You are on page 1of 11

Earth and Planetary Science Letters 277 (2009) 373383

Contents lists available at ScienceDirect

Earth and Planetary Science Letters


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

Experimental modelling of shallow magma emplacement: Application to saucer-shaped intrusions


Olivier Galland a,, Sverre Planke a,b, Else-Ragnhild Neumann a, Anders Malthe-Srenssen a
a b

Physics of Geological Processes (PGP), University of Oslo, P.O. Box 1048, Blindern, 0316 Oslo, Norway Volcanic Basin Petroleum Research, Oslo Innovation Center, Gaustadallen 21, 0349 Oslo, Norway

a r t i c l e

i n f o

a b s t r a c t
Saucer-shaped dolerite and sandstone intrusions are common in sedimentary basins world-wide. We have conducted a series of scaled experiments simulating the process of magma emplacement in sedimentary basins, with particular attention on the formation of saucer-shaped sills. The model materials were (1) cohesive negrained silica our, representing brittle crust; and (2) molten low-viscosity oil, representing magma. The experiments were performed in both homogeneous and layered models. In all the experiments, oil injection resulted in doming of the surface. In the homogeneous models, the injected oil formed cone sheets and subvertical dykes. Cone sheets formed for shallow injection (13 cm), and vertical dykes formed for deeper injection (45 cm). In layered models, the injected oil always formed saucer-shaped intrusions. Our experimental results show that (1) sill intrusion results in the formation of a dome, with melt erupting at the rim; (2) layering controls the formation of sills and saucer-shaped sills; (3) saucer-shaped sills are fed from the bottom and the uid ows upward and outward; and (4) the diameter of saucer-shaped sills increase with increasing emplacement depth. The systematic relation between domes and sills and the depth-dependence of sill diameters show that saucershaped intrusions result from the interaction between a growing at-lying shallow sill and doming of the free surface. We conclude that saucer-shaped intrusions represent fundamental geometries formed by shallow magma intrusion in stratied basins. 2008 Published by Elsevier B.V.

Article history: Received 7 March 2008 Received in revised form 27 October 2008 Accepted 1 November 2008 Available online 18 December 2008 Editor: C.P. Jaupart Keywords: saucer-shaped sills magma emplacement doming experimental modelling

1. Introduction Transport and emplacement of magma in the Earth's crust are processes with great signicance for the evolution of our planet. Magmatic intrusions are common in many sedimentary basins, and eld observations show that they mostly consist of sheet intrusions, such as sills and dykes. It has recently been realized that the intrusion of sills into sediments may have very important implications for (1) global climate by the release of large volumes of greenhouse gases from metamorphic aureoles around the sills (Svensen et al., 2004, 2007), (2) organic matter maturation and uid migration (i.e. Keating et al., 2002; Rossello et al., 2002), and (3) ground-water aquifers (Chevallier et al., 2001; Chevallier et al., 2004). Recently, eld observations in exceptionally well preserved and exposed areas in the Karoo Basin (Fig. 1; Chevallier and Woodford, 1999; Malthe-Srenssen et al., 2004; Planke et al., 2005; Planke, 2008; Polteau et al., 2008a) and seismic data (Hansen et al., 2004b; Thomson and Hutton, 2004; Hansen and Cartwright, 2006b; Hansen et al., 2008; Polteau et al., 2008a) have shown that many intrusions exhibit a saucer shape. A spectacular and well-studied example of saucer-

Corresponding author. E-mail address: olivier.galland@fys.uio.no (O. Galland). 0012-821X/$ see front matter 2008 Published by Elsevier B.V. doi:10.1016/j.epsl.2008.11.003

shaped intrusions is the Golden Valley Sill Complex, South Africa (Malthe-Srenssen et al., 2004; Galerne et al., 2008; Planke, 2008; Polteau et al., 2008b; Aarnes et al., 2008). In map view, saucer-shaped intrusions form circular to elliptic rings (Fig. 1, Chevallier and Woodford, 1999). In cross section, they typically consist of a 100 m thick horizontal inner sill, branching outward and upward to steep inclined sheets, and subsequently to atter outer sills (Fig. 1, Chevallier and Woodford, 1999; Malthe-Srenssen et al., 2004; Hansen and Cartwright, 2006b,a). Such features are common to both magma and uidized sand intrusions (Huuse and Mickelson, 2004; Duranti and Mazzini, 2005; Polteau et al., 2008a), and are present in many basins all around the world (Fig. 1a). Therefore, the saucer shape appears to represent a fundamental shape for large-scale uid intrusions in sedimentary basins. The origin of saucer-shaped intrusions is still not clear. Some models suggest that magma rises to approximately its level of neutral buoyancy, where it starts spreading laterally to form the outer sill; subsequently, magma ows downward leading to inclined sheets and inner sills (Bradley, 1965; Francis, 1982; Chevallier and Woodford, 1999). In contrast, other models assume a central conduit feeding the inner sill, with only upward and outward radial magma ow (Pollard and Johnson, 1973; Malthe-Srenssen et al., 2004). The latter models imply the formation of domes in the sill overburdens (Pollard and Johnson, 1973; Malthe-Srenssen et al., 2004). Recent seismic data

374

O. Galland et al. / Earth and Planetary Science Letters 277 (2009) 373383

Fig. 1. a. World distribution of known saucer-shaped sills (full stars) and saucer-shaped injected sands (open stars, see the review of Polteau et al. (2008a) and references therein). b. 3D-geometry of a saucer-shaped sill from the Tulipan sill, Mre Basin offshore Norway (Polteau et al., 2008a). Depths are measured from the present seabed. c. Seismic prole illustrating the relationships between saucer-shaped sills and their overburden, Rockall basin, offshore Scotland. Modied after Hansen and Cartwright (Hansen and Cartwright, 2006a). d. LANDSAT image of part of Karoo basin, South Africa. This image exhibits ring features, each corresponding to an exhumed dolerite saucer-shaped sill (S) of 10 km in diameter. e. Aerial photograph of the Golden Valley saucer-shaped sill (location on d). Bottom of the valley is the inner sill, and crests are the exposed inclined sheets. Photograph by Polteau et al. (2008b).

(Trude et al., 2003; Shoulders and Cartwright, 2004; Thomson and Hutton, 2004; Hansen and Cartwright, 2006b,a; Thomson, 2007a) and anomalies of magnetic susceptibility data (AMS; Polteau et al., 2008b) support the latter model. In such models, the saucer-shaped sills are the result of a mechanical interplay between the growing sills and their deforming overburden (Pollard and Johnson, 1973; Pollard and Holzhausen, 1979; Fialko, 2001; Malthe-Srenssen et al., 2004; AMS; Polteau et al., 2008b). However, it is still debated whether the structures overlying the sills are domes or volcanoes (Hansen and Cartwright, 2007; Thomson, 2007b). Thus, the emplacement mechanisms of saucers and their geological implications are still poorly constrained, and many fundamental questions remain to be answered. What controls the saucer evolution of sills? What is the nature and origin of the overlying structures? If they are domes, what is the effect of doming? What controls the inner sill-to-inclined sheet transition? In order to answer such questions, we resorted to scaled laboratory experiments, in which a vegetable oil, simulating magma, was injected into cohesive Coulomb silica our, simulating the brittle upper crust. Our objectives were to simulate the emplacement of saucer-shaped

intrusions and to understand the physical processes governing the saucer-shape evolution of sills. 2. Experimental setup and scaling 2.1. Model materials and scaling We used ne-grained crystalline silica our as an analogue for the brittle crust, and a low-viscosity vegetable oil for the magma. The our is produced by Sibelco, in Belgium, and sold under the name M400. The grain size is 15 m. It fails according to a MohrCoulomb criterion, and we measured its cohesion and friction coefcient at 369 44 Pa and 0.81 0.06, respectively, using a Hubbert shear box, as described for example by Hubbert (1951), Schellart (2000) and Galland et al. (2006). This value is within errors the same as that used by Galland et al. (2006). This yields an angle of internal friction 39. According to Galland et al. (2006), the tensile strength of the silica our is 100 Pa. The vegetable oil is produced by Unilever and sold in France under the name Vgtaline. It is solid at room temperature but melts at 31 C.

