magnitudes reveals that the LNS twinning shortenings strains are higher (−5.81%) than the LPS shortening strains (−3.83%); this is expected, as the Pre-Alps are a continuation of the Penninic nappes (Figure 3). The molasse sediments preserve a transport-parallel LPS fabric, as do the Jura limestones. Differential stress data are not available for all the study areas.
Figure 7. Lower hemisphere stereoplots of calcite strain data for the frontal Alps and foreland (e shortening axes; e3 = extension axes). Key: filled circles = LPS data; open circles = LNS fabric; open EV strain overprints. Strain data are in Table 1 and Table 2 and field locations in Appendix A.
Figure 7. Lower hemisphere stereoplots of calcite strain data for the frontal Alps and foreland (e shortening axes; e3 = extension axes). Key: filled circles = LPS data; open circles = LNS fabric; open EV strain overprints. Strain data are in Table 1 and Table 2 and field locations in Appendix A.
Geosciences 12 00254 g007
Figure 8. Lower hemisphere stereoplots of calcite strain data (e1 = shortening axes; e3 = extension axes) for different structural elements of the Alps. Key: filled circles are LPS fabrics, open circles are LNS fabrics.
Figure 8. Lower hemisphere stereoplots of calcite strain data (e1 = shortening axes; e3 = extension axes) for different structural elements of the Alps. Key: filled circles are LPS fabrics, open circles are LNS fabrics.
Geosciences 12 00254 g008

4.4. Alpine Foreland

Alpine orogen strains are documented in the frontal Jura folds [37,41] and foreland in Burgundy [7,119]. We have expanded the foreland coverage 1000 km northwestward in the direction of Alpine tectonic transport and eastward across the French and German foreland (Figure 1 and Figure 3) to document foreland deformation patterns and to see what role oroclinal bending played in the evolution of the Alps [41,71,120,121,122,123]. Thrust-related compressional structures are present far into the foreland, including northwest-facing ramp anticlines, southeast-verging back thrusts and some oblique faults (Figure 5; [124,125,126]). Limestones (12 samples, 15 strain analyses, n = 300 twin measurements) preserve a LPS fabric across the foreland adjacent to the Jura Mountains northwestward to the Cheddar Gorge–Bath area in Britain where Mesozoic outcrops become scarce (Figure 1 and Figure 7; Table 1 and Table 2). Six sample sites were combined into three strain analyses because of the small number of twins and the closeness of sites to each other (Chatenois + Graux, Champlitte + N. Champlitte and Taxenne + Besancon, France). This resulted in PEV and NEV data splits, where the primary shortening strain recorded is SE-NW LPS with a sub-vertical shortening (NEV) overprint. In general, strain magnitudes are small and do not preserve any particular gradient across the region, although NEV percentages diminish away from the Alpine front. Foreland calcite veins along the southern British coast at Lulworth Cove and the Isle of Wight with U-Pb ages of ~34.7 Ma are in concert with an Alpine far-field origin (Parrish et al. 2017; these authors believe the deformation is Pyrenean). Calcite strain analyses in the foreland veins (17 samples, 19 strain analyses, n = 425 grains) reveal a SE-NW sub-horizontal shortening (15 results) and a NE-SW sub-vertical strain (4 results). Two samples allowed for a NEV split, one of which preserved a SE-NW LPS fabric and a vertical overprint (Grafenberg, Germany) while near Corfe Castle, Britain, two orthogonal sub-vertical strains are preserved. All four analyses from the Corfe site are from veins in a micritic Cretaceous limestone and are anomalous. Vein strain magnitudes are higher than strain magnitudes in the hosting limestones, while the inferred twinning differential stress magnitudes are regionally ~−350 bars. Vein fabrics are nearly even between VNS and VPS interpretations.

5. Discussion

The opening of the North Atlantic is preserved on the eastern margin of the ocean as Tertiary basalts in contact with Cretaceous chalks and limestones as a combination of sills, dikes and flows. Both the basalts and limestones are offset by numerous coast-parallel normal faults (with mm-scale offsets), rare normal fault hanging wall folds (i.e., Ballintoy Harbor; [127,128]) and numerous calcite veins. The extent of this field observation seems to be ~50 km inland from the margin, although this is also the distance where outcrops are dominated by the Welsh slate belt and an absence of Mesozoic sediments. Conversely, ~100 km southeastward from the Atlantic margin is also ~1000 km northwest of the Alps, and we observe a distinctive change in the outcrop-scale structures and calcite twinning strains preserved. The absence of the Alpine LPS fabric >1000 km can be explained by the existence of previously twinned and strain-hardened calcite near the older Atlantic margin, a decrease in the differential stress magnitude of the Alpine orogen to the northwest, or both. Summaries of twinning strain data, relative to distance from an active plate boundary, are presented in Figure 9 and Figure 10. The Alpine foreland sites (Figure 9 and Figure 10) record a mix of shortening strain magnitudes, which likely preserves a series of different orogenic events, namely the Pre-Alpine Pyrenean [14], the Alpine orogen and the active CRIS rift system [81,82].
Figure 9. Plot of calcite twinning shortening strain (-e1) across the Alps, the Alpine foreland and the older (58 Ma) divergent Atlantic margin in Ireland.
Figure 9. Plot of calcite twinning shortening strain (-e1) across the Alps, the Alpine foreland and the older (58 Ma) divergent Atlantic margin in Ireland.
Geosciences 12 00254 g009
Figure 10. Plot of the angle between the calcite twining shortening axis and bedding for the Alps, Alpine foreland and the older (58 Ma) divergent Atlantic margin in Ireland.
Figure 10. Plot of the angle between the calcite twining shortening axis and bedding for the Alps, Alpine foreland and the older (58 Ma) divergent Atlantic margin in Ireland.
Geosciences 12 00254 g010

5.1. Opening the Atlantic Ocean (Northern Irish Coast)

The calcite strains in Northern Ireland record a clean, consistent pattern of vertical extension and SW-NE sub-horizontal shortening. All the strains were measured in calcite veins or normal fault fillings that have a variety of orientations and cross-cut Tertiary basalts between the ages of 70–42 Ma (Figure 4). Curiously, the small-offset local structures and kinematics, sinistral striations on veins or stepped and striated calcite in normal faults have no relation to the stress–strain field of the regional setting. Evidence of a regional pattern is provided by the variability of the calcite vein orientations and dominance of VNS fabric interpretations; veins often preserve localized slip along their margins, which results in a VPS twinning strain. We interpret these results to indicate that vertical extension and margin-parallel sub-horizontal shortening dominated the evolution of the Irish portion of the Atlantic margin, where the margin-parallel horizontal differential stress was −350 bars to a distance of ~50 km inboard from the current coast. Our results are consistent with and along strike of the Flamborough fault zone in northeastern Britain, where calcite fillings have U-Pb ages of 63–55 Ma [129]. Roberts and Walker [116] also report three sets of calcite-filled faults in the Faroe Islands with a range of U-Pb ages (44–11 Ma), and their set 3 fault kinematic interpretation (41.7 Ma) agrees with our margin-parallel shortening and margin-normal (SE-NW) extension twinning result. Our Irish strain pattern is distinctively different from the twinning strains along the active margin in Iceland, where ridge-normal sub-horizontal shortening (i.e., ridge-push; −480 bars) dominates with a ridge-parallel sub-vertical shortening overprint ([47]; see below). The Irish data (e1% and e1 angle vs. bedding) are plotted across the region to compare the Atlantic opening strain with younger Alpine deformation (Figure 9 and Figure 10); margin-parallel shortening strains associated with opening the Atlantic margin are considerably higher than some parts of the frontal Alps.

5.2. The Alpine Nappes and Foreland Deformation

Our twinning strain data in the foreland, following the twinning paleostress work of Lacombe et al. [7,130] in the Alpine foreland, were intended to extend the sample coverage ~1000 km from the frontal Jura folds northwestward in the direction of tectonic transport and along strike to the northeast toward Nurnberg, Germany. This approach allowed us to document the distance of far-field Alpine tectonic stresses preserved by twinned calcite and the orientation of the Alpine stress–strain field northwest and north of the Central Alps. Our sample suite is dominated by calcite veins (n = 16) hosted in Mesozoic limestones (n = 13; Figure 1 and Figure 5, Table 1 and Table 2), and both suites preserve a SE-NW sub-horizontal shortening strain. The limestones record a layer-parallel shortening (LPS) fabric, and the veins are a nearly equal mix of VNS and VPS fabrics, with 15 results recording SE-NW horizontal shortening and 4 preserving NE-SW sub-vertical shortening (Figure 11A–C). The veins collected near Corfe Castle, Britain, record a complex sub-vertical twinning strain that is inconsistent with the remainder of the data (Figure 11, Table 1 and Table 2). Strain overprints (high NEV%) are more common near the Alpine front, and the overprint includes one example of sub-vertical shortening (Figure 11) and a strain overprint parallel to the SE-NW Alpine shortening where the older, PEV twinning strain records N-S horizontal shortening (Iberia accretion, ECRIS extension?). The strain overprint complexities (Figure 9, Figure 10 and Figure 11) in the Alpine foreland complement earlier calcite paleostress and outcrop fabric results of Lacombe et al. [7,130], where the persistent N-S horizontal shortening and compression are attributed to formation of the Pyrenees Mountains. Combining calcite twin studies with U-Pb ages on calcite veins and fault gouge would better resolve the timing and orientations of superimposed deformation events.
Figure 11. Stereoplots of calcite strain overprint (NEV) data that do not plot within the SE-NW plane of Alpine thrust transport and shortening (AD) or of a twinning strain that does ((E,F); see Figure 1, Table 2). Key: e1 = shortening strain axis, e2 = intermediate axis, e3 = extension axis with Turner (1953) compression axes contoured. Great circles are veins.
Figure 11. Stereoplots of calcite strain overprint (NEV) data that do not plot within the SE-NW plane of Alpine thrust transport and shortening (AD) or of a twinning strain that does ((E,F); see Figure 1, Table 2). Key: e1 = shortening strain axis, e2 = intermediate axis, e3 = extension axis with Turner (1953) compression axes contoured. Great circles are veins.
Geosciences 12 00254 g011
Both the Mesozoic limestones and calcite veins (see Figure 11 for exceptions) preserve the same regional twinning fabric in marine sediments that are adjacent and autochthonous to the Alps, that is, sub-horizontal shortening parallel to the Alpine thrust transport direction. This strain is interpreted as a pre-thrusting strain imparted far into the foreland as an orogen evolved. This compares to the Appalachian collisional orogen, where the limestone and vein LPS fabric is parallel (SE-NW) in both the foreland and thrust belt [6,16,17]. In the Sevier belt in the western US, the LPS fabric in the foreland (E-W) is not parallel to the LPS fabric in the Idaho–Wyoming belt, which records a counterclockwise rotation of thrust sheets, from west to east, during the transpressive orogen [18,19,20].
Using our extended foreland data, we wanted to test this thrust sheet rotation hypothesis for the Alps, incorporating existing calcite twin strain results (Figure 1, Figure 7 and Figure 8). In structural order, from internal (SE) to external (NW): Variscan calcsilicates in the Tauern window contain abundant metamorphic calcite that preserves a distinctive sub-vertical shortening fabric as an Alpine-aged NEV data split and records burial by Helvetic–Penninic nappes [40]. Limestones in the Helvetic–Penninic nappes record a chaotic pattern of LPS and LNS fabrics that are not in the plane of Alpine tectonic transport. The Helvetic nappes preserve a combination of layer-parallel and layer-normal shortening (sub-vertical) that is generally in the plane of tectonic transport, while the vertical shortening is associated with nappe stacking. The Pre-Alps preserve a complex pattern of LPS and LNS fabrics. The molasse sediments preserve a transport-parallel LPS fabric, as do the Jura limestones. A small subset of foreland veins preserve a variety of shortening axes oriented SW-NE with steep plunges and are not of Alpine origin (Figure 11; see [7,119]). There is parallelism between the LPS fabric e1 axes for the foreland limestones, veins, Jura folds, Molasse basin sediments and the Helvetic nappes, which suggests progressive tectonic transport of the Jura–Molasse–Helvetic structures in a constant northwest direction. The Pre-Alpine and Helvetic–Penninic strains are complex, suggesting non-plane strain deformation associated with nappe burial and thrust transport that were not everywhere toward the northwest (see [131,132]). Regional relations of shortening strains (e1%) and the angle between e1 and bedding are presented in Figure 9 and Figure 10. Alpine nappes in Crete record a similar strain history with a LPS fabric in the nappes that is parallel to thrust transport, but the limestones and calcite veins in the synorogenic flysch preserve a complex, non-plane strain pattern [133,134].

