v13n1 Supplement — Putirka “Methods and Further Reading”

Methods and Further Reading by Keith Putirka 

Supplement to the February 2017 (v13n1) issue of Elements.

Disclaimer: this table was not reviewed by Elements nor were its contents verified.

Methods

Calculating Clinopyroxene and Amphibole Pressures

Clinopyroxene pressure and temperature estimates use Eqns. P1 and T2 of Putirka et al. (1996) for anhydrous systems (Hawaii) and the thermometers and barometers of Putirka et al. (2003) for hydrous systems (e.g., Cascades, Andes). For some Cascade systems (e.g., Lassen Peak) a new barometer appears to work slightly better, providing better matches for tests of equilibrium. In this new barometer, below, XiCpx are mole fractions of jadeite (Jd) and Diopside + Hedenbergite (DiHd) in clinopyroxene, and Xiliq terms are cation fractions of the indicated oxides (see Putirka 2008):

This new barometer works best when paired with the thermometer expressed as Eqn. 33 in Putirka (2008).

Pressures are converted to depths using the density model of Hill and Zucca (1987), from which the following equation is derived: depth (km) = -2.77x10-5[P4] – 2.0x10-3[P3] - 4.88x10-2[P2] + 3.6[P] - 6.34x10-2, where P is in kbar.  At the Cascades, P-T estimates make use of Putirka et al. (2003), whose models are designed for hydrous and SiO2-rich systems; the density-depth models of Mavko and Thompson (1983) and DeBari and Greene (2011) are used to convert pressure estimates to depth (km): 2.4x104[P3] - 2.11x102[P2] + 3.66 [P] + 0.46, and where P is also in kbar. Cascade data are from GEOROC; Hawaiian data are as in Putirka (2008). Two-pyroxene P-T estimates and Mauna Kea data are as in Putirka (2008); Mauna Loa gabbroic samples are from Gafney (2002), also using the two-pyroxenes thermobarometers of Putirka (2008). For the Andes, pressure is converted to depth using Prezzi et al. (2009), from which we obtain: depth (km) = 4.88 + 3.30[P] - 0.0137[P - 18.01]2.

Amphibole temperature estimates are from Eqn. 5 in Putirka (2016a), in which only amphibole compositions are used as input (independent of pressure). For the Cascades, we find that only the amphibole barometer of Anderson and Smith (1995) yields pressures that match those obtained from Cpx barometry (nearly identical mean P estimates, of ca.1 kbar); these are illustrated, converted to depth as above.

Mantle potential temperatures are calculated as in Putirka (2016b; see electronic attachment in that paper) and assume that the picritic parental magmas at arcs contain 2 wt. % H2O.

Calculating Magma Density

We use the silicate liquid density model of Lange and Carmichael (1990) for anhydrous systems, with the hydrous correction of Ochs and Lange (1999). Water contents for primitive Cascade magmas are probably closer to 1 wt. % H2O (Wallace 2005) compared to 3 wt. % as used here. But most arc basalts have 1-6 wt. % H2O (Wallace 2005, Table 1), so a middle value is used. For water-saturated magmas, bulk density is taken as a weighted fraction of silicate liquid and fluid, where fluid densities are calculated as in Holland Powell (1991). For crystalline magmas, we again take a weighted proportion of liquid and crystals, assuming all crystals are olivine, at 3.3 g/cm3.

A crucial aspect of our density calculations is that we estimate magma densities at the pressure and temperature conditions at which the magmas are stored, rather than comparing all magmas at, say, 1 atmosphere pressure. This more realistic approach is crucial in that absent such an approach, the density vs. MgO comparison yields a density minimum (at about 8 wt. % MgO). This minimum (as obtained for MORB, see Stolper and Walker 1980) disappears for the Cascades and Hawaii when density is calculated at P-T conditions of storage based on Cpx thermobarometry. This approach also shifts the absolute values for calculated density.