O. Galland et al. / Earth and Planetary Science Letters 277 (2009) 373383

375

The viscosity of the oil is poorly temperature-dependant (Galland et al., 2006), and we injected it at 50 C, where its viscosity is 2 10 2 Pa s. The scaling and the suitability of the model materials have been described in detail by Galland et al. (2006). In such settings, the scaling is challenging because (1) the ranges of geological settings and viscosities of magma are very broad, and (2) the experiments aim to simulate both the ow of low-viscosity magma and the deformation of the country rocks. The principle is to dene selected dimensionless numbers, which characterize the kinematics and the kinetics of the simulated processes. The scaling procedure is based on the standard similarity conditions as developed by Hubbert (1937) and Ramberg (1981), and used for example by Merle and Borgia (1996). The principal geometric input variable is the thickness of the overburden (D). Output geometrical variables are the sill length (l) and thickness (h). Material properties are the densities of the magma (m) and the country rock (r), the angle of internal friction () and the cohesion (C) of the country rock, and the viscosity of the magma (). The only external force is due to the gravity (g). The injection ow rate is Q. The magma ow velocity (U) can be calculated from the sill length and thickness, and the ow rate. Notice that the thickness and the radius of the intrusion are results of the experiments, and not a parameter known a priori. However, they need to be taken into account in the dimensional analysis in order to analyse the viscous stresses into the intrusions. According to the Buckingham- theorem (e.g. Barenblatt, 2003), ten variables minus three variables with independent dimensions lead to seven independent dimensionless numbers that characterize the system. The experiments are similar to nature if these seven numbers have the same values in nature and in experiments. Our experiments aim to simulate basin-scale phenomena, so that we chose a model to nature scale ratio of 10 5, i.e. 1 cm in experiments represents 1 km in nature. The rst dimensionless parameters are the geometric ratios of the considered system: 1 = h=l; 2 = l=D: 1

Table 1 Symbols, units, and values of the mechanical variables in nature and experiments Denition of parameters C D g h l U m r Cohesion of brittle material Thickness of overburden Acceleration due to gravity Intrusion thickness Intrusion radius Magma ow velocity Angle of internal friction Density of magma Density of country rock Field 107 108 1000 5000 9.81 20 200 7000 5 10 4 5 10 2 25 45 2500 2700 2500 Field 10 2 12 25 45 2.4 10 1 12.3 6.31011 6.3101 2.5 10 6 270 0.08 0 Experiment 370 0.01 0.05 9.81 1 10 3 3 10 3 0.1 4.72 10 4 3 10 3 39 892 1050 Dimension Pa m m s 2 m m m s 1 kg m 3 kg m 3

Denition of dimensionless ratios 1 2 3 4 5 6 7 Intrusion thickness/length Intrusion length/depth Angle of internal friction Gravitational stress/cohesion Viscous stresses/cohesion Inertial/viscous forces Magma/country rock densities

Experiment 10 2 1 10 39 2.8 10 1 4.3 10 4 1.2 10 2 1.4 10 2 1.3 10 1 0.15

Average dimensionless ratios for nature and experiments.

1 expresses the strain associated with the sill emplacement. In nature, sill diameter is typically 1020 km, whereas their thickness ranges between 20 and 200 m. Thus, 1 is typically of the order of 10 2. In experiments, saucer radii were typically between 5 and 10 cm, whereas their thicknesses ranged between 1 and 3 mm. Thus, 1 was also of the order of 10 2, i.e. in the same range as natural saucers. In nature, saucer-shaped sills have been emplaced at depths of a few kilometres. Thus, 2 is between 1 and 5. In our experiments, 2 ranged between 1 and 1.7 (see Fig. 6 and Table 2). Therefore, the geometry of model intrusions approximately matched with natural intrusions. The rst dimensionless parameter to scale the brittle country rock is the angle of internal friction: 3 = : 2

silica our was controlled at 1050 kg m 3, and the cohesion was 370 Pa (Table 1). Therefore, 4 in experiments yields a value of 0.28, which is towards the lower bound of the range of 4 in nature (Table 1). The experiments aim to simulate the transport of viscous uid into a deforming Coulomb material of cohesion C. The propagation of magma into the rock is controlled by a balance between the ow of magma and the deformation of the country rock. In elasto-plastic materials such as rocks and cohesive granular media, the material deforms both elastically and plastically. If the cohesion of the silica our is well known, its elastic properties are in turn challenging to estimate and difcult to scale properly. Therefore, we use the cohesion to scale the strength of the rock. Thus, the dimensionless number 5 to scale the viscosity of magma is the ratio of viscous pressure drop within the uid along the intrusion to the cohesion (i.e. strength) of the country rock. For laminar ow of a Newton uid of viscosity owing at a velocity U, the viscous pressure drop along a fracture of length l and thickness h is Pv lU / h2 (e.g. Lister and Kerr, 1991; Kavanagh et al., 2006). Thus, the expression of 5 is: 5 = lU : h2 C 4

The angle of internal friction of the silica our is about 39, which is in the range of those of natural rocks, although it is close to its upper bound (Schellart, 2000). The second dimensionless number to scale the brittle country rock is the ratio of gravitational stress to cohesion: 4 = r gD : C 3

An average density for natural rocks is about 2500 kg m 3; their cohesion spans between 107 to 108 Pa (Schellart, 2000). Sills are usually emplaced at a few kilometres depth, so that we choose D ranging from 1 to 5 km. Therefore, 4 in nature yields values ranging from 0.24 to 12.3 (Table 1). Because the model to nature scale ratio is 10 5, we varied D in experiments between 1 and 5 cm. The density of the

If we assume that 20- to 200-meter thick sills are fed by a 1-meter thick dyke, in which magma velocity U ranges between 0.1 and 1 m s 1, the estimated magma velocity in sills is expected to range between 510 4 and 510 2 m s 1 (Table 1). The viscosity range of magmas is very large, from 10 Pa s for wet crystal-poor mac magma, to 1018 Pa s for dry crystalrich felsic magma (e.g. Dingwell et al., 1993; Petford et al., 1993; Romano et al., 2003). However, we only consider common magma types such as basaltic to rhyolitic magma, for which the viscosities range between 100 Pa s to 108 Pa s. In basins, saucer-shaped sill radii are typically 5 km. Therefore, 5 in nature ranges from 6.310 11 to 6.310 1 (Table 1), so that viscous stresses are negligible compared to the strength of the country rock. In experiments, the injection ow rate was 40 ml min 1, within fractures of typical radii about 7 cm and typical thickness 13 mm; the viscosity of the oil is 210 2 Pa s at 50 C (Galland et al., 2006). Thus, the values of 5 in experiments are 4.2310 4 to 1.210 2 (Table 1). Such values are in the range of 5 in nature and much smaller than 1, which means that viscous stresses are negligible with respect to the cohesion of the country rock. This is a major difference from experiments of lithospheric deformation where the stresses in the ductile crust must be of the same order of magnitude as those in the brittle crust (e.g. Davy and Cobbold, 1991).

