• Aucun résultat trouvé

Viscoelastic relaxation: A mechanism to explain the decennial large surface displacements at the Laguna del Maule silicic volcanic complex

N/A
N/A
Protected

Academic year: 2021

Partager "Viscoelastic relaxation: A mechanism to explain the decennial large surface displacements at the Laguna del Maule silicic volcanic complex"

Copied!
50
0
0

Texte intégral

(1)

OATAO is an open access repository that collects the work of Toulouse

researchers and makes it freely available over the web where possible

Any correspondence concerning this service should be sent

to the repository administrator:

tech-oatao@listes-diff.inp-toulouse.fr

This is an author’s version published in:

http://oatao.univ-toulouse.fr/27542

To cite this version:

Novoa, Camila and Rémy, Denise and Gerbault, Muriel and Baez, Juan C. and

Tassara, Andrés and Cordova, Loreto and Cardona, Carlos and Granger, M.

and Bonvalot, Sylvain and Delgado, Francisco Viscoelastic relaxation: A

mechanism to explain the decennial large surface displacements at the Laguna

del Maule silicic volcanic complex. (2019) Earth and Planetary Science Letters,

521. 46-59. ISSN 0012-821X

(2)

Viscoelastic relaxation: A mechanism to explain the decennial large

surface displacements at the Laguna del Maule silicic volcanic complex

C.

Novoa a,b,*.

O.

Remy a,

M.

Gerbault a,

J.C.

Baez c,

A.

Tassara d,

L.

Cordova e.

C.

Cardona e.

M.

Grangera,

S.

Bonvalota,

F.

Delgado f

Gf17UMR5563 (UPS, CNRS, /RD, CNES); Obs. Midi-Pyrénées, Université P. Sabatier, Toulouse, France

b Programa Docrorado en Ciencias Geologicas, Faculrad de Ciendas Quîmicas, Universidad de Concepciôn, Otite

c Cenrro Sismolôgico Nacionat Universidad de Chile, Santiago, Otite

d Deparramenro Ciendas de la Tierra, Faculrad de Ciendas Quîmicas, Universidad de Concepciôn, Viaor Lamas 1290, Concepcion, Chile

Observarorio Volcanolôgico Otileno, Servicio Nacional de Geologia y Mineria, Rudedndo Ortega 03850, Temuco, Otite

r lnsrirur de Physique du Globe de Paris, Sorbonne Paris Ciré, Univ. Paris Diderot, UMR 7154 CNRS, Paris, France

ARTICLE

Keywords:

Andes Southern Volcanic Zone lnSAR GNSS viscoelasticicy mush reservoir volcanic unresc 1. Introduction ABSTRACT

Silicic systems generate the most explosive eruptions on Earth. ln contrast to basaltic systems, they can accumulate large volumes of magma without systematically erupting, confronting the classical interpretation that a volcano inflates when a magmatic intrusion occurs. Understanding the mechanisms of volcanic inflation and unrest is thus one of the most important challenges in volcanic risk assessment. Laguna del Maule (LdM) in the Southern Volcanic Zone (SVZ) of Chile, is one of the most active Holocene silicic complexes in the world and it has been inflating since 2007, accumulating 2 m of uplift without erupting. Severa( geophysical and geochemical studies conclude that a large crystal rich reservoir would be residing beneath LdM, in consistency with other multi-disciplinary studies showing that such crystal­ rich reservoirs ("mush zones") can be maintained beneath silicic volcanoes, fed by matie magma recharge from below. Nevertheless, the mechanical state of such reservoirs remains unclear. Here, we characterize for the first time the mechanical properties of such a mush reservoir, able to promote large surface displacements such as those measured at LdM. Using a 3D finite element method we simulate a recharge of magma at the base of a crystal rich reservoir, by assuming an overpressurized source surrounded by

a large viscoelastic shell. Inversion results show that this mode! fits the observed temporal and spatial evolution of ground displacements measured with lnSAR data and GNSS data between 2007 and 2017. We interpret the temporal behavior of ground displacement at LdM as resulting from two contributions. A magma recharge occurred within the first 4 yr of the active inflation, followed by the viscous response of