5.3. Implications for Orogenic Collisions and Thrust Mechanics

The collision of Apulia into the European plate resulted in far-field translation of tectonic stresses and formation of contractional structures (Figure 3 and Figure 5) and twinned calcite ~1200 km to the northwest of the Alps. Previously twinned calcite farther to the NW along the Atlantic margin strain hardened and did not preserve this new deformation with a twinning strain overprint, as all the NEV values are low (Table 2). The area of the critically tapered Alpine wedge includes the 1200 km of autochthonous foreland sediments and the Jura folds as well as the Helvetic and Penninic nappes. The latter have a width of 150 km, which, when palinspastically restored, augments to a width of 350 km [29,71] for a total pre-orogenic width of 1550 km. Thus, as a crude estimate, a total pre-orogenic width of 1550 km must be assumed for the initial Alpine wedge. The Alps formed by thrust shortening that progressed in time to the craton (NW), where the Molasse sediments and Jura-foreland sediments folded and cleaved at temperatures less than 100 °C, whereas the Alpine nappes are allochthonous, intensely deformed and metamorphosed with temperatures exceeding 300 °C [23,26,33,52,53,54,73,135].
During the early stages of orogenic convergence, horizontal tectonic stresses propagate into the foreland normal to the trench axis, thereby preserving a LPS twinning strain across the region; as shortening continued, the orientation of the pre-thrusting LPS fabric can be used to track motion of allochthonous rocks within the orogen. The SE-NW LPS fabric we observe in the Alpine foreland is parallel with the LPS fabric in the Jura and Helvetic allochthons, but the LPS fabric in the Molasse basin and the Helvetic–Penninic nappes is largely rotated out of the plane of thrust transport up to 90° (Figure 9 and Figure 10) to accommodate the oroclinal shape and structure of the mountain belt. The Helvetic nappes acquired a secondary LNS twinning strain that is within the plane of thrust transport and is interpreted as a burial (vertical) shortening strain. Our strain data also indicate the Pre-Alp and Helvetic–Penninic nappes acquired a secondary LNS twinning strain during nappe formation and that thrust transport was out of the plane of thrust transport (Figure 8). As a collisional orogenic system, the Alps are unlike the collisional Appalachian [17], Grenville [19] or Mazatzal [136] orogens, because the hinterland indentor, the Adriatic margin of the Apulia plate, was actively deformed into the orogen and caused nappe formation, strain overprints and nappe motions out of the plane of tectonic transport. In contrast, Laramide crystalline uplifts in the Sevier belt foreland of Wyoming have oblique-slip kinematics, small fault offsets (<5 km) and no twinning strain overprint in hanging wall drape fold limestones [137].
The Alp foreland includes the rootless, concentric Jura folds, which changed into the low-amplitude, long-wavelength Paris basin, Purbeck monocline, Weald–Artois anticline and London basin to the north (Figure 1). These folds represent far-field Alpine foreland deformation in the autochthonous footwall at a distance of ~1200 km, which is, not surprisingly, the same distance from the Alpine front where calcite is twinned (see [14]). For comparison, the Appalachian foreland includes sheared coal beds and salt-cored anticlines ~30 km inboard from the thrust front, the inversion Keweenaw–Kapuskasing thrust structure ~1220 km inboard [17] and twinned calcite in Paleozoic limestones >2000 km inboard. The Kiri uplift [138] in central Africa is ~4000 km inboard (south) and ~4000 km inboard (north) of the Caledonide and Cape orogenic belts, respectively, forming as Gondwana amalgamated in the Permian; basement inversion structures in the foreland are not part of the Alpine orogen. Orogenic forelands have different dimensions and different structural styles, including Taiwan fold belt (10 km foreland width), Zagros fold belt (300 km), Pyrenees (700 km), Alps (1200 km), Appalachian’s (>2000 km) and Sevier belt (>2000 km; [9]). Deformation is related to plate convergence rates, thicknesses, slab dip (does an arc develop?) and the presence of pre-existing crustal weaknesses (i.e., Kiri uplift).
Our data set is a compilation of the results of many researchers (Table 2), so we do not have a complete view of twinning-related differential stresses. Our Alpine foreland data preserve a differential stress magnitude of −345 bars (−34.5 MPa), considerably lower than results presented by Lacombe et al. [130] of −55 to −88 MPa. Our results are consistent with the grand compilation of Beaudoin and Lacombe [10], namely −100 MPa at a normalized depth of 1 km.

5.4. Tectonic Evolution of the European Foreland (100–0 Ma)

The tectonic evolution of western Europe was a prolonged (~100 Ma) and complex process of accretion (Iberia), extension and shearing (sinistral opening of the Atlantic Ocean), collision of Apulia and subsequent ECRIS rifting [23,31,33,44,45,46,79,80,139,140]. Lacombe et al. [7,130] was the first to report calcite twinning paleostress data in a complex pattern in the Alpine foreland. Our anomalous foreland strain data (Figure 11) support Lacombe’s results, which may be related to early, pre-Alpine N-S shortening Iberian accretion? [131,141,142], Pyrenees transpression, or post-Alpine extension (vertical shortening). A chronology of strain field data are presented in Figure 12: opening of the Atlantic Ocean (58 Ma) may have coincided with Iberian–Apulia collision as sinistral shearing dominated the Irish coast (Figure 4); the collision of the Adriatic margin of Apulia with Europe formed the Alps (starting at 100 Ma), with thrusting younging to the northwest [23]; we find calcite twinning strains are preserved ~1200 km into the foreland (Figure 7 and Figure 8). Calcite previously twinned along the Irish Coast was strain-hardened and did not twin again. The modern tectonic shortening strain field is sub-horizontal, ~parallel to the Alpine field (NW-SE), dominated by shortening normal to the Atlantic margin in Iceland [47] and parallel to the strike of the Gubbio normal fault in Italy (NW-SE; 115 U-Pb calcite ages: 230,000 and 325,000 yr. BP) as Africa converges with Europe (see [143]).

6. Conclusions

Mesozoic sediments in central Europe in the Alpine foreland preserve a complex array of calcite twinning strains associated with the collision of the Adriatic margin with Europe starting at 40 Ma. Twinning deformation is better preserved than by other brittle deformation markers [44,46,79,80] calcite vein and fault fillings (70–41 Ma U-Pb ages) in northern Ireland record a margin-parallel horizontal shortening strain with vertical extension up to ~100 km inboard (southeast) of the extensional Atlantic (58 Ma magnetic isochron) margin. The Alpine orogen (~50–5 Ma), along a traverse northwest from the Jura Mountains, is recorded by a thrust transport-parallel and horizontal shortening strain ~1200 km from the mountain front and includes thrusts and folds along the southern coast of Britain. Calcite strains become increasingly complex and bedding-normal (sub-vertical) toward the Alpine nappe stack in the Alps. Thrust sheet transport seems to be consistently normal (i.e., plane strain) to the strike of local thrust fault orientation except in the Helvetic–Penninic nappes (and Pre-Alp equivalents).

Author Contributions

J.P.C. conceptualization, writing, sampling and analysis, U.R. sampling and analysis, writing, O.A.P. analysis and writing. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

Martin Burkhard passed away in 2006, and this paper is a compilation of the efforts of many of his students and colleagues to honor his decades of inquiries in understanding the tectonics of the Alps. Martin and Adrian Pfiffner worked on understanding stress fields from fault kinematics and calcite twins with earthquake focal mechanisms [144]. Craddock and Burkhard spent time in the field in 2003, and our results were presented at the 2006 Geological Society of America meeting in Philadelphia in a session honoring Rick Groshong. Fieldwork for this study was undertaken during a sabbatical for Craddock in Erlangen, Germany, funded by the Mercator DFG program. Many of the authors participated in the foreland fieldwork, especially Cara Craddock and Jan Wood. Bernhard Schulz contributed the oriented Tauern samples and to our understanding of the complex geology in that portion of the Austrian Alps. Geochronology was conducted at the University of California, Santa Barbara, under the direction of Andrew Kylander-Clark.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Figure A1. DEM of Alpine strain study locations.
Figure A1. DEM of Alpine strain study locations.
Geosciences 12 00254 g0a1