376

O. Galland et al. / Earth and Planetary Science Letters 277 (2009) 373383

The ow regime of magma within intrusions (laminar versus turbulent) is predicted by the Reynolds number, which is the ratio between inertial and viscous forces: 6 = Re = m hU : 5

The density of magma is 2500 to 2700 kg m 3 for felsic to mac magma, respectively. Therefore, 6 in nature ranges between 2.5 10 6 and 270 (Table 1). Thus, usually inertial forces are negligible and magma ow is laminar. However, in cases of very uid magma owing at high velocity in narrow sills, some local turbulence may occur (Kavanagh et al., 2006; Menand, 2008). In experiments, 6 ranges between 1.4 10 2 and 1.3 10 1, so that inertial forces are also negligible. The nal dimensionless number is the ratio of hydrostatic forces to lithostatic forces, corresponding to the buoyancy of the magma, which can be expressed by: 7 = 1 m : r 6

If 7 b 0, the magma is heavier than the country rock, and locally negatively buoyant; in contrast if 7 N 0, the magma is lighter than the country rock, so that the magma is locally buoyant. In nature, 7 ranges between 0.08 and 0, so that magma is neutrally buoyant to negatively buoyant. In contrast in experiments, 7 = 0.15, so that the oil is positively buoyant (Table 1). Therefore, there is a discrepancy between the experiments and geological systems. However, as sills are subhorizontal conduits, the effect of the buoyancy is very small, as discussed by Lister and Kerr (1991), Kavanagh et al. (2006) and Menand (2008). We will show below that the buoyancy is indeed negligible in both nature and experiments, and is thus not a critical parameter. 2.2. Experimental setup Our apparatus is a modied version of that developed by Galland et al. (2003, 2006, 2007). The models lay in a square box, 40 cm wide and with variable thickness (Fig. 2). The model rock consisted of a pack of compacted silica our. The preparation procedure consisted of measuring a mass of our that we compacted using a high frequency compressed-air shaker sold by Houston Vibrator (model GT-25). Such a procedure allows a homogeneous, repeatable, and fast compaction of the our. We want to take into account the mechanical layering to simulate the sedimentary strata. Galland (2005) simulated layering by introducing

cohesionless layers made of ne-grained silica spheres into the cohesive silica our. However, the high frequency shaking mixes the spheres and the our, so that the layers are not preserved. In the present set of experiments, we simulated layering by introducing a at exible net into the our, of grid size 2 mm. The net simulated a weak layer as it reduced the contacts between the our grains. In all the experiments, the net was a 30 30 cm square. The advantage of using a net is that the oil can cross it easily, in contrast with a lm that prevents the oil from crossing. It is difcult to quantify the heterogeneity associated with the net, and we suspect that it is rather high. Nevertheless, cohesion contrasts in sedimentary basins can sometimes be very important, i.e. several orders of magnitude (e.g. Hoek et al., 1998; Schellart, 2000). Therefore, we consider that introducing a net is reasonable for our purpose. Finally, the net was much larger than the diameters of the intrusions, and the shape of the net (square) was different to those of the intrusions (circular). We thus infer that the boundaries of the net had no inuence on the transport of the oil, i.e. there was no boundary effect. To prepare the experiments with the net, we poured a rst layer of our; then we shook the box to atten the surface of the our. Subsequently, we placed the net and a second layer of our. We shook the box a second time until the density of the upper layer was 1050 kg m 3. Such a procedure induced a density contrast of less than 5% between the lower and upper layers. We consider this difference negligible. There are two stages in the emplacement of saucer-shaped sills: (1) the emplacement of a planar sill from a feeder, and (2) the evolution of that sill to saucer-shape. In this work, we focussed on the latter stage and we did not consider the feeder-to-sill transition. We therefore located the net right at the top of the inlet, so that the oil formed a sill from the initial stage of intrusion. In nature, feeder structures can be either planar dykes or circular pipes or plugs. The shape of intrusions would obviously depend on the feeder shape. For technical reasons, we chose a circular inlet of 1 cm diameter. Each experiment typically lasted for a minute, during which we measured the evolution of the oil overpressure (Fig. 3e). After the end of the experiments, the oil solidied within about half an hour. Then, the intrusion was excavated (Fig. 3d), and its top surface digitalized using a moir projection technique developed by Brque et al. (2004). 3. Results Here we present two series of experiments. In the rst series (series H), the models were homogeneous, made of pure silica our. In contrast, in the second series (series L), the models were heterogeneous, with a exible net simulating layering. In both series, the injection

Fig. 2. Schematic diagram of the experimental setup used in this work. Compacted silica our (white) lies in a square box. A volumetric pump injects the molten oil (grey) within the our via an inlet. A pressure gauge monitors the oil pressure during time.

O. Galland et al. / Earth and Planetary Science Letters 277 (2009) 373383

377

Fig. 3. Photographs of typical model surface evolution and some excavated model intrusions. a. Initial at surface. b. After the oil has started intruding, the surface lifts up, forming a dome. c. Eruption of oil. In all the experiments, the oil erupted at the rim of the dome. The dashed line locates the following cross-section. d. 3D view of the cross-section of a model after the end of the experiment. The grey solid body comprises the intrusion itself and an oil percolation aureole in the silica our. The intrusion exhibits a V-shape in cross-section. The excavated part of the intrusion exhibits a cone shape. e. Plot of typical oil overpressure evolution with time during experiments L3, L4 and L5. The overpressure initially builds up until the intrusion is initiated, and then drops following a hyperbolic trend. This evolution is typical of hydraulic fracturing, as discussed by Murdoch (1993a,b, 2002), Galland (2005) and Galland et al. (2007). Notice that the overpressure is larger than 1000 Pa. f. Experiment H2. The intrusion is a cone-sheet, with its centre corresponding to the inlet position. g. Experiment H4. The intrusion is a sub-vertical dyke. Beneath the surface, it branches to an elongated cone-sheet. h. Experiment L2. The intrusion is a saucer-shaped sill. A small inner sill branches outward to inclined sheets. i. Experiment L4. The intrusion is a saucer-shaped sill. As in experiment L2, it consists of an inner sill and inclined sheets.

378 Table 2 Experimental parameters and results Experiment Type Injection Intrusion depth/mm shape 14 25 34 45 55 Cone sheet Cone sheet Cone sheet Dyke Dyke (+ cone sheet) Saucer Saucer Saucer Saucer Saucer

O. Galland et al. / Earth and Planetary Science Letters 277 (2009) 373383

Diameter Intrusion Area of duration/s intrusion/ of inner sill/cm cm2 10 26 28 5 24 26 60 58 5 24

H1 H2 H3 H4 H5

Homogeneous Homogeneous Homogeneous Homogeneous Homogeneous

L1 L2 L3 L4 L5

Layered Layered Layered Layered Layered

10 20 31 41 50

5 17 24 42 83

17 29 91 117 189

1.7 2.1 5.3 6.6 7.1

In every experiment, the injection ow rate was 40 ml min 1.

depth (D) was a variable parameter, ranging from 1 cm to 5 cm (Table 2). In all experiments, the evolution of the oil overpressure was similar, with a short initial pressure build up to 500010 000 Pa, followed by a hyperbolic pressure drop down to 10002000 Pa (Fig. 3e). 3.1. Homogeneous models (series H) In the H experiments the ow rate of injection was constant at 40 ml min 1. During injection, the surface of the models locally lifted up, resulting in smooth circular doming (Fig. 3). Subsequently, the oil reached the surface at the rim of the domes (Fig. 3). The experiments terminated when the oil erupted. The resulting sheet intrusions were a few millimetres thick. This geometry is typical of dykes and sills in nature (Lister and Kerr, 1991; Rubin, 1995).

The model intrusions exhibited various shapes with respect to the injection depth. In experiments H1 to H3, the inlet was shallow (1.4 to 3.4 cm; Table 2), and the intrusions were axi-symmetric inward dipping sheets, converging towards the inlet (Figs. 3df and 4a). This geometry is typical of cone-sheets, as commonly observed at central volcanoes (e.g. Ancochea et al., 2003; Klausen, 2004; Pasquare and Tibaldi, 2007). We analysed the intrusion shapes by averaging a series of radial proles, as described in Fig. 4cd. Outward dip variations were sometimes visible, as in experiment H3: the intrusion dipped 4550 near the inlet, whereas it became gentler close to the surface (30; Figs. 4 and 5a). In addition, the deeper the injection, the later the eruption occurred (10 s to 28 s; Table 2; Fig. 6a) and the larger the map area of the intrusion (26 cm2 to 60 cm2; Fig. 6c). In contrast, in experiments H4 and H5, the inlet was deeper (4.5 and 5.5 cm, respectively) and the resulting intrusions were subvertical dykes (Fig. 3g). In experiment H5, at 12 cm depth the dyke branched to a cone sheet much smaller than those formed in experiments H1 to H3. Consequently, in H4 and H5 the oil moved to the surface more quickly (5 s to 24 s; Table 2; Fig. 6a) and the eruption occurred at an earlier stage than in H1H3. In addition, the map areas of intrusions were also much smaller (5 cm2 to 24 cm2; Fig. 6c). 3.2. Layered experiments (series L) In the L experiments the net simulated a layer. The injection ow rate was 40 ml min 1 also in these experiments. As in the H experiments oil injection resulted in doming; again, the oil erupted along the rim of the domes. Excavated intrusions exhibited sheet-like bodies a few millimetres thick. The shapes of the intrusions were a bit more complex than in the H experiments (Figs. 35). Their three-dimensional geometry was axially symmetric. Close to the inlet, the oil formed a circular horizontal inner

