+ All Categories
Home > Documents > Lithium isotope fractionation during magma degassing: Constraints from silicic differentiates and...

Lithium isotope fractionation during magma degassing: Constraints from silicic differentiates and...

Date post: 25-Nov-2023
Category:
Upload: independent
View: 0 times
Download: 0 times
Share this document with a friend
9
Author's personal copy Lithium isotope fractionation during magma degassing: Constraints from silicic differentiates and natural gas condensates from Piton de la Fournaise volcano (Réunion Island) I. Vlastélic a,b,c, , T. Staudacher d , P. Bachèlery e , P. Télouk f , D. Neuville g , M. Benbakkar a,b,c a Clermont Université, Université Blaise Pascal, Laboratoire Magmas et Volcans, BP 10448, F-63000 Clermont-Ferrand, France b CNRS, UMR 6524, LMV, F-63038 Clermont-Ferrand, France c IRD, R 163, LMV, F-63038 Clermont-Ferrand, France d Observatoire Volcanologique du Piton de la Fournaise, Institut de Physique du Globe de Paris, CNRS UMR 7154, 14 RN3, le 27 ème km, 97418, La Plaine des Cafres, La Réunion, France e Laboratoire GéoSciences Réunion, Université de La Réunion, Institut de Physique du Globe de Paris, CNRS UMR 7154, 15 Avenue René Cassin, 97715 Saint-Denis cedex 09, La Réunion, France` f Laboratoire des Sciences de la Terre, Ecole Normale Supérieure de Lyon, CNRS UMR 5570 46 Allée d'Italie, 69364 Lyon cedex 07, France g Laboratoire de Géochimie et Cosmochimie, Université Paris Diderot, Institut de Physique du Globe de Paris, CNRS UMR 7154, 1 rue Jussieu, 75238 Paris cedex 05, France abstract article info Article history: Received 30 August 2010 Received in revised form 31 January 2011 Accepted 2 February 2011 Available online 5 March 2011 Edited by: R.L. Rudnick Keywords: Lithium isotopes Isotopic fractionation Magma degassing Piton de la Fournaise volcano, Réunion Island Recent volcanic products from the Piton de la Fournaise Volcano, Reunion, show pronounced depletion or enrichment in lithium and signicant isotopic fractionation related to degassing. (1) trachytic pumices from the April 2007 eruption show extreme Li depletion (90%) and isotopic fractionation (δ 7 Li of 21). The depletion of water and volatiles (Cl, F, B, Cs) in these samples suggests that Li loss occurred in response to degassing, which most likely occurred as the small, isolated volume of magma underwent extensive differentiation near the surface. Because the pre-degassing composition is relatively well known, the composition of the degassed pumice constrains the partition coefcient to 60 b D VM b 135 and the isotopic fractionation factor, α VM , to 1.010 at magmatic temperatures. Unlike D VM , α VM does not depend on whether crystallization and degassing occurred successively or concomitantly. (2) basaltic samples from the interior wall of the long-lived 1998 Piton Kapor were extensively altered by acidic gas. They also show extreme Li depletion, but barely signicant isotopic fractionation (δ 7 Li=+4.5), suggesting that high- temperature leaching of Li by volcanic gas does not signicantly fractionate Li isotopes. (3) high-temperature (400325 °C) gas condensates formed during degassing of the thick lava ow of April 2007 display high Li contents (50100 ppm), which are consistent with Li being as volatile as Zn and Sn. Their isotopically light Li signature (average of 1.7) is consistent with their derivation from isotopically heavy vapor (+ 13.5) if the factor of isotopic fractionation between condensate and vapor is less than 0.985. A degassing- crystallization model accounts for the evolution of trace species, which, like lithium, are volatile but also moderately incompatible. © 2011 Elsevier B.V. All rights reserved. 1. Introduction Alkali metals are only moderately volatile, in the sense that only a small proportion (b 0.1%) of the initial budget of mantle derived melts is ultimately lost by magmas (Rubin, 1997). On the other hand, alkali are largely mobilized during exsolution of H 2 O-rich vapor phase in magmatic systems (Sakuyama and Kushiro, 1979). Selective vapor transport of potassium has been suggested during the late stage of crystallization of oceanic magmas (Sinton and Byerly, 1980), and from bottom to top of lava lakes (Richter and Moore, 1966) although the Kilauea data on which this idea is based were subsequently questioned (Helz et al., 1994). There has been recently a growing interest in lithium, the lightest alkali metal, but above all one of the fastest diffusing elements in silicates (Jambon and Semet, 1978; Giletti and Shanahan, 1997; Richter et al., 2003). Indeed, lithium seems to be particularly sensitive to vapor transfer, either in shallow magma conduit shortly before eruption (Berlo et al., 2004; Kent et al., 2007) or during post-eruptive degassing of thick lava ows (Kuritani and Nakamura, 2006). While the use of lithium abundance as tracer of degassing processes is still in its infancy, new perspectives arise from lithium stable isotopes ( 6 Li and 7 Li) geochemistry. The main reason is that our knowledge of the laws governing lithium isotopes fractionation in nature has improved signicantly in recent years. First, it is now well established that heavy lithium partitions preferentially into aqueous Chemical Geology 284 (2011) 2634 Corresponding author at: Laboratoire Magmas et Volcans, Observatoire de Physique du Globe de Clermont-Ferrand, UMR 6524, 5 Rue Kessler, 63038 Clermont- Ferrand, France. Tel.: +33 4 73 34 67 10; fax: +33 4 73 34 67 44. E-mail address: [email protected] (I. Vlastélic). 0009-2541/$ see front matter © 2011 Elsevier B.V. All rights reserved. doi:10.1016/j.chemgeo.2011.02.002 Contents lists available at ScienceDirect Chemical Geology journal homepage: www.elsevier.com/locate/chemgeo
Transcript

Author's personal copy

Lithium isotope fractionation during magma degassing: Constraints from silicicdifferentiates and natural gas condensates from Piton de la Fournaisevolcano (Réunion Island)

I. Vlastélic a,b,c,⁎, T. Staudacher d, P. Bachèlery e, P. Télouk f, D. Neuville g, M. Benbakkar a,b,c

a Clermont Université, Université Blaise Pascal, Laboratoire Magmas et Volcans, BP 10448, F-63000 Clermont-Ferrand, Franceb CNRS, UMR 6524, LMV, F-63038 Clermont-Ferrand, Francec IRD, R 163, LMV, F-63038 Clermont-Ferrand, Franced Observatoire Volcanologique du Piton de la Fournaise, Institut de Physique du Globe de Paris, CNRS UMR 7154, 14 RN3, le 27èmekm, 97418, La Plaine des Cafres, La Réunion, Francee Laboratoire GéoSciences Réunion, Université de La Réunion, Institut de Physique du Globe de Paris, CNRS UMR 7154, 15 Avenue René Cassin, 97715 Saint-Denis cedex 09,La Réunion, France`f Laboratoire des Sciences de la Terre, Ecole Normale Supérieure de Lyon, CNRS UMR 5570 46 Allée d'Italie, 69364 Lyon cedex 07, Franceg Laboratoire de Géochimie et Cosmochimie, Université Paris Diderot, Institut de Physique du Globe de Paris, CNRS UMR 7154, 1 rue Jussieu, 75238 Paris cedex 05, France

a b s t r a c ta r t i c l e i n f o

Article history:Received 30 August 2010Received in revised form 31 January 2011Accepted 2 February 2011Available online 5 March 2011

Edited by: R.L. Rudnick

Keywords:Lithium isotopesIsotopic fractionationMagma degassingPiton de la Fournaise volcano, Réunion Island

Recent volcanic products from the Piton de la Fournaise Volcano, Reunion, show pronounced depletion orenrichment in lithium and significant isotopic fractionation related to degassing. (1) trachytic pumices fromthe April 2007 eruption show extreme Li depletion (90%) and isotopic fractionation (δ7Li of −21‰). Thedepletion of water and volatiles (Cl, F, B, Cs) in these samples suggests that Li loss occurred in response todegassing, which most likely occurred as the small, isolated volume of magma underwent extensivedifferentiation near the surface. Because the pre-degassing composition is relatively well known, thecomposition of the degassed pumice constrains the partition coefficient to 60bDV–Mb135 and the isotopicfractionation factor, αV–M, to 1.010 at magmatic temperatures. Unlike DV–M, αV–M does not depend onwhether crystallization and degassing occurred successively or concomitantly. (2) basaltic samples from theinterior wall of the long-lived 1998 Piton Kapor were extensively altered by acidic gas. They also showextreme Li depletion, but barely significant isotopic fractionation (δ7Li=+4.5‰), suggesting that high-temperature leaching of Li by volcanic gas does not significantly fractionate Li isotopes. (3) high-temperature(400–325 °C) gas condensates formed during degassing of the thick lava flow of April 2007 display high Licontents (50–100 ppm), which are consistent with Li being as volatile as Zn and Sn. Their isotopically light Lisignature (average of −1.7‰) is consistent with their derivation from isotopically heavy vapor (+13.5‰) ifthe factor of isotopic fractionation between condensate and vapor is less than 0.985. A degassing-crystallization model accounts for the evolution of trace species, which, like lithium, are volatile but alsomoderately incompatible.