Appendix B

Table A1. U-Pb calcite ages (Ireland).
Table A1. U-Pb calcite ages (Ireland).
CommentsU ppm238U/206Pb207Pb/206PbrhoIntercept Age
PUE21-440.0082.860.700.8090.0230.0376153
PUE21-360.00040.400.450.8900.0250.03−11771964
PUE21-130.0050.900.830.8710.029−0.80−326639
PUE21-400.0233.261.000.8460.031−0.36−26150
PUE21-340.0184.042.790.8240.032−0.0523120
PUE21-310.0265.282.050.8060.033−0.374692
PUE21-70.0110.840.960.8260.0330.3793585
PUE21-410.0213.351.340.8370.0330.02−3147
PUE21-430.0173.661.950.8440.0350.14−19138
PUE21-50.0174.140.650.8410.036−0.06−11122
PUE21-300.03813.032.760.7720.041−0.084039
PUE21-220.0133.911.440.7920.043−0.1191136
PUE21-470.0258.951.850.7840.046−0.454760
PUE21-250.0081.630.880.8640.0480.62−143369
PUE21-280.0137.244.340.7690.0490.067587
PUE21-30.0185.811.270.8230.051−0.391899
PUE21-320.0103.092.680.8320.0510.2610188
PUE21-90.0100.730.820.8690.0510.47−380954
PUE21-460.0093.011.850.8520.0540.82−45204
PUE21-190.0040.651.120.8600.054−0.75−3111110
PUE21-350.0040.380.400.8080.057−0.215791647
PUE21-260.01811.857.660.8030.0580.302254
PUE21-290.0209.692.170.7770.059−0.094964
PUE21-490.00716.5211.170.8100.0640.421341
PUE21-480.0093.261.090.8020.073−0.4684218
PUE21-330.04531.033.180.6770.073−0.054222
PUE21-210.02727.277.630.6050.079−0.086932
PUE21-270.0080.490.530.8730.083−0.38−6481913
PUE21-10.0021.410.800.8390.0841.00−20565
PUE21-150.0013.662.520.7700.087−0.25145238
PUE21-160.01935.5417.250.7390.093−0.842226
PUE21-200.02533.327.600.7000.101−0.693328
PUE21-20.0092.302.030.8200.111−0.0555432
PUE21-390.0065.585.050.8100.121−0.6437193
PUE21-180.01243.4419.320.7300.1810.342036
PUE21-110.0066.904.870.9500.1910.51−136258
PUE21-170.00810.6610.660.8400.2310.31−3181
PUE21-380.00731.7026.560.8100.280−0.85773
PUE23-370.05653.807.730.4350.0530.296112
PUE23-380.02433.996.830.6370.072−0.174822
PUE23-390.03146.548.000.6600.111−0.733121
PUE23-120.03523.463.880.6640.059−0.366026
PUE23-300.0145.192.280.7750.0440.2295107
PUE23-10.0456.481.010.7940.0300.095271
PUE23-270.0111.831.550.7970.037−0.12171299
PUE23-500.0136.904.060.7990.074−0.2843107
PUE23-80.0114.512.780.8040.048−0.0557127
PUE23-310.0095.832.530.8070.0880.5040140
PUE23-280.03111.171.930.8090.0600.421956
PUE23-90.0514.790.810.8170.022−0.033291
PUE23-140.0204.191.800.8180.0390.3934122
PUE23-490.0194.460.420.8190.050−0.3430127
PUE23-70.0216.481.900.8190.053−0.462191
PUE23-110.0171.830.750.8220.023−0.2360241
PUE23-60.0382.210.590.8260.028−0.5935209
PUE23-360.0022.261.610.8260.0670.2235301
PUE23-440.0047.335.040.8400.131−0.33−5156
PUE23-260.0072.861.880.8400.1510.31−13455
PUE23-340.00613.188.000.8400.181−0.72−3116
PUE23-200.0052.861.330.8480.0820.49−35276
PUE23-320.0075.192.110.8500.151−0.69−23251
PUE23-470.0080.250.070.8710.022−0.54−12472210
PUE23-190.0104.892.240.8720.1000.08−61191
PUE23-290.0330.180.030.8840.018−0.31−26813809
PUE23-50.0480.080.010.9050.0190.2905882
PUE23-350.0010.050.100.9100.1510.13034,884
pue17-170.0600.530.320.8510.0230.68−239893
pue17-250.0562.301.800.6000.2400.688061012
pue17-20.0313.662.180.7940.0260.3592131
pue17-30.1184.221.220.8020.0280.6465108
pue17-290.1654.511.740.8100.0260.6646100
pue17-240.0605.862.880.7690.0410.669296
pue17-70.0076.900.740.8300.1610.017199
pue17-310.0586.907.310.6200.1310.66252306
pue17-110.2128.3814.360.5500.1500.66274487
pue17-120.25410.294.970.7230.0460.698965
pue17-140.27510.868.850.7300.0450.637979
pue17-100.27011.975.740.7360.0450.676854
pue17-160.06513.338.790.6910.0990.698887
pue17-280.01718.0418.320.8300.1210.39359
pue17-210.29219.885.740.6910.0490.615931
pue17-90.06635.5418.320.5800.1110.575940
pue17-270.07448.8620.380.5800.1110.544326
pue17-80.05149.0710.930.4830.0930.195821
pue17-230.02955.8510.700.3400.1000.257220
pue17-220.09166.6313.690.4300.0840.125015
pue17-190.16672.3924.170.3080.0860.485922
pue17-180.08473.3010.180.4930.0620.09389
pue17-200.05977.1633.540.2330.0640.236428
pue17-130.10484.378.660.3050.0230.30516
pue17-10.23690.2122.280.4000.1000.493913
pue17-260.149111.4810.410.3160.0710.16386
pue17-40.136113.979.690.2530.0930.15428
pue17-60.244121.4012.430.2490.0250.38394
pue17-150.255127.347.760.1770.0450.21424
pue26-30.0191.500.850.8500.0290.20−78323
pue26-50.0011.701.450.8070.0620.01136390
pue26-60.0114.694.880.8160.0860.0434174
pue26-80.08223.744.160.6940.0420.334922
pue26-90.04414.4813.230.5830.0590.36141134
pue26-100.06415.642.520.7150.0500.416335
pue26-110.02115.439.950.7600.1310.264077
pue26-120.05411.973.060.8010.0460.622445
pue26-130.0324.341.930.7780.0570.27108145
pue26-150.00212.0911.100.8700.1410.01−23104
pue26-160.0241.200.390.8490.0210.41−92381
pue26-170.0413.551.720.8330.0510.286163
pue26-180.01817.005.430.7300.1210.135163
pue26-190.0213.351.060.8710.0660.06−87209
pue26-210.0126.010.900.8590.0650.11−32113
pue26-250.0073.012.010.8100.1310.0469379
pue26-270.0181.861.150.8420.0800.11−28416
pue26-280.07668.5814.900.2650.0950.226819
pue26-290.0371.171.640.8120.0280.23163446
pue26-300.0394.510.770.8300.0300.3010105
pue26-310.0101.302.320.7710.0990.44397956
pue26-320.0195.104.210.5800.1500.45401403
pue26-340.0736.441.180.7940.0250.325368
pue26-350.1278.561.630.7910.0260.304252
pue26-370.0760.880.230.8570.0200.19−200524
pue26-380.0290.210.080.8640.0210.14−12102633
pue26-390.0243.663.090.8170.0980.1541246
pue26-400.0560.340.280.8000.1210.098122949
pue26-410.0100.970.350.8680.0360.22−277561
ASH15-10.76942.7365.660.4960.0860.5431
ASH15-10.821161.15242.540.3880.0600.0931
ASH15-101.591649.45118.370.2010.046−0.0130
ASH15-110.711714.56105.970.0820.0590.1740
ASH15-111.931719.59111.340.2510.0870.1830
ASH15-121.681411.2678.290.3040.0590.2930
ASH15-120.731633.37205.080.2640.0510.6030
ASH15-130.581670.59223.830.3100.1600.5731
ASH15-140.951651.77114.220.2560.0690.1030
ASH15-151.371804.2495.870.1940.049−0.0330
ASH15-160.961755.62176.980.2570.076−0.1530
ASH15-171.94746.9877.580.5810.0430.0731
ASH15-180.501183.41112.380.2730.0980.0941
ASH15-191.241628.83133.000.2570.0610.3130
ASH15-20.0111.7311.730.8600.8600.04−17601
ASH15-20.651370.04116.890.2760.088−0.1731
ASH15-202.721742.5892.280.1850.0140.4130
ASH15-30.681274.74195.650.3200.1100.0331
ASH15-30.921719.59177.340.2700.1700.1631
ASH15-41.201224.1764.860.4700.1100.2821
ASH15-41.141873.41139.790.1810.0680.3130
ASH15-50.831412.96149.110.3620.056−0.3630
ASH15-51.751846.86221.240.1960.0900.1331
ASH15-61.981747.78119.820.1700.0390.2330
ASH15-61.021763.54175.950.2010.0590.0830
ASH15-71.03747.9337.900.5430.0510.2231
ASH15-80.721758.2691.390.3060.0690.1120
ASH15-91.02360.8536.260.7160.032−0.3731
ASH15-91.312008.15206.820.1920.0760.2030
db-118.0064.512.990.2820.0090.04704
db-126.2068.262.100.2670.0080.51683
db-1017.6157.573.030.3010.011−0.54755
db-219.4866.265.040.2480.0070.12726
db-225.3769.351.770.2590.0070.55682
db-316.0756.933.250.3090.011−0.74755
db-320.7161.084.010.3160.0200.15696
db-410.0844.423.000.4180.017−0.66767
db-419.3247.911.920.4170.0130.53715
db-520.0050.122.560.3950.0110.32725
db-515.8858.353.400.3030.016−0.70745
db-622.7065.373.130.2840.0110.32694
db-616.3365.893.930.2600.007−0.29715
db-715.1756.113.670.3250.011−0.65746
db-729.3570.103.400.2500.0060.65684
db-816.2856.653.730.3120.017−0.54756
db-834.6073.163.020.2540.0060.59653
db-918.0365.412.970.2560.0060.32724
WC1-13.7821.801.560.1090.007−0.1226819
WC1-16.2023.510.610.1270.0030.472436
WC1-103.3320.760.780.1390.004−0.3327010
WC1-24.0220.082.950.1300.041−0.6528344
WC1-25.7223.440.590.0920.0040.402567
WC1-35.0021.640.770.1410.0040.472599
WC1-34.0122.171.820.1000.010−0.5526722
WC1-44.1523.222.750.0940.002−0.0925730
WC1-44.9423.270.760.1140.0030.442508
WC1-53.3621.801.680.0950.0040.0427421
WC1-55.2122.680.800.0940.0060.312639
WC1-64.1221.481.260.1350.005−0.1426315
WC1-66.5421.800.850.1410.0090.1425711
WC1-74.0321.281.310.1170.003−0.0727217
WC1-76.5123.891.130.1200.0040.6024211
WC1-85.6321.600.800.1330.0040.3726210
WC1-83.7521.721.470.1090.0030.5126918
WC1-93.8620.400.670.1280.004−0.272799

