magma storage and evolution of the most recent effusive ... · 1 3 contrib mineral petrol (2016)...

19
1 3 Contrib Mineral Petrol (2016) 171:30 DOI 10.1007/s00410-016-1244-x ORIGINAL PAPER Magma storage and evolution of the most recent effusive and explosive eruptions from Yellowstone Caldera Kenneth S. Befus 1 · James E. Gardner 2 Received: 3 September 2015 / Accepted: 2 February 2016 © Springer-Verlag Berlin Heidelberg 2016 to eruption. In the Tuffs of Bluff Point and Cold Moun- tain Creek, matrix glass is less evolved than most inclu- sions, which may indicate that more primitive rhyolite was injected into the reservoir just before those eruptions. The presence and dissolution of granophyre in one flow may record evidence for heating prior to eruption and also demonstrate that the Yellowstone magmatic system may undergo rapid changes. The variations in depth suggest the magmas were sourced from multiple chambers that follow similar evolutionary paths in the upper crust. Keywords Rhyolite · Obsidian · Yellowstone · Experimental petrology · Granophyre Introduction Eruptions of rhyolitic magma are capable of producing pyroclastic deposits and lava flows spanning a wide range of volumes, from <0.01 to >1000 km 3 (e.g., Walker et al. 1973; Pyle 1989). Yellowstone Caldera has been one of the most productive rhyolitic systems over the past 2 million years (Hildreth et al. 1991; Gansecki et al. 1996; Chris- tiansen 2001; Lanphere et al. 2002; Ellis et al. 2012). Mag- matism likely continues today, as evidenced by a vigorous geothermal system, dynamic surface deformation, elevated heat flow, high seismic activity, and large low velocity zones in the upper crust (Pelton and Smith 1979; Wicks et al. 2006; Chang et al. 2007; Christiansen et al. 2007; Far- rell et al. 2009, 2014; Smith et al. 2009). Rhyolitic mag- matism at Yellowstone is most well known for producing caldera-forming “supervolcano” eruptions of the Huckle- berry Ridge, Mesa Falls, and Lava Creek tuffs (Christian- sen 2001; Lowenstern et al. 2006; Lowenstern and Hurwitz 2008). The most recent volcanism at Yellowstone produced Abstract Between 70 and 175 ka, over 350 km 3 of high- silica rhyolite magma erupted both effusively and explo- sively from within the Yellowstone Caldera. Phenocrysts in all studied lavas and tuffs are remarkably homogenous at the crystal, eruption, and caldera-scale, and yield QUILF temperatures of 750 ± 25 °C. Phase equilibrium experi- ments replicate the observed phenocryst assemblage at those temperatures and suggest that the magmas were all stored in the upper crust. Quartz-hosted glass inclu- sions contain 1.0–2.5 % H 2 O and 50–600 ppm CO 2 , but some units are relatively rich in CO 2 (300–600 ppm) and some are CO 2 -poor (50–200 ppm). The CO 2 -rich mag- mas were stored at 90–150 MPa and contained a fluid that was 60–75 mol% CO 2 . CO 2 -poor magmas were stored at 50–70 MPa, with a more H 2 O-rich fluid (X CO 2 = 40–60 %). Storage pressures and volatiles do not correlate with erup- tion age, volume, or style. Trace-element contents in glass inclusions and host matrix glass preserve a systematic evolution produced by crystal fractionation, estimated to range from 36 ± 12 to 52 ± 12 wt%. Because the erupted products contain <10 wt% crystals, crystal-poor melts likely separated from evolving crystal-rich mushes prior Communicated by Gordon Moore. Electronic supplementary material The online version of this article (doi:10.1007/s00410-016-1244-x) contains supplementary material, which is available to authorized users. * Kenneth S. Befus [email protected] 1 Department of Geology, Baylor University, Waco, TX 76706, USA 2 Jackson School of Geosciences, The University of Texas at Austin, Austin, TX 78712, USA

Upload: others

Post on 30-Jul-2020

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Magma storage and evolution of the most recent effusive ... · 1 3 Contrib Mineral Petrol (2016) 171:30 DOI 10.1007/s00410-016-1244-x ORIGINAL PAPER Magma storage and evolution of

1 3

Contrib Mineral Petrol (2016) 171:30 DOI 10.1007/s00410-016-1244-x

ORIGINAL PAPER

Magma storage and evolution of the most recent effusive and explosive eruptions from Yellowstone Caldera

Kenneth S. Befus1 · James E. Gardner2

Received: 3 September 2015 / Accepted: 2 February 2016 © Springer-Verlag Berlin Heidelberg 2016

to eruption. In the Tuffs of Bluff Point and Cold Moun-tain Creek, matrix glass is less evolved than most inclu-sions, which may indicate that more primitive rhyolite was injected into the reservoir just before those eruptions. The presence and dissolution of granophyre in one flow may record evidence for heating prior to eruption and also demonstrate that the Yellowstone magmatic system may undergo rapid changes. The variations in depth suggest the magmas were sourced from multiple chambers that follow similar evolutionary paths in the upper crust.

Keywords Rhyolite · Obsidian · Yellowstone · Experimental petrology · Granophyre

Introduction

Eruptions of rhyolitic magma are capable of producing pyroclastic deposits and lava flows spanning a wide range of volumes, from <0.01 to >1000 km3 (e.g., Walker et al. 1973; Pyle 1989). Yellowstone Caldera has been one of the most productive rhyolitic systems over the past 2 million years (Hildreth et al. 1991; Gansecki et al. 1996; Chris-tiansen 2001; Lanphere et al. 2002; Ellis et al. 2012). Mag-matism likely continues today, as evidenced by a vigorous geothermal system, dynamic surface deformation, elevated heat flow, high seismic activity, and large low velocity zones in the upper crust (Pelton and Smith 1979; Wicks et al. 2006; Chang et al. 2007; Christiansen et al. 2007; Far-rell et al. 2009, 2014; Smith et al. 2009). Rhyolitic mag-matism at Yellowstone is most well known for producing caldera-forming “supervolcano” eruptions of the Huckle-berry Ridge, Mesa Falls, and Lava Creek tuffs (Christian-sen 2001; Lowenstern et al. 2006; Lowenstern and Hurwitz 2008). The most recent volcanism at Yellowstone produced

Abstract Between 70 and 175 ka, over 350 km3 of high-silica rhyolite magma erupted both effusively and explo-sively from within the Yellowstone Caldera. Phenocrysts in all studied lavas and tuffs are remarkably homogenous at the crystal, eruption, and caldera-scale, and yield QUILF temperatures of 750 ± 25 °C. Phase equilibrium experi-ments replicate the observed phenocryst assemblage at those temperatures and suggest that the magmas were all stored in the upper crust. Quartz-hosted glass inclu-sions contain 1.0–2.5 % H2O and 50–600 ppm CO2, but some units are relatively rich in CO2 (300–600 ppm) and some are CO2-poor (50–200 ppm). The CO2-rich mag-mas were stored at 90–150 MPa and contained a fluid that was 60–75 mol% CO2. CO2-poor magmas were stored at 50–70 MPa, with a more H2O-rich fluid (XCO2

= 40–60 %). Storage pressures and volatiles do not correlate with erup-tion age, volume, or style. Trace-element contents in glass inclusions and host matrix glass preserve a systematic evolution produced by crystal fractionation, estimated to range from 36 ± 12 to 52 ± 12 wt%. Because the erupted products contain <10 wt% crystals, crystal-poor melts likely separated from evolving crystal-rich mushes prior

Communicated by Gordon Moore.

Electronic supplementary material The online version of this article (doi:10.1007/s00410-016-1244-x) contains supplementary material, which is available to authorized users.

* Kenneth S. Befus [email protected]

1 Department of Geology, Baylor University, Waco, TX 76706, USA

2 Jackson School of Geosciences, The University of Texas at Austin, Austin, TX 78712, USA

Page 2: Magma storage and evolution of the most recent effusive ... · 1 3 Contrib Mineral Petrol (2016) 171:30 DOI 10.1007/s00410-016-1244-x ORIGINAL PAPER Magma storage and evolution of

Contrib Mineral Petrol (2016) 171:30

1 3

30 Page 2 of 19

the Central Plateau Member Rhyolites, which consist of effusive outpourings of high-silica rhyolite and less com-mon explosive pyroclastic eruptions between 70 and 175 ka (Fig. 1) (Christiansen 2001; Christiansen et al. 2007; Stelten et al. 2015).

The Central Plateau Member Rhyolites provide an exceptional system in which to assess the storage, frac-tionation, and mobilization of volumetrically diverse rhy-olitic magmas. Indeed, eruptive volumes of Central Plateau Member Rhyolites range from 0.01 to 70 km3 and have a

cumulative eruption volume >350 km3 (Christiansen et al. 2007). The Central Plateau Member Rhyolites have also erupted both effusively and explosively, and so they present an opportunity to evaluate a magmatic plumbing system that fed diverse eruption styles. We specifically investigate 6 lava flows and 3 tuffs (referred to as “units” throughout) that span the full range of eruption style, age, and volume for the Central Plateau Member. To help establish a com-prehensive understanding of these eruptions, we place quantitative constraints on pre-eruptive storage conditions

10 km N111° 00’

45° 15’

Pitchstone Plateau

Solfatara Plateau

West Yellowstone

Caldera rim

Summit Lake

Trischmann Knob

Douglas Knob

Yellowstone Lake

Proposed vent site of sampled lavas

Cold Mt. Creek

Pitchstone (70 km3) Douglas Knob

Trischmann Knob (each ~0.01 km3)

Solfatara (7 km3)

Cold Mt. Creek 3)

Summit Lake (37 km3)

West Yellowstone (41 km3)

(54 km3) (50 km3)

60 100 140 180Eruption age (ka)

[ [ [[ [ [ [ [

Fig. 1 Simplified geologic map of Central Plateau Member Rhyo-lites in the Yellowstone volcanic field with target flows identified, modified after Christiansen (2001). Lava flows are shown in pink, with dark pink lines on Central Plateau Member lavas representing

pressure ridges. The extent of Yellowstone Caldera is shown by the dashed black line. Pink stars indicate vent locations. Colored spheres below map present eruption volumes and ages (Christiansen 2001; Christiansen et al. 2007)

Page 3: Magma storage and evolution of the most recent effusive ... · 1 3 Contrib Mineral Petrol (2016) 171:30 DOI 10.1007/s00410-016-1244-x ORIGINAL PAPER Magma storage and evolution of

Contrib Mineral Petrol (2016) 171:30

1 3

Page 3 of 19 30

and the magmatic evolution of the rhyolites. Specifically, we determine pre-eruptive storage conditions using petro-graphic observations, phenocryst compositions, volatile contents, and phase equilibria experiments. We also use glass inclusion compositions to estimate the extent of frac-tionation and compositional evolution of the system.

Methods

Samples of rhyolitic obsidian or pumice were collected from outcrops of Pitchstone Plateau flow, Trischmann Knob dome, Summit Lake flow, West Yellowstone flow, Solfatara Plateau flow, Buffalo Lake flow, Tuff of Cold Mountain Creek, Tuff of Bluff Point, and an unnamed pyroclastic-flow deposit remnant found on Douglas Knob lava dome (Online Resource 1). Polished thin sections from each sample were prepared. For Trischmann Knob, Sum-mit Lake, West Yellowstone, Pitchstone Plateau, and Solfa-tara Plateau flows, mineral separates of magnetite, fayalite, and clinopyroxene were handpicked from crushed obsidian samples.