the large viscoelastic shell, set to a viscosity of 1017 Pas. Compared to a purely elastic solution, our model suggests that up to 50% of the accumulated surface displacement during the ten-year period can be explained by this viscous response, and predicts ongoing displacements 50 yr after the onset of inflation. This model agrees with geophysical and geochemical observations and offers a simple explanation of the temporal evolution of surface displacements. lt further allows to reconsider the mechanical behavior of large partially crystallized domains in the upper crust; such significant transient stress transfer over large viscoelastic areas should thus be accounted for in other studies of silicic volcanic complexes.

Silicic systems generate the most explosive eruptions on Earth.

flation signais over several years, without necessarily leading to an eruption. The origin of this unrest remains unclear and is subject of

debate in many volcanic complexes such as Long Valley, Campi Fle­

grei, Yellowstone and many others (Pritchard et al., 2019). To im­

prove eruptions forecasting, understanding the mechanical unrest of these silicic systems constitutes a fundamental question, if not

one of the most important challenges in volcanology (Segall, 2019;

Pritchard et al., 2019). Here, we study the unrest of Laguna del

Maule over the last decade, but the mechanism that we propose

may also explain the unrest of other silicic systems.

They can accumulate large volumes of rhyodacitic or rhyolitic mag­

mas (Degruyter and Huber, 2014) and display sometimes large

in-* Corresponding auchor ac: GET/UMRS563 (UPS, CNRS, IRD, CNES); Obs. Midi­ Pyrénées, Université P. Sabatier, Toulouse, France..

E-mail address: camila.novoa@gec.omp.eu (C. Novoa).

(3)

-35·55• -36'00' -36'05' -36"10' -36"15' -70'40' -70'35' -70'30' -10•25• Pleistocene Eruptions LJAndesite

Post Glacial Eruptions

D

Rhyolite

D

Dacite/Rhyodacite

LJRhyollte

D

Rhyodacite - National boundary

D

Andeslte

-Basalt/Mafic Andesitique

D

lgnlmbrltes

Fig. 1. Reference map of the scucly area in souchem Chile corresponding co the black box in the righc-hand inset map. A simplified geological map of the scudy area (after Hildreth et al., 2010 and Andersen et al., 2017) overlaid onco a shaded relief map, showing the emplacement of the main lava flows and their composition. (For interpretation of the colors in the figure(s). the reader is referred to the web version of this article.)

The Laguna del Maule (LdM) (Fig. 1) is a rhyolitic volcanic field located in the Southern Volcanic Zone (SVZ) of Chile that last erupted ~2000 yr ago (Andersen et al., 2012). It is consid­ ered to be an active rhyolitic complex since it displays the highest frequency of eruptions in the world (Hildreth et al., 2010) with a minimum of 50 rhyolitic eruptions in the last 26.000 yr. It is characterized by a ring of 36 post-glacial rhyolite and rhyodacite coulees and dames erupted from 24 silicic vents which are dis­ tributed over a wide basin of about 300 krn2• Lava flows from this

highest concentration of silicic vents observed in the Southern Vol­ canic Zone covered an area of 100 km2, for a total lava volume of

about 6.4 km3 (Hildreth et al., 2010).

The LdM volcanic field has fostered huge interest from the geodesy community in the last decade because it is one of the strongest deforming volcanic zone on Earth. Based on the analysis

of Synthetic Aperture Radar (InSAR) data acquired between 2001

to 2008, Fournier et al. (2010) presented the first evidence of a strong inflation up to ~18 cm/yr over an elliptical area of 26 km NNE-SSW major axis and 19 km minor axis starting sometime be­ tween February 2004 and January 2007 (the lack of well-correlated interferograms spanning this period prevents precise determina­ tion of the displacement onset). Using JnSAR and Global Navigation Satellite System (GNSS) data, Feigl et al. (2013) showed the persis­ tence of this large-wavelength inflation and provided a value of the displacement rate observed at the center of the inflation of about 28 cm/yr for the 2007-2012 period. The studies carried out by Le Mével et al. (2015, 2016) revealed a temporal change in behavior of the displacement rate with 24

±

0.6 cm/yr from 2007 to early 2010 decaying to 22 cm/yr in 2014 (consistent with a nearly linear temporal behavior after 2010).