References

  1. Chinn, A.A.; Konig, R.H. Stress inferred from calcite twin lamellae in relation to regional structure of northwest Arkansas. Geol. Soc. Am. Bull. 1973, 84, 3731–3736. [Google Scholar] [CrossRef]
  2. Nickelsen, R.P. Fossil distortion and penetrative rock deformation in the Appalachian plateau, Pennsylvania. J. Geol. 1966, 74, 924–931. [Google Scholar] [CrossRef]
  3. Engelder, T.; Engelder, R. Fossil distortion and decollement tectonics of the Appalachian plateau. Geology 1977, 5, 457–460. [Google Scholar] [CrossRef]
  4. Engelder, T.; Geiser, P. One the use of regional joint sets as trajectories of paleostress fields during the development of the Appalachian plateau, New York. J. Geophys. Res. 1980, 85, 6319–6341. [Google Scholar] [CrossRef]
  5. Geiser, P.; Engelder, T. The distribution of layer-parallel shortening fabrics in the Appalachian foreland of New York and Pennsylvania: Evidence for two non-coaxial phases of the Alleghanian orogen. In Contributions to the Tectonics and Geophysics of Mountain Chains; Geological Society of America Memoirs: Boulder, CO, USA, 1983; Volume 158, pp. 161–176. [Google Scholar]
  6. Craddock, J.P.; van der Pluijm, B.A. Late Paleozoic deformation of the cratonic carbonate cover of eastern North America. Geology 1989, 17, 416–419. [Google Scholar] [CrossRef]
  7. Lacombe, O.; Angelier, J.; Laurent, P.; Bergerat, F.; Tourneret, C. Joint analyses of calcite twins and fault slips as a key for deciphering polyphase tectonics: Burgundy as a case study. Tectonophysics 1990, 182, 279–300. [Google Scholar] [CrossRef]
  8. Lacombe, O.; Amrouch, K.; Mouthereau, F.; Dissez, L. Calcite twinning constraints on late Neogene stress patterns and deformation mechanisms in the active Zagros collision belt. Geology 2007, 35, 263–266. [Google Scholar] [CrossRef]
  9. Lacombe, O. Calcite twins, a tool for tectonic studies in thrust belts and stable orogenic forelands. Oil Gas Sci. Technol. 2010, 65, 809–838. [Google Scholar] [CrossRef] [Green Version]
  10. Beaudoin, N.; Lacombe, O. Recent and future trends in paleopiezometry in the diagenetic domain: Insights into the tectonic paleostress and burial depth history of fold-and-thrust belts and sedimentary basins. J. Struct. Geol. 2018, 114, 357–365. [Google Scholar] [CrossRef]
  11. Chapple, W.M. Mechanics of thin-skinned fold-and-thrust belts. Geol. Soc. Am. Bull. 1978, 89, 1189–1198. [Google Scholar] [CrossRef]
  12. Davis, D.; Suppe, J.; Dahlen, F.A. Mechanics of fold-and-thrust belts and accretionary wedges. J. Geoph. Res. 1983, 88, 1153–1172. [Google Scholar] [CrossRef]
  13. Dahlen, F.A.; Suppe, J.; Davis, D. Mechanics of fold-and-thrust belts and Accretionary Wedges: Cohesive Coulomb theory. J. Geophys. Res. 1984, 89, 10087–10101. [Google Scholar] [CrossRef]
  14. Lacombe, O.; Mouthereau, F. What is the real front of orogens? The Pyrenean orogen as a case study. C.R. Acad. Sci. Ser. IIA Earth Planet. Sci. 1999, 329, 889–896. [Google Scholar]
  15. Wiltschko, D.V.; Medwedeff, D.A.; Millson, H.E. Distribution and mechanisms of strain within rocks on the northwest ramp of Pine Mountain block, southern Appalachian foreland: A field test of theory. Geol. Soc. Am. Bull. 1985, 96, 426–435. [Google Scholar] [CrossRef]
  16. Kilsdonk, W.; Wiltschko, D.V. Deformation mechanisms in the southeastern ramp region of the Pine Mountain block, Tennessee. Geol. Soc. Am. Bull. 1988, 100, 644–653. [Google Scholar] [CrossRef]
  17. Craddock, J.P.; Malone, D.H.; Porter, R.; Luczaj, J.; Konstantinou, A.; Day, J.E.; Johnston, S.T. Paleozoic reactivation structures in the Appalachian-Ouachita-Marathon foreland: Far-field deformation across Pangea. Earth Sci. Rev. 2017, 169, 1–34. [Google Scholar] [CrossRef]
  18. Craddock, J.P. Transpression during tectonic evolution of the Idaho-Wyoming fold-and-thrust belt. Reg. Geol. East. Ida. West. Wyo. Geol. Soc. Am. Mem. 1992, 179, 125–139. [Google Scholar]
  19. Craddock, J.P.; van der Pluijm, B.A. Regional stress-strain fields of Sevier-Laramide tectonism from calcite twinning data, west-central North America. Tectonophys. Spec. 1999, 305, 275–286. [Google Scholar] [CrossRef] [Green Version]
  20. Craddock, J.P.; Malone, D.H. An overview of strains in the Sevier thin-skinned thrust belt, Idaho and Wyoming, USA (latitude 42° N). In Tectonic Evolution of the Sevier-Laramide Hinterland, Thrust Belt, and Foreland, and Postorogenic Slab Rollback (180–20 Ma); Craddock, J.P., Malone, D.H., Foreman, B.Z., Konstantinou, A., Eds.; Geological Society of America: Boulder, CO, USA, 2021; Volume 555, pp. 133–148. [Google Scholar] [CrossRef]
  21. Schmid, S.M.; Fugenschuh, B.; Kissling, E.; Schuster, R. Tectonic map and overall architecture of the Alpine orogen. Eclogae Geol. Helv. 2004, 97, 93–117. [Google Scholar] [CrossRef]
  22. Rosenberg, C.L.; Kissling, E. Three-dimensional insight into Central-Alpine collision: Lower-plate or Upper-Plate indentation? Geology 2013, 41, 1219–1222. [Google Scholar] [CrossRef]
  23. Pfiffner, O.A. Geology of the Alps; Wiley Publishers: Hoboken, NJ, USA, 2014; p. 392. [Google Scholar]
  24. Price, J.B.; Wernicke, B.P.; Cosca, M.A.; Farley, K.A. Thermochronometry across the Austroalpine-Pennine boundary, Central Alps, Switzerland: Orogen—Perpendicular Normal Fault Slip on a Major “Overthrust” and its implications for orogenesis. Tectonics 2018, 37, 724–757. [Google Scholar] [CrossRef] [Green Version]
  25. Schmid, S.M. The Glarus overthrust: Field evidence and mechanical model. Eclogae Geol. Helv. 1975, 68, 251–284. [Google Scholar]
  26. Pfiffner, O.A.; Lehner, P.; Heitzmann, P.; Müller, S.; Steck, A. (Eds.) Deep Structure of the Swiss Alps: Results of NRP 20; Birkhäuser: Basel, Switzerland, 1997; pp. 101–114. [Google Scholar]
  27. Pfiffner, O.A.; Schlunegger, F.; Buiter, S.J.H. The Swiss Alps and their peripheral foreland basin: Stratigraphic response to deep crustal processes. Tectonics 2002, 21, 1009. [Google Scholar] [CrossRef] [Green Version]
  28. Kuhlemann, J.; Kempf, O. Post-Eocene evolution of the North Alpine Foreland Basin and its response to Alpine tectonics. Sediment. Geol. 2002, 152, 45–78. [Google Scholar] [CrossRef]
  29. Burkhard, M.; Sommurga, A. Evolution of the western Swiss Molasse basin: Structural relations with the Alps and Jura belt: In Cenozoic Foreland Basins of Western Europe. Geol. Soc. Spec. Publ. 1998, 134, 279–298. [Google Scholar] [CrossRef]
  30. Van der Pluijm, B.A.; Craddock, J.P.; Graham, B.R.; Harris, J.H. Paleostress in cratonic North America: Implications for deformation of continental interiors. Science 1997, 277, 792–796. [Google Scholar] [CrossRef] [Green Version]
  31. Tavani, S.; Storti, F.; Lacombe, O.; Corradetti, A.; Munoz, J.A.; Mazzoli, S. A review of deformation pattern templates in foreland basin systems and fold-and-thrust belts: Implications for the stress in the frontal regions of thrust wedges. Earth-Sci. Rev. 2015, 141, 82–104. [Google Scholar] [CrossRef]
  32. Lacombe, O.; Bellahsen, N. Thick-Skinned tectonics and basement-involved fold-thrust belts: Insight from selected Cenozoic orogens. Geol. Mag. 2016, 153, 763–810. [Google Scholar] [CrossRef] [Green Version]
  33. Pfiffner, O.A. Thick-skinned and thin-skinned Tectonics: A global perspective. Geosciences 2017, 7, 71. [Google Scholar] [CrossRef] [Green Version]
  34. Ring, U.; Bohar, R. Tilting, uplift volcanism and disintegration of the South German block. Tectonophysics 2020, 795, 228611. [Google Scholar] [CrossRef]
  35. Craddock, J.P.; Burkhard, M. Alpine deformation from the internal Austrian Tauren window northwestward to the British foreland and Irish Tertiary province. Geol. Soc. Am. Abstr. Programs 2006, 38, 480. [Google Scholar]
  36. Tschanz, X.; Sommaruga, A. Deformation associated with folding above frontal and oblique ramps around the rhomb-shaped Val-de-Ruz basin (Jura Mountains. Ann. Tecton. 1993, 3, 53–70. [Google Scholar]
  37. Ferrill, D.A.; Groshong, R.H. Deformation conditions in the northern Subalpine chain, France, estimated from deformation modes in coarse-grain limestone. J. Struct. Geol. 1993, 15, 995–1006. [Google Scholar] [CrossRef]
  38. Hindle, D. Quantifying Stresses and Strains from the Jura Arc, and Their Usefulness in Choosing a Deformation Model for the Region. Ph.D. Thesis, Neuchatel University, Neuchâtel, Switzerland, 1996. [Google Scholar]
  39. Mosar, J. Internal deformation in the Prealpes Medianes, Switzerland. Ecol. Geol. Helv. 1989, 82, 765–793. [Google Scholar]
  40. Burkhard, M. Deformation des calcaires de L’Helvetique de la Suisse occidentale (phenomènes, mécanismes et interprétations tectoniques). Rev. Geol. Dyn. Geogr. Phys. 1986, 27, 281–301. [Google Scholar]
  41. Hindle, D.; Burkhard, M. Strain, displacement and rotation associated with the formation of curvature in fold belts: The example of the Jura arc. J. Struct. Geol. 1999, 21, 1089–1101. [Google Scholar] [CrossRef] [Green Version]
  42. Groshong, R.H., Jr.; Teufel, L.W.; Gasteiger, C.M. Precision and accuracy of the calcite strain-gage technique. Bull. Geol. Soc. Am. 1984, 95, 357–363. [Google Scholar] [CrossRef]
  43. Handy, M.R.; Zing, A. The tectonic and rheological evolution of an Attenuated cross section of the continental crust: Ivrea crustal section, southern Alps, northwestern Italy and southern Switzerland. Geol. Soc. Am. Bull. 1991, 103, 236–253. [Google Scholar] [CrossRef]
  44. Kley, J.; Voigt, T. Late Cretaceous intraplate thrusting in central Europe: Effect of Africa-Iberia-Europe convergence, not Alpine collision. Geology 2008, 36, 839–842. [Google Scholar] [CrossRef]
  45. Ring, U.; Gerdes, A. Kinematics of the Alpenrhein—Bodensee graben system in the Central Alps: Oligocene/Miocene transtension due to formation of the Western Alps arc. Tectonics 2016, 35, 1367–1391. [Google Scholar] [CrossRef]
  46. Navabpour, P.; Malz, A.; Kley, J.; Sieburg, M.; Kasch, N.; Ustaszewski, K. Intraplate brittle deformation and states of paleostress constrained by fault kinematics in the central German platform. Tectonophysiscs 2017, 694, 146–163. [Google Scholar] [CrossRef]