The best approach for correcting liquid densities for thermal expansion or pressure-induced contraction would be to use P-T estimates from Cpx + liq thermobarometers. To calculate density for a much wider array of compositions, though, thermobarometers are used that do not rely on the composition of equilibrium crystals.   In Figure 3, T is first calculated using the P-independent (liquid composition only) Eqn. 14 of Putirka (2008) as an initial estimate; this T estimate is then used to estimate P (see below), and this P estimate is then used to refine our T estimate, using Eqn. 15 (P-dependent, but again using liquid compositions only) from Putirka (2008). These two thermometers (Eqns. 14 and 15 from Putirka 2008) assume that such liquids are saturated with olivine ± other phases at the liquidus. The model appears to slightly over-estimate temperatures for rhyolites (estimates range from 850-1000oC) but should be roughly representative of liquidus temperatures for andesitic to basaltic liquids. To estimate P, the only analytic barometers available for liquids are based on Si-activity (Putirka 2008), but application of these yields negative pressures for a great many compositions examined here. However, the highest T estimates for Cpx (e.g., Fig. 1, but also in other studies, e.g., Armienti et al. 2013) yield rather coherent P-T trends, with a near constant dP/dT; these Cpx saturation conditions are judged likely to provide a useful approximation regarding the depths at which magmas stall and partially crystallize. So P is thus estimated using a simple regression high-T Cpx + liquid P-T estimates: P(kbar) = 0.0811[T(oC)] - 85.787, applying a P-estimate of 2 kbar for all cases where the model yields a pressure less than this value. Yet another method, not applied here, would be to regress Cpx-derived P-T estimates against liquid composition, and then extrapolate such a model to obtain P and T for compositions where Cpx compositions are not available.

The variety of basalts produced by mantle partial melting

It is now well demonstrated by experiments that a wide range of basalt compositions are produced by partially melting peridotitic mantle. We emphasize that picrities will pond and partially crystallize in the lower crust and upper mantle, but mantle partial melts can range down to as little as 6-8% MgO, according to some experimental studies and so can pond at a wide range of depths in the crust – a range made wider still by adding water. See Takahashi and Kushiro (1983), O’Hara (1968), Stolper (1980), Takahashi’s (1986), Kinzler and Grove (1992), Walter (1998), Gaetani and Grove (1998), and Pickering-Witter and Johnston (2000), which illustrate the wide range of basalt compositions that can be produced, depending upon, T, P and bulk composition, including water.

Deep Seismic Activity

We highlight Aki and Koyanagi (1981) for their relevance to our Hawaiian example, but more recent studies also indicate magmatically induced seismic activity at depths extending at least into the lower crust, notably Shelly and Hill (2011) at Mammoth Mountain in California and Okubo and Wolfe (2008) at Mauna Loa. And in a fascinating study, Greenfield and White (2015) show seismic evidence not only of magmatic movement in the lower crust beneath Iceland, but the possible seismic signal as magma is transferred through a dike from a deeper sill to a more shallow one.

Mantle Xenoliths

There are many studies of mantle xenoliths, which are important for understanding mantle composition (e.g., McDonough 1990) and mantle processes (Ducea and Saleeby 1996), and are an important part of the magma transport story (e.g., Spera, 1984). Many studies concern xenoliths erupted in continental interiors, but Arai and Ishimaru (2008) provide a nice review of mantle xenoliths from arc localities. Ross et al (1954) may have been the first to recognize that ultramafic xenoliths in volcanic rocks may represent mantle material, and so by implication, at least some volcanic rocks are sourced in the upper mantle.