(4)

Fournier et al. (2010), Feigl et al. (2013) and Le Mével et al. (2015) proposedthatthisupliftobservedatLdMwascausedbyan inflatingsill locatedatabout5kmdepth.Thesestudies assumed thatthecrustsurroundingthedisplacementsources behavedasa pureelastic solid(Okada,1985).OnlyLe Méveletal. (2016) con-sidereda morecomplexviscoelastic responseof therocks inthe vicinityofthepressurizedsource.Theseauthorsproposetoexplain thechangeintemporal behaviorofsurface displacementsby vis-cousmagmainfluxfromadeepchamber intoashallowchamber connected by a vertical conduit. In this model the magma’s in-fluxasymptoticallyapproacheszeroasthepressurebalanceofthis plumbingsystemisreached.Itallowsthemtocorrelatethechange inobservedsurfaceupliftratewithabi-modalevolutionofmagma volumeflowratefrom0.03km3/yrfrom2007tomid-2009to0.02

km3/yrfrommid-2009tolate2014.

Multiple lines of evidence suggest the presence of a crystal rich reservoir at shallow crustal level beneath the LdM volcanic field. Field observations and geochemical data indicate that the huge concentration ofsilicic ventsobserved inthe area is rather duetoalargeupper-crustalmagmachamberevolving throughout postglacial times, instead of a deeper (i.e. 20 km) crustal melt-ing zone releasing recurrent batches of ascending silicic magma (Hildrethetal., 2010).From geochemicalanalysis,Andersen etal. (2017) arguedthatalleruptionssince26kyrwerefedbythesame largecrystallizedreservoir,despitedifferencesbetweenthe north-ernandsoutherndomainsoftheLdM.Accordingtotheseauthors, thecrustalreservoirsystemwouldbethinnerinthenorthleading to a lower residence time of the magmatic intrusion and induc-ing andesiticvent compositions, incontrast to thesouthern part ofLdMwhereprincipallyrhyoliticeruptionsoccurred. Magnetotel-luric(MT)studiesinturn(Singeretal.,2014; Wagneretal.,2017; Cordell et al., 2018) indicate the presence of a conductive body inthenorthernpartofLdM atabout5kmdepth,whichwas in-terpreted asa hot partially molten body. On the other hand,in thesouthern partofLdM, Bouguer gravityanomalies indicate an anomalyof−19mGalcenteredbelowtheobservedon-going infla-tion(Nieblasarea,Fig.1),whichwouldbecausedbyalowdensity mostly crystallizedmagmabody (∼115km3)containing some 30 km3 of crystal poor rhyolite magma (>85 % of melt) (Miller et

al., 2017).Cordell et al. (2018) support the hypothesis ofa large crystalrich reservoirbeneaththeobservedon-going inflation,but questiontheexistenceof30km3 ofmoltenmagma,arguingthatit

wouldthenbedetectedbytheirmagnetotelluricstudy(MT).From thesedifferentstudies,itappears reasonable toconsiderthat be-neaththeentirevolcaniccomplexofLdMalargepartiallymolten body may have been residing since at leastthe last 26 kyr, but boththegeometryandthestoragedurationofthismagmasource remains unclear. Here we willinvestigatehow such a hypothesis mayhelpto interpretthestrong surfacedisplacements measured intheNieblasarea.