© 2011 Elsevier B.V. All rights reserved.

1. Introduction

Alkali metals are only moderately volatile, in the sense that only asmall proportion (b0.1%) of the initial budget of mantle derived meltsis ultimately lost by magmas (Rubin, 1997). On the other hand, alkaliare largely mobilized during exsolution of H2O-rich vapor phase inmagmatic systems (Sakuyama and Kushiro, 1979). Selective vaportransport of potassium has been suggested during the late stage ofcrystallization of oceanic magmas (Sinton and Byerly, 1980), and frombottom to top of lava lakes (Richter and Moore, 1966) although the

Kilauea data on which this idea is based were subsequentlyquestioned (Helz et al., 1994). There has been recently a growinginterest in lithium, the lightest alkali metal, but above all one of thefastest diffusing elements in silicates (Jambon and Semet, 1978;Giletti and Shanahan, 1997; Richter et al., 2003). Indeed, lithiumseems to be particularly sensitive to vapor transfer, either in shallowmagma conduit shortly before eruption (Berlo et al., 2004; Kent et al.,2007) or during post-eruptive degassing of thick lava flows (Kuritaniand Nakamura, 2006).

While the use of lithium abundance as tracer of degassingprocesses is still in its infancy, new perspectives arise from lithiumstable isotopes (6Li and 7Li) geochemistry. The main reason is that ourknowledge of the laws governing lithium isotopes fractionation innature has improved significantly in recent years. First, it is now wellestablished that heavy lithium partitions preferentially into aqueous

Chemical Geology 284 (2011) 26–34

⁎ Corresponding author at: Laboratoire Magmas et Volcans, Observatoire dePhysique du Globe de Clermont-Ferrand, UMR 6524, 5 Rue Kessler, 63038 Clermont-Ferrand, France. Tel.: +33 4 73 34 67 10; fax: +33 4 73 34 67 44.

E-mail address: [email protected] (I. Vlastélic).

0009-2541/$ – see front matter © 2011 Elsevier B.V. All rights reserved.doi:10.1016/j.chemgeo.2011.02.002

Contents lists available at ScienceDirect

Chemical Geology

j ourna l homepage: www.e lsev ie r.com/ locate /chemgeo

Author's personal copy

fluids over silicate rocks, and that the magnitude of this phenomenondecreases with increasing temperature (e.g., Wunder et al., 2007;Millot et al., 2010). Second, as 6Li diffuses faster than 7Li, large isotopicfractionation arises during diffusion (e.g., Richter et al., 2003). Thesetwo types of isotopic fractionation, which, for convenience, will besimply referred to as “chemical” and “kinetic”, respectively, may occurduring magma degassing (Beck et al., 2004; Rowe et al., 2008; Schiaviet al., 2010), although little is yet known on this topic.

Taking advantage of the recent intense volcanic activity of Piton dela Fournaise (Réunion Island), a Li isotopic study of late-stagemagmatic processes was undertaken. This paper focuses on exten-sively devolatilized samples (silicic differentiates) and gas conden-sates (natural sublimates), which provide new constraints on thebehavior of Li isotopes during magma degassing.

2. Geological setting and samples

Piton de la Fournaise (Réunion Island, Indian Ocean) in anintraplate shield volcano that has erupted on average once a yearsince 1930, and seemingly more frequently since 1998 (see a reviewof the most recent activity in Peltier et al. (2009)). The volcanoproduces dominantly transitional basalts with little compositionalvariability (these basalts are referred to as Steady State Basalts (SSB)following Albarède et al., 1997) and, every 15–30 years and morefrequently since 2001, olivine-rich basalts (the most magnesian beingcommonly referred to as “oceanites”). Geophysical data support theexistence of a shallow magma reservoir near sea level (2.5 km depth)and a deeper reservoir at the crust-mantle interface (7.5 km depth)

(Peltier et al., 2008). The volume of the storage system, in the range of0.1–0.35 km3, and magma residence time of 15–30 years wereestimated based on geochemical data (Albarède, 1993; Sigmarssonet al., 2005; Vlastélic et al., 2009a).

Among the recent products of Piton de la Fournaise, this studyfocused on samples that underwent pronounced depletion orenrichment in lithium in relation with degassing processes (Fig. 1).They include trachytic pumices recovered near the main vent of April2007 eruption. The pumices are almost completely glassy, highlyvesicular (80–90%) and often coated by basaltic glass. These silicicdifferentiates, the only known at Piton de la Fournaise, are thought tooriginate from small batches of magmas trapped at shallow depthwithin the volcanic edifice (Bachèlery, in prep.). Comparing the Pbisotopic signature of the pumices to the detailed Pb isotope temporalrecord (Vlastélic et al., 2009a) suggests derivation from liquidserupted between 1977 and 1986. Theses pockets of differentiatedliquids may have been disrupted and entrained during the paroxys-mal phase of the eruption, on April 5th (see review of the eruption byStaudacher et al. (2009)). To confirm this hypothesis we also analyzedthe olivine-bearing glass coating the pumices. The second type ofsample is gas-altered rocks from the interior wall of the 1998 Kaporcrater. These rocks, which have been extensively leached by acidic gasduring the unusually long 1998 eruption (nearly 6 months), havebeen selected to evaluate how Li isotopes behave during gas-rockinteraction. The last type of sample is gas condensates that formedduring degassing of the April 2007 lava flow. Post-eruptive degassingprocesses, which include formation of segregation veins and gas-filterpressing (Martin and Sigmarsson, 2007), have been shown to

Fig. 1. Map of Piton de la Fournaise volcano showing the eruptions of March 1998 (Piton Kapor) and April 2007 (Piton Tremblet). Locations and photos of the studied samples areindicated by numbers: (1) gas altered basalt from the interior wall of the 1998 Kapor crater. (2) trachytic pumice and its basaltic coating (recovered in May 2007 near PitonTremblet) were probably emitted during the paroxysmal phase of April 2007 eruption (Bachèlery et al., in prep.) (3) gas condensates from a fumarole that was active between 2007and 2009 in the region where April 2007 lava flow is the thickest (ca. 50 m).

27I. Vlastélic et al. / Chemical Geology 284 (2011) 26–34

Author's personal copy

efficiently mobilize lithium (Kuritani and Nakamura, 2006). SincePiton de la Fournaise volcano lacks long-lived fumaroles, thevoluminous (130 Mm3) and slowly cooling lava flow of April 2007(Staudacher et al., 2009) offers a rare possibility to sample magmaticgas, in particular in its thickest part where an active fumarole wasdiscovered. Natural gas condensates were sampled four times at thissite between August 2008 and November 2009 as the venttemperature decreased from ~400 to ~300 °C. In addition, a largenumber of steady-state basalts (n=31) and olivine-rich basalts(n=30) erupted between 1927 and 2007 were also analyzed forcomparison.