  47. Craddock, J.P.; Farris, D.; Roberson, A. Calcite-twinning constraints on stress-strain fields along the Mid-Atlantic Ridge, Iceland. Geology 2004, 32, 49–52. [Google Scholar] [CrossRef]
  48. Dewey, J.F.; Helman, M.L.; Turco, E.; Hutton, D.H.W.; Knott, S.D. Alpine tectonics. In Alpine Tectonics; Coward, M.P., Dietrich, D., Park, R.G., Eds.; Special Publications; Geological Society: London, UK, 1989; Volume 45, pp. 265–283. [Google Scholar]
  49. Platt, J.P.; Behrmann, J.H.; Cunningham, P.C.; Wallis, S.; Western, P.J. Kinematics of the Alpine arc and the motion history of Adria. Nature 1993, 337, 158–161. [Google Scholar] [CrossRef]
  50. Ziegler, P.A.; Cloetingh, S.; van Wees, J.-D. Dynamics of intra-plate Compressional deformation: The Alpine foreland and other examples. Tectonophysics 1995, 252, 7–59. [Google Scholar] [CrossRef]
  51. Frisch, W. The plate tectonic evolution of the Alps. Tectonophysics 1979, 60, 121–134. [Google Scholar] [CrossRef]
  52. Escher, A.; Hunziker, J.C.; Marthaler, M.; Masson, H.; Sartori, M.; Steck, A. Geologic framework and structural evolution of the western Swiss-Italian Alps. In Deep Structure of the Swiss Alps: Results of NRP 20; Birkhauser: Basel, Switzerland, 1997; pp. 205–222. [Google Scholar]
  53. Pfiffner, A.O.; Ellis, S.C.; Beaumont, C. Collisional tectonics in the Swiss Alps: Insights from geodynamic modeling. Tectonics 2000, 19, 1065–1094. [Google Scholar] [CrossRef] [Green Version]
  54. Ring, U.; Brandon, M.T.; Ramthun, A. Soultion-mass-transfer deformation adjacent to the Glarus Thrust, with implications for the tectonic evolution of the Alpine wedge in eastern Switzerland. J. Struct. Geol. 2001, 23, 1491–1505. [Google Scholar] [CrossRef]
  55. Carrapa, B. Tracing exhumation and orogenic wedge dynamics in the Alps via detrital thermochronology. Geology 2009, 37, 1127–1130. [Google Scholar] [CrossRef]
  56. Carrapa, B.; DiGiulio, A.; Mancin, N.; Gupta, S.; Fantoni, R.; Stockli, D. Tectonic significance of Cenozoic deformation, exhumation and basin history in the Western Alps. Tectonics 2016, 35, 1892–1912. [Google Scholar] [CrossRef] [Green Version]
  57. Mock, S.; von Hagke, C.; Schlunegger, F.; Dunkl, I.; Herwegh, M. Long-wavelength late Miocene thrusting in the north Alpine foreland: Implications for orogenic processes. Solid Earth 2020, 11, 1823–1847. [Google Scholar] [CrossRef]
  58. Mancktelow, N.S. Neogene lateral extension during convergence in the Central Alps: Evidence from interrelated faulting and backfolding around Simplonpass (Switzerland). Tectonophysics 1992, 215, 295–317. [Google Scholar] [CrossRef]
  59. Homewood, P.; Allen, P.A.; Williams, G.D. Dynamics of the Molasse Basin of Western Switzerland. In Foreland Basins; Special Publications of the International Association of Sedimentologists; Wiley Online Library: Hoboken, NJ, USA, 1986; Volume 8, pp. 199–217. [Google Scholar]
  60. Schmid, S.M.; Pfiffner, O.A.; Schönborn, G.; Froitzheim, N.; Kissling, E. Integrated cross section and tectonic evolution of the Alps along the Eastern Traverse. In Deep structure of the Swiss Alps: Results of NRP, 20; Birkhauser Verlag: Basel, Switzerland, 1997; pp. 289–304. [Google Scholar]
  61. Illies, J.H. Recent and paleo-intraplate tectonics in stable Europe and the Rhinegraben rift system. Tectonophysics 1975, 29, 251–264. [Google Scholar] [CrossRef]
  62. Illies, J.H.; Greiner, G. Rhinegraben and the Alpine system. Bull. Geol. Soc. Am. 1978, 85, 770–782. [Google Scholar] [CrossRef]
  63. Ziegler, P.A. Cenozoic rift of western and central Europe: An overview. Geol. Mijnb. 1994, 73, 99–127. [Google Scholar]
  64. Platt, J.P. Dynamics of orogenic wedges and the uplift of high-pressure metamorphic rocks. Geol. Soc. Am. Bull. 1986, 97, 1037–1053. [Google Scholar] [CrossRef]
  65. Royden, L.H. Evolution of retreating subduction boundaries formed during continental collision. Tectonics 1993, 12, 629–638. [Google Scholar] [CrossRef]
  66. Wells, M.L.; Hoisch, T.D.; Cruz-Uribe, A.M.; Vervoort, J.D. Geodynamics of synconvergen extension and tectonic mode switching: Constraints from the Sevier-Laramide Orogen. Tectonics 2012, 31, TC1002. [Google Scholar] [CrossRef] [Green Version]
  67. Molnar, P. Gravitational instability of mantle lithosphere and core complexes. Tectonics 2015, 34, 478–487. [Google Scholar] [CrossRef]
  68. Siddans, A.W.B. Deformation, metamorphism and texture development in Permian mudstones of the Glarus Alps (eastern Switzerland). Eclogae Geol. Helv. 1979, 72, 601–621. [Google Scholar]
  69. Ramsay, J.G. Tectonics of the Helvetic Nappes. In Thrust and Nappe Tectonics; The Geological Society of London: London, UK, 1981; pp. 293–309. [Google Scholar]
  70. Ramsay, J.G.; Huber, M.I. The Techniques of Modern Structural Geology, Volume 1: Strain Analysis; Academic Press: Cambridge, MA, USA, 1983; p. 307. [Google Scholar]
  71. Pfiffner, A.O. Kinematics and intrabed-strain in mesoscopically folded limestone layers: Examples from the Jura and Helvetic zone of the Alps. Eclogae Geol. Helv. 1990, 83, 585–602. [Google Scholar]
  72. Badertscher, N.P.; Beaudoin, G.; Therrien, R.; Burkhard, M. Glarus overthrust: A major pathway for the escape of fluids out of the Alpine orogen. Geology 2002, 30, 875–878. [Google Scholar] [CrossRef] [Green Version]
  73. Groshong, R.H.; Pfiffner, A.O.; Pringle, L.R. Strain partitioning in the Helvetic thrust belt of eastern Switzerland from leading edge to the internal zone. J. Struct. Geol. 1984, 6, 5–18. [Google Scholar] [CrossRef]
  74. Kligfield, R.; Owens, W.H.; Lowrie, W. Magnetic susceptibility anisotropy, strain and progressive deformation in Permian sediments from the Maritime Alps (France). Earth Planet. Sci. Lett. 1981, 55, 181–189. [Google Scholar] [CrossRef]
  75. Schmid, S.M.; Casey, M.; Starkey, J. The microfabric of calcite from the Helvetic nappes (Swiss Alps). In Thrust and Nappe Tectonics; McClay, K.R., Price, N.J., Eds.; Special Publications; Geological Society: London, UK, 1981; Volume 9, pp. 151–158. [Google Scholar]
  76. Heitzmann, P. Calcite mylonites in the central alpine “root zone”. Tectonophysics 1987, 135, 207–215. [Google Scholar] [CrossRef]
  77. Marret, R.; Allmendinger, R.W. Kinematic analysis of fault-slip data. J. Struct. Geol. 1990, 12, 973–986. [Google Scholar] [CrossRef]
  78. Allmendinger, R.W.; Cardozo, N.C.; Fisher, D. Structural Geology Algorithms: Vectors & Tensors; Cambridge University Press: Cambridge, UK, 2012; p. 289. [Google Scholar]
  79. Rocher, M.; Cushing, M.; Lemeille, F.; Lozac’h, Y.; Angelier, J. Intraplate paleostresses reconstructed with calcite twinning and faulting: Improved method and application to the eastern Paris Basin (Lorraine, France). Tectonophysics 2004, 387, 1–21. [Google Scholar] [CrossRef]
  80. Rocher, M.; Cushing, M.; Lemeille, F.; Baize, S. Stress induced by the Mio-Pliocene Alpine collision in northern France. Bull. Soc. Géol. Fr. 2005, 176, 319–328. [Google Scholar] [CrossRef]
  81. Larroque, J.M.; Laurent, P. Evolution of the stress field pattern in the south Rhine Graben from the Eocene to the Present. Tectonophysics 1988, 148, 41–58. [Google Scholar] [CrossRef]
  82. Lacombe, O.; Angelier, J.; Byrne, D.; Dupin, J.M. Eocene-Oligocene Tectonics and Kinematics of the Rhine-Saone Continental Transform Zone (Eastern France). Tectonics 1993, 12, 874–888. [Google Scholar] [CrossRef] [Green Version]
  83. Groshong, R.H., Jr. Strain calculated from twinning in calcite. Bull. Geol. Soc. Am. 1972, 83, 2025–2038. [Google Scholar] [CrossRef]
  84. Turner, F.J. Nature and dynamic interpretation of deformation lamellae in calcite of three marbles. Am. J. Sci. 1953, 251, 276–298. [Google Scholar] [CrossRef]
  85. Turner, F.J. Compression and tension axes deduced from (0112) Twinning in calcite. J. Geophys. Res. 1962, 67, 1660. [Google Scholar]
  86. Paulsen, T.S.; Wilson, T.J.; Demosthenous, C.; Millan, C.; Jarrad, R.; Laufer, A. Kinematics of the Neogene Terror rift: Constraints from calcite twinning strains in the ANDRILL McMurdo Ice Shelf (AND-1B) core, Victoria Land Basin, Antarctica. Geosphere 2014, 10, 828–841. [Google Scholar] [CrossRef]
  87. Engelder, T. The nature of deformation within the outer limits of the central Appalachian foreland fold-and-thrust belt in New York state. Tectonophysics 1979, 55, 289–310. [Google Scholar] [CrossRef]
  88. Spang, J.H.; Groshong, R.H., Jr. Deformation mechanisms and strain history of a minor fold from the Appalachian Valley and Ridge Province. Tectonophysics 1981, 72, 323–342. [Google Scholar] [CrossRef]
  89. Groshong, R.H., Jr. Strain, fractures, and pressure solution in natural single-layer folds. Bull. Geol. Soc. Am. 1975, 86, 1363–1376. [Google Scholar] [CrossRef]
  90. Groshong, R.H., Jr. Strain and pressure solution in the Martinsburg slate, Delaware Water Gap, New Jersey. Am. J. Sci. 1976, 276, 1131–1146. [Google Scholar] [CrossRef]
  91. Ferrill, D.A. Calcite twin widths and intensities as metamorphic indicators in natural low-temperature deformation of limestone. J. Struct. Geol. 1991, 13, 675–677. [Google Scholar] [CrossRef]
  92. Craddock, J.P.; Jackson, M.; van der Pluijm, B.A.; Versical, R. Regional shortening fabrics in eastern North America: Far-field stress transmission from the Appalachian–Ouachita orogenic belt. Tectonics 1993, 12, 257–264. [Google Scholar] [CrossRef]
  93. Craddock, J.P.; Neilson, K.J.; Malone, D.H. Calcite twinning strain constraints on Heart Mountain detachment kinematics, Wyoming. J. Struct. Geol. 2000, 22, 983–991. [Google Scholar] [CrossRef]
  94. Craddock, J.P.; McKiernan, A.; DeWit, M. Calcite twin analysis in synorogenic calcite, Cape Fold Belt: Implications for fold rotation and cleavage formation. J. Struct. Geol. 2007, 27, 1100–1113. [Google Scholar] [CrossRef]
  95. Amrouch, K.; Lacombe, O.; Bellahsen, N.; Daniel, J.-M.; Callot, J.-P. Stress and strain patterns, kinematics and deformation mechanisms in a basement-cored anticline: Sheep Mountain Anticline, Wyoming. Tectonics 2010, 29, TC1005. [Google Scholar] [CrossRef] [Green Version]