In this study, we benefit from a long InSAR time series of groundsurfacedisplacementatLdMspanningaperiodoftenyears to analyze the characteristic of the surface displacement source anditstime-dependent behavior. First,to constrainthe geometry ofthesource,weinverttheInSARtimeseriesconsideringthat sur-facedisplacementsaretriggeredbyanincreaseininternalpressure within a massive source assuming that the crust acts asa pure elasticmedium.Next,inordertoaccountforamorecomplex rhe-ology related to the presence of a mush reservoir beneathLdM, weusea3D finiteelementapproachto studytheviscoelastic re-sponseofalargecrystalrichreservoirtoaninflatingbasalmagma source.WeinverttheInSARandGNSStimeserieswiththismodel toconstrainthedurationofthesource’sinflationandtheviscosity associated tothis crystalrich reservoir beneath LdM. The results showthatbothsetsofelasticandvisco-elasticmodelsareableto explain themeasuredgroundsurface displacementatLdM.

How-ever they both lead to different conclusions, and provide useful informationforabettercharacterizationandunderstandingofthe present-daymagmaticactivityatLdM.

2. GNSSandInSARdata

GrounddeformationatLdMwascharacterizedfromdaily solu-tionsatfivecontinuousGNSSstationsinstalledbytheObservatorio Volcanológico de losAndesdelSur (OVDAS) coveringa periodof about five years (May 2012 to April 2017) (Fig. 2). SAR imagery overLdMwasacquiredbytwosatellitemissions:C-band Interfero-metricWideSwathmodeimagesfromtheEuropeanSpaceAgency (ESA) Sentinel-1A/B (Table 1) andL-band Stripmapmode images from ALOS1and 2operated by theJapan Aerospace Exploration Agency (JAXA).Full detailsonGNSS andInSAR processingcanbe foundinSupplementalSectionS1.

3. Timeseriesanalysis

The good agreement between the GNSS and InSAR data con-firms thereliability of theInSAR processing (Fig. 3B). Thesubset of inverted InSAR time series (Fig. 3) indicates that C-band and L-band data exhibit a good coherence (>0.6) over most of the studyarea. The timeseries mapsalsoreveal a displacement pat-ternaffectinganellipticalareawitha26kmNNE-SSWmajoraxis anda19kmminoraxiswithadifferencebetweenascendingand descending tracksdue to their difference in satellite viewing ge-ometry. The resulting GNSS and InSAR time series are analyzed by applying a principal componentdecomposition using the de-composition algorithm available inPCAIM (Perfettinietal., 2010; Remyetal.,2014)(see detailsinSupplementalsection S2.1).This analysissuggeststhattheshapeofthedisplacementfieldremains nearly constantover thethree-yearperiodandthatonlythe am-plitudeofthesignalevolveslinearlyovertime,indicatingthatthe observedsignalismainlytheresultofauniqueprocess.

3.1. SpatialandtemporalpatternsofthegrounddisplacementatLdM

We benefit from the ascending and descending Sentinel-1 tracks to retrievetheeasting andthevertical componentsof the grounddisplacement,allowingustomapthedisplacementfieldat LdM(seeTextS2.2foradetailedexplanation).Theresultingmaps ofverticalandhorizontaldisplacementratesderivedfromSentinel datafromOctober2014toJuly2017areshowninFig.4A.Vertical displacement rates reach a maximum of 20 cm/yr, while maxi-mumeastern displacementratesarelowerreaching10cm/yrand 5cm/yr forthedisplacementtothewestandtotheeast, respec-tively.Theeastingdisplacementmapexhibitsclearlyadouble-lobe patternalongaN10◦–N30◦ axiswithapositive lobepattern (dis-placement toward the east)much smaller than the negative one (displacement toward the west). Assuming ahomogeneous crust, suchabehaviorsuggeststhatthesourceroofisnothorizontal. An-otherhypothesisisthatthisasymmetryisduetospatialvariations inthemechanicalpropertiesoftherocks.However,inabsenceof morepreciseinformationonrocklayeringatdepth,andgiventhat a number of non-unique rock heterogeneities can explain simi-lardisplacementpatterns(e.g.seereviewbyMastelark (2007),we choose to condense the information ofthis patterninto a single meaningful property, e.g. the roof ofthe magmatic source(there maybea gradationaltransitionoftherockmagmaboundary,e.g. seediscussioninTownsendetal.,2019).