Sr/Y ratios and Garnet Stability

Chapman et al. (2015) propose that high magmatic Sr/Y ratios might reflect partial crystallization at great crustal depths. But the issue is not simple. Basaltic magmas require pressures that are quite high to place garnet on the liquids. For example, experiments by Baker and Eggler (1983) show that Aleutian high-Al basalts can precipitate garnet at the liquidus P>17 kbar (they state 19 kbar in their abstract, but their phase diagram places the boundary at about 17 kbar), where it replaces plagioclase. Plagioclase is the liquidus phase at all lower pressures. But their basalts are quite low in MgO (<10%), and high in Al. More mafic magmas, as might characterize the picritic lavas that precipitate high Fo olivine, are much more likely to have lower Al2O3 and precipitate garnet at even greater pressures. Maaloe (2004) obtains garnet (not on the liquidus) at P>22 kbar. And Eggins (1992) did not find garnet on the liquidus of his high MgO liquids at any pressure. The moderate MgO basalts (4-6% MgO) that Baker and Eggler (1983) examined, are much more likely to rise to middle or upper crust depths, and so will precipitate plagioclase, amphibole or olivine, depending upon water contents and pressure. Garnet fractionation, and hence high Sr/Y ratios, are much more likely to occur by partial melting of pre-existing crust, rather than by fractionation of primitive, mantle-derived basalts.

Lower Crust

The composition of the lower crust is usually considered to be mafic (Rudnick and Gao 2003) but has recently become a topic of debate. Hacker et al. (2015), for example, suggest that the lower crust is much more felsic than traditionally understood. The idea stems, in part, from an inference that lower-crust P-wave velocities of 7.0 km/s may represent andesitic, rather than basaltic, bulk compositions (Behn and Kelemen 2003). This appears to be contradicted in a review of experimental data by Huang et al. (2013); they show that only basaltic compositions are likely to have such high P-wave velocities. In any case, the Hacker et al. (2015) interpretation feeds into models of lower crust evolution. For example, Ducea and Saleeby (1996) emphasize the precipitation of dense, mafic cumulates as being important for the formation of granitic curst, and these cumulates are themselves denser the underlying mantle and eventually “delaminate”, i.e., sink into the mantle. This delamination leads to passive upwelling of asthenosphere, which then initiates a new round of mantle melting. In contrast, Kelemen and Behn (2016), while they also accept some degree of delamination, suggest that “relamination” is more important for explaining lower crust composition at arcs. In this model, andesitic composition materials (sediments, plutons from the arc) from subaerial parts of the arc are fed into the trench and forcefully subducted, but only partially, eventually rising upwards to re-laminate the base of the arc crust (see Hacker et al. 2015, their Figure. 15).

Magma Transport and Eruption Triggering

We refer to Tait et al. (1989) in the main text for convenience, as they nicely build on the earlier work of Blake (1984) and Blake (1981), and present other ideas of eruption triggering that are still relevant today. But Daly (1911) almost certainly deserves some credit for appreciating the importance of volatile saturation as a triggering mechanism. And subsequent studies by Snyder (2000) are also important, while Fowler and Spera (2008) provide a nice follow up to Tait et al. (1989), using more detailed and realistic phase equilibria, for four model basalt compositions—their Figure 2 shows that magma overpressures increase at about 850-900oC, or in other words, just after the onset of amphibole saturation for many systems. However, it is not clear that amphibole, or any other phase necessarily acts as a “phase equilibria trigger”, but rather the triggering mechanism is a function of total crystallization, as inferred by Tait et al. (1989). Also, while we again refer to Sparks et al. (1977), in part for the plethora of ideas contained therein regarding recharge and eruption processes, Eichelberger (1980) provides another early study that was key in demonstrating how mafic recharge magmas, upon crystallization and by reaching vapor saturation, may induce magma mixing, and by implication, trigger volcanic eruptions. Eichelberger (1995) also provides a very nice overview of possible eruption triggering mechanisms. Also, discussions of magma mobility and transport are utterly incomplete without references to Bruce Marsh’s (1981) ideas about crystallinity; it is his recognition that magmas with >50-55% crystals are immobile that have vitally influenced ideas of magma transport.

Finally, while melt production in the mantle is discounted as a useful “ultimate” cause of melting in the main text, recent work by Poland et al. (2012) at Kilauea, Hawaii suggests otherwise. They suggest that melt production rates in the mantle may indeed provide a useful ultimate or even proximal cause of eruptions there, as tracked by deep-seated CO2 degassing. An interesting issue here is whether CO2 degassing represents magmas ponded in subcrustal sills (and so not necessarily connected to mantle melt production), or freshly delivered from even greater mantle depths.