  96. Craddock, J.P.; Geary, J.; Malone, D.H. Vertical Injectites of Detachment Carbonate Ultracataclasite at White Mountain, Heart Mountain Detachment, Wyoming. Geology 2012, 40, 463–466. [Google Scholar] [CrossRef]
  97. Craddock, J.P.; Craddock, S.D.; Konstantinou, A.; Kylander-Clark, A.; Malone, D.H. Calcite Twinning Strain Variations across the Proterozoic Grenville Orogen and Keweenaw-Kapuskasing Inverted Foreland, USA and Canada. Geosci. Front. 2017, 8, 1357–1384. [Google Scholar] [CrossRef]
  98. Craddock, J.P.; Pearson, A. Non-coaxial horizontal shortening strains preserved in amygdule calcite, DSDP Hole 433C, Suiko Seamount. J. Struct. Geol. 1994, 16, 719–724. [Google Scholar] [CrossRef]
  99. Craddock, J.P.; Pearson, A.; McGovern, M.; Kropf, E.P.; Moshoian, A.; Donnelly, K. Post-extension shortening strains preserved in calcites of the Keweenawan rift. In Middle Proterozoic to Cambrian Rifting, Central North America; Geological Society of America Special Paper, 213; Ojakgangas, R.W., Dickas, A.B., Green, J.C., Eds.; Springer: Dordrecht, The Netherlands, 1997; pp. 115–126. [Google Scholar]
  100. Craddock, J.P.; Anziano, J.; Wirth, K.R.; Vervoort, J.D.; Singer, B.; Zhang, X. Structure, geochemistry and geochronology of a lamprophyre dike swarm, Archean Wawa terrane, Michigan, USA. Precambrian Res. 2007, 157, 50–70. [Google Scholar] [CrossRef]
  101. Nuriel, P.; Weinberger, R.; Kylander-Clark, A.R.C.; Hacker, B.C.; Craddock, J.P. Calcite U-Pb ages constrain the evolution of the Dead Sea Fault. Geology 2017, 45, 587–590. [Google Scholar] [CrossRef] [Green Version]
  102. Weinberger, R.; Nuriel, P.; Kylander-Clark, A.R.; Craddock, J.P. Temporal and spatial relations between large-scale fault systems: Evidence from the Sinai-Negev shear zone and the Dead Sea Fault. Earth-Sci. Rev. 2020, 211, 103377. [Google Scholar] [CrossRef]
  103. Craddock, J.P.; Nuriel, P.; Kylander-Clark, A.R.; Hacker, B.R.; Luczaj, J.; Weinberger, R. Long-term (7 Ma) Strain Fluctuations within the Dead Sea Transform from High-resolution U-Pb Dating of a Calcite Vein. Bulletin 2022, 134, 1231–1246. [Google Scholar] [CrossRef]
  104. Lacombe, O.; Parlangeau, C.; Beaudoin, N.; Amrouch, K. Calcite twin formation, measurement ad use as stress-strain indicators: A review of progress over the last decade. Geosciences 2021, 11, 445. [Google Scholar] [CrossRef]
  105. Wenk, H.-R.; Takeshita, T.; Bechler, E.; Erskine, B.G.; Matthies, S. Pure Shear and Simple Shear Calcite Textures. Comparison of Experimental, Theoretical and Natural Data. J. Struct. Geol. 1987, 9, 731–745. [Google Scholar] [CrossRef]
  106. Burkhard, M. Calcite twins, their geometry, appearance and significance as stress-strain markers and indicators of tectonic regime: A review. J. Struct. Geol. 1993, 15, 351–368. [Google Scholar] [CrossRef] [Green Version]
  107. Lacombe, O.; Laurent, P. Determination of deviatoric stress tensors based on inversion of calcite twin data from experimentally deformed monophase samples: Preliminary results. Tectonophysics 1996, 255, 189–202. [Google Scholar] [CrossRef]
  108. Ferrill, D.A. Critical re-evaluation of differential stress estimates from calcite twins in coarse-grained limestone. Tectonophysics 1998, 285, 77–86. [Google Scholar] [CrossRef]
  109. Groshong, R.H., Jr. Experimental test of least-squares strain calculations using twinned calcite. Bull. Geol. Soc. Am. 1974, 85, 1855–1864. [Google Scholar] [CrossRef]
  110. Rowe, K.J.; Rutter, E.H. Paleostress estimation using calcite twinning: Experimental calibration and application to nature. J. Struct. Geol. 1990, 12, 1–17. [Google Scholar] [CrossRef]
  111. Teufel, L.W. Strain analysis of experimental superposed deformation using Calcite twin lamellae. Tectonophysics 1980, 65, 291–309. [Google Scholar] [CrossRef]
  112. Evans, M.A.; Groshong, R.H. A Computer Program for the Calcite Strain-Gage Technique. J. Struct. Geol. 1994, 16, 277–281. [Google Scholar] [CrossRef]
  113. Ferrill, D.A.; Morris, A.P.; Evans, M.A.; Burkhard, M.; Groshong, R.H.; Onasch, C.M. Calcite Twin Morphology: A Low-Temperature Deformation Geothermometer. J. Struct. Geol. 2004, 26, 1521–1529. [Google Scholar] [CrossRef] [Green Version]
  114. Gray, M.B.; Stamatakos, J.A.; Ferrill, D.A.; Evans, M.A. Fault-Zone Deformation in Welded Tuffs at Yucca Mountain, Nevada, USA. J. Struct. Geol. 2005, 27, 1873–1891. [Google Scholar] [CrossRef]
  115. Craddock, J.P.; Malone, D.H.; Wartman, J.; Kelly, M.J.; LJunlai, L.; Bussolotto, M.; Invernizzi, C.; Knott, J.; Porter, R. Calcite Twinning Strains from Syn-faulting Calcite Gouge: Small-Offset Strike-Slip, Normal and Thrust Faults. Int. J. Earth Sci. 2020, 109, 1–42. [Google Scholar] [CrossRef]
  116. Roberts, N.M.W.; Walker, R.J. U-Pb geochronology of calcite-mineralized faults: Absolute timing of rift-related fault events on the northeast Atlantic margin. Geology 2016, 44, 531–534. [Google Scholar] [CrossRef]
  117. Schulz, B. Deformation, Metamorphose and Petrographie im Ostalpinen Altkristallin Sudlich des Taurenfensters. Ph.D. Thesis, University of Erlangen-Nuremberg, Erlangen, Germany, 1988; p. 133. [Google Scholar]
  118. Rosenberg, C.L.; Brun, J.-P.; Gapais, D. Indentation model of the eastern Alps and the origin of the Tauren Window. Geology 2004, 32, 997–1000. [Google Scholar] [CrossRef]
  119. Lacombe, O.; Angelier, J.; Laurent, P. Determining paleostress orientations from faults and calcite twins: A case study near the Sainte-Victoire Range (southern France). Tectonophysics 1992, 201, 141–156. [Google Scholar] [CrossRef]
  120. Burkhard, M. Aspects of large-scale Miocene deformation in the external part of the Swiss Alps (Subalpine Molasse to Jura fold belt). Eclogae Geol. Helv. 1990, 83, 559–583. [Google Scholar]
  121. Homberg, C.; Hu, J.C.; Angelier, J.; Bergerat, F.; Lacombe, O. Characterization of stress perturbations near major fault zones: Insights from 2-D distinct-element numerical modeling and field studies (Jura Mountains). J. Struct. Geol. 1997, 19, 703–718. [Google Scholar] [CrossRef]
  122. Becker, A. In Situ stress data from the Jura Mountains-new results and interpretation. Terra Nova 1999, 11, 9–15. [Google Scholar] [CrossRef]
  123. Homberg, C.; Lacombe, O.; Angelier, J.; Begerat, F. New constraints for indentation mechanisms in arcuate belts from the Jura Mountains, France. Geology 1999, 27, 827–830. [Google Scholar] [CrossRef]
  124. House, M.R. The Structure of Weymouth Anticline. Proc. Geol. Assoc. 1961, 72, 211–238. [Google Scholar] [CrossRef]
  125. House, M.R. Geology of the Dorset Coast. Geologists’ Association Guide, 22; Geologists’ Association: London, UK, 1986. [Google Scholar]
  126. Sanderson, D.J.; Dix, J.K.; Westhead, K.R.; Collier, J.S. Bathymetric mapping of the coastal and offshore geology and structure of the Jurassic Coast, Weymouth Bay, UK. J. Geol. Soc. Lond. 2017, 174, 498–508. [Google Scholar] [CrossRef]
  127. Lyle, P. A Geological Excursion Guide to the Giant’s Causeway; Baird Publishers: London, UK, 1998; ISBN 0-9528258-1-3. [Google Scholar]
  128. Lyle, P. The North of Ireland; Terra Publishing: Arnhem, The Netherlands, 2003; ISBN 1-903544-08-4. [Google Scholar]
  129. Roberts, N.M.W.; Lee, J.K.; Holdsworth, R.E.; Jeans, C.; Farrant, A.R.; Haslam, R. Near-surface Palaeocene fluid flow, mineralisation and faulting at Flamborough Head, UK: New field observations and U-Pb calcite dating constraints. Solid Earth 2020, 11, 1931–1945. [Google Scholar] [CrossRef]
  130. Lacombe, O.; Laurent, P.; Angelier, J. Calcite twins as a key to paleostresses in sedimentary basins: Preliminary results from drill cores of the Paris basin. In Peri-Tethyan Platforms; Roure, F., Ed.; Technip: Paris, France, 1994; pp. 197–210. [Google Scholar]
  131. Ratschbacher, L. Kinematics of Austro-Alpine cover nappes: Changing translation path due to transpression. Tectonophysics 1986, 125, 335–356. [Google Scholar] [CrossRef]
  132. Beltrando, M.; Lister, G.; Hermann, J.; Forster, M.; Compagnoni, R. Deformation mode switches in the Penninic units of the Urtier Valley (Western Alps): Evidence for a dynamic orogen. J. Struct. Geol. 2008, 30, 194–219. [Google Scholar] [CrossRef]
  133. Craddock, J.P.; Klein, T.; Kowalczyk, G.; Zulauf, G. Calcite twinning strains in Alpine orogen flysch: Implications for thrust-nappe mechanics and the geodynamics of Crete. Lithosphere 2009, 1, 174–191. [Google Scholar] [CrossRef] [Green Version]
  134. Klein, T.; Craddock, J.P.; Zulauf, G. Constraints on the geodynamical evolution of Crete: Insights from illite crystallinity, Raman spectroscopy and calcite twinning above and below the “Cretan detachment”. Int. J. Earth Sci. 2013, 102, 139–182. [Google Scholar] [CrossRef]
  135. Burkhard, M. L’Helvétique de la bordure occidentale du massif de l’Aar (évoltion tectonique et métamorphique). Eclogae Geol. Helv. 1988, 81, 63–114. [Google Scholar]
  136. Craddock, J.P.; McKiernan, A.W. Tectonic implications of finite strain gradient in Baraboo-interval quartzites (ca. 1700 Ma), Mazatzal orogen, Wisconsin and Minnesota, USA. Precambrian Res. 2007, 156, 175–194. [Google Scholar] [CrossRef]
  137. Craddock, J.P.; Malone, D.H.; Konstantinou, A.; Spruell, J.; Porter, R. Calcite twinning strains associated with Laramide uplifts, Wyoming Province. In Tectonic Evolution of the Sevier-Laramide Hinterland, Thrust Belt, and Foreland, and Postorogenic Slab Rollback (180–20 Ma); Special Paper 555; Craddock, J.P., Malone, D.H., Foreman, B.Z., Konstantinou, A., Eds.; Geological Society of America: Boulder, CO, USA, 2022; pp. 149–192. [Google Scholar] [CrossRef]
  138. Daly, M.C.; Lawrence, S.R.; Kimun’a, D.; Binga, M. Late Paleozoic deformation in central Africa: A result of distant collision? Nature 1991, 350, 605–607. [Google Scholar] [CrossRef]
  139. Stampfli, G.M.; Hochard, C. Plate Tectonics of the Alpine Realm: Ancient Orogens and Modern Analogues; The Geological Society of London: London, UK, 2009; Volume 327, pp. 89–111. [Google Scholar]