3. Analytical methods

The bulk major element chemistry was determined by ICP-OES(rocks) and ion chromatography (condensates). Scanning electronmicroscopy (SEM) was used to investigate the compositionalheterogeneity of the altered basalt and to determine the mineralogyof gas condensates. Water content was determined in the trachyticpumice by Raman spectroscopy (IPGP, Paris) using a recentlyimproved calibration curve (Le Losq et al., in press). The Ramanspectra were recorded using a T64000 Jobin-Yvon spectrometer, withan excitation line at 514.532 nm. Acquisition times and laser powerwere set to 900s and 20 mW, respectively. Data treatment andaccuracy of the method are detailed in Le Losq et al. (in press). Whole-rock analysis of Cl, F and B concentrations has been assigned to theService d'Analyse des Roches et Minéraux (CRPG-Nancy). Chlorinecontent was determined by a spectrophotometric method based onthe formation of ferrithiocyanate, while F content was determined bypotentiometry using an ion-selective electrode. The precision of F andCl determination is 10–20% (2σ) for the concentration range (100–500 ppm) relevant to our study (Vernet et al., 1987). Followingsodium carbonate fusion, B was purified on ion-exchange resin and itsconcentration was determined by the colorimetric method usingCarmin with a precision of ca. 25%. Other trace elements and Liisotopic compositions were analyzed on same dissolutions. Between50 and 100 mg of sample were dissolved in HF-HNO3 (rocks) ordiluted HNO3 (gas condensates). Once digested, solutions wereevaporated to near dryness and re-dissolved in 7M HNO3. Thedissolved samples were then split into two aliquots. The first wasintended for determination of trace element abundance by quadru-pole ICPMS (Agilent 7500, Laboratoire Magmas et Volcans). To thisend, the aliquots were evaporated to near dryness and subsequentlydiluted in HNO3 0.4 M to reach a total dilution factor ranging from5000 (rocks) to 20,000 (gas condensates). The reaction cell (Hemode)was used to reduce interferences on masses ranging from 45 (Sc) to75 (As). The signal was calibrated externally with a reference basalticstandard (BHVO-2) dissolved as samples (rocks) or with a syntheticstandard (gas condensates). One of these two standards and pureHNO3 0.4 M were measured every 4 samples. The external reproduc-ibility of the method, as estimated by running repeatedly differentstandards (BCR-2, BIR, BEN) is b5% for most lithophile elements andb15% for most chalcophile and siderophile elements. The secondaliquot of dissolved sample was used for measurement of lithiumisotopic composition. Following evaporation and conversion of ions tochloride form (with 1 ml of HCl 6 N), samples were successfully fullydissolved in 1 ml of HCl 0.5 N, and Li was purified on AG50W-X8cation exchange resin following the two-step method described inVlastélic et al. (2009b). The lithium chemistry blank measured byquadrupole ICP-MS is less than 0.1 ng. Given the amount of lithiumrequired for isotopic analysis (N50 ng) and the loading limit of thefirst-step column (the equivalent of 50 mg of rock), the Li depletedpumices were processed through two columns and Li fractions weremerged before further purification on the second-step column. Lithiumisotopic compositions were measured by MC-ICPMS (Nu 500) at theEcole Normale Supérieure de Lyon. Samples were introduced through a

desolvator (Nu DSN) at a rate of 100 μl/min, yielding a total Li beam of4 to 6 V for 70 ng/g Li solutions. Standard operating conditions wereused (RF power of 1350W, Ar cool gas flow of 13 L/min, Ar auxiliary gasflow of 1 L/min, Ar sample gas flow of 0.8 L/min, and accelerationvoltage of 4000 volts). Lithium isotopes were measured in static modein L5 and H6 faraday cups. Measurements (60 cycles of 10s) wereperformed following an uptake-stabilization time of 90s. The washoutprocedure (300s with 0.65 N HNO3 and 300s with 0.05 N HNO3)reduced Li signal by a factor of 104. Mass fractionation was monitoredexternally with IRMM-016 standard using a sample-standard bracket-ing technique. Depending on the amount of lithium extracted, multiplemeasurements of each sample were performed. Repeated analysisof the USGS BHVO-2 standard (batch 759) over 4 years yielded δ7Li=+4.2‰±0.5 (2σ, 17 dissolutions). Repeated measurement (n=12) ofa L-SVEC solution (not processed through separation columns) duringa single session yielded δ7Li=−0.20‰ ±0.04 (S.E.), raising the pos-sibility that L-SVEC is isotopically slightly lighter than IRMM-016, aspreviously proposed (Millot et al., 2004).

4. Results

Lithium isotopic compositions and trace element concentrations ofrocks and gas condensates are reported in Table 1, together with theaverage composition of steady-state basalts and oceanites, the mostcommon types of basalts erupted at Piton de la Fournaise (a completelithium data set is provided in Supplementary Table 1). Major elementchemistry of the studied samples is reported in SupplementaryTable 2.

4.1. Silicic differentiates

The pumice samples 0704-PS and 0704-PN have total alkali andSiO2 contents (ca. 11 wt.% and 62 wt.%, respectively) that plot withinthe field of trachytes (Bachelery, in prep.). Loss-on-ignition at 1000 °Cis barely significant, indicating that water content in these differen-tiated samples is probably less than 0.5 wt.% (i.e., the maximumweight gain due to oxidation of iron). Consistently, precise determi-nation of water content by Raman spectroscopy yielded 0.2±0.1 wt.%.By comparison to steady-state basalts, the 2007 trachytic pumices areenriched inmost incompatible trace elements (Fig. 2a,b). Thedegree ofenrichment is the highest for Nb and Ta (ca. 7) and decreasesprogressively from light REE (ca. 5) to heavy REE (ca. 2). The increaseof Zr and Hf concentration with digestion strength (Table 1) suggeststhe occurrence of zircons. There are also a number of negativeanomalies,most of them (Eu, Sr, Ba, and Pb) being consistentwith low-pressure fractionation of plagioclase feldspar. Lithium, B and Cs,together with Cl and F (not reported on the plots), all beingincompatible in plagioclase, show the most pronounced anomalies,however. The lithium depletion is accompanied by strong isotopicfractionation, the two pumice samples analyzed showing extremelylight lithium compositions (−21bδ7Lib−17‰) when compared tohistorical lavas of Piton de la Fournaise (+2.7bδ7Lib+4.9‰), andterrestrial rocks in general. The olivine-bearing glass (0704-BAS)coating the pumices has trace element pattern (Fig. 2c) and Liisotopic signature (δ7Li of +4.2‰) indistinguishable from typicaloceanites.

4.2. Gas-altered basalt

Whole rock analysis (ICP-OES) of the altered basalt 9809–301revealed that it is essentially made of SiO2 (58 wt.%), CaO (11 wt.%)and a volatile component lost during sample fusing. Compared tofresh basalts, this specimen shows a marked enrichment in silica anddepletion in Al2O3, Fe2O3, and MgO, which have concentrations thatdo not exceed 2 wt.%. Inspection of the sample by SEM revealed theoccurrence of a silica-rich phase (SiO2 up to 87%vol.%) and anhydrite

28 I. Vlastélic et al. / Chemical Geology 284 (2011) 26–34

Author's personal copy

(CaSO4). Silicification in fumarolic environment was shown to be theresult of extensive leaching of major cations by acidic gas condensates(Africano and Bernard, 2000). Replicate analysis of this sample pointsto heterogeneous distribution of trace elements, the two extremepatterns being reported in Fig. 3. The alteration resulted in no loss in

Ba, Ta and Nb, small to significant (up to 50% in sample 9809-301a)loss of rare earth elements, Th, Pb, Zr, Hf and Y, and major to extreme(N90%) loss of Cs, Rb, U, Be and Li. In contrast with the pumice, lithiumloss is accompanied here by barely significant isotopic fractionation(+4.1bδ7Lib+5.1‰).

Table 1Trace element concentrations and Li isotopic compositions of rocks and gas condensates.

Main rock types Devolatilized trachytic pumice Gas altered basalt Gas condensates

SSBa Oceanitesb 0704-PS dupc 0704-PN 0704-BAS 9809-301-a 9809-301-b 9809-301-c 07.05-a 07.05-b 08.10.281 09.06.261 09.11.201

Long. (°S) 55°46.47′ 55°46.32′ 55°42.84′ 55°47.57′

Lat. (°E) 21°16.84′ 21°16.93′ 21°13.65′ 21°17.25'

Mean compositionsd T(°C) ~400° ~400° 384° 345° 325°

Material powder powder powder powder powder chip chip chip chip chip chip chip chip chip