Fig. 4. 3D diagrams of the upper surface of typical model intrusions. a. Experiment H3. The intrusion exhibits a typical axi-symmetric cone-sheet shape. b. Experiment L5. The intrusion exhibits a typical axi-symmetric saucer shape, with a at basal sill, steep inclined sheets, and at outer sills. c. Map view of the intrusion of experiment L5. The scale is in millimetres. The radial white lines locate serial topographic proles. The origin of the proles is centre of the basal sill. d. Plots of the serial radial sections (light gray) of the intrusion of experiment L5. The average prole (red) gives the whole shape of the intrusion.

O. Galland et al. / Earth and Planetary Science Letters 277 (2009) 373383

379

Fig. 5. Plots of average proles of model intrusions. a. Proles from homogeneous models (series H). The intrusions are inward dipping cone-sheets. Notice that experiments H1, H2 and H3 only are presented. Intrusions in experiments H4 and H5 were vertical dykes, and so not possible to measure. b. Proles from layered models (series L). The intrusions consist of at inner sill following a heterogeneity, outward inclined sheets turning into at outer sills. Such a geometry is typical of saucer-shaped intrusions in sedimentary basins. The arrows locate the position of the inlet.

sill that propagated along the net. Subsequently, the inner sill turned into steeply inclined sheets (60), whose inclinations became gentler towards the surface and sometimes ended in sub-horizontal outer sills (Figs. 4 and 5). The inner sill-to-inclined sheet transition controlled the nal shape of the inner sill. In all L experiments the inner sills were close to circular (Figs. 4 and 5). Although the saucer-shaped intrusions in the experiments were axi-symmetric, the centre of the inner sill was not systematically consistent with the location of the inlet (Fig. 5). In L3, the inlet was even next to the rim of the inner sill. The variable parameters in the L experiments were the depth D of the inlet and the depth of the net, varying between 1 and 5 cm (Table 2). The model intrusions all exhibited similar saucer shapes (Fig. 5); nevertheless the deeper the injection, the wider the inner sills (Table 2; Figs. 5 and 6). The ratio of the inner sill diameter to the emplacement depth (W/D) was close to constant about 1.5 in all experiments (Fig. 6c). Positive correlations between the depth of emplacement and intrusion duration, the map area of the intrusions and the diameter of the inner sills illustrate a depth-dependence of the intrusions (Fig. 6). 4. Discussion 4.1. Validity of the scaling 4.1.1. Viscous stresses A major assumption in the experiments is that the viscous stresses are negligible with respect to the strength of the country rock. This assumption can be veried if the uid overpressure is larger than the viscous dissipation along the fracture. In the experiments, the measured oil overpressure was larger than 1000 Pa (Fig. 3e), whereas the viscous dissipation along the crack, calculated from Pv lU / h2

Fig. 6. Plots of the intrusion duration (a), the inner sill diameter (b), the inner sill diameter to inlet depth ratio (l/D, c) and the intrusion area in map view (d) for homogeneous models (gray full circles) and layered models (black triangles). Notice that the inner sill diameter is an average calculated from the area of the inner sill, assuming it is circular.

380

O. Galland et al. / Earth and Planetary Science Letters 277 (2009) 373383

(e.g. Lister and Kerr, 1991; Kavanagh et al., 2006), was 1.5 10 1 Pa. Therefore, the viscous dissipation in the experiments was indeed negligible with respect to the driving overpressure, and so to the strength of the host material. In nature, the estimates of magma driving overpressure within magma conduits range from a few MPa to a few tens MPa (e.g. Pollard and Segall, 1987; Rubin, 1995; Hogan et al., 1998). Considering 20- to 200-meter thick sills with between 100 and 107 Pa, l being 5 km, and U between 5 10 4 m s 1 and 5 10 2 m s 1 (Table 1), the viscous drag along the sills is 6.3 10 3 Pa to 6 MPa. Therefore, as in experiments, the viscous drop along sills in usual cases is also negligible with respect to the magma driving overpressure. However, this assumption is not true for very viscous magma owing at high velocity in narrow sills. 4.1.2. Buoyancy forces A second critical assumption is the negligible effect of buoyancy on the emplacement of saucers. The buoyancy forces (i.e. hydrostatic forces PH) of an intrusion of thickness h is given by PH = gh, where = m r. As discussed by Lister and Kerr (1991), because sills are sub-horizontal intrusions, h is very small and the buoyancy forces are likely to be negligible. In experiments, h was typically 1 to 3 mm and was 158 kg m 3, so that the buoyancy forces of the magma within the intrusions ranged between 1.5 Pa and 4.5 Pa. Thus, as the oil driving overpressure was larger than 1000 Pa, the buoyancy of the intrusion was negligible compared to the driving pressure. In addition, we even observed signicant downward ow of the oil in the models below the sills, although the oil is lighter than the surrounding silica our. This conrms that the buoyancy forces are indeed negligible in the experiments. In nature, sills are often considered to have emplaced along the level of neutral buoyancy (e.g. Francis, 1982; Lister and Kerr, 1991). However, observations show that this assumption is not obvious because heavy magma can rise through lighter rocks such as sediments. In addition, the formation of domes overlying sills and the formation of inclined sheets show that the magma is able to lift up its overburden and to keep rising after being emplaced into sills, so that it is over-pressured. Indeed, the magma driving overpressure results from the local buoyancy integrated along the whole magmatic system, so that the driving pressure of a heavy magma intruding through lighter rocks can be very high (Lister and Kerr, 1991; Hogan et al., 1998; Gerya and Burg, 2007). Usual estimates of the driving overpressure range from 106 to 108 Pa (e.g. Pollard and Segall, 1987; Rubin, 1995; Hogan et al., 1998). In contrast, the buoyancy of a 20- to 200-meter thick sills ranges from 3 105 Pa to zero Pa. Thus, usually, in sills as well as in our experiments, the buoyancy forces are negligible with respect to the magma driving overpressure and have negligible effect on the inner sill-to-inclined sheet transition. In nature and in experiments, the dip angles of the inclined sheets strongly differ. In both case, the dip angle of the inclined sheets is acquired during the very early stage of inclined sheet formation, i.e. when they are very small. In such a conguration, the buoyancy forces are also negligible. Thus, we infer that the difference in the dip angles of the inclined sheets in nature and experiments is not due to differences in buoyancy forces. Nevertheless, we cannot rule out that in the later stages of saucer-shaped sill emplacement where the inclined sheets can be up to 1000 m high, buoyancy forces can be signicant. 4.2. General considerations The overall intrusion geometries in our experiments match major features of intrusions observed in nature. Our homogeneous experiments reproduced sub-vertical dykes and cone sheets, as observed at central volcanoes. Our layered experiments reproduced the threedimensional shape of saucer-shaped intrusions with a horizontal inner sill along a horizontal weakness, a sharp transition from the inner sill to steep inclined sheets, and a progressive transition from