  140. Le Breton, E.; Handy, M.R.; Molli, G.; Ustaszewski, K. Post-20 Ma motion of the Adriatic plate: New constraints from surrounding Orogens and implications for crust-mantle decoupling. Tectonics 2017, 36, 3135–3154. [Google Scholar] [CrossRef] [Green Version]
  141. Ring, U.; Ratschbacher, L.; Frisch, W.; Biehler, D.; Kralik, M. Kinematics of the Alpine plate margin: Structural styles, strain and motion along the Penninic-Austroalpine boundary in the Swiss-Austrian Alps. J. Geol. Soc. Lond. 1989, 146, 835–849. [Google Scholar] [CrossRef]
  142. Neubauer, F.; Genser, J.; Handler, R. The Eastern Alps: Result of a two-stage collision process. Mitt. Osterr. Geol. Ges. 1999, 92, 117–134. [Google Scholar]
  143. Howe, T.C.; Bird, P. Exploratory models of long-term crustal flow and resulting seismicity in the Alpine-Aegean orogen. Tectonics 2010, 29, TC4023. [Google Scholar] [CrossRef]
  144. Pfiffner, O.A.; Burkhard, M. Determination of paleo-stress axes Orientations from fault, twin and earthquake data. Ann. Tecton. 1987, 1, 48–57. [Google Scholar]
Figure 12. Northwestern Europe with the Mid-Atlantic Ridge (orange line) defining the active divergent margin. Calcite twinning strain axes are presented by color (oldest to youngest): opening of the Atlantic Ocean (58 Ma; purple), and Alpine shortening (30 Ma, blue). Faroe Islands data from Roberts and Walker (2016) with Atlantic opening calcite ages ranging from 44–37 Ma. Flamborough Head preserves deformation ~63–55 Ma Roberts et al. [129]. Key: Colored lines represent layer-parallel shortening, circles are layer-normal shortening in the Alpine nappes (Table 2, Figure 7 and Figure 8).
Figure 12. Northwestern Europe with the Mid-Atlantic Ridge (orange line) defining the active divergent margin. Calcite twinning strain axes are presented by color (oldest to youngest): opening of the Atlantic Ocean (58 Ma; purple), and Alpine shortening (30 Ma, blue). Faroe Islands data from Roberts and Walker (2016) with Atlantic opening calcite ages ranging from 44–37 Ma. Flamborough Head preserves deformation ~63–55 Ma Roberts et al. [129]. Key: Colored lines represent layer-parallel shortening, circles are layer-normal shortening in the Alpine nappes (Table 2, Figure 7 and Figure 8).
Geosciences 12 00254 g012
Table 1. Summary Calcite Twinning Strain Data.
Table 1. Summary Calcite Twinning Strain Data.
LocationAgeSamplesLS/VeinAvg.
NEV (%)
Avg.—e1 (%)Avg. e1
(Tr and Plunge)
Avg. ds (Bars)LNS/LPSReferences
Atlantic Margin
(N. Ireland)
58 Ma16
(n = 454)
0/168−3.4150°, 1°−3870/16This Study; U-Pb ages (Figure 4)
Alp Foreland30 Ma32
(n = 1648)
11/1614−4.98310°, 5°−3434/28This Study (Figure 5)
Jura Mountains30 Ma10
(n = 612)
10/026−0.32(See Figure 7)No Data0/28[36]
Molasse Basin30 Ma26
(n = 1292)
26/021−0.6(See Figure 7)No Data1/25[36,38]
Pre-Alps30 Ma39
(n = 2081)
26/1330−4.88(See Figure 7)No Data20/19[39]
Helvetic Nappes30 Ma17
(n = 112)
13/416−3.66(See Figure 7)−37210/7[45,73]
Helvetic–Penninic Nappes30 Ma29
(n = 1450)
29/026−6.6(See Figure 7)No Data13/16[40]
Austroalpine Nappes (Tauren)30 Ma4
(n = 1000)
Calcsilicate26−3.84Vertical−3384/0This Study (Figure 6 and Figure 8)
IcelandActive19
(n = 430)
0/1915−6.02Ridge-normal−4803/19[47]
Gubbio Fault, ItalyActive17
(n = 840)
0/174−4.23340°, 5°−3720/17[115]
TOTALS 209
(n = 9919)
119/90
Key: LPS is layer-parallel shortening; LNS is layer-normal shortening.
Table 2. Calcite strain data.
Table 2. Calcite strain data.
SampleSection IDLocationRock TypeBeddingVeinGrains (n = )e1e2e3e1(%)NEV (%)Twins/mmDs (Bars)Fabric (Bedding)Fabric (Vein)Orogenic Distance
Atlantic Margin (N. Ireland)
24 (North)PUE22Port Rush, IrelandVein in BasaltHorizontalN-S, 90°1258°, 0°328°, 2°174°, 87°−6.38360385.1277LPSVNS1250 km
24PUE23Port Rush, IrelandVein in BasaltHorizontalN-S, 90°2226°, 2°296°, 4°132°, 86°−3.418456402.683LPSVNS1250
24PUE24Port Rush, IrelandVein in BasaltHorizontal
24PUE25Port Rush, IrelandVein in BasaltHorizontalN-S, 90°24186°, 3°276°, 3°48°, 85°−1.98446401.0363LPSVPS1250
24PUE26Port Rush, IrelandVein in BasaltHorizontalN-S, 90°2456°, 4°146°, 2°261°, 85°−3.44389390.8814LPSVNS1250
2422, 23, 25Port Rush, IrelandVein in BasaltHorizontalN-S, 90°580°, 0°270°, 1°155°, 89°−2.72432398.6677LPSVPS1250
2422, 26Port Rush, IrelandVein in BasaltHorizontalN-S, 90°369°, 1°279°, 2°140°, 87°−2.511416395.865LPSVPS1250
25PUE17Ballintoy Harbor, IrelandLS/Vein87°, 31° NN-S, 90°26257°, 1°347°, 1°104°, 88°−3.68421396.7522LPSVNS1250
25PUE18Ballintoy Harbor, IrelandLS/Vein87°, 31° NE-W, 90°20209°, 1°300°, 2°88°, 86°−4.720295370.3396LPSVNS1250
25PUE19Ballintoy Harbor, IrelandVein in LS87°, 31° NN-S, 90°25211°, 1°302°, 1°68°, 88°−6.20278365.9317LPSVNS1250
25PUE20Ballintoy Harbor, IrelandVein in BasaltHorizontalE-W, 90°2540°, 2°311°, 1°181°, 88°−4.212548416.3315LPSVNS1250
25PUE21Ballintoy Harbor, IrelandVein in BasaltHorizontalE-W, 90°2212°, 1°101°, 3°273°, 86°−4.40469404.7706LPSVNS1250
2520,21Ballintoy Harbor, IrelandVein in BasaltHorizontalE-W, 90°4728°, 2°301°, 3°143°, 86°−4.36508410.7027LPSVNS1250
23PUE13Moneymore, IrelandLS/VeinHorizontal0°, 90°23250°, 19°159°, 1°65°, 70°−1.911407394.2406LPSVNS1200
23PUE14Moneymore, IrelandLS/VeinHorizontal
23PUE15Moneymore, IrelandLS/VeinHorizontal0°, 90°20252°, 4°162°, 1°67°, 85°−1.620237354.082LPSVPS1200
23PUE16Moneymore, IrelandLS/VeinHorizontal0°, 90°25266°, 1°354°, 1°152°, 88°−1.84261361.2455LPSVNS1200
2315 & 16Moneymore, IrelandVeinsHorizontal0°, 90°45218°, 1°308°, 2°85°, 87°−1.70252358.6395LPSVNS1200
454 −3.41258.25 387.9561
Alpine Foreland (NW)
PUE1Hull, UKLS/VeinHorizontal No twins
22OSE15Gloucester, UKLS22°, 12° SE 1222°, 8°112°, 1°206°, 81°−2.325312374.5004LPS 800
22OSE16Gloucester, UKVein in LS22°, 12° SE72°, 90°18340°, 23°79°, 5°175°, 65°−8.511385390.1138LPSVNS800
2215 & 16Gloucester, UKLS/Vein22°, 12° SE72°, 90°30350°, 10°266°, 1°176°, 80°−13.320373387.7622LPSVNS800
21OSE13Bath, UKLS45°, 7° SE No twins
21OSE14Bath, UKLS45°, 7° SE 13310°, 21213°, 12°95°, 65°−4.323162325.8271LPS 750
20PUE11Cheddar Gorge, UKVein12°, 15° SHorizontal23219°, 23°128°, 1°38°, 66°−5.911262361.5295LPSVPS715
20PUE12Cheddar Gorge, UKLS12°, 15° SHorizontal27334°, 4°64°, 1°183°, 86°−3.111337380.2247LPS 715
20PUE12Cheddar Gorge, UKVein12°, 15° S61°, 48° S26158°, 1°68°, 2°261°, 88°−1.93357384.5063LPSVNS715
19PUE10Birdport, UKLSHorizontal No twins
18PUE8Lulworth Cove, UKVeinHorizontalHorizontal27176°, 4°84°, 1°303°, 85°−5.50307373.3007LPSVPS680
18PUE9Lulworth Cove, UKLS90°, 23° N No twins
176&7 PEVCorfe Castle, UKVeinHorizontal110°, 82° S39120°, 76°260°, 10°352°, 8°−1.90164326.7383LNSVNS680
176&7 NEVCorfe Castle, UKVeinHorizontal110°, 82° S23210°, 75°358°, 11°90°, 7°−4.8100125306.5716LPSVNS680
15PUE2W. Isle of Wight, UKLS/VeinHorizontalHorizontal20153°, 19°58°, 7°318°, 55°−8.515213346.1529LPSVPS610
15PUE3W. Isle of Wight, UKLS/VeinHorizontalHorizontal9143°, 6°234°, 5°1°, 81°−7.90250358.0477LPSVPS610
16PUE2&3W. Isle of Wight, UKLS/VeinHorizontalHorizontal29153°, 11°61°, 5°315°, 71°−9.431220348.5543LPSVPS610
PUE4W. Isle of Wight, UKLS/VeinHorizontal No twins
PUE5E. Isle of Wight, UKLS/VeinHorizontal No twins
14OSE17Audreselles, FranceVeinHorizontal341°, 90°18260°, 23°95°, 65°351°, 5°−2.80229351.5319LPSVNS510
14OSE18Audreselles, FranceVeinHorizontal0°, 90°18132°, 2°41°, 10°238°, 79°−2.816245356.5474LPSVNS510
14OSE19Audreselles, FranceVeinHorizontal0°, 90°20175°, 29°79°, 12°339°, 62°−5.720173330.7059LPSVPS510
1418 &19Audreselles, FranceVeinsHorizontal0°, 90°54145°, 18°53°, 4°310°, 70°−6.231208344.3888LPSVNS480
14OSE19Audreselles, FranceLSHorizontal 18125°, 1°28°, 3°223°, 86°−1.80246356.8499LPS 480
13OSE20Boulogne, FranceVeinHorizontal0°, 90°23170°, 11°269°, 1°3°, 78°−1.80259360.6743LPSVPS480
13OSE21Boulogne, FranceVeinHorizontal0°, 90°20201°, 5°291°, 1°32°, 84°−5.910172330.2754LPSVPS480
1320&21Boulogne, FranceVeinsHorizontal0°, 90°43193°, 6°103°, 11°, 82°−5.72224349.8924LPSVPS480
OSE22Paris, FranceLSHorizontal No twins
OSE23Paris, FranceLSHorizontal No twins
12OSE12Pottenstein, GermanyLSHorizontal No twins
11OSE11PEVGrafenberg, GermanyVein in LSHorizontal343°, 90°31345°, 9°77°, 14°223°, 73°−1.30194339.2141LPSVPS230
11OSE11NEVGrafenberg, GermanyVein in LSHorizontal343°, 90°19358°, 62°99°, 5°192°, 27°−9.6100333379.338LNSVNS230
5OSE5Saints-Geosmes, FranceLSHorizontal 24181°, 1°271°, 7°81°, 82°−1.10210345.0995LPS 160
6OSE6&7NChatenois, FranceLSHorizontal 2111°, 76°181°, 12°272°, 2°−2.6100164326.7383LNS 170
7OSE6&7PGraux, FranceLSHorizontal 26172°, 7°262°, 4°22°, 81°−13.430204342.9468LPS 140
8OSE8Nancy, FranceLSHorizontal 18132°, 5°40°, 10°250°, 78°−3.15110297.0781LPS 130
9OSE9Wiesensteig, GermanyLSHorizontal No twins 100
10OSE10Solnhofen, GermanyLSHorizontal No twins 145
4OSE3 & 4NN. Champlitte, FranceLSHorizontal 19167°, 63°16°, 23°281°, 11°−3.14100100290LNS 140
2OSE3 & 4PChamplitte, FranceLSHorizontal 25330°, 30°48°, 24°166°, 46°−2.940117301.6598LPS 100
2OSE1 & 2NTaxenne, FranceLSHorizontal 27136, 13°226°, 1°351°, 56°−1.56100131310.0534LPS 30
1(South)OSE1 & 2PBesancon, FranceLSHorizontal 20348°, 3°252°, 1°147°, 86°−2.540164326.7383LPS 5
1648 −4.8814.25 345.276
Jura Mtns.[36]
5 Mamirolle, FranceJurassic LS 6220°, 17°177°, 71°288°, 7°−0.527 LPS
6 Mamirolle, FranceJurassic LS 4261°, 40°173°, 23°284°, 41°−1.933 LPS
11 Mamirolle, FranceJurassic LS 63345°, 13°248°, 27°98°, 58°−0.830 LPS
19 Mamirolle, FranceJurassic LS 578°, 3°270°, 27°137°, 86°−0.610 LPS
21 Mamirolle, FranceJurassic LS 60190°, 37°22°, 53°284°, 6°−3.415 LPS
5 Dole, FranceJurassic 43143°, 59245°, 6°339°, 29°1.118 LPS
8b Dole, FranceJurassic 73344°, 14°252°, 1°162°, 16°1.731 LPS
8c Dole, FranceJurassic 63192°, 13°292°, 35°85°, 51°0.731 LPS
9 Dole, FranceJurassic 71199°, 20°316°, 51°96°, 31°5.142 LPS
10 Dole, FranceJurassic 78191°, 4753°, 33°308°, 22°1.826 LPS
612 0.3226.3
Val-de-Ruz area ([36], Molasse Basin)
1 Val-de-Ruz, SwitzerlandValanginian80, 12S 30166°, 13°262°, 24°45°, 31°1.235 LPS
2 Val-de-Ruz, SwitzerlandMalm52, 47S 645°, 44°110°, 16°215°, 42°2.828 LNS
3 Val-de-Ruz, SwitzerlandMalm92, 8S 50157°, 16°248°, 4°32°, 61°2.122 LPS
4 Val-de-Ruz, SwitzerlandDogger73, 82S 57177°, 64°4°, 26°312°, 9°2.017 LPS
5 Val-de-Ruz, SwitzerlandMalm71, 28S 45207°, 18°60°, 70°129°, 12°1.120 LPS