Li 5.85 3.70 0.99 1.05 1.41 4.01 0.356 0.599 0.450 101 71.7 52.7 92.5 70.4Be 1.04 0.501 3.87 4.58 4.95 0.69 0.029 0.092 0.049 2.98 2.44 1.91 2.88 2.25B 3.7 2.0 2.0 n.d. 2.0 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d.Sc 32.6 20.7 5.61 7.47 4.28 20.39 5.76 10.0 9.95 2.80 3.27 3.16 2.20 1.67Ti 16,130 n.d. 1276 1318 1076 9040 11,566 9391 9163 818 941 762 1139 637V 300 169 10.8 12.6 3.18 172 14.2 26.2 17.8 469 564 339 1048 395Cr 225 n.d. 20.1 151 9.80 443 11.0 15.6 15.7 84.8 140 101 125 76.0Co 44.9 125 4.14 5.03 2.78 98.4 1.80 3.32 2.24 19.85 21.75 17.04 19.03 21.57Ni 92.1 1126 29.2 38.7 16.7 887 3.76 6.44 4.15 98.95 109 85.49 93.26 108Cu 104 56.6 8.47 10.2 14.6 59.6 5.74 11.8 7.33 7821 6693 5098 8446 5598Zn 110 111 32.2 36.2 9.27 104 12.7 16.7 16.2 638 523 431 600 523Ga 22.29 7.63 31.4 32.9 35.1 13.1 2.84 3.30 2.46 1.90 2.24 1.71 2.66 1.47As 0.82 n.d. 1.49 1.85 1.63 0.432 1.37 1.19 1.21 36.51 26.71 29.59 17.47 35.11Rb 17.5 8.48 52.6 55.4 58.0 9.92 2.92 3.74 3.71 361 289 280 430 348Sr 353 175 318 331 264 197 196 403 214 11.4 15.2 9.59 24.8 5.67Y 28.8 15.2 51.2 61.6 46.9 16.3 10.8 24.9 12.8 0.205 0.314 0.286 0.588 0.176Zr 193 113 456 736 267 116 107 101 102 25.1 29.8 23.4 26.9 20.3Nb 22.2 10.8 150 170 154 14.1 25.3 22.3 22.7 0.947 1.10 0.814 1.68 0.877Cd 0.065 n.d. 0.116 0.174 0.065 0.041 0.114 0.120 0.088 503 443 357 461 359In 0.089 n.d. 0.101 0.107 0.143 0.056 0.007 0.012 0.006 22.7 22.9 20.1 26.0 19.8Sn 1.83 n.d. 3.04 3.22 4.93 1.04 0.903 0.490 0.498 18.3 22.4 17.5 38.8 27.5Sb 0.065 n.d. 0.026 0.029 0.026 0.040 0.039 0.024 0.018 1.22 1.52 1.27 3.34 1.59Cs 0.261 0.123 0.042 0.040 0.017 0.143 0.016 0.033 0.030 11.2 10.2 6.63 18.3 8.14Ba 136 64.6 223 229 229 77.2 150 134 119 7.91 8.73 6.45 8.22 4.82La 19.6 9.10 89.9 102 95.6 11.7 9.29 17.2 10.3 0.084 0.117 0.124 0.437 0.069Ce 44.5 20.8 176 182 188 26.6 21.3 41.2 26.1 0.170 0.284 0.253 0.947 0.151Pr 5.94 2.79 20.1 20.7 21.7 3.52 2.83 5.40 3.57 0.022 0.042 0.033 0.127 0.022Nd 26.0 12.3 69.8 71.8 75.9 15.3 12.4 24.2 15.8 0.111 0.173 0.146 0.573 0.096Sm 6.28 3.02 12.6 13.0 13.7 3.68 2.96 6.31 3.73 0.029 0.068 0.039 0.163 0.024Eu 2.11 1.01 1.41 1.46 1.49 1.20 0.989 2.08 1.23 0.009 0.020 0.017 0.048 0.008Tb 0.977 0.487 1.69 1.80 1.68 0.576 0.465 1.03 0.516 0.006 0.007 0.009 0.024 0.004Gd 6.55 3.01 10.6 11.2 11.0 3.87 3.03 6.49 3.61 0.031 0.050 0.055 0.136 0.025Dy 5.70 2.70 9.92 10.5 9.31 3.39 2.62 5.80 2.87 0.040 0.050 0.047 0.126 0.029Ho 1.07 0.522 1.89 2.01 1.72 0.635 0.468 1.05 0.533 0.008 0.011 0.012 0.025 0.007Er 2.79 1.40 5.20 5.63 4.73 1.65 1.20 2.77 1.42 0.020 0.037 0.037 0.071 0.028Tm 0.367 0.184 0.733 0.826 0.657 0.218 0.157 0.349 0.192 0.005 0.005 0.009 0.010 0.003Yb 2.24 1.09 4.57 5.35 4.14 1.35 0.862 2.02 1.19 0.030 0.039 0.037 0.062 0.023Lu 0.308 0.154 0.63 0.73 0.58 0.186 0.123 0.267 0.167 0.005 0.007 0.007 0.009 0.004Hf 4.72 2.58 10.6 15.2 7.57 2.83 3.32 2.84 3.05 0.585 0.760 0.575 0.680 0.465Ta 1.39 0.759 8.25 8.49 8.07 0.882 1.75 1.50 1.54 b0.005 b0.005 b0.005 b0.005 b0.005W 0.295 n.d. 0.384 0.375 0.292 0.168 0.463 0.360 0.289 90.0 120 49.8 234 64.1Tl 0.046 0.023 0.102 0.104 0.088 0.026 0.036 0.030 0.018 378 396 409 624 443Pb 1.70 0.802 2.65 2.68 1.94 0.900 0.603 0.741 0.502 437 460 454 836 341Bi 0.020 n.d. 0.040 0.043 0.033 0.023 0.074 0.081 0.077 39.1 33.2 25.2 60.5 31.9Th 2.26 1.04 12.7 13.6 12.5 1.41 1.07 1.76 1.07 0.022 0.029 0.022 0.033 0.011U 0.560 0.257 2.32 2.65 0.988 0.336 0.029 0.058 0.043 0.407 0.365 0.249 0.601 0.314Cl 215 118 80 n.d. 117 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d.F 464 226 100 n.d. 60 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d.δ7Lie +3.6 +3.9 −21.0 −19.7 −17.0 +4.2 +4.4 +4.1 +5.1 −1.3 −1.3 −2.8 −1.4 −1.82σ ± 0.8 ± 0.8 − ± 0.4 – ± 0.0 – – ± 0.6 ± 0.2 ± 0.2 ± 0.4 ± 0.4n 31f 21f 2g 2g 10g 2g 3g 3g 4g

All concentration in g/g (ppm).n.d.: not determined.a SSB: Steady State Basalts. Mean composition inferred from 1931 to 2007 lavas with MgOb9%wt.b Océanites: Lavas rich in cumulative olivine. Mean composition inferred from 1931 to 2007 lavas with MgON20%wt.c PARR bomb dissolution.d A complete lithium data set is provided in Supplementary Table 1. Source of trace element data: Vlastélic et al. (2005, 2007) and unpublished data.e δ7Li=((7Li /6Li)sample / (7Li/6Li)IRMM−1)⁎1000.f number of sample analyzed.g Repeated analysis of the same dissolution.

29I. Vlastélic et al. / Chemical Geology 284 (2011) 26–34

Author's personal copy

4.3. Gas condensate

Analysis of the gas condensates by SEM identified K–Na sulfateswith K/Na suggesting the main occurrence of aphthitalite ((K,Na)3Na(SO4)2). The distribution of trace elements in gas condensates isshown on Fig. 4 where the elements are sorted according to theirenrichment relative to lavas. Enrichment factor (EF) is defined as:

EFX = X=XRð Þgas condensate = X=XRð Þlava ð1Þ

where X is the element of interest and XR the refractory element ofreference. Beryllium was chosen here because of its low volatility andlow abundance in lavas (Moune et al., 2006). The gas condensates aredepleted (EFb1) in rare earth elements and high field strengthelements and enriched (EFN1) in alkali and chalcophile elements,which pattern is consistent with element volatility at mafic volcanoes(Toutain et al., 1990; Rubin, 1997). Lithium (5bEFb6), like Zn, Sn, Rb,Sb, and As, displays amoderately volatile behavior during degassing ofthe April 2007 lava flow. The condensates have light and ratheruniform Li isotopic signatures (−2.8bδ7Lib−1.3‰), the lowest δ7Li(−2.8‰) being measured in the sample (08.10.281) with the lowestLi content (52.7 ppm, against N70 ppm in other samples).

5. Discussion

5.1. Origin of lithium depletion and isotopic fractionation in silicicdifferentiates

The trachytic pumices display very low lithium contents andextremely light Li isotope compositions that have not yet beenreported in fresh volcanic rocks. Such a combination rules out diffusiveinflux of lithium,which best accounts for very low δ7Li signatures in Li-rich rocks only (Rudnick and Ionov, 2007; Marschall et al., 2007).Hydrothermal alteration,which preferentially removes 7Li from solids,could yield such light signatures, but only in low-temperatureenvironments (Wunder et al., 2007; Millot et al., 2010), which isinconsistentwith themagmatic origin of pumices. Independently, the

TaLa

CeBe

PrPb

NaNd

SrSm

ZrHf

EuGd

TbDy Y

Ho ErTm

LiYb

Lu

Pb Sr Eu Li

0704-PS

0704-PN

0704-BAS

0.1

1

10

0.01

0.1

1

10

0.1

1

10

CsRb

BaTh

UK

Nb

Con

cent

ratio

n / S

SB

Con

cent

ratio

n / S

SB

Con

cent

ratio

n / O

cean

ite

Cs Ba Ua

b

c

B

B

Fig. 2. Trace element patterns of the April 2007 trachytic pumice (a,b) and its basalticcoating (c). Concentrations are normalized to the average concentrations of steady-state basalts (a,b) and oceanites (c), both being given in Table 1. Vertical lines are forelements showing negative anomalies.

Con

cent

ratio

n / S

SB

0.01

0.

1

1

10

9809-301a

Con

cent

ratio

n / S

SB

0.01

1

0.1

10

CsRb

BaTh

UNb

TaLa

CeBe

PrPb

NdSr

SmZr

HfEu

GdTb

Dy YHo Er

TmLi

YbLu

9809-301b

a

b

Fig. 3. Trace element patterns of the gas-altered basalts from the 1998 Piton Kapor.Concentrations are normalized to the average concentrations of steady-state basaltsgiven in Table 1. Vertical lines indicate elements that underwent the largest depletion.Amongst the three samples analyzed (see Table 1), the two extreme patterns are shownhere.

0.001

0.01

0.1

1

10

100

1 000

10 000

Ce

Pr

Er

Ho

Y

Th

Dy

Tb

Sm

Eu

Gd

La

Nd

Yb

Tm

Lu

Sr

Nb

Ba

Ti

Sc

Ga

Hf

Zr

Co

Cr

UNi

V

Be

ZnSn

Rb

Sb

As

Cs

Cu

In

Pb

W

Bi

CdTl

Enr

ichm

ent F

acto

r

Li

Gas condensates

Fig. 4. Enrichment factors of trace elements in 2008–2009 gas condensates. Enrichmentfactors are calculated according to Eq. (1), using beryllium as reference. Normalizationis made to the average composition of steady-state basalts (see Table 1). Elements aresorted according to their mean enrichment factor. For each element, the range ofvariation of the enrichment factor within the five samples analyzed is shown.

30 I. Vlastélic et al. / Chemical Geology 284 (2011) 26–34

Author's personal copy

mantle-like oxygen isotope signature of pumices (Bachèlery et al.,in prep.) does not support input of meteoric water. Fluxing byvapors released by melt degassing deeper in the plumbing system(Blundy et al., 2010) is another possibility, which could alsoexplain the less dramatic loss of uranium, which is fluid-mobile whenoxidized but non-volatile. Although this possibility cannot be definitelyruled out, our attempt to test it suggests that high-temperature leachingof lithium by magmatic gas does not significantly fractionate 7Li/6Li. Ontheother hand, the concomitant loss of B, Cs, Cl andF suggests that Liwaslost during devolatilization of the magma, as it underwent extensivecrystallization at shallow depth. Indeed, the water content of pumicesamples (0.2 wt.%) is unexpectingly low given the largemass fraction ofanhydrous crystals removed. Both major and trace elements areconsistent with a crystallizing mineral assemblage made dominantlyof clinopyroxene, plagioclase, olivine, Fe–Ti oxides, apatite, with noevidence for hydrous minerals such as amphibole or phlogopite.Elements (e.g., Th, Nb) that are highly incompatible in these mineralsare strongly enriched (a factor of 6.2 on average) in the trachytic sampleswith respect to steady-state basalts. Assuming these elements do notenter the crystallizing phases (bulk DS–M ~0), the fraction ofmelt havingcrystallized during the basalt–trachyte differentiation is estimated at ca.84%. From the water content of primitive melt (0.7–1.0%wt% accordingto Bureau et al. (1999)), and assuming that H2O is as incompatible asCe (Michael, 1995), awater content of 2.9–4.1 wt.%wouldbeexpected inthe trachytic melt if it evolved in a closed system. The residual watercontent of pumices (0.2±0.1 wt.%) suggests that between 2.6 and4.0 wt.% of water left the system and entrained a major fraction of theinitial budget of lithium.

5.2. Modeling lithium loss and isotopic fractionation during coupleddegassing and crystallization

To explain the composition of the dehydrated silicic differentiates,we resorted to the open system degassing-crystallization modelpreviously used for chlorine (Villemant and Boudon, 1999; Villemantet al., 2008). However, the model has to be modified to take intoaccount that, unlike chlorine, lithium significantly partitions intocrystallizing phases. The mass balance equations in the system madeof melt (M), crystals (S) and vapor (V) are:

dmS + dmV = −dmM ð2Þ

and, for lithium,

CS:dmS + CV :dmV = −d CM :mMð Þ ð3Þ

where C is the concentration of lithium. Following Villemant andBoudon (1999), we assume that the masses of crystallizing melt andexsolving vapor are proportional:

dmS = k:dmV ð4Þ

and, from Eq. (2), dmV=−dmM/(1+k) and dmS=−dmM/(1+1/k).Substituting in Eq. (3) yields:

dCMCM

=dmM

mMβ−1ð Þ with β =

DV−M + kDS−M

k + 1ð5Þ

where DV–M and DS–M are the vapor–melt and crystal–melt partitioncoefficients of lithium, respectively. We then set dmM/mM=dF/F, Fbeing the mass fraction of residual melt expressed as:

F = 1− 1 + kð ÞΔH20 ð6Þ

where ΔH20 is the mass fraction of vapor lost. Integrating Eq. (5) yieldsthe expression of lithium concentration in the residual melt (CM) as afunction of the initial melt content (C0M):

CM = CM0 :F β−1ð Þ ð7Þ

Note that if DS–M~0, β=DV–M/(1+k) as inferred for chlorine(Villemant et al., 2008). Considering lithium isotopes as individualspecies, Eq. (7) may also be used to compute the evolution of lithiumisotopic ratio of the residual melt:

7Li6Li

!M

=7Li6Li

!M

0

:Fk DS−M

7 −DS−M6ð Þ + DV−M

7 −DV−M6

1 + k

� �ð8Þ

where (7Li /7Li)0 is the initial isotopic ratio, and subscripts 7 and 6denote for 7Li and 6Li species. If lithium isotopes partition equally intothe crystallizing phases (D7

S–M=D6S–M):

7Li6Li

!M

=7Li6Li

!M

0

:FDV−M6

1 + k: αV−M−1ð Þ� �

ð9Þ

with

αV−M =7Li=6Li� �V7Li=6Li� �M =

M6 =M7ð Þ CV7 = CV

6

� �M6 =M7ð Þ CM

7 = CM6

� � =DV−M7

DV−M6

and

DV−M6 = DV−M

:1 + M7 =M6ð Þ 7Li=6Li

� �M1 + αV−M M7 =M6ð Þ 7Li=6Li

� �M0B@

1CA

where αV–M is the coefficient of isotopic fractionation between vaporand melt and M the molar mass. When there is only degassing (k=0)and DV–MNN1, it is convenient to use the simplified Rayleigh equationdescribing the isotopic evolution as a function of C/C0 and α:

7Li6Li

!M

=7Li6Li

!M

0

:F DV−M−1ð Þ αV−M−1ð Þ =7Li6Li

!M

0

:CM

CM0

! αV−M−1ð Þð10Þ

We stress that Eq. (10) only applies when elements stronglypartition into the vapor phase, and is thus not valid when elementsenter the crystalline assemblage significantly. The evolution of Licontent and Li isotopic composition of the residual melt was modeledconsidering that crystallization and degassing occurred eithersuccessively (case A) or simultaneously (case B). In both cases, thestarting composition is assumed to be that of steady-state basalts(Li=5.85 ppm, δ7Li=+3.6‰, see Table 1). It is also assumed that Liisotopes do not fractionate significantly during crystallization(Tomascak et al., 1999; Schuessler et al., 2009). In case A, Li is assumedto behave like heavy rare earth elements during low-pressurecrystallization (Ryan and Langmuir, 1987). Its concentration in theundegassed trachyticmelt is thus calculatedby considering smooth SSB-normalized trace element pattern (Li=LiSSB.[Tm/TmSSB+Yb/YbSSB]/2=11.8 ppm), which, in turn, implies a solid-melt partition coefficientof ca. 0.6. As discussed above, the differentiated melt must have lostbetween 2.6 and 4.0 wt.% of water during degassing. In case B, the initialwater content is assumed to be that of primitivemelts (~1 wt.%). Thus, kon the order of 100 is inferred from the fraction of vapor lost (~0.8 wt.%)and the fraction of melt having crystallized (~84 wt.%). Lithium lossduring degassing was modeled using Eq. (7) (Fig. 5), setting k=0 (nocrystallization) in caseA. The low lithiumcontent of thepumice requires60bDV–Mb95 in case A and β~1.9 in case B (DV–M=135 for DS–M=0.6

31I. Vlastélic et al. / Chemical Geology 284 (2011) 26–34

Author's personal copy

and k=100). Note that if β=1 (DV–M=k(1−DS–M)+1), thecompeting effects of crystallization and degassing cancel out andthe Li content of the melt remains constant. Taking 0.6 as an upperbound for DS–M, it can be estimated that lithium content in meltdecreases only if DV–MN~40. The values of DV–M required to explainLi depletion in pumice are larger than those estimated for thepartitioning of lithium between fluids and melts, although trulyrelevant data lack in the literature. For granitic–rhyolitic composi-tions, DV–M was shown to increase from 0 to 13 with decreasingpressure from 400 MPa to 50 MPa (London et al., 1988; Webster etal., 1989), while 1bDV–Mb30 was inferred by mass balancefollowing degassing experiments (Koga et al., 2008). The elevatedDV–M value inferred in this study could be a consequence of thebasalt–trachyte differentiation process, which generally involveshigh concentrations of water (Ringwood, 1959). However, thisdoes not explain the scarcity of the Li signature of the 2007trachyte. The most comparable case, even if Li loss (~20%) is lessextreme, is that of post-eruptive internal differentiation of thethick lava flow of Rishiri Volcano (Japan) (Kuritani and Nakamura,2006). Based on these observations, it is suggested that DV–M couldbe very elevated during low-pressure, vapor-saturated crystalliza-tion, in particular if there is a long interaction (possibly 20 years forthe 2007 trachyte) between exsolved vapor and melt. Anotherpossibility is that fluxing by vapors released by deeper meltsaccompanied degassing (Blundy et al., 2010) and enhanced lithium

entrainment into the vapor phase. In this scenario, our calculatedDV–M values are overestimated.

Fractionation of Li isotopes during degassing was modeledusing Eq. (10) in case A (DV–M≫1) and Eq. (9) in case B assuming(D6)V–M~DV–M (since αV–M ~1) (Fig. 6). The end-member pumicecomposition (Li=1 ppm and δ7Li=−21‰) is consistent with αV–M of1.0100 (case A) and 1.0099 (case B). The similarity ofαV–M in both casesis at first glance surprising given the difference in the startingcomposition. It is consistent with αV–M being dependent only on theamount of lithium lost, which is the same in cases A and B. These valuesof αV–M are unexpectingly large (isotopic fractionation of ~10‰) for ahigh-temperature process. Large fractionations (1.005bαV–Mb1.015)during magma degassing were also proposed to explain Li isotopesystematics in 2003 ashes from Stromboli, suggesting that degassingfractionates lithium isotopesmore efficiently than fluid-rock interaction(Schiavi et al., 2010).

5.3. Lithium isotopic composition of volcanic vapor

The isotopic composition of Piton de la Fournaise volcanic gas hasnot been measured, but can reasonably be inferred from αV–M

calculated for the dehydrated trachytic melt. The instantaneouscomposition of the vapor is given by:

7Li6Li

!V

= αV−M7Li6Li

!M

ð11Þ

bcrystallization + degassing

0

2

4

6

8

0 20 40 60 80

0 20 40 60 800 1 2 3 4

Li in pumice

0

2

4

6

8

10

12

14

Li (

pp

m)

Li (

pp

m)

crystallizationDs-m=0.62

degassing

Fcrystal (%)

Fcrystal (%)

Fvapor (%)

Li in pumic

a

e

Dv-m=60

Dv-m=95

β=1.9

β=1.0

β=0.8

β = Dv-m + k.Ds-m

1 + k

Fig. 5. Modeling Li content in residual melt during crystallization and degassing. Twocases are considered: (A) crystallization occurs first and degassing occurs subsequently,and (B) crystallization and degassing occur concomitantly. In both cases, the average Licontent of steady-state basalts (5.85 ppm) is used for the starting composition. Thetarget composition is that of the trachytic pumice (1 ppm Li). In case A, the Li content ofmelt (11.8 ppm) after differentiation and before degassing, is estimated assuming thatLi displays no anomaly along trace element patterns when it is positioned between Tmand Yb. The fraction of melt having crystallized (Fcrystal) is inferred from the enrichmentof the most incompatible elements, while the fraction of vapor lost (Fvapor) isconstrained from the initial (assumed) and final (measured) water content. DS–M andDV–M are the crystal-melt and vapor-melt partition coefficients, respectively. FollowingVillemant and Boudon (1999), k is the mass ratio of crystal removed to vapor exsolved.See text for details.

δδ7L

i (m

elt)

δ7L

i vap

or)

-15

-10

-5

0

5

10

15

0 2 4 6 8 10 12 14

Li in melt (ppm)

B

A

vapor instantaneous

vapor accumulated

-25

-20

-15

-10

-5

0

5

10

0 2 4 6 8 10 12 14

crystallization

degassing(α=1.0100)

Undegassedtrachytic melt

Pumice (basaltic glass coating)

Pumice (trachytic glass)

Common lavas (SSB + oceanites)

AverageSSB

α=(7Li/6Li)vapor/(7Li/6Li)melt

A

Bcrystallization+ degassing(α=1.0099)

Pumice (bulk analysis)

Fig. 6. Lithium isotopic composition of the melt (upper panel) and vapor (lower panel)plotted against Li content in the residual melt. A and B refer to the two crystallization-degassing scenarios described in Fig. 5. Upper panel: all points are data points, with theexception of the undegassed trachtytic composition which is calculated as explained intext. The straight dashed line, onwhich plot the pumice samples (0704-PS and 0704-PN),the coating glass (0704-BAS) as well as a bulk analysis not reported in Table 1(Li=3.36 ppm; δ7Li=−1.8‰±0.3), suggests that the rock ismade of two homogeneouscomponents (basaltic and trachytic). Curved lines showtheevolutionofδ7Li in the residualmelt during degassing (Eqs. (10) and (9) for cases A and B). Lower panel: vaporcomposition is calculated for cases A and B using Eq. (12) and Eq. (13), respectively).

32 I. Vlastélic et al. / Chemical Geology 284 (2011) 26–34

Author's personal copy

The average isotopic ratio of the accumulated (ACC) vapor isobtained by integrating Eq. (11). Knowing that the average value of afunction f(x) that is continuous over the interval a≤x≤b is 1/(b–a).∫f(x)dx, we infer:

7Li6Li

!V

ACC

=7Li6Li

!M

0

:1− CM

=CM0

� �αV−M

1− CM = CM0

� � ð12Þ

and

7Li6Li

!V

ACC

=7Li6Li

!M

0

:αV−M 1−F ε + 1ð Þ

� �ε + 1ð Þ 1−Fð Þ with ε =

DV−M6 αV−M−1

� �1 + k

ð13Þ

for cases A and B, respectively. During degassing of a finite batch ofmagma, the lithium isotopic composition of the accumulated vaporvaries widely, from ~13.5‰ in the first vapors down to about ~5‰ inthe case of extreme degassing, as for the April 2007 pumice (Fig. 6).The 7Li /6Li ratio of the gas phase will subsequently increase duringpartial condensation, as suggested by the isotopically light Licomposition of the April 2007 fumarolic condensates. While thecoefficient of isotopic fractionation during condensation (αCond-V) isconstrained to be less than unity, it cannot be determined preciselybecause both the isotopic composition of the condensing gas and thefraction of gas condensed are poorly known. To explain thecomposition of the April 2007 gas condensates (average of −1.7‰),we calculated thatαCond-V≤0.99 is required if the condensing gas hadδ7LiN8‰, or if more than 50% of the vapor condensed. If the vapor hadδ7Li=13.5‰, then αCond-V≤0.985.

The trachytic pumice recorded an extreme process that is notrepresentative of the degassing pattern of Piton de la Fournaisemagmas,however. Steady state basalts display rather uniform Li content andisotopic composition (Supplementary Table 1), suggesting that lithiumloss in normal degassing mode is of only a few percents and,consequently, that those volcanic emanationshave very heavy Li isotopicsignatures. The slightly heavier Li signature of oceanites (average of+3.9‰, Table 1) is due to the incorporationof cumulative olivine crystalshaving distinctly heavy composition (δ7Li=+5.2±1.4‰(2σ, N=10),Supplementary Table 1), whichmost likely reflect mineral-melt isotopicfractionation (Seitz et al., 2004) rather thana lessdegassed signature.Ourattempt to find a distinctive lithium signature in naturally quenchedsamples (Pele's hairs) failed (Li=5.6±0.4 ppm, δ7Li=+3.4±0.3‰(2σ, N=4), Supplementary Table 1). However, amongst lavas free ofcumulative olivine, the highest δ7Li (+4.0 to 4.5‰) coincide with theeruption of more primitive, less degassed magmas (as during the 1998Hudson eruptionor during the paroxysmal phase of April 2007 eruption)suggesting that degassing may influence whole-rock Li signature to ameasurable degree. We calculated that decreasing δ7Li from +4.5 to+3.6‰ (average of SSB) by degassing isotopically heavy lithium (αV–M

of 1.010) is consistent with 9% loss of lithium with δ7Li of +14.1‰.

6. Concluding remarks

The preliminary results obtained show that 7Li partitions preferen-tially into the vapor phase during low-pressure, vapor-saturatedcrystallization. This strongly suggests that Li isotopic fractionation iscontrolled by chemical rather than by kinetic effects. The composition ofthe degassed trachytic pumice places strong constraints on themagnitude of the vapor-melt isotopic fractionation (αV–M of 1.010)because the pre-degassing composition is relatively well known. Sincethe fraction of Li volatilized in normal degassing mode is very small, afewpercent at themost, a very heavy Li isotopic signature (δ7LiN+13‰)is expected in volcanic emanations. In addition, the Li composition of thevapor may subsequently increase as isotopically light lithium conden-sates. These findings and expectations are summarized in Fig. 7.

This behavior of lithium isotopes is consistent with that inferredfor boron isotopes, but in magmatic systems only (You, 1994; Gillis etal., 2003; Kuritani and Nakamura, 2006). It is possible that Li, as with B(Rose-Koga et al., 2006), obeys somewhat different laws duringevaporation and condensation processes in the ocean–atmospheresystem. By comparison with heavier isotopic systems, Li isotopesseem to behave rather like Zn isotopes during degassing (preferentialevaporation of heavy isotopes (Toutain et al., 2008)), but rather followTl isotopes during condensation (preferential removal of lightisotopes (Baker et al., 2009)). However, because these isotopicsystems were studied independently in different volcanic settings, itis unclear whether the contrasting isotope behaviors are fundamentalor just reflect distinct degassing contexts. A more comprehensiveview of trace isotopes behavior during degassingwill probably requiremulti-isotopic approaches.

Supplementarymaterials related to this article can be found onlineat doi:10.1016/j.chemgeo.2011.02.002.

Acknowledgements

The authors are grateful to A.J.R. Kent and an anonymous reviewerfor their constructive comments, and to R. Rudnick for editorialhandling. E. Rose-Koga, G. Falco and K. Koga are thanked fordiscussions. Thanks also to N. Bolfan-Casanova, C. Bosq, K. David, J.-L.Devidal, J.-L. Piro (LMV, Clermont-Ferrand), C. Douchet (LST, Lyon) andC. Le Losq (IPGP, Paris) for technical assistance in the lab. Thanks toSARM (CRPG, Nancy) and L. Bouvier (LaMP, Clermont-Ferrand) forproviding major element analyses. This study benefited financialsupport from the Institut National des Sciences de l'Univers (ST,A02009 and AO2010).

References

Africano, F., Bernard, A., 2000. Acidic alteration in the fumarolic environment of Usuvolcano, Hokkaido, Japan. Journal of Volcanology and Geothermal Research 97,475–495.

Albarède, F., 1993. Residence time analysis of geochemical fluctuations in volcanicseries. Geochimica et Cosmochimica Acta 57, 615–621.

Albarède, F., Luais, B., Fitton, G., Semet, M., Kaminski, E., Upton, B.G.J., Bachèlery, P.,Cheminée, J.L., 1997. The geochemical regimes of Piton de la Fournaise volcano(Réunion) during the last 530 000 years. Journal of Petrology 38, 171–201.

Baker, R.G.A., Rehkämper, M., Hinkley, T.K., Nielsen, S.G., Toutain, J.-P., 2009.Investigation of thallium fluxes from subaerial volcanism — implication for thepresent and past mass balance of thallium in the oceans. Geochimica etCosmochimica Acta 73, 6340–6359.

Measured

Inferred

-25

-20

-15

-10

-5

0

5

10

δ7L

i (‰

)

15

Gas condensates300-400°C

Common lavas

Vapor-differentiatedlava

Cooled gas ?

20

25

Exsolved fluids High-temperaturegas

Degassing Condensation

Fig. 7. Summary of lithium isotope fractionation during magma degassing and gascondensation. Measured compositions are shown by squares while inferred composi-tions are shown by circles.

33I. Vlastélic et al. / Chemical Geology 284 (2011) 26–34

Author's personal copy

Beck, P., Barrat, J.A., Chaussidon, M., Gillet, P., Bohn, M., 2004. Li isotopic variations insingle pyroxenes from the Northwest Africa 480 shergottite (NWA 480): a record ofdegassing of Martian magmas ? Geochimica et Cosmochimica Acta 68, 2925–2933.

Berlo, K., Blundy, J., Turner, S., Cashman, K., Hawkesworth, C., Black, S., 2004.Geochemical precursors to volcanic activity at Mount St. Helens, USA. Science306, 1167–1169.

Blundy, J., Cashman, K.V., Rust, A., Witham, F., 2010. A case for CO2-rich arc magmas.Earth and Planetary Sciences Letters 290, 289–301.

Bureau, H., Métrich, N., Semet, M., Staudacher, T., 1999. Fluid-magma decoupling in ahot-spot volcano. Geophysical Research Letters 23, 3501–3504.

Giletti, B.J., Shanahan, T.M., 1997. Alkali diffusion in plagioclase feldspar. ChemicalGeology 139, 3–20.

Gillis, K.M., Coogan, L.A., Chaussidon, M., 2003. Volatile element (B, Cl, F) behaviour inthe roof of an axial magma chamber from the East Pacific Rise. Earth and PlanetarySciences Letters 213, 447–462.

Helz, R.T., Kirschenbaum, H.K., Marinenko, J.W., Qian, R., 1994. Whole-rock analyses ofcore samples from the 1967, 1975, 1979 and 1981 drillings of Kilauea Iki lava lake,Hawaii. U.S. Geological Survey Open-File Report 94–684. .

Jambon, A., Semet, M.P., 1978. Lithium diffusion in silicate glasses of albite, orthoclase,and obsidian composition: an ion-microprobe determination. Earth and PlanetarySciences Letters 37, 445–450.

Kent, A.J.R., Blundy, J., Cashman, K.V., Cooper, K.M., Donnelly, C., Pallister, J.S., Reagan,M., Rowe, M.C., Thornber, C.R., 2007. Vapor transfer prior to the October 2004eruption of Mount St. Helens, Washington. Geology 35, 231–234.

Koga, K.T., Cluzel, N., Rose-Koga, E., Laporte, D., Shimizu, N., 2008. Experimentaldemonstration of lithium-boron depletion during magma degassing. EOS.Transactions of the American Geophysical Union 89 (53) Fall Meeting Supplement,Abstract V31B-2143.

Kuritani, T., Nakamura, E., 2006. Elemental fractionation in lavas during post-eruptivedegassing: evidence from trachytic lavas, Rishiri Volcano, Japan. Journal ofVolcanology and Geothermal Research 149, 124–138.

Le Losq, C., Neuville, D.R., Morretti, R., Roux, J., Behrens, H. in press Determination ofwater content in silicate glasses using Raman spectrometry: implications formaterials and volcanic activities. American Mineralogist.

London, D., Hervig, R.L., Morgan, G.B., 1988. Melt-vapor solubilities and elementalpartitioning in peraluminous granite-pegmatite systems: experimental results withMacusani glass at 200 MPa. Contribution to Mineralogy and Petrology 99, 360–373.

Marschall, H.R., Pogge von Strandmann, P.A.E., Seitz, H.-M., Elliott, T., Niu, Y., 2007. Thelithium isotopic composition of orogenic eclogites and deep subducted slabs. Earthand Planetary Sciences Letters 262, 563–580.

Martin, E., Sigmarsson, O., 2007. Low-pressure differentiation of tholeiitic lavas asrecorded in segregation veins from Reykjanes (Iceland), Lanzarote (Canary Islands)andMasaya (Nicaragua). Contributions to Mineralogy and Petrology 154, 559–573.