Compositional analyses

Major-element compositions of minerals and glasses from natural samples and experimental products were analyzed using the JEOL JXA-8200 electron microprobe at the Uni-versity of Texas at Austin. All analyses were made using a 10 nA beam current with an 15 keV accelerating voltage. Magnetite, fayalite, and clinopyroxene were analyzed with a focused beam. The beam was defocused during sanidine and glass analyses to 2 and 10 μm diameter to minimize Na migration (Nielsen and Sigurdsson 1981). During glass analyses, Na migration was corrected using the Na-migration capability of Probe for Windows™. During all sessions, working standards were analyzed repeatedly to monitor analytical quality and instrumental drift. Crystals too small to analyze using wavelength-dispersive spectrom-etry (WDS) were identified using energy-dispersive spectra (EDS).

To supplement matrix glass and phenocryst datasets, we collected major-element, trace-element, and volatile abun-dances in quartz-hosted glass inclusions and matrix glasses from all units. About 10–15 inclusions were analyzed from each unit, for a total of 125 inclusions (Online Resource 2). Most inclusions originally contained a vapor bubble and crystallites, and so all were rehomogenized. Homogeniza-tion occurred by loading the host crystals in a Pt capsule, which was then placed inside a TZM pressure vessel and furnace at 850 °C and 150 MPa for 24 h, following the methods of Skirius et al. (1990). Most inclusions homog-enized to clear glass after a single 24-h homogenization

experiment. A second 24-h homogenization was used on the remaining inclusions that did not homogenize during the first 24-h experiment. F and Cl in glass inclusions were analyzed by electron microprobe using a 40 nA beam cur-rent with a 15 keV accelerating voltage and a 10 μm diam-eter defocused beam.

We used Fourier transform infrared (FTIR) spectros-copy to determine the dissolved H2O and CO2 contents in doubly-exposed, homogenized, quartz-hosted glass inclu-sions, and matrix glasses. FTIR spectra were collected in mid-IR and near-IR using a ThermoElectron Nicolet 6700 spectrometer and Continuμm IR microscope. Reported H2O contents in glass inclusions are the sums of dis-solved molecular and hydroxyl water determined from absorbances at ~5200 and ~4500 cm−1, respectively, using the model of Zhang et al. (1997). We measured absorb-ance at ~2350 cm−1 to determine dissolved molecular carbon dioxide (CO2) contents in glass inclusions, using the Beer–Lambert Law and an absorption coefficient of 1214 ± 16 L cm−1 mol−1 (Behrens et al. 2004). Inclusion thickness was measured optically by focusing on the top and bottom of the inclusions using a petrographic micro-scope with a Heidenhain linear encoder that measures stage height. Uncertainties in thickness ranged from 4 to 11 μm and are the most significant source of error in quantifying volatile abundances. Thus, we report error bars for H2O and CO2 using the standard deviation of inclusion thickness measurements.

Trace-element concentrations in glass inclusions and matrix glasses were measured by LA-ICP-MS at the Uni-versity of Texas at Austin, using a New Wave Research UP 193-FX fast excimer laser system with an Agilent 7500ce ICP-MS. For each target lava or pyroclastic unit, the matrix glass was analyzed 3 times, and each inclusion was ana-lyzed once. We used the laser to ablate a 30 μm diameter spot in the matrix glass or near the center of the inclu-sions, with analytical duration of 30–50 s. Trace-element concentrations were calculated assuming a concentration of 77 wt% SiO2 in the glass. NIST 612 was used at the primary calibration standard for all analytes and was run against NIST 610 as the secondary standard. Recoveries among analytes were better than 1.5 %, which are calcu-lated using the 1σ deviations relative to GeoREM preferred values. We report trace-element uncertainties with 2σ standard errors.

Experimental techniques

A portion of sample Y82 collected from Solfatara Pla-teau lava flow was crushed to a <50 μm powder and used as starting material for high-temperature phase equi-libria experiments. That sample, considered to be com-positionally representative of Central Plateau Member

Page 4: Magma storage and evolution of the most recent effusive ... · 1 3 Contrib Mineral Petrol (2016) 171:30 DOI 10.1007/s00410-016-1244-x ORIGINAL PAPER Magma storage and evolution of

Contrib Mineral Petrol (2016) 171:30

1 3

30 Page 4 of 19

Rhyolites, is composed of 5–10 % phenocrysts of quartz, sanidine, clinopyroxene, fayalite, and magnetite, all hosted in a glassy rhyolitic matrix. The whole-rock composition of sample Y82 is identical to the matrix glass composition provided in Table 2.

Phase equilibria experiments were prepared by par-tially filling 2–5 mm O.D. Ag70Pd30 capsules that had been welded on one end with starting material powder and dis-tilled water. Enough water was added to each capsule to ensure that each experiment was water-saturated at experi-mental conditions (typically 2–5 mg H2O). The capsules were then welded shut. Reversal experiments used aliquots of materials from previous experiments. Crystallization experiments used material run previously at hotter tempera-tures and/or higher water pressures, whereas melting exper-iments used material previously run at lower temperatures and/or water pressures.

Experiments were performed by loading capsules and filler rods into externally heated, cold-seal pressure vessels of Nickel-based alloy, which were then inserted into fur-naces. Temperature was gauged in the vessel using K-type thermocouples that are precise to ±5 °C. Pressure was gen-erated with a hydraulic pressure system, and measured to ±0.1 MPa. Oxygen fugacity was largely controlled during experiments by the composition of a filler rod placed inside the pressure vessel. Two sets of experiments were per-formed. The first used rods of Ni that maintained an oxy-gen fugacity at approximately one log unit above the Ni–NiO oxygen buffer (NNO+1) (e.g., Gardner et al. 1995). A second set used stainless steel rods, which significantly reduced oxygen fugacity. To ensure a controlled oxygen fugacity during those experiments, 2–3 mm O.D. experi-mental capsules were placed inside 5 mm O.D. Ag70Pd30 capsules that also contained a stoichiometric mixture of water, quartz, magnetite, and fayalite powders. When all phases are present, the oxygen fugacity is buffered at the quartz–magnetite–fayalite buffer (QFM).

Experimental runs typically lasted 7 days (Table 1). Some experiments were run for longer time (up to 14 days) and demonstrate that 7 days was long enough for phases and rhyolite melt to equilibrate. Equilibrium tests included compositional similarity in experiments run at identical conditions and observations of euhedral, distributed crys-tals. Samples were quenched in 1–2 min by removing the pressure vessels from the furnaces, and first blowing on the vessels with compressed air, and then submerging the ves-sels in water. Once cool, the experimental capsules were removed from the pressure vessels and weighed to check for leaks. We only consider experiments that did not lose water and thus remained water-saturated. The capsules were then opened, and the samples were removed. A por-tion of each sample was mounted on a glass slide and prepared for petrographic and microprobe analysis. In

QFM-buffered experiments, a portion of the buffer pow-der was examined petrographically to verify that quartz, fayalite, and magnetite were present. We report only those QFM-buffered experiments in which all three phases still occurred. Of our 62 phase equilibrium experiments, 8 were run at QFM conditions because that oxygen fugac-ity was prohibitively costly for materials and equipment. Ultimately, we conclude that cost was largely unnecessary because the differences between NNO+1 can be used to consistently and predictably understand QFM equilibria.

Results

Phenocryst and glass compositions

Glass and phenocryst major element compositions from the rhyolites are remarkably homogenous (Table 2), and our results agree with and complement previously pub-lished data (Christiansen 2001; Vazquez et al. 2009; Girard and Stix 2010). Phenocrysts account for <10 vol% of each thin section or hand sample from each unit. Of that, quartz and sanidine dominate (90–95 %) and occur in approximately equal proportions. Quartz occurs as subhe-dral crystals, ~500–2500 μm in diameter, and commonly contains glass inclusions. Sanidine occurs as euhedral to subhedral crystals and crystal fragments that reach up to 6 mm in size. They are compositionally indistinguish-able amongst the units and are unzoned from core to rim (Or50±2Ab47±2An2±1). Plagioclase was not identified in Trischmann Knob, Summit Lake, West Yellowstone, or Pitchstone Plateau, but it was found in 1–3 sanidine phe-nocrysts in a few samples (<2 %) from Solfatara Plateau; it occurs as cores of oligoclase (Or9±1Ab72±1An19±1). The compositions of those plagioclase cores are similar in com-position to plagioclase phenocrysts in the ~160 ka Central Plateau Member Rhyolites (Dry Creek, Buffalo Lake, Mal-lard Lake) (Girard and Stix 2010).

Fayalite, magnetite, and clinopyroxene are far less com-mon, and together account for <2 % of the total crystal pop-ulation. Magnetite occurs as euhedral, equant phenocrysts that are ~50–500 μm across, found as individual crys-tals or in glomerocrysts with clinopyroxene, fayalite, and other trace phases (zircon, allanite). Magnetite is unzoned and compositionally similar in all units (Mgt48±2Usp52±2), except Pitchstone Plateau, in which it is ~1 wt% richer in FeO. Ilmenite was never identified, although it has been previously reported in Central Plateau Member Rhyolites (e.g., Vazquez et al. 2009). Clinopyroxene occurs as euhe-dral crystals up to ~1000 μm long. They are composition-ally unzoned ferroaugite (En13±1Fs46±1Wo41±1), except again those in Pitchstone Plateau, which are slightly more iron-rich. Fayalite is the least common mafic phase in

Page 5: Magma storage and evolution of the most recent effusive ... · 1 3 Contrib Mineral Petrol (2016) 171:30 DOI 10.1007/s00410-016-1244-x ORIGINAL PAPER Magma storage and evolution of

Contrib Mineral Petrol (2016) 171:30

1 3

Page 5 of 19 30

Table 1 Phase equilibria experimental conditions and products

Experiment Starting materiala Duration (h) P (MPa) T (°C) % Glass fO2 Stable phasesb

1 Y82 167 25 700 0 NNO+1 gl, qtz, pyx, mgt, san (Or48±1)

2 Y82 167 25 850 100 NNO+1 gl, mgt

3 Y82 168 50 850 100 NNO+1 gl, mgt

4 Y82 168 50 775 90 NNO+1 gl, mgt, san (Or51±3)

6 Y82 168 50 700 15 NNO+1 qtz, pyx, mgt, san (Or42±1)

7 Y82 168 50 750 60 NNO+1 gl, qtz, pyx, mgt, san (Or54±5)

9 Y82 168 25 750 na NNO+1 gl, qtz, pyx, mgt, san (Or52±3)

11 Y82 170 125 700 80 NNO+1 gl, qtz, pyx, mgt, san (Or66±4)

12 Y82 170 125 725 98 NNO+1 gl, pyx, mgt, san (na)

13 7 169 100 750 100 NNO+1 gl, mgt

14 9 168 25 800 60 NNO+1 gl, pyx, mgt, san (Or40±8)

15 2 168 25 800 80 NNO+1 gl, pyx, mgt, san (Or44±4)

16 2 167 75 750 na NNO+1 gl, pyx, mgt, san (na)

17 Y82 167 75 800 100 NNO+1 gl, mgt

18 Y82 160 100 700 na NNO+1 gl, qtz, pyx, mgt, san (na)

19 17 160 75 725 na NNO+1 gl, qtz, pyx, mgt, san (Or62±6)

25 9 168 25 850 100 NNO+1 gl, mgt

29 17 168 75 750 90 NNO+1 gl, pyx, mgt, san (Or60±3)

30 Y82 336 75 675 0 NNO+1 qtz, pyx, mgt, san (na)

32 7 168 50 825 100 NNO+1 gl, mgt

33 Y82 168 200 650 na NNO+1 gl, qtz, pyx, mgt, san (na), plag (na)

36 3 168 50 775 100 NNO+1 gl, pyx, mgt, san (Or52±6)