6 Val-de-Ruz, SwitzerlandMalmHorizontal 43137°, 5°47°, 3°302°, 15°0.930 LPS
7 Val-de-Ruz, SwitzerlandValanginian83, 12S 44311°, 40°74°, 33°156°, 6°2.634 LPS
Molasse Basin[38]
dh4 Neuchatel, SwitzerlandMioceneHorizontal 5187°, 15°353°, 14°222°, 69°0.1520 LPS
dh5 Neuchatel, SwitzerlandMioceneHorizontal 53145°, 4°235°, 1°341°, 86°0.1423 LPS
dh8 Neuchatel, SwitzerlandMioceneHorizontal 51183°, 25°321°, 58°84°, 19°0.1627 LPS
dh12 Neuchatel, SwitzerlandMioceneHorizontal 51150°, 49°241°, 1°332°, 41°0.1220 −128LPS
dh15 Neuchatel, SwitzerlandMioceneHorizontal 5118°, 8°286°, 9°149°, 78°0.1416 −137LPS
dh17 Neuchatel, SwitzerlandMioceneHorizontal 5174°, 1°344°, 27°165°, 63°0.0929 LPS
dh2l Neuchatel, SwitzerlandMioceneHorizontal 51341°, 15°84°, 41°235°, 45°0.1512 −191LPS
dh2 Neuchatel, SwitzerlandMioceneHorizontal 51153°, 3°249°, 64°62°, 26°0.1331 LPS
dh22 Neuchatel, SwitzerlandMioceneHorizontal 51164°, 20°41°, 56°264°, 26°0.1418 LPS
dh24 Neuchatel, SwitzerlandMioceneHorizontal 51165°, 19°257°, 6°5°, 70°0.1122 −161LPS
dh26 Neuchatel, SwitzerlandMioceneHorizontal 51190°, 34°89°, 13°344°, 48°0.0929 LPS
dh28 Neuchatel, SwitzerlandMioceneHorizontal 51145°, 4°237°, 24°47°, 66°0.142 LPS
dh3l Neuchatel, SwitzerlandMioceneHorizontal 51145°, 17°343°, 72°237°, 5°0.0725 LPS
dh32 Neuchatel, SwitzerlandMioceneHorizontal 51144°, 8°54°, 4°297°, 81°0.2410 LPS
dh37 Neuchatel, SwitzerlandMioceneHorizontal 51116°, 6°19°, 50°211°, 40°0.1412 LPS
dh39 Neuchatel, SwitzerlandMioceneHorizontal 51108°, 14°18°, 2°281°, 76°0.1125 LPS
dh40 Neuchatel, SwitzerlandMioceneHorizontal 39260°, 12°352°, 7°109°, 76°0.1121 LPS
dh4l Neuchatel, SwitzerlandMioceneHorizontal 5140°, 2°309°, 4°153°, 86°0.1112 LPS
dh42 Neuchatel, SwitzerlandMioceneHorizontal 51232°, 24°141°, 3°44°, 86°0.110 −132LPS
1292 0.621.15385
Pre-Alps[39]
41 Pre-Alp Medianes nappeMalm38°, 47° N 53155°, 78°322°, 41°41°, 8°−5.1234 LNS
47 Pre-Alp Medianes nappeMalm21°, 46° N 51312°, 33°71°, 80°212°, 22°−2.7143 LPS
81 Pre-Alp Medianes nappeMalm, vein300°, 70° S?52120°, 23°45°, 3°302°, 62°−2.8129 LPS?
82 Pre-Alp Medianes nappeDogger, vein302°, 60° S?53161°, 12°261°, 5°348°, 81°−2.532 LNS?
84 Pre-Alp Medianes nappeFlysch105°, 88° N 54205°, 31°22°, 47°111°, 3°−3.7330 LNS
85V Pre-Alp Medianes nappeFlysch, vein ?58 −3.931 LPS?
85 Pre-Alp Medianes nappeFlysch71°, 45°N 52352°, 41°142°, 28°251°, 4°−3.8744 LPS
86 Pre-Alp Medianes nappeNeocomHorizontal 55201°, 11°311°, 41°92°, 46°−2.8818 LPS
87 Pre-Alp Medianes nappeNeocom88°, 47° S 55222°, 18°20°, 82° S142°, 3°−1.8646 LPS
88 Pre-Alp Medianes nappeNeocom, vein45°, 22°N?56118°, 72°2°, 3°122°, 81°−7.6820 LNS?
89 Pre-Alp Medianes nappeLias65°, 85°N 59120°, 821°, 1°270°, 5°−5.3636 LPS
90 Pre-Alp Medianes nappeMalm, vein300°, 45° S?73240°, 38°121°, 11°12°, 41°−2.8133 LPS?
91 Pre-Alp Medianes nappeDogger, vein100°, 70° S?52112°, 22°10°, 12°255°, 42°−3.9925 LPS?
92 Pre-Alp Medianes nappeDogger, vein100°, 70° S?57181°, 88°330°, 2°64°, 1°−1.9519 LPS?
93 Pre-Alp Medianes nappeDoggerHorizontal 54271°, 58°92°, 11°190°, 1°−3.2132 LNS
98 Pre-Alp Medianes nappeDogger52°, 88°N 5151°, 35°202°, 41°305°, 12°−4.2329 LPS
100 Pre-Alp Medianes nappeLias, veinHorizontal?46185°, 72°88°, 4°351°, 5°−3.3535 LNS?
102 Pre-Alp Medianes nappeDogger81°, 70°N 53168°, 5°271°, 45°72°, 5°−4.9111 LNS
105 Pre-Alp Medianes nappeFlysch91°, 80°N 99182°, 7°351°, 79°101°, 4°−5.6124 LNS
105V Pre-Alp Medianes nappeFlysch, vein ?51 −6.1710 LPS?
53 Pre-Alp Medianes nappeVein5°, 70°W?57123°, 47°227°, 44°352°, 41°−2.4639 LNS?
94 Pre-Alp Medianes nappeLias338°, 75° 50331°, 51°153°, 62°247°, 5°−6.2628 LNS
64 Pre-Alp Rigid nappeMalm38°, 31° S 4141°, 35°242°, 31°138°, 1°−3.7544 LNS
72 Pre-Alp Rigid nappeTrias51°, 20° S 52162°, 31°270°, 1°2°, 47°−1.8525 LPS
73 Pre-Alp Rigid nappeTrias, vein44°, 21° S?51122°, 20°203°, 2°303°, 62°−2.7233 LPS?
76 Pre-Alp Rigid nappeFlysch, vein21°, 62° S?57242°, 18°342°, 31°106°, 58°−232 LNS?
106 Pre-Alp Rigid nappeTriasHorizontal 50135°, 6°267°, 82°44°, 2°−5.4834 LPS
109 Pre-Alp Rigid nappeMalm, vein45°, 6° S?46135°, 88°272°, 2°22°, 3°−13.0535 LNS?
110 Pre-Alp Rigid nappeMalm45°, 23° S 45342°, 31°182°, 41°83°, 5°−8.3842 LPS
129 Pre-Alp Rigid nappeDogger2°, 47° E 47273°, 14°182°, 1°88°, 42°−9.721 LNS
132 Pre-Alp Rigid nappeMalm18°, 41° E 48233°, 51°26°, 21°131°, 3°−9.0523 LNS
133 Pre-Alp Rigid nappeMalm21°, 12° E 48282°, 62°177°, 6°83°, 12°−9.4813 LNS
135 Pre-Alp Rigid nappeMalmHorizontal 48186°, 9°311°, 41°82°, 22°−7.3338 LPS
136 Pre-Alp Rigid nappeTrias, vein38°, 28° S?48251°, 22°347°, 12°117°, 33°−3.0431 LNS?
137 Pre-Alp Rigid nappeTrias41°, 25° E 46252°, 47°87°, 30°352°, 8°−7.5426 LNS
78 Gurnigel nappeFlysch 53 −2.1726 LPS
74 Gurnigel nappeBreche nappeHorizontal 54136°, 2°52°, 5°302°, 83°−3.235 LPS
61 Gurnigel nappeFlysch51°, 46° S 50350°, 43°161°, 44°258°, 2°−7.0535 LNS
69 Gurnigel nappeFlysch12°, 41° E 56183°, 41°73°, 11°322°, 26°−7.1427 LNS
2081 −4.879529.94872 LNS = 20/
LPS = 19
Helvetic Nappes[73]
80-21 Santis ThrustK limestone 181°, 1° −4.220 LNS
80-23 Santis ThrustK limestone 172°, 5° −0.519 LPS
80-24 Santis ThrustK limestone 205°, 22° −2.99 LNS
80-25 Santis ThrustK limestone 172°, 12° −3.413 LPS
80-15 Santis ThrustK limestone 153°, 30° −216 LPS
80-14 Glarus ThrustK limestone 342°, 82° −1.112 LNS
80-11 Murtschen ThrustK limestone 0°, 82° −5.78 LNS
80-12 Murtschen ThrustK limestone 340°, 81° −5.725 LNS
80-13 Murtschen ThrustK limestone 276°, 84° −8.917 LPS
80-16 Glarus FootwallK limestone 284°, 77° −1.911 LPS
80-17 Glarus FootwallK limestone 78°,82° −5.316 LPS
80-19 Glarus FootwallK limestone 330°, 45° −318 LPS
80-20 Glarus FootwallK limestone 271°, 33° −327 LNS
−3.661516.23077 LNS = 6/LPS = 7
Helvetic Nappes (this study)
BE-2416BE-2416Alpenrhein GrabenVein (striated)48° E, 42° S6°, 82°W22348°, 5°269°, 4°153°, 88°−6.39276365.395LNSVNS0
H2H2Alpenrhein GrabenVein (striated)73° E, 33° S63°, 79° E36171°, 3°274°, 13°47°, 78°−1.6311179333.238LNSVNS0
H3H3Alpenrhein GrabenVein (striated)69° E, 24° S349°, 86°W18157°, 4°27°, 7°246°, 86°−6.4322366386.355LNSVNS0
H4H4Alpenrhein GrabenVein (striated)87° E, 16° S7°, 55° E36168°, 8°281°, 27°96°, 41°−3.348483406.955LNSVPS0
n = 112 −4.4312.5 −372.25LNS = 4/LPS = 0
Helvetic–Penninic Nappes [40]
J20 Dolden–Wildhorn nappesValanginian0, 0 50352°, 16°234°, 16°90°, 26°5.025 No plot; LPS
J30V Dolden–Wildhorn nappesValanginian180, 30 50298°, 8°40°, 56°203°, 33°6.010 No plot; LPS
J30M Dolden–Wildhorn nappesValanginian180, 30 50195°, 24°99°, 14°340°, 62°5.025 No plot; LPS
J31V Dolden–Wildhorn nappesValanginian180, 30 50313°, 37°186°, 38°69°, 29°3.029 No plot; LPS
G1M Dolden–Wildhorn nappesValanginian240, 22 50303°, 45°151°, 42°47°, 14°6.018 LPS
G1V Dolden–Wildhorn nappesValanginian240, 22 50271°, 39°132°, 43°20°, 22°5.018 LPS
Ge1 Dolden–Wildhorn nappesMalm290, 20 5064°, 41°320°, 16°213°, 45°6.030 LNS
A3 Dolden–Wildhorn nappesdogger308, 46 508°, 16°115°, 45°265°, 40°3.042 LPS
L13V Dolden–Wildhorn nappesMalm190, 40 5016°, 25°240°, 56°116°, 20°7.043 LNS
L13M Dolden–Wildhorn nappesMalm190, 40 50118°, 40°14°, 17°265°, 45°9.030 LPS
W17 Dolden–Wildhorn nappesMalm234, 14 50126°, 13°321°, 76°217°, 3°3.017 LPS
W50 Dolden–Wildhorn nappesMalm240, 18 50273°, 6°19°, 70°181°, 20°5.032 LPS
W68 Dolden–Wildhorn nappesMalm310, 56 50265°, 67°134°, 16°39°, 17°9.032 LPS
W72 Dolden–Wildhorn nappesBarremian225, 50 50276°, 77°82°, 13°172°, 3°4.035 LPS
W77 Dolden–Wildhorn nappesBarremian110, 75 50359°, 41°169°, 49°265°, 5°2.022 LNS
W90 Dolden–Wildhorn nappesBarremian310, 56 50354°, 48°258°, 6°163°, 41°15.016 LPS
M129 Dolden–Wildhorn nappesValanginan350, 9 50274°, 70°156°, 9°63°, 17°18.020 LNS
M130 Dolden–Wildhorn nappesValanginan300, 16 50134°, 58°30°, 8°295°, 30°2.029 LNS
M133 Dolden–Wildhorn nappesEocene290, 40 50312°, 32°143°, 58°45°, 58°2.030 LPS
M156 Dolden–Wildhorn nappesEocene316, 44 50325°, 37°186°, 45°73°, 22°6.08 LPS
M170 Dolden–Wildhorn nappesMalm328, 35 50106°, 46°235°, 31°344°, 27°8.020 LNS
M174 Dolden–Wildhorn nappesBarremian28, 50 50182°, 52°334°, 35°73°, 14°6.029 LNS
M175 Dolden–Wildhorn nappesBarremian30, 30 50201°, 49°0°, 40°99°, 11°4.030 LNS
M186 Dolden–Wildhorn nappesBarremian0,0 50142°, 40°33°, 21°282°, 43°6.017 LPS
M191 Dolden–Wildhorn nappesValanginan286, 25 50354°, 51°145°, 35°245°, 15°14.025 LNS
M199 Dolden–Wildhorn nappesValanginian296, 11 5037°, 78°269°, 10°177°, 7°9.027 LNS
M220 Dolden–Wildhorn nappesDogger304, 40 50152°, 68°56°, 3°325°, 21°9.028 LNS
M245 Dolden–Wildhorn nappesEocene160, 40 50136°, 71°324°, 19°234°, 3°5.029 LNS
M356 Dolden–Wildhorn nappesDogger0,0 50190°, 71°299°, 3°30°, 11°8.035 LNS
1450 6.625.89655 LNS = 13/
LPS = 16
Tauren Window, Austria
17,651/17,653NEVTauren Window, AustriaCalcsilicateN-S Sheath fold781°, 72°192°, 1°263°, 8°−2.21100140314.988LNS
289a6/298a7NEVTauren Window, AustriaCalcsilicateN-S Sheath fold6238°, 83°171°, 12°84°, 2°−1.74100108295.715LNS
72,664/74,664NEVTauren Window, AustriaCalcsilicateN-S Sheath fold73183°, 87°145°,6°86°, 5°−1.83100221348.891LNS
801a1/801a2NEVTauren Window, AustriaCalcsilicateN-S Sheath fold66231°, 69°133°,6°88°, 4°−2.57100240355.016LNS
279 −2.08 −328
Avg. of all Tauren samples 1000 −3.61229.1 −337.351
Key: E1 = maximum shortening axis; E2 = intermediate axis; E3 = extension axis. LPS = layer-parallel shortening; LNS = layer-normal shortening; VPS = vein-parallel shortening; VNS = vein-normal shortening; Δσ = −51 + log 171 (#twins/mm).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Craddock, J.P.; Ring, U.; Pfiffner, O.A. Deformation of the European Plate (58-0 Ma): Evidence from Calcite Twinning Strains. Geosciences 2022, 12, 254. https://0-doi-org.brum.beds.ac.uk/10.3390/geosciences12060254

AMA Style

Craddock JP, Ring U, Pfiffner OA. Deformation of the European Plate (58-0 Ma): Evidence from Calcite Twinning Strains. Geosciences. 2022; 12(6):254. https://0-doi-org.brum.beds.ac.uk/10.3390/geosciences12060254

Chicago/Turabian Style

Craddock, John P., Uwe Ring, and O. Adrian Pfiffner. 2022. "Deformation of the European Plate (58-0 Ma): Evidence from Calcite Twinning Strains" Geosciences 12, no. 6: 254. https://0-doi-org.brum.beds.ac.uk/10.3390/geosciences12060254

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop