July 2014 LIP of the Month

Magma emplacement at Large Igneous Provinces

Nicolas Le Corvec

Lunar and Planetary Institute, USRA, 3600 Bay Area Blvd., Houston, TX, 77058, USA

(lecorvec@lpi.usra.edu)

Large Igneous Provinces (LIPs) are products of some of the largest magmatic (both volcanic and intrusive) events known on Earth [e.g., Coffin and Eldholm, 1994; Bryan and Ferrari, 2013; Ernst, 2014]. They have punctuated the Earth’s history and led to catastrophic mass extinctions through the injection of volatiles (CO2 and SO2) into the atmosphere during prolonged phases of LIP volcanic activity [e.g., Bond and Wignall, 2014]. LIPs are also associated with the dismembering of supercontinents through continental rifting and the formation of new ocean basins [Courtillot et al., 1999; Dalziel et al., 2000].

LIPs are commonly defined by having areal extents >0.1 Mkm2 and are relatively short lived, with maximum lifespans of ~50 Myr, but typically restricted to one or more pulses of just a few millions years [Bryan and Ernst, 2008]. These intraplate voluminous magmatic events (>0.1 Mkm3) produce an extended range of material from high-level basaltic to silicic extrusive and intrusive materials (lava flow fields, sills, dikes including radiating dike swarms, Fig. 1) [Bryan et al., 2010; Ernst and Bleeker, 2010; Muirhead et al., 2012; Ernst, 2014].


Figure 1: Types of extrusive and intrusive materials in LIPs. a. Deccan lavas of the Wai Sub-group exposed in the Western Ghats of India near Mahabaleshwar. These lavas are part of the sheet-lobe dominated Ambenali (lower part of cliffs) and Mahabaleshwar Formations (upper part). Scarp face seen here is approximately 450 m high (http://www.largeigneousprovinces.org/12may); b. A quarry in southeastern Botswana exposes a weathering profile into a Mesoproterozoic sill of the Umkondo Large Igneous Province (http://www.swanson-hysell.org/field-work-photos/); c. A 20-km-long dike in the western coastal region of the Deccan flood basalt province, where most dike swarms have N-S to NNW-SSE trends. The dike forms a dam across the Surya River near the coastal town of Dahanu (http://www.geology.sdsu.edu/how_volcanoes_work/Fissure.html); and d. 2500-2450 Matachewan and Mistassini radiating swarms, Superior Province, Canada (http://www.mantleplumes.org/GiantRadDykeSwarms.html).

Emplacement of magmatic systems

The formation of LIPs is believed to result from the upwelling of mushroom-shaped regions of solid mantle material, a.k.a. mantle plumes , but their existence is still an ongoing matter of debate [e.g., Anderson and Natland, 2005; Campbell and Kerr, 2007; Bryan and Ernst, 2008; Foulger, 2011; Ernst, 2014]. The rapid transport of magma toward the surface and formation of intrusive complexes are mainly controlled by the fracturing of the host rock through extensional fractures, also called hydrofracturing [Wadsworth, 1982]. The orientation of those magma-filled cracks is mainly controlled by the orientation of the principal stresses within the lithosphere [Rubin, 1995], thus the formation of dikes (i.e., discordant and generally subvertical, Fig. 1c) requires the minimum compressional stress (σ3) to be horizontal, and the maximum compressional stress (σ1) to be vertical or horizontal, and vice-versa for the formation of sills (i.e., bedding concordant, subhorizontal, Fig. 1b). Non-subhorizontal and non-subvertical intrusions are referred to sheets and require more complex stress orientations [Pasquarè and Tibaldi, 2007]. In addition, heterogeneities within the Earth have been recognized to play an important role in the behavior of magma propagation: e.g., numerical and analogue models have shown the influence of mechanical layering (Fig. 2 a, b, and c) [Kavanagh et al., 2006; Gudmundsson, 2011; Maccaferri et al., 2011; Barnett and Gudmundsson, 2014] or pre-existing fractures [Gaffney et al., 2007; Le Corvec et al., 2013] on direction of magma propagation.


Figure 2: a. Example of analogue modeling: A vertically propagating dyke has turned to intrude as a sill along the interface separating upper rigid layer from lower layer. Note the protrusions from the sill periphery into the lower less rigid layer (from Kavanagh et al. 2006), b. Example of analogue modeling: Formation of a dyke–sill hybrid. The dyke has reached the interface and has intruded horizontally as a sill and into the upper rigid layer (from Kavanagh et al. 2006), c. Example of numerical modeling: Successive “snapshots” of the propagating dike reaching a mechanical boundary. The Coulomb failure function variation induced on horizontal planes is shaded in background (blue : negative values; red : positive values), the dashed line represents the energetically preferred path and the shape of the dike is drawn exaggerated by a factor 1400 (from Maccaferri and al. 2011), and d. Network of Ferrar Dolerite sills south of Terra Cotta Mountain, Transantarctic Mountains, Antarctica (http://www.otago.ac.nz/geology/people/students/airoldi/)

Natural examples of giant sills can be observed in many LIP locations, e.g., the “Ferrar Dolerite” intrusions in Antarctica (Fig. 2d) [Airoldi et al., 2012; Muirhead et al., 2012]. Transition from feeding dikes to sills may occur at boundaries between two distinct layers. This reorientation mechanism is controlled by several parameters: e.g., the difference in stiffness between the two layers [Rivalta et al., 2005; Kavanagh et al., 2006; Maccaferri et al., 2011; Ritter et al., 2013; Barnett and Gudmundsson, 2014], stress barriers [Gretener, 1969; Gudmundsson, 1990; Barnett and Gudmundsson, 2014], the difference in fracture toughness [Maccaferri et al., 2011], the intrusion temperature [Chanceaux and Menand, 2014], and the rock interface strength (a.k.a., Cook-Gordon debonding) [Barnett and Gudmundsson, 2014; Kavanagh and Pavier, 2014].

Emplacement of large amounts of magma in the lithosphere requires space. Observation at the Ferrar Dolerite intrusions shows that in addition to the thickest sills (e.g., Finger Mountain Sill), other sill intrusions, although thinner and shorter in length, are found to ascend the stratigraphy in ‘step-wise’ fashion. These lead to a configuration of sills connected by sheets (Fig. 3a) [Muirhead et al., 2012]. As the dike rises toward the surface, it may change direction of propagation due to the potential mechanisms cited earlier. Depending of its size, a sill may bend its overburden which will influence the local stress and allow the formation of sheets. These new sheets may reach an interface along their path and transform into sills, which will lead to the formation of interconnected sills such as the ones observed by Muirhead and al. (2012) at Allan Hills and Terra Cotta Mountain at the Ferrar Dolerite intrusions. An alternative mechanism to the formation of interconnected sills is described by Barnett and Gudmundsson (2014) where a succession of stiff and soft layers acts as stress barriers (Fig. 3b).


Figure 3: a. Distribution of the basement sill in South Victoria Land (from Muirhead et al., 2012), b. Finite element model (FEM) of a layered crustal segment where both dyke and sill follow the maximum compressive principal stress (σ1) trajectories represented by the red ticks. The black line shows the paths which the dyke and later the sill would follow as they reach different layers with varying stiffness (from Barnett and Gudmundsson 2014).

Giant radiating dike swarms

Another peculiar type of magma intrusion in LIPs produces radiating dike swarms [Halls, 1982; Fahrig, 1987; Ernst and Baragar, 1992; Ernst et al., 1995; Hoek, 1995; Ernst et al., 2001]. Radiating dike swarm are seen on Earth, Mars and Venus [Grosfils and Head, 1994a; Grosfils and Head, 1994b; Mège and Masson, 1996; Ernst et al. 2003; Galgana et al., 2013] and have lengths ranging from tens up to 2200 km from the focal point of the swarm [Ernst et al., 2001]. Giant dike swarms, are defined to be longer than 300 km. Their formation may involve magma chambers at levels of neutral buoyancy [Grosfils and Head, 1995], or could be explained by a combination of tectonic loads applied to a homogeneous lithosphere [Galgana et al., 2013]. Despite the importance of the tectonic state of stress, heterogeneities within the lithosphere also have a strong influence on the behavior of magmatic systems as shown earlier. The presence of crustal low density rock above mantle higher density material may trap magmas of intermediate density near the level of neutral buoyancy, which favors the formation of magma chambers at the interface [e.g., Takada, 1989; Walker, 1989; Lister, 1991]. The stability of such chambers and their failure can vary depending on the tectonic environment and location of the chamber within the lithosphere. Magma chamber failure is controlled by the magma chamber wall in-plane tangential (στ) and out-of-plane hoop (σθ) stress components in the context of an axisymmetric model [Galgana et al., 2011; Galgana et al., 2013]. Depending on the stress reaching the tensile failure of the host rock, different magmatic intrusions will be favored: στ will favor vertical and circumferential dikes or lateral sill intrusions while σθ favors radial dike intrusions [Grosfils, 2007].

Using FEMs (finite element models), Le Corvec et al. [2014] have also studied the influence of upward flexure due to a rising mantle plume on the stability of magma chambers, the type of intrusion and their distance of propagation. Their result shows that failure of a magma reservoir in the upper part of the lithosphere is controlled by the hoop stress (σθ), which will lead to the formation of radial dikes (Fig. 4a and b). The failure of the magma chamber is also controlled by the thickness of the crust (i.e., the depth of the magma chamber within the lithosphere). The resulting radial dike can travel to distances up to ~300 km depending on the thickness of the lithosphere and the depth of the magma chamber (Fig. 4c). A cross-section of the lithosphere shows that the minimum deviatoric compressional stress is out of the plane favoring lateral dike propagation (Fig. 4d).


Figure 4: a. Amount of overpressure at failure and type and location of failure affecting the magma chamber for the upward flexure environment. The dotted, dash and dots and dashed lines represents FEMs with a lithosphere thickness of 20, 30 and 40 km, respectively. The symbols in green and orange represent a failure mechanism controlled by the tangential stress (στ) and hoop stress (σθ), respectively. The diamond and triangle symbols represent the failure location at the center and bottom of the magma chamber’s wall, respectively. Open and filled symbols represent stable and unstable magma chambers, respectively. Magma chambers that fail at their bottom are considered unviable. The grey area represents the appropriate conditions for the formation of radial dikes with radial distances > 200 km. b. Overpressure and stresses controlling the failure of the magma chamber for three different lithosphere thicknesses (20, 30 and 40 km). For the upward flexure environment, the left side of each figure represents the stresses (στ and σθ) along the lower half of the magma chamber. In some cases, no overpressure is needed to reach failure as the magma chamber is unstable. c. Distance favorable to the propagation of radial dikes within lithospheres subjected to upward flexure. Each symbol represents FEMs with their magma chamber failure controlled by the hoop stress (i.e., favoring radial dike formation). The symbol surrounded by a circle is the example of the deviatoric stress orientation for a lithosphere subjected to upward flexure shown in d. for a lithosphere subjected to downward flexure. The deformation in d. is increased by a factor of 5.

Conclusion

LIP eruptions lead to the emplacement of exceptionally large-scale magmatic intrusions and extrusions, during which large surfaces of the Earth or other planets are affected. The understanding of such phenomena lies in the understanding of magma propagation from source to surface. New evidence has shown the impact of crustal heterogeneities on the direction of magma propagation from dikes to sills to sheet to sills or dikes as well as the impact of the intrusive complex on the local state of stress, which can overcome the regional tectonic stress. The combination of fieldwork and analogue and numerical models has better constrained our understanding on how magma propagates within the lithosphere. More complex models will have to be developed in the future to constrain the thermal activity of these mechanisms as well as the temporal accommodation of the lithosphere to the intrusive and extrusive loads, since LIPs are related to attempted or successful continental breakup.

References

Airoldi, G., J. D. Muirhead, E. Zanella, and J. D. L. White (2012), Emplacement process of Ferrar Dolerite sheets at Allan Hills (South Victoria Land, Antarctica) inferred from magnetic fabric, Geophysical Journal International, 188(3), 1046-1060.

Anderson, D. L., and J. H. Natland (2005), A brief history of the plume hypothesis and its competitors: Concept and controversy, Geological Society of America Special Papers, 388, 119-145.

Barnett, Z. A., and A. Gudmundsson (2014), Numerical modelling of dykes deflected into sills to form a magma chamber, Journal of Volcanology and Geothermal Research, 281(0), 1-11.

Bond, D. P. G., and P. B. Wignall (2014), Large igneous provinces and mass extinctions: An update, Geological Society of America Special Papers, 505.