inclined sheets to atter outer sills (Figs. 4 and 5). Furthermore, the experiments simulate dome formation above the intrusions; the geometrical relationship between dome rims and the inclined sheets was striking (Figs. 1 and 3). In addition, our 3D experimental results also match results from 2D numerical simulations of saucer-shaped sill formation (Malthe-Srenssen et al., 2004). We thus suggest that our experiments simulate well the processes governing the emplacement of saucer-shaped sills in sedimentary basins. In addition, our results suggest that dome-like structures overlying sills are forced folds as proposed by Hansen and Cartwright (2006a, 2007). The geometry of intrusions is a good indicator of the processes that govern the emplacement of magma. In our experiments, the intrusions formed horizontal sheets with a thickness to width aspect ratio of 0.02. Such aspect ratios are characteristic of dykes and sills formed by hydraulic fracturing in nature (Hubbert and Willis, 1957; Lister and Kerr, 1991; Clemens and Mawer, 1992; Rubin, 1993, 1995). In previous experiments, also using vegetable oil and ne-grained silica our, Galland et al. (2007) showed that the oil intruded into the silica our by hydraulic fracturing. 4.3. Emplacement of the inner sill The rheological properties of the host rock are important parameters that control the emplacement of magma. In experiments with a weak layer, the oil initially intruded along the layer and formed a sill (Figs. 4 and 5). In contrast, in experiments without a layer, oil injection always resulted in the formation of cone sheets or subvertical dykes (Figs. 3 and 5). Therefore, the experiments strongly suggest that the formation of at-lying sheets require layering in the country rock. In nature, sills mostly form in sedimentary basins where strata typically represent strong mechanical layering (e.g. Hoek et al., 1998; Schellart, 2000). Our experiments suggest that the inner sill forms by following a sedimentary layer. This conclusion is supported by previous experimental modelling (Pollard and Johnson, 1973; Hyndman and Alt, 1987; Rivalta et al., 2005; Kavanagh et al., 2006; Menand, 2008), and numerical calculations (Gudmundsson and Brenner, 2001). The geometrical relationship between the sills and their feeders provide important constraints on the mechanisms of sill emplacement. In our layered experiments, the inlet was always connected to the inner sills, although it was not systematically at the centre (Fig. 5). Therefore, the intrusions were always fed from the inner sills and the oil owed upward and outward within the inclined sheets and the outer sills (Fig. 5). Several models propose that saucer-shaped sills in nature result from feeding from the outer sills or from the inclined sheets, followed by buoyancy-controlled downward ow of the magma to the inner sills (Bradley, 1965; Francis, 1982; Chevallier and Woodford, 1999). However, the models of sills fed from the sides are not consistent with recent AMS (Polteau et al., 2008b) and seismic (e.g. Thomson and Hutton, 2004; Hansen and Cartwright, 2006b; Thomson, 2007a) data, which suggest that magma ows outward and upward in saucer-shaped sill complexes and that the outer parts of the saucers are fed from the inner sills. Our 3D experimental results support the model of saucers fed from the inner sill, and are consistent with previous 2D models (Pollard and Johnson, 1973; MaltheSrenssen et al., 2004). 4.4. Inner sill-to-inclined sheet transition The sharp inner sill-to-inclined sheet (IS-ISH) transition is the most characteristic feature of saucer-shaped intrusions. In our experiments, the IS-ISH transition documents the transition between (1) the emplacement of the inner sill following a layer and (2) the emplacement of the inclined sheet propagating away from the layer. Former models of sill intrusion in layered media also reproduced similar sharp transitions (Johnson and Pollard, 1973; Malthe-

O. Galland et al. / Earth and Planetary Science Letters 277 (2009) 373383

381

Fig. 7. Schematic cross-section diagram explaining the mechanism of saucer-shaped intrusions in our experiments. a. A shallow sill spreads into strata close to the surface. Consequently, the overlying sediments lift up and a dome forms. b. The sill keeps spreading and thickening and the dome keeps expanding. Differential uplift at the dome edges generates shear stresses. The stresses are likely to exceed the yield stress, so that faulting may occur. c. When the sill diameter is 1.5 times the overburden thickness, the shear stresses and/or faults deect the sill, resulting in inclined sheets. Subsequently, the melt forms the outer sill, and locally rises along the dome rim, where it reaches the surface and erupts.

Srenssen et al., 2004; Galland et al., 2006, 2007). In contrast, former models of sill intrusion in homogeneous media produced smooth transitions from horizontal to inclined fractures (Pollard and Holzhausen, 1979; Fialko, 2001; Bunger and Detournay, 2005; Bunger et al., 2005). That suggests that the layering controls the formation of sharp inner sill-to-inclined sheet transitions. In nature, eld observation show that such sharp IS-ISH transitions occur (e.g. Chevallier and Woodford, 1999; Polteau et al., 2008a), suggesting that the layering plays a major role in sill emplacement. In our experiments (1) the IS-ISH transitions exhibited sub-circular shapes similar to those of the overlying domes (Figs. 4 and 5), (2) the emplacement depths and the diameters of the inner sills are positively correlated (Fig. 6), and (3) the inclined sheets always formed directly underneath the rims of the domes (Fig. 3). Such systematic relationships imply that the formation of the domes and the inclined sheets is genetically linked. In nature, saucer-shaped intrusions exhibit similar three-dimensional shapes, depth-dependence trends, and geometrical links with overlying domes (Fig. 1; Hansen et al., 2004a; MaltheSrenssen et al., 2004; Polteau et al., 2008a). The detailed processes governing the IS-ISH transition and the role of the overlying dome are poorly understood. In our experiments, we cannot observe the in situ evolution of the structures because the silica our is opaque. We have demonstrated earlier that the buoyancy forces are negligible both in nature and in our experiments, so that the IS-ISH transition is not controlled by the buoyancy of the magma. Former work has addressed similar processes in relationship to growing cracks interacting with a free surface (Johnson and Pollard, 1973; Pollard and Johnson, 1973; Pollard and Holzhausen, 1979; Fialko, 2001; Fialko et al., 2001; Malthe-Srenssen et al., 2004; Bunger and Detournay, 2005; Bunger et al., 2005; Goulty and Schoeld, 2008). On the basis of these studies, we infer the following scenario for the emplacement of saucer-shaped intrusions (Fig. 7): (1) a shallow sill grows and spreads along a layer parallel to the surface, (2) the sill deforms its overburden by doming (Fig. 7a), (3) shear stresses at the edges of the dome result in asymmetric stresses at the crack tip (Fig. 7b; Malthe-Srenssen et al., 2004), and (4) when the sill reaches a

diameter of approximately 1.5 times the thickness of the overburden in our experiments (Fig. 6c), an inclined sheet is initiated (Fig. 7c). Malthe-Srenssen et al. (2004) suggested that such a transition occurs when the asymmetric stresses reach a critical value and deect the crack tip upward. The deformation mode of the overburden may play a major role for the location of the IS-ISH transition (Fig. 7). In our experiments, the silica our was an elasto-plastic material (Galland et al., 2006; Galland et al., 2007), and plastic deformation, i.e. faulting, was likely to occur at the rim of the domes and to trigger the formation of inclined sheets. In contrast, previous works considered an elastic overburden, and control of the IS-ISH transition by elastic stresses (Johnson and Pollard, 1973; Pollard and Johnson, 1973; Pollard and Holzhausen, 1979; Fialko, 2001; Fialko et al., 2001; Malthe-Srenssen et al., 2004; Bunger and Detournay, 2005; Bunger et al., 2005; Goulty and Schoeld, 2008). In nature, seismic data sometimes show that the seismic reectors are offset at the rims of overlying domes (Fig. 1c; Hansen and Cartwright, 2006a), suggesting that plastic deformation plays a signicant role in the IS-ISH transition (Galland, 2005). The role of plastic deformation of the country rock has important implications for the emplacement of shallow sills in very weak unconsolidated sediments. In such sediments, magma does not intrude by hydraulic fracturing and the resulting intrusions are likely to exhibit lobate shapes (e.g. Hallot et al., 1996). In addition, the thermal effect of sills on such water-rich sediments often leads to the formation of peperites (e.g. Busby-Spera and White, 1987; Skilling et al., 2002; Galerne et al., 2006). The rheological architecture of the country rock may also play a major role on the formation of inclined sheets. In our experiments, where sill overburden was homogeneous, the inner sill diameter to emplacement depth aspect ratio (W/D) was 1.5 (Figs. 4 and 5) and the dip angles of the inclined sheets were 4050. In contrast, in sedimentary basins where sill overburdens are typically layered, the W/D ratios for natural saucers are larger (34; Fig. 8; MaltheSrenssen et al., 2004; Polteau et al., 2008a) and the dip of the inclined sheets is gentler (1030) than in our experiments. As we demonstrated above, the buoyancy of the magma is negligible in both nature and experiments, so that such a discrepancy cannot be inferred from a buoyancy effect. In contrast, Galland (2005, Chapter 5) suggested that the mechanical layering of the overburden plays a major role in the formation of the IS-ISH transition. In the experiments of Galland (2005, Chapter 5), experiments with strong homogeneous overburden resulted in small inner sills with steep inclined sheets, whereas experiments with thin weak layers resulted in wider inner sills and gentler inclined sheets. Such a difference comes from the different elastic stiffness of the overburden, where a homogeneous

Fig. 8. Plots of average proles of experimental saucers in full lines, and in dashed lines are the average proles of the Golden Valley, South Africa, and the Tulipan saucer, Mre basin offshore Norway (Fig. 1b; Polteau et al., 2008a). Proles are normalized by the emplacement depth of the inner sills. The proles of the experimental saucers exhibit similar shapes, whereas the proles of natural saucers are much wider relative to depth, their inclined sheets less steep, and the inner sill-to-inclined sheet less sharp. Notice that the normalized depth of inner sills of the experimental saucers is N 1. Indeed, oil percolation aureole around the intrusion is a few millimeters thick, so that the measured surface is slightly shallower than the net, and so relative to the emplacement level.

382

O. Galland et al. / Earth and Planetary Science Letters 277 (2009) 373383

medium is elastically much stiffer than a multilayered one (e.g. Pollard and Johnson, 1973; Galland, 2005). Thus, we infer that a growing sill may spread easily underneath a low stiffness layered overburden, like in a sedimentary basins, whereas a stiff homogeneous overburden will prevent the sill from spreading, like in our experiments. 4.5. Geological implications The understanding of the emplacement of saucer-shaped sills has important implications for oil exploration (Hansen and Cartwright, 2006a). Indeed, the emplacement of large sills into organic-rich layers can trigger or enhance the maturation of organic matter (e.g. Galushkin, 1997; Svensen et al., 2004; Fjeldskaar et al., 2008). In addition, the domes overlying saucer-shaped sills represent potential traps for hydrocarbons (Trude et al., 2003; Hansen and Cartwright, 2006a; Tulipan sill, Mre Basin; Polteau et al., 2008a). Thus, the systematic association of saucer-shaped sills and overlying domes may generate local petroleum prospects. In this study, we have shown that saucer-shaped sills result from a fundamental interaction between intruding shallow sills and doming of the overburden in sedimentary basins. The understanding of such processes also has wide implications for the understanding of subvolcanic systems. Intrusions similar to saucer-shaped sills are observed in active and exhumed volcanoes (e.g. O'Driscoll et al., 2006; Cup-shaped intrusions of Mathieu et al., 2008). Those intrusions are often associated with deformation structures of the volcanic edices, including doming (Chang et al., 2007; Mathieu and Van Wyk de Vries, in press) and caldera collapse (e.g. Walter and Troll, 2001; Troll et al., 2002). Thus, at-lying and saucer-shaped intrusions are likely to play a major role in the volcanological and deformation histories of volcanic systems. In addition, we suspect that sill intrusions represent more realistic magma reservoir geometry than the usually assumed spherical or elliptical magma chambers (e.g. Gudmundsson, 1998, 2006; Gudmundsson and Philipp, 2006). Our results show that experimental modelling is a very useful method to understand magmatic and sub-volcanic systems. 5. Conclusions In this paper, we have described the results of experimental modelling of shallow magma intrusion in sedimentary basins. The model magma (molten vegetable oil) was injected at a constant ow rate into a model crust (cohesive silica our). Our main experimental results are the following. 1. The oil injection systematically induced doming of the surface of the models right above the intrusions. The domes were sub-circular with a few millimetres of relief. The oil always erupted at the rim of the domes. 2. The oil intruded by hydraulic fracturing and resulted in sheet intrusions (dykes, cone sheets, and saucer-shaped sills). 3. In homogeneous experiments, intrusions were cone sheets and sub-vertical dykes, whereas in layered experiments, sills formed along the layer. The horizontal spreading of magma along inner sills seems to be enhanced by layering of the country rock. 4. In layered experiments, the sills developed three-dimensional saucer-shaped geometries consisting of a at inner sill, steep inclined sheets, and atter outer sills. The inner sill to inclined sheet transition was sharp, whereas the inclined sheet to outer sill transition was progressive. The inner sills were circular. 5. Saucer-shaped intrusions were fed from the inner sill, and the melt owed upward and outward. However, the inlet was not systematically at the centre of the inner sill; it was sometimes even close to its rim. 6. The diameter of the inner sills increased with increasing emplacement depth. In addition, the inclined sheets always formed

underneath the dome rims. Those observations suggest that saucer-shaped sills result from the mechanical interactions between the growing sills and the doming of the overburden. In general, our experimental results show that at-lying shallow intrusions are likely to evolve naturally to saucer shape. Saucershaped intrusions are thus fundamental geometries of intrusions in stratied, consolidated shallow sedimentary sequences. The detailed understanding of the processes governing their emplacement has implications for hydrocarbon exploration and volcanic hazards. Acknowledgments We gratefully acknowledge the technical support of Olav Gundersen in the development of the experimental setup. The staff of the workshop of the Physics Institute at the University of Oslo built the experimental box. This work beneted from stimulating discussions with the researchers and students of PGP. This study was supported by a Center of Excellence grant from the Norwegian Research Council. This work was also partly funded by the Norwegian Research Council (NFR project 159824 Emplacement mechanisms and magma ow in sheet intrusions in sedimentary basins). We acknowledge very constructive reviews from T. Menand and two anonymous reviewers. References
Aarnes, I., Podladchikov, Y.Y. and Neumann, E.-R., in press. Formation of D-shaped geochemical proles in mac sill intrusions due to post-emplacement melt-ow induced by thermal stresse. Earth Planet. Sci. Lett. 276, 152166. Ancochea, E., Brndle, J.L., Huertas, M.J., Cubas, C.R., Hernan, F., 2003. The felsic dikes of La Gomera (Canary Islands): identication of cone sheet and radial dike swarms. J. Volcanol. Geotherm. Res. 120, 197206. Barenblatt, G.I., 2003. Scaling. Cambridge University Press, Cambridge. 171 pp. Bradley, J., 1965. Intrusion of major dolerite sills. Trans. R. Soc. N.Z., Earth. Sci. 3, 2755. Brque, C., Dupr, J.-C., Brmand, F., 2004. Calibration of a system of projection moire for relief measuring: biomechanical applications. Opt. Lasers Eng. 41, 241260. Bunger, A.P., Detournay, E., 2005. Asymptotic solution for a penny-shaped near-surface hydraulic fracture. Eng. Fract. Mech. 72, 24682486. Bunger, A.P., Detournay, E., Jeffrey, R.G., 2005. Crack tip behavior in near-surface uiddriven fracture experiments. C. R. Mec. 333, 299304. Busby-Spera, C.J., White, J.D.L., 1987. Variation in peperite textures associated with differing host-sediment properties. B. Volcanol. 49, 765775. Chang, W.-L., Smith, R.B., Wicks, C., Farrell, J.M., Puskas, C.M., 2007. Accelerated uplift and magmatic intrusion of the Yellowstone Caldera, 2004 to 2006. Science 318, 952956. Chevallier, L., Woodford, A., 1999. Morpho-tectonics and mechanism of emplacement of the dolerite rings and sills of the western Karoo, South Africa. S. Afr. J. Geol. 102, 4354. Chevallier, L., Goedhart, M., Woodford, A., 2001. The inuence of dolerite sill and ring complexes on the occurrence of groundwater in the Karoo fractured aquifers: a morpho-tectonic approach. Water Res. Com., S. Afr., vol. 937. 143 pp. Chevallier, L., Gibson, L.A., HNhleko, L.O., Woodford, A.C., Nomquphu, W., Kippie, I., 2004. Hydrogeology of fractured-rock aquifers and related ecosystems within the Qoqodala dolerite ring and sill complex, Great Kei catchment, Eastern Cape. Water Res. Com., S. Afr., vol. 1238. 127 pp. Clemens, J.D., Mawer, C.K., 1992. Granitic magma transport by fracture propagation. Tectonophysics 204, 339360. Davy, P., Cobbold, P.R., 1991. Experiments on shortening of a 4-layer model of the continental lithosphere. Tectonophysics 188, 125. Dingwell, D.B., Bagdassarov, N.S., Bussov, G.Y., Webb, S.L., 1993. Magma rheology. In: Luth, R.W. (Ed.), Experiments at high pressure and applications to the Earth's mantle. Mineralogists Association of Canada Short Course Handbook, pp. 131196. Duranti, D., Mazzini, A., 2005. Large-scale hydrocarbon-driven sand injection in the Paleogene of the North Sea. Earth Planet. Sci. Lett. 239, 327335. Fialko, Y., 2001. On origin of near-axis volcanism and faulting at fast spreading midocean ridges. Earth Planet. Sci. Lett. 190, 3139. Fialko, Y., Khazan, Y., Simons, M., 2001. Deformation due to a pressurized horizontal circular crack in an elastic half space, with applications to volcano geodesy. Geophys. J. Int. 146, 181190. Fjeldskaar, W., Helset, H.M., Johansen, H., Grunnaleite, I., Horstad, I., 2008. Thermal modelling of magmatic intrusions in the Gjallar Ridge, Norwegian Sea: implications for vitrinite reectance and hydrocarbon maturation. Basin Res. 20, 143459. doi:10.1111/j.1365-2117.2007.00347.x. Francis, E.H., 1982. Magma and sediment I. Emplacement mechanism of late Carboniferous tholeiite sills in northern Britain. J. Geol. Soc. Lond. 139, 120. Galerne, C., Caroff, M., Rolet, J., Le Gall, B., 2006. Magma-sediment mingling in an Ordovician rift basin: the Plouezec-Plourivo half-graben, Armorican Massif, France. J. Volcanol. Geotherm. Res. 155, 164178.