Crystal Residence Times

The result that crystal residence times are in the range 103-105 years is nicely illustrated in Cooper and Kent (2014) and other works cited in that paper. Cooper (2015) shows that these resident times (age dates) might be no longer than the ages of the volcanic centers from which the zircon are derived. The issue is perhaps no better illustrated than at the Lassen Volcanic center, where Klemetti and Clynne (2013) show that almost no zircons yield age dates that match eruption ages, but few to none actually precede the formation of the Lassen Volcanic center. This indicates that recharge events probably heat felsic systems beyond temperatures of zircon saturation.

Summary

A key thesis of the main text is that progress will be made by better understanding the links between storage depths, cooling rates, and recharge input rates, and by using observations of such to better understand the various causes of eruptions. Snyder (2000) provides a nice summary of cooling rates and the effects of mafic/felsic magma ratios with regard to cooling of one and heating of the other. He argues that heating of the felsic magma by mafic recharge is a key trigger for large explosive eruptions, since the heat added to a felsic system (from the recharge magma) significantly decreases the solubility of water in such magmas. Hawkesworth et al. (2000) also provide an excellent summary of tools that can be applied to understand the time scales of magma storage and crystallization, which can be directly applied with thermometers and barometers to better understand eruption triggering mechanisms.

REFERENCES

Aki K, Koyanagi R (1981) Deep volcanic tremor and magma ascent mechanism under Kilauea Hawaii. Journal of Geophysical Research 86: 7095-7109

Anderson JL, Smith DR (1995) The effects of temperature and fO2 on the Al-in-hornblende barometer. American Mineralogist 80: 549-559

Arai S, Ishimaru S (2008) Insights into petrological characteristics of the lithosphere of mantle wedge beneath arcs through peridotite xenoliths: a review. Journal of Petrology 49: 665-695

Armienti P, Perinelli C, Putirka KD (2013) A new model to estimate deep-level magma ascent rates, with applications to Mt. Etna (Sicily, Italy). Journal of Petrology 54: 795-813

Baker DR, Eggler DH (1983) Fractionation paths of Atka (Aleutians) High-alumina basalts: constraints from phase relations. Journal of Volcanology and Geothermal Research 18: 387-404

Behn MD, Kelemen PB (2003) Relationship between seismic P-wave velocity and the composition of anhydrous igneous and meta-igneous rocks. Geochemistry, Geophysics, Geosystems 4: doi:10.1029/2002gc000393

Blake S (1981) Volcanism and the dynamics of open magma chambers. Nature 289: 783-785

Blake S (1984) Volatile ovesrsaturation during the evolution of silicic magma chambers as an eruption trigger. Journal of Geophysical Research 89: 8237-8244

Chapman JB, Ducea MN, DeCelles PG, Profeta L (2015) Tracking changes in crustal thickness during orogenic evolution with Sr/Y: an example from the North American Cordillera. Geology doi:10.1130/G36996.1

Cooper KM (2015) Timescales of crustal magma reservoir processes: insights from U-series crystal ages. In: Caricchi L, Blundy JD, Chemical, Physical and Temporal Evolution of Magmatic Systems, Geological Society, London, Special Publications 422: 141-174

Cooper KM, Kent A (2014) Rapid remobilization of magmatic crystals kept in cold storage. Nature 506: 480-483

Daly RA (1911) The nature of volcanic action. Proceedings of the American Academy of Arts and Sciences 47: 47-122

DeBari SM, Greene AR (2011) Vertical stratification of composition, density, and inferred magmatic processes in exposed arc crustal sections. In: Brown D, Ryan PD, Arc-Continent Collision, Frontiers in Earth Sciences, Springer-Verlag, Berlin, doi: 10.1007/978-3-540-88558-0-5

Ducea MN, Saleeby JB (1998) Buoyancy source for a larger, unrooted mountain range, the Sierra Nevada, California: evidence form xenolith thermobarometry. Journal of Geophysical Research 101: 8229-8244