Bryan, S. E., and R. E. Ernst (2008), Revised definition of Large Igneous Provinces (LIPs), Earth-Science Reviews, 86(1–4), 175-202.

Bryan, S. E., and L. Ferrari (2013), Large igneous provinces and silicic large igneous provinces: Progress in our understanding over the last 25 years, Geological Society of America Bulletin.

Bryan, S. E., I. U. Peate, D. W. Peate, S. Self, D. A. Jerram, M. R. Mawby, J. S. Marsh, and J. A. Miller (2010), The largest volcanic eruptions on Earth, Earth-Science Reviews, 102(3–4), 207-229.

Campbell, I. H., and A. C. Kerr (2007), The Great Plume Debate: Testing the plume theory, Chemical Geology, 241(3–4), 149-152.

Chanceaux, L., and T. Menand (2014), Solidification effects on sill formation: An experimental approach, Earth and Planetary Science Letters, 403(0), 79-88.

Coffin, M. F., and O. Eldholm (1994), Large igneous provinces: Crustal structure, dimensions, and external consequences, Reviews of Geophysics, 32(1), 1-36.

Courtillot, V., C. Jaupart, I. Manighetti, P. Tapponnier, and J. Besse (1999), On causal links between flood basalts and continental breakup, Earth and Planetary Science Letters, 166(3–4), 177-195.

Dalziel, I. W. D., L. A. Lawver, and J. B. Murphy (2000), Plumes, orogenesis, and supercontinental fragmentation, Earth and Planetary Science Letters, 178(1–2), 1-11.

Ernst, R. E. (2014), Large Igneous Provinces, Cambridge University Press, in press.

Ernst, R., and W. Bleeker (2010), Large igneous provinces (LIPs), giant dyke swarms, and mantle plumes: significance for breakup events within Canada and adjacent regions from 2.5 Ga to the Present<i></i>, Canadian Journal of Earth Sciences, 47(5), 695-739.

Ernst, R. E., and W. R. A. Baragar (1992), Evidence from magnetic fabric for the flow pattern of magma in the Mackenzie giant radiating dyke swarm, Nature, 356(6369), 511-513.

Ernst, R. E., J. W. Head, E. Parfitt, E. Grosfils, and L. Wilson (1995), Giant radiating dyke swarms on Earth and Venus, Earth-Science Reviews, 39(1–2), 1-58.

Ernst, R., E. Grosfils, and D. Mège (2001), Giant Dike Swarms: Earth, Venus, and Mars, Annual Review of Earth and Planetary Sciences, 29(1), 489-534.

Ernst, R.E.,D.W. Desnoyers, J.W. ,Head, and E.B.,Grosfils, E.B. (2003),.Graben–fissure systems in Guinevere Planitia and Beta Regio (264–312º E, 24–60º N), Venus, and implications for regional stratigraphy and mantle plumes/diapirs. Icarus, 164: 282–316.

Fahrig, W. (1987), The tectonic settings of continental mafic dyke swarms: failed arm and early passive margin, Mafic dyke swarms. Edited by HC Halls and WF Fahrig. Geological Association of Canada, Special Paper, 34, 331-348.

Foulger, G. R. (2011), Plates vs plumes: A geological controversy, John Wiley & Sons.

Gaffney, E. S., B. Damjanac, and G. A. Valentine (2007), Localization of volcanic activity: 2. Effects of pre-

existing structure, Earth and Planetary Science Letters, 263(3-4), 323-338.

Galgana, G. A., P. J. McGovern, and E. B. Grosfils (2011), Evolution of large Venusian volcanoes: Insights from coupled models of lithospheric flexure and magma reservoir pressurization, J. Geophys. Res., 116(E3), E03009.

Galgana, G. A., E. B. Grosfils, and P. J. McGovern (2013), Radial dike formation on Venus: Insights from models of uplift, flexure and magmatism, Icarus, 225(1), 538-547.

Gretener, P. E. (1969), On the mechanics of the intrusion of sills, Canadian Journal of Earth Sciences, 6(6), 1415-1419.

Grosfils, E., and J. Head (1994a), Emplacement of a radiating dike swarm in western Vinmara Planitia, Venus: interpretation of the regional stress field orientation and subsurface magmatic configuration, Earth Moon Planet, 66(2), 153-171.

Grosfils, E. B. (2007), Magma reservoir failure on the terrestrial planets: Assessing the importance of gravitational loading in simple elastic models, Journal of Volcanology and Geothermal Research, 166(2), 47-75.

Grosfils, E. B., and J. W. Head (1994b), The global distribution of giant radiating dike swarms on Venus: Implications for the global stress state, Geophysical Research Letters, 21(8), 701-704.

Grosfils, E. B., and J. W. Head (1995), Radiating dike swarms on Venus: evidence for emplacement at zones of neutral buoyancy, Planetary and Space Science, 43(12), 1555-1560.

Gudmundsson, A. (1990), Emplacement of dikes, sills and crustal magma chambers at divergent plate boundaries, Tectonophysics, 176(3-4), 257-275.

Gudmundsson, A. (2011), Deflection of dykes into sills at discontinuities and magma-chamber formation, Tectonophysics, 500(1–4), 50-64.

Halls, H. C. (1982), The Importance and Potential of Mafic Dyke Swarms in Studies of Geodynamic Processes, 1982.

Hoek, J. D. (1995), Dyke propagation and arrest in Proterozoic tholeiitic dyke swarms, Vestfold Hills, East Antarctica, Physics and Chemistry of Dykes, 79-93.

Kavanagh, J. L., and M. J. Pavier (2014), Rock interface strength influences fluid-filled fracture propagation pathways in the crust, Journal of Structural Geology, 63(0), 68-75.

Kavanagh, J. L., T. Menand, and R. S. J. Sparks (2006), An experimental investigation of sill formation and propagation in layered elastic media, Earth and Planetary Science Letters, 245(3-4), 799-813.

Le Corvec, N., T. Menand, and J. Lindsay (2013), Interaction of ascending magma with pre-existing crustal fractures in monogenetic basaltic volcanism: an experimental approach, Journal of Geophysical Research: Solid Earth, 118(3), 968-984.

Le Corvec, N., P. J. McGovern, and E. Grosfils (2014), Effects of crustal-scale mechanical layering on magma chamber failure and magma propagation within the Venusian lithosphere, Lunar Planet. Sci. 45th, Abstract 2330.

Lister, J. R. (1991), Steady solutions for feeder dykes in a density-stratified lithosphere, Earth and Planetary Science Letters, 107(2), 233-242.

Maccaferri, F., M. Bonafede, and E. Rivalta (2011), A quantitative study of the mechanisms governing dike propagation, dike arrest and sill formation, Journal of Volcanology and Geothermal Research, 208(1-2), 39-50.

Mège, D., and P. Masson (1996), Stress models for Tharsis formation, Mars, Planetary and Space Science, 44(12), 1471-1497.

Muirhead, J. D., G. Airoldi, J. V. Rowland, and J. D. L. White (2012), Interconnected sills and inclined sheet intrusions control shallow magma transport in the Ferrar large igneous province, Antarctica, Geological Society of America Bulletin, 124(1-2), 162-180.

Pasquarè, F., and A. Tibaldi (2007), Structure of a sheet-laccolith system revealing the interplay between tectonic and magma stresses at Stardalur Volcano, Iceland, Journal of Volcanology and Geothermal Research, 161(1–2), 131-150.

Ritter, M. C., V. Acocella, J. Ruch, and S. L. Philipp (2013), Conditions and threshold for magma transfer in the layered upper crust: Insights from experimental models, Geophysical Research Letters, 40(23), 2013GL058199.

Rivalta, E., M. Bottinger, and T. Dahm (2005), Buoyancy-driven fracture ascent: Experiments in layered gelatine, Journal of Volcanology and Geothermal Research, 144(1-4 SPEC. ISS.), 273-285.

Rubin, A. M. (1995), Propagation of magma-filled cracks, Annual Review of Earth & Planetary Sciences, 23, 287-336.

Takada, A. (1989), Magma transport and reservoir formation by a system of propagating cracks, Bull Volcanol, 52(2), 118-126.

Wadsworth, W. J. (1982), Physics of Magmatic Processes : R.B. Hargraves (Editor). Princeton University Press, Princeton, N.J., 1980, 585 pp, Physics of the Earth and Planetary Interiors, 28(3), 273-274.

Walker, G. P. L. (1989), Gravitational (density) controls on volcanism, magma chambers and intrusions, Australian Journal of Earth Sciences, 36(2), 149-165.