41 4 168 50 850 100 NNO+1 gl, mgt

42 17 168 20 850 100 NNO+1 gl, mgt

46 18 168 75 800 100 NNO+1 gl, mgt

47 Y82 168 75 825 100 NNO+1 gl, mgt

50 Y82 126 50 780 na NNO+1 gl, pyx, mgt, san (na)

53 Y82 122 50 780 90 NNO+1 gl, pyx, mgt, san (na)

57 Y82 120 130 720 na NNO+1 gl, pyx, mgt, san (Or63±4)

58 Y82 120 90 780 100 NNO+1 gl, mgt

61 47 336 25 725 85 NNO+1 gl, qtz, pyx, mgt, san (Or52±3)

62 47 336 25 750 85 NNO+1 gl, qtz, pyx, mgt, san (Or52±3)

63 47 336 50 750 85 NNO+1 gl, qtz, pyx, mgt, san (Or57±3)

64 50 336 100 700 0 NNO+1 gl, qtz, pyx, mgt, san (Or50±3)

67 50 336 75 675 0 NNO+1 qtz, pyx, mgt, san (Or51±5)

68 50 336 200 650 na NNO+1 gl, qtz, pyx, mgt, san (Or47±2), plag (Ab80)

71 Y82 120 50 700 100 NNO+1 gl, qtz, pyx, mgt, san (Or44±6)

72 Y82 120 100 800 100 NNO+1 gl, mgt

75 72 167 100 750 100 NNO+1 gl, mgt

76 72 167 125 725 92 NNO+1 gl, pyx, mgt, san (Or64±4)

77 72 167 125 700 80 NNO+1 gl, pyx, mgt, san (Or63±5)

80 72 147 100 675 50 NNO+1 gl, qtz, pyx, mgt, san (Or58±5)

81 72 333 150 675 70 NNO+1 gl, qtz, pyx, mgt, san (Or66±5), plag (Ab81±2)

85 58 169 150 725 98 NNO+1 gl, pyx, mgt, san (Or63±9)

88 57 120 75 750 40 QFM gl, qtz, pyx, fay, mgt, san (Or52±2)

96 Y82 116 100 750 90 QFM gl, pyx, fay, mgt, san (Or51±1)

98 96 120 50 750 90 QFM gl, qtz, pyx, fay, mgt, san (Or50)

99 96 120 100 700 70 QFM gl, qtz, pyx, fay, mgt, san (Or58)

Page 6: Magma storage and evolution of the most recent effusive ... · 1 3 Contrib Mineral Petrol (2016) 171:30 DOI 10.1007/s00410-016-1244-x ORIGINAL PAPER Magma storage and evolution of

Contrib Mineral Petrol (2016) 171:30

1 3

30 Page 6 of 19

the units. It occurs as euhedral crystals ~200–800 μm in diameter. As with the other mineral phases, it is also com-positionally unzoned, but those in Pitchstone Plateau are slightly more iron-rich (Fa95±1) than those in all other flows (Fa92±1).

Granophyre, composed of intergrown, vermicular to micrographic quartz and sanidine (Or46±6Ab51±6An2±1), was identified in some samples from Solfatara Plateau (Fig. 2). It occurs as subhedral to anhedral overgrowths up to ~1000 μm thick on sanidine and quartz phenocrysts, and as individual rounded globules that reach ~1000 μm in diameter. In some instances, granophyre occurs as over-growths that connect sanidine and quartz phenocrysts to form glomerocrysts. Discontinuous bands of abundant microlite-sized sanidine and quartz fragments occur in samples with granophyre and can sometimes be traced to the rounded granophyre clasts. Such bands are absent in samples that lack granophyre. To understand the distribu-tion of granophyre, we examined thin sections from >60 samples collected across Solfatara Plateau. We found that granophyre was generally restricted to areas within 2 km of the flow front (Fig. 3). We searched for granophyre in samples collected across many of the other units and only found it in one sample from the periphery of Pitchstone Plateau. It has also been identified in samples from Hayden Valley flow (Stelten, personal communication).

Water contents in glass inclusions from all units are similar and overall range from 1.0 to 2.5 wt% H2O (Fig. 4) (Online Resource 2). Dissolved CO2 contents vary from 50 to 600 ppm. Individual units have more restricted vol-atile contents, and define regions in H2O and CO2 space

with internal variability of approximately ±0.5 wt% and ±100 ppm, respectively. Volatile concentrations in both lavas and pyroclastic units have similar, relatively low water contents, but CO2 contents vary and can be used to separate the lavas and tuffs into CO2-rich and CO2-poor populations. CO2-rich lavas and tuffs contain 300–600 ppm CO2 (Tuff of Bluff Point, Pitchstone Plateau, Trischmann Knob, Buffalo Lake, Summit Lake). CO2-poor units con-tain 50–200 ppm CO2 (Solfatara Plateau, Tuff of Cold Mountain Creek, unnamed remnant of pyroclastic deposit on Douglas Knob, West Yellowstone). Concentrations of F and Cl in glass inclusions are indistinguishable amongst the units, and average 1900 ± 100 and 1100 ± 100 ppm, respectively.

The majority of the volatile contents are derived from quartz-hosted glass inclusions extracted from slowly cooled lavas. Such data should be treated initially with caution as diffusion or cracking may cause the inclusions to not be conservative. We argue, however, that most of our inclu-sions preserve magmatic values. First, volatile contents of individual units cluster in tight domains that display limited evidence for degassing and gas-saturated crystallization (Fig. 4). Second, volatile contents from the lavas are indis-tinguishable from those measured in the rapidly cooled pyroclastic units. If diffusion had significantly modified volatile contents, then years of slow cooling of the host lava at the surface should generate dry inclusions.

Trace-element concentrations in glass inclusions and rhyolitic groundmass range greatly (Online Resource 3). We can infer the behavior of each element by calculat-ing its bulk distribution coefficient (D) using published

a Starting material is the powdered material used in the experimental charge. Numbers refer to starting material using crushed material from previously run experimentsb Stable phases are glass (gl), magnetite (mgt), quartz (qtz), clinopyroxene (pyx), fayalite (fay), and sanidine (san). na stands for not analyzed

Table 1 continued

Experiment Starting materiala Duration (h) P (MPa) T (°C) % Glass fO2 Stable phasesb

100 96 120 100 800 100 QFM gl, mgt

101 96 120 50 825 100 QFM gl, mgt

102 96 120 150 725 90 QFM gl, pyx, fay, mgt, san (Or59±10)

103 96 120 50 700 10 QFM gl, qtz, pyx, fay, mgt, san (Or56±3)

109 47 168 150 700 92 NNO+1 gl, qtz, pyx, mgt, san (Or75±2)

110 47 168 250 700 100 NNO+1 gl, pyx, mgt

111 47 168 175 725 100 NNO+1 gl, mgt

112 47 335 200 675 55 NNO+1 gl, pyx, mgt, san (Or69±3), plag (Ab79±5)

113 Y82 168 200 750 100 NNO+1 gl, mgt

115 113 335 175 725 100 NNO+1 gl, mgt

116 113 168 225 700 100 NNO+1 gl, pyx, mgt, san (Or73±4)

117 113 168 225 725 100 NNO+1 gl, mgt

118 113 169 250 675 99 NNO+1 gl, pyx, mgt, san (Or72±3)

119 113 169 25 850 100 NNO+1 gl, mgt

Page 7: Magma storage and evolution of the most recent effusive ... · 1 3 Contrib Mineral Petrol (2016) 171:30 DOI 10.1007/s00410-016-1244-x ORIGINAL PAPER Magma storage and evolution of

Contrib Mineral Petrol (2016) 171:30

1 3

Page 7 of 19 30

Tabl

e 2

Com

posi

tions

of

mat

rix

glas

s an

d m

ajor

phe

nocr

yst p

hase

s

Ana

lyse

s by

ele

ctro

n m

icro

prob

e. M

ajor

oxi

des

repo

rted

in w

eigh

t per

cent

and

are

ave

rage

s of

n s

ampl

es. V

alue

s in

par

enth

eses

rep

rese

nt s

tand

ard

devi

atio

n

“–”

oxid

e w

as n

ot a

naly

zed

a Tot

al ir

on r

epor

ted

as F

eO

Phas

eU

nit

Sam

ple

#Si

O2

TiO

2A

l 2O

3Fe

Oa

MgO

MnO

CaO

K2O

Na 2

OTo

tal

n

Mat

rix

Gla

ssSo

lfat

ara

Plat

eau

Y82

76.8

5 (1

.84)

0.12

(0.

06)

12.0

6 (0

.22)

1.57

(0.

25)

0.02

(0.

02)

0.09

(0.

06)

0.47

(0.

04)

5.12

(0.

19)

4.08

(0.

15)

100.

36 (

1.54

)3

Wes

t Yel

low

ston

eY

216

77.5

0 (0

.72)

0.12

(0.

03)

11.9

8 (0

.11)

1.21

(0.

21)

0.04

(0.

01)

0.03

(0.

02)

0.43

(0.

01)

5.17

(0.

14)

3.73

(0.

51)

100.

20 (

1.21

)6

Sum

mit

Lak

eY

7476

.78

(0.3

4)0.

11 (

0.03

)11

.96

(0.1

2)1.

23 (

0.10

)0.

02 (

0.02

)0.

04 (

0.03

)0.

47 (

0.02

)5.

11 (

0.11

)4.

10 (

0.16

)99

.81

(0.3

6)6

Pitc

hsto

ne P

late

auY

114

77.5

5 (0

.89)

0.12

(0.

05)

12.0

5 (0

.17)

1.17

(0.

30)

0.01

(0.

01)

0.03

(0.

02)

0.25

(0.

03)

5.26

(0.

03)

4.10

(0.

09)

100.

52 (

0.58

)4

Tri

schm

ann

Kno

bY

9776

.95

(0.4

0)0.

13 (

0.06

)11

.85

(0.1

5)1.

31 (

0.19

)0.

01 (

0.01

)0.

02 (

0.01

)0.

48 (

0.03

)5.

20 (

0.08

)3.

60 (

0.48

)99

.55

(0.9

5)4

Solf

atar

a Pl

atea

uY

8230

.09

(0.3

3)0.

04 (

0.05

)0.

01 (

0.01

)64

.88

(0.9

1)2.

75 (

0.10

)1.

62 (

0.08

)0.

28 (

0.01

)–

–99

.69

(1.1

4)25

Wes

t Yel

low

ston

eY

216

30.4

8 (0

.24)

0.02

(0.

03)

0.01

(0.

01)

64.6

5 (0

.46)

3.06

(0.

09)

2.24

(0.

11)

0.26

(0.

02)

––

100.

70 (

0.47

)32

Faya

lite

Sum

mit

Lak

eY

7430

.31

(0.3

3)0.

02 (

0.02

)0.

01 (

0.01

)64

.34

(0.5

8)3.

11 (

0.08

)2.

23 (

0.10

)0.

27 (

0.03

)–

–10

0.57

(0.

68)

31

Pitc

hsto

ne P

late

auY

114

30.0

9 (0

.30)

0.02

(0.

03)

0.01

(0.

01)

65.8

4 (0

.49)

2.03

(0.

07)

2.33

(0.

03)

0.28

(0.

01)

––

100.

80 (

0.57

)38

Tri

schm

ann

Kno

bY

9730

.38

(0.1

5)0.

02 (

0.03

)0.

01 (

0.01

)64

.90

(0.7

2)2.

94 (

0.08

)2.

25 (

0.06

)0.

27 (

0.01

)–

–10

0.74

(0.

74)

7

Solf

atar

a Pl

atea

uY

8249

.45

(0.7

6)0.

21 (

0.08

)0.

45 (

0.15

)26

.70

(0.6

1)4.

12 (

0.39

)0.

73 (

0.05

)18

.68

(0.2

4)–

0.31

(0.

03)

100.

65 (

0.89

)55

Wes

t Yel

low

ston

eY

216

49.2

6 (0

.61)

0.21

(0.