Eichelberger JC (1980) Vesiculation of mafic magma during replenishment of silicic reservoirs. Nature 288: 446-450

Eichelberger JC (1995) Silicic volcanism: ascent of viscous magmas from crustal reservoirs. Annual Review of Earth and Planetary Sciences 23: 41-63

Eggins SM (1992) Petrogenesis of Hawaiian tholeiites, 1, Phase equilibria constraints. Contributions to Mineralogy and Petrology 110: 387–397

Fowler SJ, Spera FJ (2008) Phase equilibria trigger for explosive volcanic eruptions. Geophysical Research Letters 35: doi:10.1029/2008FL033665

Gaetani GA, Grove T (1998) The influence of melting of mantle peridotite. Contributions to Mineralogy and Petrology 131: 323-346

Gafney A (2002) Environments of crystallization and compositional diversity of Mauna Loa xenoliths. Journal of Petrology 43: 963-980

Greenfield T, White RS (2015) Building Icelandic igneous crust by repeated melt injections. Journal of Geophysical Research, doi: 10.1002/2015JB012009

Hacker BR, Kelemen PB, Behn MD (2015) Continental lower crust. Annual Reviews of Earth and Planetary Science 43: 167-205

Hawkesworth C, Blake S, Ebans P, Hughes R, MacDonald R, Thomas LE, Turner SP, Zellmer G (2000) Time scales of crystal fractionation in magma chambers—physical, isotopic and geochemical perspectives. Journal of Petrology 41: 991-1006

Holland TJB, Powell R (1991) A compensated-Redlich-Kwong (CORK) equation for volumes and fugacities of CO2 and H2O in the range 1 bar to 50 kbar and 100-1600oC. Contributions to Mineralogy and Petrology 109: 265-273

Huang Y, Chubakov V, Mantovani F, Rudnick RL, McDonough WF (2013) A reference Earth model for the heat-producing elements and associated geoneutrino flux. Geochemistry, Geophysics, Geosystems 14: doi:10.1002/ggge20129

Kelemen P, Behn, MD (2016) Formation of lower continental crust by relamination of buoyant arc lavas and plutons. Nature Geoscience 9: doi:10.1038/NGE02662

Kinzler RJ, Grove TL (1992) Primary magmas of mid-ocean ridge basalts 1. Experiments and methods. Journal of Geophysical Research 97: 6885-6906

Klemetti E, Clynne M (2014) Localized rejuvination of a crystal mush recorded in zircon temporal and compositional variation at the Lassen Volcanic Center, Northern California. Plos One, doi: 10.1371/journal.pone.0133157

Lange RL, Carmichael ISE (1990) Thermodynamic properties of silicate liquids with emphasis on density, thermal expansion and compressibility. In: Nicholls J, Russell JK, Modern Methods of Igneous Petrology. Mineralogical Society of America, Reviews in Mineralogy 24: 25-64

Maaloe S (2004) The P-T-phase relations of an MgO-rich Hawaiian tholeiite: the compositions of primary Hawaiian tholeiites. Contributions to Mineralogy and Petrology 148: 236-246

Marsh B (1981) On the crystallinity, probability of occurrence, and rheology of lava and magma. Contributions to Mineralogy and Petrology 78: 85-98

Mavko BB, Thompson GA (1983) Crustal and upper mantle structure of the northern and central Sierra Nevada. Journal of Geophysical Research 88: 5874-5892

McDonough WF (1990) Constraints on the composition of the continental lithospheric mantle. Earth and Planetary Science Letters 101: 1-18

Ochs FA III, Lange RA (1999) The density of hydrous magmatic liquids. Science 283: 1314-1317

O’Hara MJ (1968) Are ocean floor basalts primary magmas? Nature 220: 683-686

Okubo PG, Wolfe CJ (2008) Swarms of similar long-period earthquakes in the mantle beneath Mauna Loa Volcano. Journal of Volcanology and Geothermal Research 178: 787-794