O. Galland et al. / Earth and Planetary Science Letters 277 (2009) 373383 Galerne, C., Neumann, E.-R., Planke, S., 2008. Geochemical architecture of the Golden Valley Sill Complex, South Africa: implications for saucer-shaped sill emplacement in sedimentary basins, in press. J. Volcanol. Geotherm. Res. 177, 425477. Galland, O., 2005. Interactions mcaniques entre la tectonique compressive et le magmatisme: expriences analogiques et exemple naturel. PhD Thesis, Universit de Rennes1, Mmoires de Gosciences-Rennes, n116, 411 pp, url: http://www. geosciences.univ-rennes1.fr/IMG/pdf/Galland.pdf. Galland, O., de Bremond d'Ars, J., Cobbold, P.R., Hallot, E., 2003. Physical models of magmatic intrusion during thrusting. Terra Nova 15, 405409. Galland, O., Cobbold, P.R., de Bremond d'Ars, J., Hallot, E., 2007. Rise and emplacement of magma during horizontal shortening of the brittle crust: insights from experimental modeling. J. Geophys. Res. 112, B06402. doi:10.1029/2006JB004604. Galland, O., Cobbold, P.R., Hallot, E., de Bremond d'Ars, J., Delavaud, G., 2006. Use of vegetable oil and silica powder for scale modelling of magmatic intrusion in a deforming brittle crust. Earth Planet. Sci. Lett. 243, 786804. Galushkin, Y.I., 1997. Thermal effects of igneous intrusions on maturity of organic matter: a possible mechanism of intrusion. Organ. Geochem. 26, 645658. Gerya, T.V., Burg, J.-P., 2007. Intrusion of ultramac magmatic bodies into the continental crust: numerical simulation. Phys. Earth Planet. Inter. 160, 124142. Goulty, N.R., Schoeld, N., 2008. Implications of simple exure theory for the formation of saucer-shaped sills. J. Struct. Geol. 30, 812817. Gudmundsson, A., 1998. Magma chambers modeled as cavities explain the formation of rift zone central volcanoes and their eruption and intrusion statistics. J. Geophys. Res. 103, 74017412. Gudmundsson, A., 2006. How local stresses control magma-chamber ruptures, dyke injections, and eruptions in composite volcanoes. Earth Sci. Rev. 79, 131. Gudmundsson, A., Brenner, S.L., 2001. How hydrofractures become arrested. Terra Nova 13, 456462. Gudmundsson, A., Philipp, S.L., 2006. How local stress elds prevent volcanic eruptions. J. Volcanol. Geotherm. Res. 158, 257268. Hallot, E., Davy, P., de Bremond d'Ars, J., Auvray, B., Martin, H., Van Damme, H., 1996. NonNewtonian effects during injection in partially crystallised magmas. J. Volcanol. Geotherm. Res. 71, 3144. Hansen, D.M., Cartwright, J.A., 2006a. The three-dimensional geometry and growth of forced folds above saucer-shaped igneous sills. J. Struct. Geol. 28, 15201535. Hansen, D.M., Cartwright, J.A., 2006b. Saucer-shaped sill with lobate morphology revealed by 3D seismic data: implications for resolving a shallow-level sill emplacement mechanism. J. Geol. Soc. Lond. 163, 509523. Hansen, D.M., Cartwright, J.A., 2007. Reply to comment by K. Thomson on The threedimensional geometry and growth of forced folds above saucer-shaped sills by D.M. Hansen and J. Cartwright. J. Struct. Geol. 29, 741744. Hansen, D.M., Cartwright, J.A., Thomas, D., 2004a. 3D seismic analysis of the geometry of igneous sills and sill junctions relationships. In: Davies, R.J., Cartwright, J.A., Stewart, S.A., Lappin, M., Underhill, J.R. (Eds.), 3D Seismic Technology: Application to the Exploration of Sedimentary Basins. . Memoirs. Geological Society, London, pp. 199208. Hansen, D.M., Cartwright, J.A., Thomas, D., 2004b. 3D seismic analysis of the geometry of igneous sills and sill junction relationships. In: Davies, R.J., Cartwright, J., Stewardt, S.A., Lappin, M., Underhill, J.R. (Eds.), 3D seismic technology: application to the exploration of sedimentary basins. Geological Society of London Memoir, London. Hansen, D.M., Redfern, J., Federici, F., di Biase, D., Bertozzi, G., 2008. Miocene igneous activity in the Northern Subbasin, offshore Senegal, NW Africa. Mar. Petrol. Geol. 25, 115. Hoek, E., Marinos, P., Benissi, M., 1998. Applicability of the geological strength index (GSI) classication for very weak and sheared rock masses. The case of the Athens Schist Formation. Bull. Eng. Geol. Environ. 57, 151160. Hogan, J.P., Price, J.D., Gilbert, M.C., 1998. Magma traps and driving pressure: consequences for pluton shape and emplacement in an extensional regime. J. Struct. Geol. 20, 11551168. Hubbert, M.K., 1937. Theory of scale models as applied to the study of geologic structures. Geol. Soc. Am. Bull. 48, 14591520. Hubbert, M.K., 1951. Mechanical basis for certain familiar geologic structures. Geol. Soc. Am. Bull. 62, 355372. Hubbert, M.K., Willis, D.G., 1957. Mechanics of hydraulic fracturing. In: Hubbert, M.K. (Ed.), Structural Geology. Hafner Publishing Company, New York, pp. 175190. Huuse, M., Mickelson, M., 2004. Eocene sandstone intrusions in the Tampen Spur area (Norwegian North Sea Quad 34) imaged by 3D seismic data. Mar. Petrol. Geol. 21, 141155. Hyndman, D.W., Alt, D., 1987. Radial dikes, lacolliths, and gelatin models. J. Geol. 95, 763774. Johnson, A.M., Pollard, D.D., 1973. Mechanics of growth of some laccolithic intrusions in the Henry Mountains, Utah, I. Field observations, Gilbert's model, physical properties and ow of the magma. Tectonophysics 18, 261309. Kavanagh, J.L., Menand, T., Sparks, R.S.J., 2006. An experimental investigation of sill formation and propagation in layered elastic media. Earth Planet. Sci. Lett. 245, 799813. Keating, G.N., Geissman, J.W., Zyvoloski, G.A., 2002. Multiphase modeling of contact metamorphic systems and application to transitional geomagnetic elds. Earth Planet. Sci. Lett. 198, 429448. Klausen, M.B., 2004. Geometry and mode of emplacement of the Thverartindur cone sheet swarm, SE Iceland. J. Volcanol. Geotherm. Res. 138, 185204. Lister, J.R., Kerr, R.C., 1991. Fluid-mechanical models of crack propagation and their application to magma transport in dykes. J. Geophys. Res. 96, 1004910077. Malthe-Srenssen, A., Planke, S., Svensen, H., Jamtveit, B., 2004. Formation of saucershaped sills. In: Breitkreuz, C., Petford, N. (Eds.), Physical geology of high-level magmatic systems. Geol. Soc. Lond. Spec. Pub., pp. 215227.