05)

0.49

(0.

03)

26.0

0 (0

.70)

4.33

(0.

41)

1.07

(0.

07)

18.8

2 (0

.27)

–0.

32 (

0.04

)10

0.49

(0.

71)

24

Clin

opyr

oxen

eSu

mm

it L

ake

Y74

49.2

5 (0

.59)

0.23

(0.

05)

0.52

(0.

16)

26.2

1 (0

.63)

4.21

(0.

39)

1.07

(0.

07)

18.7

9 (0

.26)

–0.

32 (

0.04

)10

0.60

(0.

58)

33

Pitc

hsto

ne P

late

auY

114

48.9

7 (0

.53)

0.22

(0.

04)

0.54

(0.

27)

27.4

8 (0

.45)

3.03

(0.

17)

1.07

(0.

08)

18.9

7 (0

.30)

–0.

33 (

0.04

)10

0.60

(0.

57)

39

Tri

schm

ann

Kno

bY

9749

.41

(0.4

2)0.

22 (

0.04

)0.

48 (

0.14

)26

.09

(0.5

7)4.

28 (

0.44

)1.

03 (

0.05

)18

.65

(0.1

7)–

0.31

(0.

05)

100.

49 (

0.54

)35

Solf

atar

a Pl

atea

uY

820.

09 (

0.03

)18

.53

(0.6

5)0.

85 (

0.08

)76

.39

(1.1

2)0.

18 (

0.03

)0.

87 (

0.04

)0.

01 (

0.01

)–

–96

.86

(1.3

7)78

Wes

t Yel

low

ston

eY

216

0.09

(0.

03)

18.4

5 (0

.56)

0.78

(0.

09)

77.8

6 (1

.41)

0.20

(0.

03)

0.87

(0.

06)

0.01

(0.

01)

––

98.2

5 (1

.45)

59

Mag

netit

eSu

mm

it L

ake

Y74

0.10

(0.

02)

18.5

3 (0

.43)

0.82

(0.

05)

78.0

7 (1

.05)

0.22

(0.

03)

0.90

(0.

06)

0.01

(0.

01)

––

98.6

2 (1

.11)

69

Pitc

hsto

ne P

late

auY

114

0.10

(0.

02)

17.6

7 (0

.33)

0.75

(0.

07)

79.3

3 (0

.70)

0.12

(0.

04)

0.86

(0.

06)

0.01

(0.

01)

––

98.8

5 (0

.84)

85

Tri

schm

ann

Kno

bY

970.

10 (

0.03

)18

.92

(0.4

2)0.

86 (

0.05

)76

.84

(1.0

2)0.

21 (

0.04

)0.

90 (

0.05

)0.

01 (

0.01

)–

–97

.85

(1.1

4)35

Solf

atar

a Pl

atea

uY

8266

.42

(0.5

9)–

18.9

4 (0

.22)

0.09

(0.

07)

––

0.39

(0.

18)

8.28

(0.

61)

5.57

(0.

49)

100.

07 (

0.92

)13

6

Wes

t Yel

low

ston

eY

216

66.6

0 (0

.36)

–18

.97

(0.1

9)0.

13 (

0.03

)–

–0.

51 (

0.04

)8.

58 (

0.17

)5.

09 (

0.13

)99

.87

(0.5

9)17

Sani

dine

Sum

mit

Lak

eY

7466

.14

(0.6

7)–

18.7

6 (0

.26)

0.14

(0.

03)

––

0.52

(0.

05)

8.49

(0.

12)

5.14

(0.

09)

99.1

8 (0

.97)

28

Pitc

hsto

ne P

late

auY

114

66.1

8 (1

.25)

–18

.81

(0.3

6)0.

15 (

0.03

)–

–0.

46 (

0.04

)8.

63 (

0.11

)4.

97 (

0.07

)99

.19

(1.7

2)10

Tri

schm

ann

Kno

bY

9765

.65

(0.9

4)–

18.7

4 (0

.17)

0.15

(0.

02)

––

0.56

(0.

05)

8.26

(0.

10)

5.13

(0.

10)

98.5

0 (1

.15)

5

Page 8: Magma storage and evolution of the most recent effusive ... · 1 3 Contrib Mineral Petrol (2016) 171:30 DOI 10.1007/s00410-016-1244-x ORIGINAL PAPER Magma storage and evolution of

Contrib Mineral Petrol (2016) 171:30

1 3

30 Page 8 of 19

partition coefficients and estimates of the modal abun-dances of major and minor mineral phases present (Online Resource 4). Compatible trace elements include Sc, Mn, Zn, Sr, Ba, and Eu. The remainder are incompatible, with most of those being very incompatible (D < 0.2) (Online Resource 4). Nb is thought to be one of the most incom-patible elements. When the compatible elements are plot-ted against Nb, negative correlations are observed (Fig. 5). Conversely, strong positive correlations are observed when incompatible elements are plotted against Nb (Fig. 5). Pop-ulations of glass inclusions from individual units contain incompatible trace-element concentrations that range from relatively primitive to evolved (Online Resource 3). The most primitive inclusions from each rhyolite as a group are compositionally similar, with the least evolved concentra-tions of those elements differing by <20 %. Incompatible trace-element concentrations in matrix glasses are typi-cally more enriched than those in evolved inclusions from the same unit (Fig. 5). When plotted against one another, incompatible elements within each unit increase systemati-cally from primitive inclusions, to more evolved inclusions, and finally to highly evolved matrix (Fig. 5). Inclusions

from the Tuffs of Bluff Point and Cold Mountain Creek display similar progressive evolution, except that matrix glass in those tuffs is less evolved than some or most of the inclusions.

Phase equilibrium experiments

Bubbles were present in all experiments, demonstrating that all were water-saturated. Glass occurs in all samples above 725 °C. Mineral stability fields are based on crystal mor-phology (e.g., anhedral, euhedral) and crystal content rela-tive to the starting material. In both NNO+1- and QFM-buffered experiments, magnetite is the liquidus phase, and is stable under all water pressure and temperature (P–T) conditions investigated (Fig. 6, 7). At NNO+1 conditions above 50 MPa, the crystallization sequence with decreas-ing temperature is sanidine, clinopyroxene, quartz, and pla-gioclase, with the solidus at <700 °C. At pressures below 50 MPa, the crystallization sequence with decreasing tem-perature is similarly sanidine, clinopyroxene, and quartz. At those low pressures, plagioclase does not form before the solidus is reached at 700–725 °C. Overall, phase stabilities

1 mm

SanidineQtz

Qtz

San

Plagioclasecore

GranophyreQtz

1 mm

Qtz

Qtz

QtzSan

Granophyre

(a) (c)

(d)(b)

Opaqueglass

Qtz

void

Glass matrix

San

500 μm

200 μm

Fig. 2 a, b Photomicrographs and c, d backscattered images of quartz, sanidine, and granophyre in Solfatara Plateau. Granophyre occurs in association with glomerocrysts of sanidine and quartz, as overgrowths on sanidine and quartz, or as individual clasts

Page 9: Magma storage and evolution of the most recent effusive ... · 1 3 Contrib Mineral Petrol (2016) 171:30 DOI 10.1007/s00410-016-1244-x ORIGINAL PAPER Magma storage and evolution of

Contrib Mineral Petrol (2016) 171:30

1 3

Page 9 of 19 30

between the different fO2 conditions are similar, although some differences exist (Fig. 7). The reduced experiments contain much less magnetite at equivalent P–T conditions, and they also contain fayalite, which was not observed in oxidized experiments. In QFM experiments, the crystalli-zation sequence with decreasing temperature is sanidine, fayalite, clinopyroxene, and quartz. Sanidine, fayalite, and clinopyroxene stabilize over a very narrow temperature

window. The solidus was not determined. Sanidine, clino-pyroxene, and quartz are stabilized at temperatures 5°–20° hotter than in NNO+1 experiments at equivalent pressures. Plagioclase was not stabilized in QFM experiments.

In NNO+1 experiments, glass abundance gradu-ally decreases from ~100 % above the sanidine in-curve, to ~80 % glass only 20 °C above the solidus (Fig. 6; Table 1). Although fewer experiments limit our resolu-tion, we observe similar changes in crystal abundance and % glass in QFM experiments. In the final 20 °C above the solidus, the melt crystallized significantly, forming fine-grained intergrowths of sanidine and quartz crystals that are 1–5 μm in size. Glass composition changes systematically with temperature (Online Resource 5). Concentration of SiO2 increases with the decrease in temperature, whereas those of K2O and FeO decrease. Sanidine ranges in com-position from Or66±5Ab34±4An1±1 to Or44±4Ab52±3An3±1 and is more potassic at higher PH2O (Fig. 6). Clinopyrox-ene crystals are often only 1 μm in diameter and thus dif-ficult to analyze quantitatively using WDS. When WDS analyses were possible, compositions in NNO+1-buffered

4 km

Grand Loop Road

Norris Canyon RoadLOS in b

(a)

(b)

Granophyre-bearing

Granophyre-free

v.e.~10:1

1 km

Fig. 3 a Simplified geologic map of Solfatara Plateau, modified after Christiansen (2001). Sample locations are shown by circles. White circles indicate sample contained granophyre, whereas black cir-cles represent samples without granophyre. The location of sample Y82 (experimental starting material) is marked by a yellow star. The dashed black line separates Solfatara Plateau into granophyre-bearing and granophyre-free domains. The distribution of granophyre largely corresponds to a change in topography across the flow as highlighted using the DEM overlay from open source LIDAR data from NSF OpenTopography. b Aerial view of topographic change at Solfatara Plateau with the view looking to the southwest as shown by the line of sight (LOS) arrow shown in (a). As before, the dashed line sepa-rates granophyre-bearing and granophyre-free domains

0

200

400

600

1 2 3 4

20/8

0

40/6

0

50/5

0

60/40

CO2 (

ppm

)

H2O (wt.%)

200 MPa

150 MPa

100 MPa

50 MPa

25 MPa 80/20

Pyro. remnant on DKPitchstone PlateauTrischmann Knob

West YellowstoneSummit LakeDouglas Knob Solfatara Plateau

Cold Mt. Creek

Lava CreekMesa Falls

Fig. 4 Dissolved volatile contents in quartz-hosted glass inclusions. Squares are effusive eruptions and circles are for explosive eruptions. Colored fields highlight the general distribution for each unit. Error bars (1σ) are shown in gray when larger than the symbol size. Equi-librium solubility at 25, 50, 100, 150, and 200 MPa and 750 °C are shown in black. Dashed gray lines are fluid compositions in percent H2O and CO2 for rhyolite in equilibrium with a mixed fluid (Liu et al. 2005). Data for Tuff of Bluff Point, Douglas Knob dome (DK), Lava Creek Tuff, and Mesa Falls Tuff are from Befus et al. (2012, 2014) and Gansecki (1998), respectively

Page 10: Magma storage and evolution of the most recent effusive ... · 1 3 Contrib Mineral Petrol (2016) 171:30 DOI 10.1007/s00410-016-1244-x ORIGINAL PAPER Magma storage and evolution of

Contrib Mineral Petrol (2016) 171:30

1 3

30 Page 10 of 19

experiments are En14±8Fs40±7Wo46±4 and are slightly more Fe-rich in lower temperature experiments. Clinopy-roxene compositions in QFM-buffered experiments are En7±3Fs52±4Wo41±4. EDS analyses confirm the presence of Fe-rich clinopyroxene in samples in which quantitative WDS analyses were not possible. When present, plagio-clase (Or12±3Ab81±2An7±2) and fayalite (Fa96±2) composi-tions do not measurably vary with pressure or temperature.

Discussion

The Central Plateau Member Rhyolites we studied are nearly identical petrographically, containing a phenocryst assemblage of quartz, sanidine, magnetite, clinopyroxene, and fayalite, all set in a high-silica rhyolitic glass, despite differing by 3–4 orders of magnitude in volume (Fig. 1). Compositionally, matrix glass and phenocryst phases are internally homogenous with respect to major elements, and are nearly indistinguishable between units. The only recognizable differences in major elements are in mag-netite, clinopyroxene, and fayalite in Pitchstone Plateau, which are slightly more Fe-rich when compared with those phases in other units, consistent with compositional analyses performed by Christiansen (2001), Vazquez et al.

(2009), and Girard and Stix (2010). To constrain condi-tions of phenocryst growth, we apply the QUILF geother-mometer using compositions of coexisting natural assem-blages of quartz, fayalite, clinopyroxene, and magnetite (Andersen et al. 1993). Average phenocryst compositions in Trischmann Knob, West Yellowstone, Summit Lake, and Solfatara Plateau each yield temperatures of 755 ± 20 °C, whereas those in Pitchstone Plateau provide an estimate of 745 ± 15 °C. All flows return oxygen-fugacity values rang-ing 0.4–0.5 log units below the QFM buffer.

The phenocryst assemblage is generally reproduced by both NNO+1- and QFM-buffered phase equilibria experi-ments (Figs. 6, 7). Experiments run at NNO+1 gener-ate sanidine, quartz, magnetite, and clinopyroxene, but were too oxidized to stabilize fayalite. Fayalite joined the equilibrium assemblage in QFM-buffered experiments. The only other effect of oxygen fugacity was that QFM experiments were found to stabilize phases at tempera-tures 5°–20° hotter than in NNO+1 experiments. Because phase equilibria are more tightly constrained in P–T space at NNO+1, we use it to evaluate the natural phase assem-blage, but use insights from QFM experiments when pos-sible. The natural phase assemblage is only reproduced experimentally at conditions below the quartz-in curve at 715 ± 10 °C at 150 MPa and at 800 ± 15 °C at 25 MPa, but

Elem

ent (

ppm

)

Nb (ppm)

5

10

15

20

Elem

ent (

ppm

)

Pr

Sm

Zr

Y

70

Nb (ppm)70

Nb (ppm)70

Nb (ppm)40 50 60 50 60 50 60 50 60 70

100

200

300

Hf

4

8

12

Elem

ent (

ppm

) Sr

Sc

Fig. 5 Correlations between compatible (Sr, Sc) and incompatible elements (Zr, Y, Pr, Sm, Hf, Nb) in representative Central Plateau Member Rhyolites. Inclusion compositions are shown as squares, whereas matrix glasses are shown as circles. Error bars (1σ) are shown in gray when larger than the symbol size. Black lines are Ray-

leigh fractionation models; tick marks represent 10 % intervals of fractional crystallization. Fractionation models are not shown for the Tuff of Cold Mountain Creek because the groundmass is less evolved than the inclusions

Page 11: Magma storage and evolution of the most recent effusive ... · 1 3 Contrib Mineral Petrol (2016) 171:30 DOI 10.1007/s00410-016-1244-x ORIGINAL PAPER Magma storage and evolution of

Contrib Mineral Petrol (2016) 171:30

1 3

Page 11 of 19 30

above the solidus or plagioclase-in curve. Such constraints generate a ~75 °C stability window that narrows with increasing water pressure. That window shifts to ~730–815 °C, given that fO2 is similar to QFM. Because experi-mental sanidine is more potassic than natural sanidine (Or50±2Ab47±2An2±1) at higher water pressures, the rhyo-lites were probably stored at water pressures <120 MPa.

We can further infer pressures and the composition of magmatic fluids using the volatile contents in homog-enized, quartz-hosted glass inclusions (Liu et al. 2005). We assume the magmas were volatile-saturated at depth because H2O and CO2 do not correlate positively with incompatible elements. If the magmas were undersaturated, then H2O and CO2 would have increased in concentration during crystal fractionation because they were not parti-tioned into any phase. That behavior is not observed. Such

an inference is also supported by textural observations of glass re-entrants in quartz phenocrysts by Girard and Stix (2010). The CO2-rich magmas were in equilibrium with fluids composed of 60–75 mol% CO2, and were stored at 90–150 MPa. The CO2-poor magmas contained fluids com-posed of 40–60 mol% CO2 and were stored at 50–70 MPa.

Pressure estimates from inclusion values are not directly comparable to pressures from phase equilibria experi-ments, because experiments were run at water-saturated conditions (Ptotal = PH2O). The experimental condition that Ptotal = PH2O should have very little effect on phase equilibria temperatures because CO2 is a trace element (0.005–0.06 wt%) that would have little effect on mineral stabilities (Gardner et al. 2014). We emphasize that the connection between our experiments and the inclusions is water content. By extension, the connection to other mixed

Fig. 6 NNO+1 phase equilibria diagram assuming Ptotal = PH2O

. Squares are initial experiments using powder from sample Y82, whereas right and left pointing triangles represent melting and crystallization experiments, respectively, using powders from initial experi-ments. Lines represent mineral stability curves. The dashed gray lines mark contours for sanidine composition. The gray dotted lines are H2O solubility isopleths from Liu et al. (2005). Shaded regions show amount of melt present. Most crystalliza-tion occurs at temperatures just above the solidus

25

50

75

100

125

P Tota

l = P

H2O

(MPa

)

0700 750 800 850

Temperature (°C)

Sanidine

Clinopyroxene

Quartz

Solidus

150

Or60Ab40

Plagioclase

melt+vapor+magnetite

175

200

225

250

Or70Ab30

Percent Melt

>90 %

80-89 %

1-79 %6 wt.% H2

O

4 wt.% H2O

2 wt.% H2O

Page 12: Magma storage and evolution of the most recent effusive ... · 1 3 Contrib Mineral Petrol (2016) 171:30 DOI 10.1007/s00410-016-1244-x ORIGINAL PAPER Magma storage and evolution of

Contrib Mineral Petrol (2016) 171:30

1 3

30 Page 12 of 19

fluid, H2O-undersaturated, and H2O-saturated experimen-tal work on high silica rhyolites is also by water content. Indeed, our results coincide with Almeev et al. (2012) and Bolte et al. (2015), which demonstrate that low H2O con-tents experimentally replicate natural Yellowstone hot spot mineral assemblages. Glass inclusions indicate the Central Plateau Member magmas contained 1.0–2.5 wt% H2O. That range is equivalent to experiments run at water pres-sures up to 40 MPa. At those water pressures, the natural assemblage is consistent with our experiments at tempera-tures ranging from 700 to 760 °C.

Together, phase equilibria, glass inclusions, and geo-thermometry indicate the magmas existed in the upper crust at relatively cool conditions. We assume phenocryst growth occurred during magma storage. Both the CO2-rich and CO2-poor magmas were stored at 750 ± 25 °C. Phase equilibria experiments demonstrate that it is not possible to crystallize the natural assemblage of quartz and sanidine from the hydrous melt at temperatures greater than 800 °C. Our estimates for magmatic temperatures directly over-lap estimates using Ti-in-zircon by Stelten et al. (2015), which range from approximately 700 to 800 °C. Both our new estimates and those from Stelten et al. (2015) are 50–150 °C cooler than most previous temperature estimates for Yellowstone and Snake River Plain rhyolite volcanism, which range from 800 °C to nearly 1000 °C (Honjo et al. 1992; Bindeman and Valley 2000; Branney et al. 2008; Cathey and Nash 2009; Vazquez et al. 2009;

Ellis et al. 2010; Wolff et al. 2011; Pritchard and Larson 2012; Watts et al. 2012; Almeev et al. 2012; Bolte et al. 2015; Till et al. 2015). In part, the wide distribution in temperature can be attributed to the well-established trend that younger and more eastern magmas were stored at cooler temperatures.

Some of those hotter temperature estimates for Cen-tral Plateau Member Rhyolites can be reconciled with our results, whereas others cannot. For example, we are unable to reconcile crystallization temperatures for clino-pyroxene of ~900 °C based on oxygen isotope fractiona-tion (Loewen and Bindeman 2015). Our phase equilibria demonstrate clinopyroxene crystallizes at temperatures <825 °C. We also suggest that geothermometry results from Zr concentrations in glass and whole-rock, which have ranged from 820 to 900 °C, produce inaccurate tem-peratures (e.g., Bindeman and Valley 2001; Watts et al. 2012). Most other temperature estimates can be readily reconciled. Bindeman and Valley (2001) and Watts et al. (2012) estimated storage temperatures using the liquidus temperature calculated from MELTS. When compared to our phase equilibria results, those liquidus temperatures of 850 ± 25 °C are merely a few 10’s of degrees hotter than our experimental sanidine- and quartz-in-curves at low pressures, which must be colder than the experimental liquidus established by magnetite (Figs. 6, 7). Addition-ally, Ti-in-quartz estimates from Vazquez et al. (2009) and Girard and Stix (2010) were performed using the initial

Fig. 7 QFM phase equilibria diagram assuming Ptotal = PH2O

. Squares and triangles are the same as in Fig. 6. NNO+1 phase equilibria are reproduced in gray for comparison. The order of the mineral stability curves for sanidine, clinopy-roxene, and fayalite cannot be distinguished. Refer to Fig. 6 for the position of H2O isopleths. The best estimate for conditions of phenocryst growth is highlighted in yellow

700 750 800 850

25

50

75

100

125

Temperature (°C)

0

150

melt+vapor+magnetite

Sanidine

Clinopyroxene

Quartz

Solidus

Plagioclase

Sanidine, Clinopyroxene, Fayalite

Quartz

P Tota

l = P

H2O

(MPa

)

Page 13: Magma storage and evolution of the most recent effusive ... · 1 3 Contrib Mineral Petrol (2016) 171:30 DOI 10.1007/s00410-016-1244-x ORIGINAL PAPER Magma storage and evolution of

Contrib Mineral Petrol (2016) 171:30

1 3

Page 13 of 19 30

calibration of Wark and Watson (2006). More recent cali-brations of that model indicate that it is very sensitive to storage pressures (e.g., ~1 kb variations change tempera-ture by ~60 °C) (Thomas et al. 2010; Huang and Audetat 2012; Girard and Stix 2012; Thomas et al. 2015). When reapplied to the new calibrations, Ti concentrations in quartz generate temperatures ranging from 650 to 800 °C at pressures consistent with both phase equilibria and vola-tile contents in glass inclusions.

The CO2-poor magmas were stored at pressures of 50–70 MPa, whereas the CO2-rich magmas were stored at 90–150 MPa. Using a regional Yellowstone crustal den-sity of 2500–2670 kg m−3 (DeNosaquo et al. 2009), the storage pressure estimates from CO2-poor and CO2-rich melts correspond to storage depths of 2–3 and 3–6 km, respectively. Interestingly, storage depth does not corre-late with eruption age, volume, style, or geographic posi-tion in the caldera. Magma from Pitchstone Plateau, the most recent eruption at Yellowstone, was stored at ~5 km. That depth is consistent with estimates for modern low-velocity zones across Yellowstone Caldera from high-res-olution seismic tomography (Smith et al. 2009; Chu et al. 2010; Farrell et al. 2014). Similar upper crustal depths are also consistent with isotopic constraints on source rocks, which indicate Central Plateau Member magmas were generated by melting of hydrothermally altered material. Some of that material may have been intracal-dera Lava Creek and Huckleberry Ridge Tuffs, which would only exist in the shallow crust (Bindeman and Val-ley 2000, 2001; Bindeman et al. 2001; Girard and Stix 2010; Pritchard and Larson 2012; Watts et al. 2012). Pb isotopes demonstrate the Central Plateau Member Rhy-olites could not solely have been derived from melting those intracaldera tuffs and instead require a hybridized source modified with other silicic sources (Vazquez et al. 2009; Stelten et al. 2015).

Magmatic evolution

The Central Plateau Member Rhyolites we studied are monotonous with respect to phenocryst content and major-element composition, indicating similar magmatic condi-tions over episode of explosive and effusive volcanism. Our trace-element data for glass inclusions and matrix glass help define the temporal evolution of the magmatic sys-tem. Trace-element contents in quartz-hosted glass inclu-sions preserve snapshots of melt composition sequentially trapped in growing quartz crystals. Hence, inclusions and matrix glass preserve the process of crystal-liquid frac-tionation. We find that Rb/Sr ratios in groundmass glasses range from 20 to 80 and progressively increase with time (Fig. 8a), similar to Vazquez and Reid (2002) and Girard and Stix (2010). Ratios of other incompatible to compatible trace elements (e.g., Gd/Ba) also increase as eruption age decreases (Fig. 8b). Such patterns support the conclusion that crystal fractionation is important in the evolving sys-tem (Vazquez and Reid 2002; Girard and Stix 2010; Watts et al. 2012). Local crystallization of the quartz host dur-ing entrapment did not produce the range in trace-element concentrations. If the trace elements were concentrated by growth of the quartz host, then almost all elements should behave incompatibly, which is not observed (Fig. 5). Simi-larly, boundary layer crystallization at the chamber scale is also not impactful. The trace-element fractionation trends closely follow those predicted by the modal assemblage; thus, crystallization by a non-modal assemblage is not sup-ported by the data.

Focusing on the trace-element compositions of inclu-sions, we find that those ratios (Rb/Sr and Gd/Ba) vary between each magma that feeds the individual flows. The most primitive inclusions for each unit have similar low ratios. The incompatible to compatible element ratios increase towards the evolved ratio in the groundmass. The

Rb/S

r

0

30

60

90

Eruption Age (ka)180

Gd/

Ba

0

0.1

0.2

0.3

Eruption Age (ka)60 100 140 100 140180 60

Pyro. remnant on DKPitchstone PlateauTrischmann Knob West YellowstoneSummit LakeSolfatara PlateauCold Mountain Creek

(a) (b)

Fig. 8 Ratios of incompatible to compatible trace elements are shown as a measure of magmatic evolution and crystal fractiona-tion. Ratios are plotted against eruption ages from Christiansen et al. (2007). West Yellowstone, Douglas Knob, and Trischmann Knob

erupted at ~115 kya and are separated slightly on this graph for leg-ibility. Squares and circles are for groundmass glass and inclusions, respectively

Page 14: Magma storage and evolution of the most recent effusive ... · 1 3 Contrib Mineral Petrol (2016) 171:30 DOI 10.1007/s00410-016-1244-x ORIGINAL PAPER Magma storage and evolution of

Contrib Mineral Petrol (2016) 171:30

1 3

30 Page 14 of 19

most evolved inclusions have ratios that approximate those found in the groundmass glass (Fig. 5). The Tuff of Cold Mountain Creek, Tuff of Bluff Point, and Trischmann Knob do not follow that pattern. In those units, the inclusions increase from the primitive values towards the enriched values that would align with the groundmass fractiona-tion trend (Fig. 8a, b). The groundmass compositions are less evolved than some of the inclusions. To obtain primi-tive groundmass, yet maintain the compositional evolution observed in the inclusions, less-evolved magma (likely still rhyolitic) mixed with the resident high-silica rhyolite after most quartz growth ceased yet prior to eruption. The mix-ing must have been thorough as no other evidence of mix-ing is preserved.

We use the range of incompatible trace-element concen-trations from the most primitive inclusions to the evolved groundmass to estimate the percent crystal fractionation. To make the estimates, we use a Rayleigh fractional crys-tallization model:

where f is the mass fraction of melt, Cp is the trace ele-ment concentration in the most primitive inclusion, Cgm is the trace element concentration in the groundmass, and D is the bulk distribution coefficient (Online Resource 4). We acknowledge that our estimates use bulk distribution coefficients that have unavoidable uncertainties, which we minimized by carefully selecting the most appropriate published values. The fraction of crystals formed is equal to 1−f. We calculate the percent fractionation using more than 15 incompatible trace elements and average those values to provide a single estimate for each eruptive unit. Fractionation estimates for those units range from 36 ± 12 to 52 ± 12 wt %, which are similar to previously published estimates of ~40 wt% (Vazquez and Reid 2002; Leeman and Phelps 1981; Hildreth et al. 1991; Bindeman and Val-ley 2001), and slightly larger than the 25–30 wt% estimate from Watts et al. (2012), and 10–15 % using the differ-ence between whole-rock and matrix glass by Loewen and Bindeman (2015). We acknowledge that Rayleigh frac-tionation is a simple model that does not fully replicate nature. We use it to gain a first-order approximation of the extent of the process. Interestingly, if we were to relax the assumption of perfect separation, then even greater degrees of crystallization would be required to produce the spread in trace element concentration.

The erupted products contain only 5–10 vol% phe-nocrysts (~6–12 wt% as quartz and sanidine are more dense than hydrous rhyolite melt), whereas fractionation estimates range up to ~50 wt%. Short-duration tempera-ture fluctuations provide one mechanism to reconcile the difference in crystallinity. Phase equilibria experiments

Cgm

Cp= f D−1

demonstrate that increase in temperature by a few 10’s of degrees can cause large changes in crystal content. Rare cores of plagioclase in Solfatara Plateau suggest that batch of magma was colder than 700 °C for a time, but then was reheated. Similar heating events may have occurred in other units, but cannot be confirmed as they lack plagio-clase. Loewen and Bindeman (2015) argue that elevated concentrations of Zr surrounding zircon crystals indicates heating events indeed preceded eruptions of Central Pla-teau Member lavas. If temperatures did increase, then the timing of that thermal event must have been short-lived because temperature-sensitive elements do not increase in concentration in the rims of phenocrysts from most flows (e.g., Ti-in-zircon) (Stelten et al. 2015). The crystallinity problem can also be reconciled through pre-eruptive melt separation. During melt separation, crystal-poor melts are extracted from a crystal-rich mush (e.g., Bachmann and Bergantz 2004), and form a crystal-poor layer near the top of the chamber (e.g., Huber et al. 2012). We suggest that the eruptions tapped that crystal-poor layer, which may also have been affected by temperature fluctuations.

Our dataset requires a model that combines both a single evolving chamber as well as separate isolated chambers. Recall that Rb/Sr demonstrates the Central Plateau Member magmas evolve progressively with time (Fig. 8; Vazquez et al. 2009). Despite that, estimates for crystal fractionation using trace elements in inclusions and matrix glass do not change progressively with eruption age. Instead, fractiona-tion estimates vary randomly with eruption age and posi-tion. In addition, the distinct CO2-poor and CO2-rich flu-ids preserved in inclusions from separate units are also not consistent with a single shared evolving chamber, as they do not correlate with eruption age. Together, volatiles and trace elements in inclusions and matrix glasses suggest that the magmas were sourced from multiple chambers that fol-lowed similar evolutionary paths, which is consistent with models of localized magma chambers (as discussed in Bindeman and Valley 2001; Bindeman et al. 2008; Vazquez et al. 2009; Girard and Stix 2010; Loewen and Bindman 2015; Stelten et al. 2015). We suggest that our data favor the model proposed by Stelten et al. (2015) and provide one additional constraint. They suggest that Central Pla-teau Member melts are extracted from a shared, long-lived source (as discussed in Vazquez and Reid 2002; Vazquez et al. 2009; Watts et al. 2012; Stelten et al. 2013), which then integrate to form isolated pots of crystal-poor melt containing only rare antecrysts. Phenocrysts form in those isolated chambers, which are ephemeral and erupt in less than ~10,000 years after forming (Stelten et al. 2015). Our data supplement this model in suggesting that significant fractional crystallization also operates in these isolated pots, with only the crystal-poor portions being erupted. It is possible that the isolated pots even occur within the larger

Page 15: Magma storage and evolution of the most recent effusive ... · 1 3 Contrib Mineral Petrol (2016) 171:30 DOI 10.1007/s00410-016-1244-x ORIGINAL PAPER Magma storage and evolution of

Contrib Mineral Petrol (2016) 171:30

1 3

Page 15 of 19 30

evolving and near-solidus chamber (e.g., Long Valley sys-tem, Hildreth 2004). To visualize one possible distribution of the eruptible portions of the magmas in the subsurface, we calculated chamber dimensions assuming the following: thickness was equal to the range in volatile saturation pres-sures from glass inclusions, the chambers were cylinders centered under the vent, and eruption volume was equal to chamber volume (Fig. 9). We find that the chambers would not have overlapped in that configuration, or even if the chambers were only 200 m thick. Although clearly specu-lative, those geometric and geographic constraints support the model of discrete magma batches evolving in tandem.

Granophyre and Solfatara Plateau heterogeneity

Granophyre had not been identified previously in Solfatara Plateau, or any other Central Plateau Member Rhyolite, and its discovery presents a unique opportunity to gain new insights to the magmatic system and/or eruption dynamics. The composition of sanidine in the granophyre is indis-tinguishable from phenocrystic sanidine, which suggests they formed under similar storage conditions. In the field, the transition from granophyre-bearing to granophyre-free portions sometimes corresponds with a sharp topographic break, with the granophyre-bearing portions at lower

10 km N111° 00’

45° 15’

Pitchstone Plateau

Solfatara Plateau

West Yellowstone

Caldera rim

Summit Lake

Trischmann Knob

Douglas Knob

Yellowstone Lake

Proposed vent site of sampled lavas

Cold Mt. Creek

0

1000

2000

3000

4000

5000

6000

Dep

th (m

)

?

(a)

(b)

Fig. 9 Simplified geologic map of Central Plateau Member Rhyo-lites in Yellowstone Caldera and the inferred spatial relationship of their magma chambers, assuming they are cylinders. Red circles in a show 2D distribution. Chamber thickness for each unit is estimated using the range in depth inferred from volatile saturation in analyzed

glass inclusions. Chamber radius is set to ensure total eruption vol-ume matches the volume of the cylindrical magma chambers. b In 3D note that no magma chambers overlap. They do not overlap even when thickness is assumed to be 200 m, forcing a much wider radius

Page 16: Magma storage and evolution of the most recent effusive ... · 1 3 Contrib Mineral Petrol (2016) 171:30 DOI 10.1007/s00410-016-1244-x ORIGINAL PAPER Magma storage and evolution of

Contrib Mineral Petrol (2016) 171:30

1 3

30 Page 16 of 19

elevations on the flow (Fig. 3). The topographic break is considered to be unrelated to post-emplacement ero-sion because it and other surface features such as pressure ridges and pumiceous carapace remain preserved. Because the distribution of granophyre switches near a topographic break, we argue that Solfatara Plateau is a composite lava comprised of 2 separate flows whereby the granophyre-bearing lava as the product of an earlier effusion (~5 km3) that was overridden by a younger, granophyre-free flow (~2 km3). Other large-volume Central Plateau Member lavas may also be composite flows. Indeed, we observe similar topographic discontinuities on West Yellowstone and Pitchstone Plateau, but have not observed petrologic differences in samples from those domains.