Pickering-Witter J, Johnston AD (2000) The effects of variable bulk composition on the melting systematics of fertile peridotitic assemblages. Contributions to Mineralogy and Petrology 140: 190-211

Poland MP, Miklius A, Sutton AJ, Thornber CR (2012) A mantle-driven surge in magma supply to Kilauea Volcano during 2003-2007. Nature Geoscience doi:10.1038/NGE01426

Prezzi CB, Gotze H-J, Schmidt S (2009) 3D density model of the central Andes. Physics of the Earth and Planetary Interiors 177: 217-234

Putirka K (2008) Thermometers and Barometers for Volcanic Systems. Reviews in Mineralogy and Geochemistry 69: 61-120

Putirka K, Johnson M, Kinzler R, Longhi J, Walker, D. (1996) Thermobarometry of mafic igneous rocks based on clinopyroxene-liquid eqilibria, 0-30 kbar. Contributions to Mineralogy and Petrology 123: 92-108

Putirka K (2016a) Amphibole thermometers and barometers for igneous systems and some implications for eruption mechanisms of felsic magmas at arc volcanoes. American Mineralogist 101: 841-858

Putirka K (2016a). Rates and styles of planetary cooling on Earth, Moon, Mars and Vesta, using new models for oxygen fugacity, ferric-ferrous ratios, olivine-liquid Fe-Mg exchange, and mantle potential temperature. American Mineralogist 101: 819-840.

Putirka KD, Mikaelian H, Ryerson F, Shaw H (2003) New clinopyroxene-liquid thermobarometers for mafic, evolved, and volatile-bearing lava compositions, with applications to lavas from Tibet and the Snake River Plain, Idaho. American Mineralogist 88: 1542-1554

Ross CS, Foster MD, Myers AT (1954) Origin of dunites and of olivine-rich inclusions in basaltic rocks. American Mineralogist 39: 693-737

Rudnick RL, Gao S (2003) Composition of the continental crust. In: Rudnick, RL, The Crust, Treatise on Geochemistry, Oxford 3: 164

Shelly DR, Hill DP (2011) Migrating swarms of brittle-failure earthquakes in the lower crust beneath Mammoth Mountain, California. Geophysical Research Letters 38: doi 10.1029/2011GL049336

Snyder D (2000) Thermal effects of the intrusion of basaltic magma into a more silicic magma chamber and implications for eruption triggering. Earth and Planetary Science Letters 175: 257-273

Sparks RJS, Sigurdsson H, Wilson L (1977) Magma mixing: a mechanism for triggering acid explosive eruptions. Nature 267: 315-318

Spera FJ (1984) Carbon-dioxide in petrogenesis 3. Role of volatiles in the ascent of alkaline magma with special reference to xenolith-bearing mafic lavas. Contributions to Mineralogy and Petrology 88: 217-232

Stolper E (1980) A phase diagram for mid-ocean ridge basalts: preliminary results and implications for petrogenesis. Contributions to Mineralogy and Petrology 74: 13-27

Stolper E, Walker D (1980) Melt density and the average composition of basalt. Contributions to Mineralogy and Petrology 74: 7-12

Tait S, Jaupert C, Vergniolle S. (1989) Pressure, gas content and eruption periodicity of a shallow, crystallizing magma chamber. Earth and Planetary Science Letters 92: 107-123

Takahashi E (1986) Melting of dry peridotite KLB-1 up to 14 GPa: implications on the origin of peridotitic upper mantle. Journal of Geophysical Research 91: 9367-9382

Takahashi E, Kushiro I (1983) Melting of dry peridotite at high pressures and basalt magma genesis. American Mineralogist 68: 859-879

Wallace P (2005) Volatiles in subduction zone magmas: concentrations and fluxes based on melt inclusion and volcanic gas data. Journal of Volcanology and Geothermal Research 140: 217-240

Walter MJ (1998) Melting of garnet peridotite and the origin of komatiite and depleted lithosphere of komatiite and depleted lithosphere.