Michael, P., 1995. Regionally distinctive source of depleted MORB: evidence from traceelements and H2O. Earth and Planetary Sciences Letters 131, 301–320.

Millot, R., Guerrot, C., Vigier, N., 2004. Accurate and high-precision measurement oflithium isotopes in two reference materials by MC-ICP-MS. Geostandards andGeoanalytical Research 28, 153–159.

Millot, R., Scaillet, B., Sanjuan, B., 2010. Lithium isotopes in island arc geothermalsystems: Guadeloupe, Martinique (French West Indies) and experimentalapproach. Geochimica et Cosmochimica Acta 74, 1852–1871.

Moune, S., Gauthier, P.-J., Gislason, S.R., Sigmarsson, O., 2006. Trace element degassingand enrichment in the eruptive plume of the 2000 eruption of Hekla volcano,Iceland. Geochimica et Cosmochimica Acta 70, 461–479.

Peltier, A., Famin, V., Bachèlery, P., Cayol, V., Fukushima, Y., Staudacher, T., 2008. Cyclicmagma storages and transfers at Piton de la Fournaise volcano (La Réunionhotspot) inferred from deformation and geochemical data. Earth and PlanetarySciences Letters 270, 180–188.

Peltier, A., Bachèlery, P., Staudacher, T., 2009. Magma transport and storage at Piton dela Fournaise (La Réunion) between 1972 and 2007: a review of geophysical andgeochemical data. Journal of Volcanology and Geothermal Research 184, 93–108.

Richter, D.H., Moore, J.G., 1966. Petrology of the Kilauea Iki lava lake, Hawaii. U.S.Geological Survey Professional Paper 537B. .

Richter, F.M., Davis, A.M., DePaolo, D.J., Watson, E.B., 2003. Isotope fractionation bychemical diffusion between molten basalt and rhyolite. Geochimica et Cosmochi-mica Acta 67, 3905–3923.

Ringwood, A.E., 1959. Genesis of the basalt–trachyte association. Beiträge zurMineralogie und Petrographie 6, 346–351.

Rose-Koga, E.F., Sheppard, S.M.F., Chaussidon, M., Carignan, J., 2006. Boron isotopiccomposition of atmospheric precipitations and liquid-vapour fractionations.Geochimica et Cosmochimica Acta 70, 1603–1615.

Rowe, M.C., Kent, A.J.R., Thornber, C.R., 2008. Using amphibole phenocrysts to trackvapor transfer during magma crystallization and transport: an example from

Mount St. Helens, Washington. Journal of Volcanology and Geothermal Research178, 593–607.

Rubin, K., 1997. Degassing of metals and metalloids from erupting seamount and mid-ocean ridge volcanoes: observations and predictions. Geochimica et CosmochimicaActa 61, 3525–3542.

Rudnick, R.L., Ionov, D.A., 2007. Lithium elemental and isotopic disequilibrium inminerals from peridotite xenoliths from far-east Russia: product of recent melt/fluid-rock reaction. Earth and Planetary Sciences Letters 256, 278–293.

Ryan, J.G., Langmuir, C.H., 1987. The systematics of lithium abundances in youngvolcanic rocks. Geochimica et Cosmochimica Acta 51, 1727–1741.

Schiavi, F., Kobayashi, K., Moriguti, T., Nakamura, E., Pompilio, M., Tiepolo, M., Vannuci,R., 2010. Degassing, crystallization and eruption dynamics at Stromboli: traceelement and lithium isotopic evidence from 2003 ashes. Contributions toMineralogy and Petrology 159. doi:10.1007/s00410-009-441-2.

Sakuyama, M., Kushiro, I., 1979. Vesiculation of hydrous andesitic melt and transport ofalkalies by separated vapor phase. Contributions to Mineralogy and Petrology 71,61–66.

Schuessler, J.A., Schoenberg, R., Sigmarsson, O., 2009. Iron and lithium isotopesystematics of the Hekla volcano, Iceland — evidence for Fe isotope fractionationduring magma differentiation. Chemical Geology 258, 78–91.

Seitz, H.-M., Brey, G.P., Lahaye, Y., Durali, S., Weyer, S., 2004. Lithium isotopic signaturesof peridotite xenoliths and isotopic fractionation at high temperature betweenolivine and pyroxenes. Chemical Geology 212, 163–177.

Sigmarsson, O., Condomines, M., Bachèlery, P., 2005. Magma residence time beneaththe Piton de la Fournaise volcano, Réunion Island, from U-series disequilibria. Earthand Planetary Science Letters 234, 223–234.

Sinton, J.M., Byerly, G.R., 1980. Silicic differentiates of abyssal oceanic magmas:evidence for late-magmatic vapor transport of Potassium. Earth and PlanetarySciences Letters 47, 423–430.

Staudacher, T., Ferrazzini, V., Peltier, A., Kowalski, P., Boissier, P., Catherine, P., Lauret, F.,Massin, F., 2009. The April 2007 eruption and the Dolomieu crater collapse, twomajor events at Piton de la Fournaise (La Réunion Island, Indian Ocean). Journal ofVolcanology and Geothermal Research 184, 126–137.

Tomascak, P.B., Tera, F., Helz, R.T., Walker, R.J., 1999. The absence of lithium isotopefractionation during basalt differentiation: new measurements by multicollectorICP-MS. Geochimica et Cosmochimica Acta 63, 907–910.

Toutain, J.-P., Aloupogiannis, P., Delorme, H., Person, A., Blanc, P., Robaye, 1990. Vapordeposition of trace elements from degassed basaltic lava, Piton de la Fournaisevolcano, Reunion Island. Journal of Volcanology and Geothermal Research 40,257–268.

Toutain, J.-P., Sonke, J., Munoz, M., Nonell, A., Polvé, M., Viers, J., Freydier, R., Sortino, F.,Joron, J.-L., Sumarti, S., 2008. Evidence for Zn isotopic fractionation at Merapivolcano. Chemical Geology 253, 74–82.

Vernet, M., Marin, L., Boulmier, S., Lhomme, J., Demange, J.C., 1987. Dosage du Fluor etdu Chlore dans les matériaux géologiques y compris les échantillons hyperalumi-neux. Analysis 15, 490–498.

Villemant, B., Boudon, G., 1999. H20 and halogen (F, Cl, Br) behaviour during shallowmagma degassing processes. Earth and Planetary Science Letters 168, 271–286.

Villemant, B., Mouatt, J., Michel, A., 2008. Andesitic magma degassing investigatedthrough H2O vapour-melt partitioning of halogens at Soufrière Hills Volcano,Montserrat (Lesser Antilles). Earth and Planetary Science Letters 269, 212–229.

Vlastélic, I., Staudacher, T., Semet, M., 2005. Rapid change of lava composition from1998 to 2002 at Piton de la Fournaise (Réunion) inferred from Pb isotopes and traceelements: evidence for variable crustal contamination. Journal of Petrology 46,79–107.

Vlastélic, I., Peltier, A., Staudacher, T., 2007. Short-term (1998–2006) fluctuations of Pbisotopes at Piton de la Fournaise volcano (Réunion Island): Origins and constraintson the size and shape of the magma reservoir. Chemical Geology 244, 202–220.

Vlastélic, I., Deniel, C., Bosq, C., Télouk, P., Boivin, P., Bachèlery, P., Famin, V., Staudacher,T., 2009a. Pb isotope geochemistry of Piton de la Fournaise historical lavas. Journalof Volcanology and Geothermal Research 184, 63–78.

Vlastélic, I., Koga, K., Chauvel, C., Jacques, G., Télouk, P., 2009b. Survival of lithiumisotopic heterogeneities in the mantle supported by HIMU-lavas from RurutuIsland, Austral Chain. Earth and Planetary Science Letters 286, 456–466.

Webster, J.D., Holloway, J.R., Hervig, R.L., 1989. Partitioning of lithophile trace elementsbetween H2O and H2O+CO2 fluids and topaz rhyolite melt. Economic Geology 84,116–134.

Wunder, B., Meixner, A., Romer, R.L., Feenstra, A., Schettler, G., Heinrich, W., 2007.Lithium isotope fractionation between Li-bearing staurolite, Li-mica and aqueousfluids: an experimental study. Chemical Geology 238, 277–290.

You, C.-F., 1994. Comment on “Boron content and isotopic composition of oceanicbasalts: geochemical and cosmochemical implications” by M. Chaussidon and A.Jambon. Earth and Planetary Science Letters 128, 727–730.

34 I. Vlastélic et al. / Chemical Geology 284 (2011) 26–34


Recommended