Granophyric intergrowths are usually thought to repre-sent rapid crystallization in response to significant under-cooling at low pressures (Barker 1970; Lipman et al. 1997; Lowenstern et al. 1997; Morgan and London 2012). Most likely, granophyre within large bodies of magma reflects sudden undercooling in response to degassing. Quartz-hosted glass inclusions from Solfatara Plateau contain up to 2.0 wt% H2O and 240 ppm CO2, consistent with storage and quartz crystallization at depths of 2–3 km. Because the inclusions indicate shallow storage, we suggest the grano-phyre formed in response to a significant degassing event when the magma was already at shallow depth. Although the interior of the granophyric intergrowths display clas-sic texture for rapid crystallization, the rounded, anhedral to subhedral exterior of the granophyre and discontinuous bands of small quartz and alkali feldspar fragments suggest that granophyre started to dissolve just prior to eruption. We interpret dissolution as a response to elevated tempera-tures initiated by a magmatic heating event that preceded (and possibly triggered) the effusive eruption (e.g., Bach-mann et al. 2002; Girard and Stix 2010; Morgan and Lon-don 2012). The heating event was likely initiated by under-plating of a hotter magma just prior to eruption, because phenocrysts are compositionally unzoned, and do not preserve evidence of pre-eruptive thermal disequilibrium, magma-mixing, or convective overturn. This suggests the Central Plateau Member magmatic systems may undergo rapid changes that are not preserved by crystal textures or compositions (e.g., crystallization, heating, degassing, mixing).

Solfatara Plateau also preserves a limited geochemical heterogeneity. Our petrographic and compositional analy-ses of minerals and matrix glasses indicate that Solfatara Plateau is largely similar, if not identical, to the other Central Plateau Member Rhyolites. This conclusion dif-fers with results from Vazquez and Reid (2002), Vazquez et al. (2009), and Stelten et al. (2013), who each demon-strate Solfatara Plateau is compositionally distinct from other lavas with respect to trace elements and isotopes.

For example, they demonstrate enriched Ba concentrations in sanidine (0.6 ± 0.2 wt%) and matrix glass (595 ppm). The compositional distinction, however, is based on a sin-gle sample (YCV04) reused by each of those three pro-jects. Analyses of matrix glass and sanidine by Girard and Stix (2010) and this study use multiple samples collected across a large portion of Solfatara Plateau, thus providing important context for interpreting sample YCV04 (Online Resource 6). Girard and Stix (2010) suggest YCV04 is anomalous for Solfatara Plateau because matrix glass from their samples contains ~60 ppm Ba, whereas YCV04 contains 595 ppm Ba. The Solfatara Plateau sample from Bindeman and Valley (2001) is also anomalous and has a whole-rock Ba concentration of 870 ppm. We report a Ba concentration of 76.6 ± 6.2 ppm for the near-vent sample Y82, which is comparable to the compositions reported by Girard and Stix (2010) for Solfatara Plateau and many of the other Central Plateau Member Rhyolites. In addition, 5 of the 6 samples with our new sanidine analyses (collected over a ~20 km2 area) contain sanidine that are Or49.2±2.7 and contain 0.19 ± 0.06 wt% BaO. Sanidine from our remaining new sample (Y36) are Or42.8±2.9 and contain 0.58 ± 0.14 wt% BaO, both of which are indistinguishable from sample YCV04 (Vazquez and Reid 2002; Vazquez et al. 2009; Stelten et al. 2013). Interestingly, Ba-rich sam-ples, Y36, YL96-16, and YCV04, were collected over an area of <1 km2 on the western flow front. Following the work by Girard and Stix (2010), we conclude that Solfa-tara Plateau is indeed a “normal” Central Plateau Member Rhyolite, but it likely preserves a small-scale heterogene-ity. Our conclusion does not obviate the important results from Vazquez and Reid (2002), Vazquez et al. (2009), and Stelten et al. (2013). Instead, it may imply that those authors discovered a rarely preserved parcel of melt com-posed of a mixture of extra-caldera and “normal” rhyolite (Stelten et al. 2013), or parental rhyolite (Christiansen and McMurry 2008), that did not have time to homogenize with the “normal” magma prior to the eruption, and may have acted as an eruption trigger. Similar concentrations of Ba in sanidine rims (0.6–0.7 wt%) occur in Scaup Lake flow and were inferred to represent a thermal rejuvenation event prior to eruption of that flow (Till et al. 2015).

Controls on eruption style

To first order, eruption style is known to be controlled by ascent rate (Rutherford 2008). In turn, ascent rate is likely controlled by a variety of factors, including pre-eruptive conditions such as temperature, pressure, volatile content, and viscosity. Using a diverse set of geochemical tools, we characterized the pre-eruption conditions for phenocryst growth for both the effusive and explosive eruptions. We conclude that during phenocryst growth, all of the magmas

Page 17: Magma storage and evolution of the most recent effusive ... · 1 3 Contrib Mineral Petrol (2016) 171:30 DOI 10.1007/s00410-016-1244-x ORIGINAL PAPER Magma storage and evolution of

Contrib Mineral Petrol (2016) 171:30

1 3

Page 17 of 19 30

were likely stored at similar temperatures (±25 °C), and their water contents are indistinguishable, ranging from ~1.5 to ~2.5 wt%. Because melt composition, crystal con-tent, temperature, and dissolved H2O are similar amongst all units, the pre-eruptive magmatic viscosities were simi-lar and are estimated to have been 107.0±0.5 Pa s (Giordano et al. 2008). The units can be separated into distinct pop-ulations based on CO2 content, but those populations do not clearly discriminate based on eruption style or volume. High-CO2 magmas fed the pyroclastic eruptions of the Tuff of Bluff Point and the remnant pyroclastic deposit on Douglas Knob. They were also the source of the large-vol-ume effusive lavas (>30 km3) of Pitchstone Plateau, Sum-mit Lake, and Buffalo Lake, as well as the small-volume (<0.1 km3) Trischmann Knob. Low-CO2 magmas also generated a pyroclastic eruption, the Tuff of Cold Moun-tain Creek, and sourced both the small- and large-volume effusive lavas (<0.1 to >30 km3) of Douglas Knob, Sol-fatara Plateau, and West Yellowstone. It seems clear that pre-eruptive storage conditions did not control eruptive volume or style. Instead, conduit dimensions or chamber overpressures could have controlled those eruption vari-ables, but we acknowledge this is speculative. Interest-ingly, it is clear that the Central Plateau Member magmas were compositionally distinct with respect to volatile con-tents from the Lava Creek Tuff magma, which contained 3.2–4.0 wt% H2O and 100–600 ppm CO2 (Fig. 4) (Gan-secki and Lowenstern 1995; Gansecki 1998). Such evi-dence indicates magma genesis and the overall Yellow-stone plumbing system may behave differently leading up to supereruptions.

Trace element contents in matrix glass and quartz-hosted glass inclusions are an insightful tracer to changes in the magmatic system. In general, trace elements indicate the magmas experienced similar fractionation histories irrespec-tive of eruption volume. The Tuffs of Cold Mountain Creek and Bluff Point deviate from such a behavior. Trace elements in the groundmass glass of those units are more primitive than in inclusions, providing strong evidence for pre-eruptive mixing with a less-evolved batch of rhyolite, perhaps similar to the low silica rhyolite of East Biscuit Basin (Girard and Stix 2009; Watts et al. 2012). It is possible that mixing events triggered the explosive eruptions of the Tuff of Cold Moun-tain Creek and the Tuff of Bluff Point.

Conclusion

We present four complementary datasets that provide insight into the most recent episode of volcanism at Yellowstone. The eruptions that emplaced the Central Plateau Member Rhyolites were sourced from crystal-poor melts that were at least temporarily stored as separate batches of magma with

volumes up to 70 km3. The rhyolite magmas experienced 36 ± 12–52 ± 12 % fractional crystallization prior to melt segregation events and eruption. The magmas were saturated with a mixed fluid, and the individual eruptive units can be subdivided based on fluid composition into CO2-poor and CO2-rich populations. The magmas crystallized their phe-nocrysts at 750 ± 25 °C, with the CO2-poor and CO2-rich magmas occurring in the upper crust at depths of 2–3 and 3–6 km, respectively. The discovery of granophyre-bearing domains at Solfatara Plateau provides a unique petrologic clue about the storage and/or eruption of Central Plateau Member lavas. It demonstrates that rapid magmatic pro-cesses, such as systematic dissolution of granophyre in response to a heating event, can occur in this system.

Acknowledgments We thank Robert Zinke, Matt Williams, Ryan Cahalan, Kevin Befus, Tim Prather, Jake Jordan, Brent Jackson, Gail Mahood and Jim Watkins for scientific conversations and their assis-tance in the field. We also thank Renat Almeev, Mark Stelten, Jorge Vazquez, Kathryn Watts, and 3 anonymous reviewers for helpful reviews. Nathan Miller operated the LA-ICP-MS and helped process the trace-element data. This research was made possible by a grant from the National Science Foundation (EAR-1049829) to J.E.G and a National Park Service research permit (YELL-05678).

References

Almeev RR, Bolte T, Nash BP, Holtz F (2012) High-temperature, low-H2O silicic magmas of the yellowstone hotspot: an experi-mental study of rhyolite from the Bruneau-Jarbidge Eruptive Center, Central Snake River Plain, USA. J Petrol 53:1837–1866

Andersen DJ, Lindsley DH, Davidson PM (1993) QUILF: a pas-cal program to assess equilibria among Fe–Mg–Mn–Ti oxides, pyroxenes, olivine, and quartz. Comp Geosci 19:1333–1350

Bachmann O, Bergantz GW (2004) On the origin of crystal-poor rhyolites: extracted from batholithic crystal mushes. J Petrol 45:1565–1582

Bachmann O, Dungan MA, Lipman PW (2002) The fish canyon magma body, San Juan volcanic field, Colorado: rejuvenation and eruption of an upper-crustal batholith. J Petrol 43:1469–1503

Barker DS (1970) Compositions of granophyre, myrmekite, and graphic granite. Geol Soc Am Bull 81:3339–3350

Befus KS, Gardner JE, Zinke RW (2012) Analyzing water contents in unexposed glass inclusions in quartz crystals. Am Mineral 97:1898–1904

Befus KS, Zinke RW, Jordan JS, Manga M, Gardner JE (2014) Pre-eruptive storage conditions and eruption dynamics of a small rhyolite dome: Douglas Knob, Yellowstone volcanic field, USA. Bull Volcanol 76:1–12

Behrens H, Tamic N, Holtz F (2004) Determination of the molar absorption coefficient for the infrared absorption band of CO2 in rhyolitic glasses. Am Mineral 89:301–306

Bindeman IN, Valley JW (2000) Formation of low-δ18O rhyolites after caldera collapse at Yellowstone, Wyoming, USA. Geology 28:719–722

Bindeman IN, Valley JW (2001) Low-δ18O rhyolites from Yellow-stone: magmatic evolution based on analyses of zircons and indi-vidual phenocrysts. J Petrol 42:1491–1517

Bindeman IN, Valley JW, Wooden JL, Persing HM (2001) Post-cal-dera volcanism: in situ measurement of U–Pb age and oxygen

Page 18: Magma storage and evolution of the most recent effusive ... · 1 3 Contrib Mineral Petrol (2016) 171:30 DOI 10.1007/s00410-016-1244-x ORIGINAL PAPER Magma storage and evolution of

Contrib Mineral Petrol (2016) 171:30

1 3

30 Page 18 of 19

isotope ratio in Pleistocene zircons from Yellowstone Caldera. Earth Planet Sci Lett 189:197–206

Bindeman IN, Fu B, Kita, NT, Valley JW (2008) Origin and evolution of silicic magmatism at Yellowstone based on ion microprobe analysis of isotopically zoned zircons. J Petrol 49:163–193

Bolte M, Holtz F, Almeev R, Nash B (2015) The Blacktail Creek Tuff: an analytical and experimental study of rhyolites from the Heise volcanic field, Yellowstone hotspot system. Contrib Mineral Pet-rol 169:1–24

Branney MJ, Bonnichsen B, Andrews GDM, Ellis B, Barry TL, McCurry M (2008) “Snake River (SR)-type” volcanism at the Yellowstone hotspot track: distinctive products from unu-sual, high-temperature silicic super-eruptions. Bull Volcanol 70:293–314