383

Mathieu, L. and Van Wyk de Vries, B., in press. Edice and substrata deformation induced by intrusive complexes and gravitational loading in the Mull volcano (Scotland). Earth Planet. Sci. Lett. doi:10.1029/2008JB005620. Mathieu, L., van Wyk de Vries, B., Holohan, E.P., Troll, V.R., 2008. Dykes, cups, saucers and sills: analogue experiments on magma intrusion into brittle rocks. Earth Planet. Sci. Lett. 271, 113. Menand, T., 2008. The mechanics and dynamics of sills in layered elastic rocks and their implications for the growth of laccoliths and other igneous complexes. Earth Planet. Sci. Lett. 267, 9399. Merle, O., Borgia, A., 1996. Scaled experiments of volcanic spreading. J. Geophys. Res. 101, 13,80513,817. Murdoch, L.C., 1993a. Hydraulic fracturing of soil during laboratory experiments. Part 2. Propagation. Gotechnique 43, 267276. Murdoch, L.C., 1993b. Hydraulic fracturing of soil during laboratory experiments. Part 3. Theoretical analysis. Gotechnique 43, 277287. Murdoch, L.C., 2002. Mechanical analysis of idealized shallow hydraulic fracture. J. Geotech. Geoenviron. Eng. 128. O'Driscoll, B., Troll, V.R., Reavy, R.J., Turner, P., 2006. The Great Eucrite intrusion of Ardnamurchan, Scotland: reevaluating the ring-dike concept. Geology 34, 189192. Pasquare, F., Tibaldi, A., 2007. Structure of a sheet-laccolith system revealing the interplay between tectonic and magma stresses at Stardalur Volcano, Iceland. J. Volcanol. Geotherm. Res. 161, 131150. Petford, N., Kerr, R.C., Lister, J.R., 1993. Dike transport of granitoid magmas. Geology 21, 845848. Planke, S., 2008. Revealing the secrets of volcanic sedimentary basins. Geo 11, 1622. Planke, S., Rasmussen, T., Rey, S.S., Myklebust, R., 2005. Seismic characteristics and distribution of volcanic intrusions and hydrothermal vent complexes in the Vring and Mre basins. In: Dor, A.G., Vining, B.A. (Eds.), Petroleum geology: Northwestern Europe and global perspectives Proceedings of the 6th Petroleum Geology Conference. Geological Society, London. Pollard, D.D., Johnson, A.M., 1973. Mechanics of growth of some laccolithic intrusions in the Henry Mountains, Utah, II. Bending and failure of overburden layers and sill formation. Tectonophysics 18, 311354. Pollard, D.D., Holzhausen, G., 1979. On the mechanical interaction between a uid-lled fracture and the Earth's surface. Tectonophysics 53, 2757. Pollard, D.D., Segall, P., 1987. Theoretical displacements and stresses near fractures in rock: with applications to faults, joins, veins, dikes, and solution surfaces. In: Atkinson, B.K. (Ed.), Fracture Mechanics of Rock. Academic Press, London, pp. 277349. Polteau, S., Mazzini, A., Galland, O., Planke, S., Malthe-Srenssen, A., 2008a. Saucershaped intrusions: occurrences, emplacement and implications. Earth Planet. Sci. Lett. 266, 195204. Polteau, S., Planke, S., Ferr, E.C., Svensen, H., Malthe-Sorenssen, A., Neumann, E.-R. and Podladchikov, Y., 2008b. Emplacement mechanisms of saucer-shaped sill intrusion as inferred from detailed eld work and AMS analyses of the Golden Valley Sill Complex, South Africa, in press. J. Geophys. Res. Ramberg, H., 1981. Gravity, deformation and the Earth's crust. Academic Press, New York. 452 pp. Rivalta, E., Bttinger, M., Dahm, T., 2005. Buoyancy-driven fracture ascent: experiments in layered gelatine. J. Volcanol. Geotherm. Res. 144, 273285. Romano, C., Giordano, D., Papale, P., Mincione, V., Dingwell, D.B., Rosi, M., 2003. The dry and hydrous viscosities of alkaline melts from Vesuvius and Phlegrean Fields. Chem. Geol. 202, 2338. Rossello, E.A., Cobbold, P.R., Diraison, M., Arnaud, N., 2002. Auca Mahuida (Neuqun basin, Argentina): a Quaternary shield volcano on a hydrocarbon-producing substrate. 5th International Symposium on Andean Geodynamics (ISAG 2002), Extended Abstracts, Toulouse, pp. 549552. Rubin, A.M., 1993. Dikes vs. diapirs in viscoelastic rock. Earth Planet. Sci. Lett. 119, 641659. Rubin, A.M., 1995. Propagation of magma-lled cracks. Annu. Rev. Earth Planet. Sci. 23, 287336. Schellart, W.P., 2000. Shear test results for cohesion and friction coefcients for different materials: scaling implications for their usage in analogue modelling. Tectonophysics 324, 116. Shoulders, S.J., Cartwright, J., 2004. Constraining the depth and timing of large-scale conical sandstone intrusions. Geology 32, 661664. Skilling, I.P., White, J.D.L., McPhie, J., 2002. Peperite: a review of magma-sediment mingling. J. Volcanol. Geotherm. Res. 114, 117. Svensen, H., Planke, S., Chevallier, L., Malthe-Sorenssen, A., Corfu, F., Jamtveit, B., 2007. Hydrothermal venting of greenhouse gases triggering Early Jurassic global warming. Earth Planet. Sci. Lett. 256, 554566. Svensen, H., Planke, S., Malthe-Sorenssen, A., Jamtvelt, B., Myklebust, R., Eldem, T.R., Rey, S.S., 2004. Release of methane from a volcanic basin as a mechanism for initial Eocene global warming. Nature 429, 542545. Thomson, K., 2007a. Determining magma ow in sills, dykes and laccoliths and their implications for sill emplacement mechanisms. Bull. Volcanol. 70, 183201. Thomson, K., 2007b. Comment on The three-dimensional geometry and growth of forced folds above saucer-shaped igneous sills by Hansen and Cartwright. J. Struct. Geol. 29, 736740. Thomson, K., Hutton, D., 2004. Geometry and growth of sill complexes: insights using 3D seismic from the North Rockall Trough. Bull. Volcanol. 66, 364375. Troll, V.R., Walter, T.R., Schmincke, H.U., 2002. Cyclic caldera collapse: piston or piecemeal subsidence? Field and experimental evidence. Geology 30, 135138. Trude, J., Cartwright, J., Davies, R.J., Smallwood, J., 2003. New technique for dating igneous sills. Geology 31, 813816. Walter, T.R., Troll, V.R., 2001. Formation of caldera periphery faults: an experimental study. Bull. Volcanol. 63, 191203. doi:10.1007/s004450100135.

You might also like