Cathey HE, Nash BP (2009) Pyroxene thermometry of rhyolite lavas of the Bruneau-Jarbidge eruptive center, Central Snake River Plain. J Volcanol Geotherm Res 188:173–185

Chang W-L, Smith RB, Wicks C, Farrell JM, Puskas CM (2007) Accelerated uplift and magmatic intrusion of the Yellowstone Caldera, 2004 to 2006. Science 318:952–956

Christiansen RL (2001) The quaternary and pliocene Yellowstone pla-teau volcanic field of Wyoming, Idaho, and Montana. USGS Prof Paper 729-G

Christiansen EH, McMurry M (2008) Constrasting origins of Ceno-zoic silicic volcanic rocks from the western Cordillera of the United States. Bull Volcanol 70:251–267

Christiansen RL, Lowenstern JB, Smith RB, Heasler H, Morgan LA, Nathenson M, Mastin LG, Muffler LJP, Robinson JE (2007) Pre-liminary assessment of volcanic and hydrothermal hazards in Yellowstone National Park and vicinity. USGS Open-file report 2007-1071

Chu R, Helmberger DV, Sun D, Jackson JM, Zhu L (2010) Mushy magma beneath Yellowstone. Geophys Res Lett. doi:10.1029/2009GL041656

DeNosaquo KR, Smith RB, Lowry AR (2009) Density and lith-ospheric strength models of the Yellowstone-Snake River Plain volcanic system from gravity and heat flow data. J Volcanol Geo-therm Res 188:108–127. doi:10.1016/jjvolgeores200908006

Ellis BS, Barry T, Branney MJ (2010) Petrologic constraints on the development of a large-volume, high temperature, silicic magma system: the Twin Falls eruptive centre, central Snake River Plain. Lithos 120:475–489

Ellis BS, Mark DF, Pritchard CJ, Wolff JA (2012) Temporal dissec-tion of the Huckleberry Ridge Tuff using the 40Ar/39Ar dating technique. Quat Geochronol 9:34–41

Farrell J, Husen S, Smith RB (2009) Earthquake swarm and b-value characterization of the Yellowstone volcano-tectonic system. J Volcanol Geotherm Res 188:260–276

Farrell J, Smith RB, Husen S, Diehl T (2014) Tomography from 26 years of seismicity revealing that the spatial extent of Yellow-stone crustal magma reservoir extends well beyond the Yellow-stone Caldera. Geophys Res Lett 41:3068–3073

Gansecki CA (1998) 40Ar/39Ar geochronology and pre-eruptive geo-chemistry of the Yellowstone Plateau volcanic field rhyolites. UMI Dissertation Services, Stanford

Gansecki CA, Lowenstern JB (1995) Pre-eruptive volatile composi-tions of the Lava Creek Tuff magma, Yellowstone Plateau Vol-canic Field. Trans Am Geophys Union Eos 76:F665

Gansecki CA, Mahood GA, McWilliams MO (1996) 40Ar39Ar geo-chronology of rhyolites erupted following collapse of the Yel-lowstone Caldera, Yellowstone Plateau volcanic field: implica-tions for crustal contamination. Earth Planet Sci Lett 142:91–107

Gardner JE, Rutherford M, Carey S, Sigurdsson H (1995) Experi-mental constraints on pre-eruptive water contents and changing magma storage prior to explosive eruptions of Mount St Helens volcano. Bull Volcanol 57:1–17

Gardner JE, Befus KS, Gualda GAR, Ghiorso MS (2014) Experi-mental constraints on rhyolite-MELTS and the Late Bishop Tuff magma body. Contrib Mineral Petrol 168:1051

Giordano D, Russell DK, Dingwell DB (2008) Viscosity of magmatic liquids: a model. Earth Planet Sci Lett 271:123–134

Girard G, Stix J (2009) Magma recharge and crystal mush rejuvena-tion associated with early post-collapse Upper Basin Member Rhyolites, Yellowstone Caldera, Wyoming. J Petrol. doi:10.1093/petrology/egp070

Girard G, Stix J (2010) Rapid extraction of discrete magma batches from a large differentiating magma chamber: the Central Plateau Member rhyolites, Yellowstone Caldera, Wyoming. Contrib Min-eral Petrol 160:441–465

Girard G, Stix J (2012) Future volcanism at Yellowstone Caldera: insights from geochemistry of young volcanic units and monitor-ing of volcanic unrest. GSA Today. doi:10.1130/GSATG143A1

Hildreth W (2004) Volcanological perspectives on Long Valley, Mam-moth Mountain, and Mono Craters: several contiguous but dis-crete systems. J Volcanol Geotherm Res 136:169–198

Hildreth W, Halliday AN, Christiansen RL (1991) Isotopic and chem-ical evidence concerning the genesis and contamination of basal-tic and rhyolitic magma beneath the Yellowstone Plateau vol-canic field. J Petrol 32:63–138

Honjo N, Bonnichsen B, Leeman WP (1992) Mineralogy and geo-thermometry of high-temperature rhyolites from the central and western Snake River Plain. Bull Volcanol 54:220–237

Huang R, Audetat A (2012) The titanium-in-quartz (TitaniQ) thermo-barometer: a critical examination and re-calibration. Geochim Cosmochim Acta 84:75–89

Huber C, Bachmann O, Dufek J (2012) Crystal-poor versus crystal-rich ignimbrites: a competition between stirring and reactivation. Geology 40:115–118

Lanphere MA, Champion DE, Christiansen RL, Izett GA, Obradovich JD (2002) Revised ages for tuffs of the Yellowstone Plateau vol-canic field: assignment of the Huckleberry Ridge Tuff to a new geomagnetic polarity event. Geol Soc Am Bull 114:559–568

Leeman WP, Phelps DW (1981) Partitioning of rare earths and other trace elements between sanidine and coexisting volcanic glass. J Geophys Res 86:10193–10199

Lipman P, Dungan M, Bachmann O (1997) Comagmatic granophy-ric granite in the Fish Canyon Tuff, Colorado: implications for magma-chamber processes during a large ash-flow eruption. Geology 25:915–918

Liu Y, Zhang Y, Behrens H (2005) Solubility of H2O in rhyolitic melts at low pressures and a new empirical model for mixed H2O–CO2 solubility in rhyolitic melts. J Volcanol Geotherm Res 143:219–235

Loewen MW, Bindeman IN (2015) Oxygen isotope and trace element evidence for three-stage petrogenesis of the youngest episode (260–79 ka) of Yellowstone rhyolitic volcanism. Contrib Mineral Petrol 170:1–25

Lowenstern JB, Hurwitz S (2008) Monitoring a supervolcano in repose: heat and volatile flux at the Yellowstone Caldera. Ele-ments 40:35–40

Lowenstern JB, Clynne MA, Bullen TD (1997) Comagmatic A-type granophyre and rhyolite from the Alid Volcanic Center, Eritrea, Northeast Africa. J Petrol 38:1707–1721

Lowenstern JB, Smith RB, Hill DP (2006) Monitoring super-volcanoes: geophysical and geochemical signals at Yellowstone and other large caldera systems. Philos Trans R Soc Lon 364:2055–2072

Morgan GB, London D (2012) Process of granophyre crystallization in the Long Mountain Granite, southern Oklahoma. Geol Soc Am Bull 124:1251–1261

Nielsen CH, Sigurdsson H (1981) Quantitative methods for electron microprobe analysis of sodium in natural and synthetic glasses. Am Mineral 66:547–552

Page 19: Magma storage and evolution of the most recent effusive ... · 1 3 Contrib Mineral Petrol (2016) 171:30 DOI 10.1007/s00410-016-1244-x ORIGINAL PAPER Magma storage and evolution of

Contrib Mineral Petrol (2016) 171:30

1 3

Page 19 of 19 30

Pelton JR, Smith RB (1979) Recent crustal uplift in Yellowstone National Park. Science 206:1179–1182

Pritchard CJ, Larson PB (2012) Genesis of the post-caldera eastern Upper Basin Member rhyolites, Yellowstone, WY: from volcanic stratigraphy, geochemistry, and radiogenic isotope modeling. Contrib Mineral Petrol 164:205–228

Pyle DM (1989) The thickness, volume and grainsize of tephra fall deposits. Bull Volcanol 51:1–15

Rutherford MJ (2008) Magma ascent rates. Rev Mineral Geochem 69:241–271

Skirius CM, Peterson JW, Anderson AT (1990) Homogenizing rhyolitic glass inclusions from the Bishop Tuff. Am Mineral 75:1381–1398

Smith RB, Jordan M, Steinberger B, Puskas CM, Farrell J, Waite GP, Husen S, Chang W-L, O’Connell R (2009) Geodynamics of the Yellowstone hotspot and mantle plume: seismic and GPS imaging, kinematics, and mantle flow. J Volcanol Geotherm Res 188:26–56

Stelten ME, Cooper KM, Vazquez JA, Reid MR, Garfod GH, Wimpenny J, Qing-zhu Y (2013) Magma mixing and the gen-eration of isotopically juvenile silicic magma at Yellowstone Caldera inferred from coupling 238U–230Th ages with trace ele-ments and Hf and O isotopes in zircon and Pb isotopes in sani-dine. Contrib Mineral Petrol 166:587–613

Stelten ME, Cooper KM, Vazquez JA, Calvert AT, Glessner JJG (2015) Mechanisms and timescales of generating eruptible rhy-olitic magmas at Yellowstone Caldera from zircon and sanidine geochronology and geochemistry. J Petrol 56:1607–1642

Thomas JB, Watson BE, Spear FS, Shemella PT, Nayak SK, Lanzi-rotti A (2010) TitaniQ under pressure: the effect of pressure and temperature on the solubility of Ti in quartz. Contrib Mineral Petrol 160:743–759

Thomas JB, Watson BE, Spear FS, Wark DA (2015) TitaniQ recrys-tallized: experimental confirmation of the original Ti-in-quartz calibrations. Contrib Mineral Petrol 169:1–16

Till CB, Vazquez JA, Boyce JW (2015) Months between rejuvenation and volcanic eruption at Yellowstone Caldera, Wyoming. Geol-ogy 43:695–698

Vazquez JA, Reid MR (2002) Time scales of magma storage and dif-ferentiation of voluminous high-silica rhyolites at Yellowstone Caldera, Wyoming. Contrib Mineral Petrol 144:274–285

Vazquez JA, Kyriazis SF, Reid MR, Sehler RC, Ramos FC (2009) Thermochemical evolution of young rhyolites at Yellowstone: evidence for a cooling but periodically replenished postcaldera magma reservoir. J Volcanol Geotherm Res 188:186–196

Walker GPL, Huntingdon AT, Sanders AT, Dinsdale JL (1973) Lengths of lava flows [and discussion]. Phils Trans R Soc Lond 274:107–118

Wark DA, Watson EB (2006) TitaniQ: a titanium-in-quartz geother-mometer. Contrib Mineral Petrol 152:743–754

Watts KE, Bindeman IN, Schmitt AK (2012) Crystal scale anatomy of a dying supervolcano: an isotope and geochronology study of individual phenocrysts from voluminous rhyolites of the Yellow-stone Caldera. Contrib Mineral Petrol 164:45–67

Wicks CW, Thatcher W, Dzurisin D, Svarc J (2006) Uplift, thermal unrest and magma intrusion at Yellowstone Caldera. Nature 440:72–75

Wolff JA, Ellis BS, Ramos FC (2011) Strontium isotopes and magma dynamics: insights from high-temperature rhyolites. Geology 39:931–934

Zhang Y, Belcher R, Ihinger PD, Wan L, Xu Z, Newman S (1997) New calibration of infrared measurement of dissolved water in rhyolitic glasses. Geochim Cosmochim Acta 61:3089–3100