WetakeadvantageoftheavailabilityoftheInSARdataacquired since 2007 by ALOS-1 satellite to extent the period of ground displacement measurements at the location of the GNSS station MAU2, whichrecordsisthe highestrateofdisplacement atLdM. It was not possible to use ENVISAT data for this purpose since

(5)

LDMP MAU2

]""

·:1 .. �

l �

..,

-0.2

·:ï•'

Il

·:i--..

-0.2 -0.2 ��

li

-0.2

·:

., 1 .. �

1 "

_:

0 --0.2 PUEL

J

..,

..

1

ii"

1

'

f

:

_

l

]

::2:J

]"•

1

-0.2 2014 2015 2016 2014 2015 2016 2014 2015 2016 NIE2 COLO

't•

1�

·:1 ....

..

-0.2

I!

-0.2

Il

·r�

�:c°�

-0.2

·:

_

j

l'

1

-0.2 -0.2 2014 2015 2016 2014 2015 2016

Fig. 2. Ground surface displacement ac the five GNSS stations operating ac LdM. The red circles and associated name show the location of the GNSS stations. Blue dots with

error bars are the observed GNSS data with their la uncertaincy. Red line shows the reconstructed GNSS data using the lirst component of the decomposition. Table 1

Summary of the processed SAR data.

Satellite Mode Orbit Track Mean incidence angle No. of SAR images Scare End (·) Envisat IM Ascending 10 21 ALOS-1 IM Ascending 112 38.2 ALOS-1 IM Ascending 113 41.4 AL0S2 SM3 Descending 130 33 SENTINELl Wv Ascending 18 33.3 SENTINELl IW Descending 83 33.3

only one interferogram formed from the C-band Envisat images ac­ quired in March 2003 and February 2004 is coherent, and it does not reveal any displacements. Consequently, we generate two in­ dependent ALOS-1 lnSAR time series from tracks 112 and 113, using a similar approach to the one used previously. As the LOS are very close for these two tracks, we can compare the two resulting lnSAR time series. This comparison shows a good agreement be­ tween the two independent datasets providing confidence in our lnSAR processing (Fig. 4B). ln order to make the GPS vectors geo­ metrically comparable with the lnSAR data, we projected the GPS vector observed at MAU2 into the LOS of the track 112. To fill the gap between the end of lnSAR measurement and the beginning of the acquisition at MAU2 we used the average velocity observed at MAU2 using ALOSl data spanning the one-year period 2010-2011.

13 20030325 20061219 8 20070126 20101222 9 20070212 20110108 5 20150312 20170705 33 20141030 20170728 32 20141023 20170721

ln agreement with Le Mével et al. (2016), the reconstructed dis­

placement time history recorded at MAU2 from January 2007 to July 2017 highlights a clear decrease of the displacement rate at MAU2 from 24.4 cm/yr for the period spanning January 2007 to January 2011 and 15 cm/yr for the period spanning October 2014

to January 2017 (Fig. 4B).

4. Mechanical modeling strategy

We first verify our assumption that the surface displacements measured by lnSAR and GNSS could be related to a pressure in­ crease in a massive source. To determine the most appropriate geometry of such a source, we consider a variety of classical an­ alytical source geometries to mode! the surface displacements at

(6)

0.0

Sentine! Asc. Track

Sentine! Ose. Track

g

g

ALOS2 Ose. Track

g

LOS Displacement (m) 0.1 0.2 0.3 0.4 ��---�----�-� 200 •�A �2-0�1�5---2�01_6 _____ 2�0�1=7� •200

_.r"?

2015 �B � 0.3 �:t;i; 0.2 2016 Tlme (year) 0.1 :i: lnSAR - GNSS 0.0 2017 ��---�----�-� 200 2015 r"' .. 2015

\

:;_...-2015 2016 2016 Time (year) A 2017 B lnSAR

-

GNSS 2017 A

·

7

2016 2016 Time (year) 2017 B 3E :I: lnSAR GNSS 2017 0 ·200 0.3 0.2 0.1 0.0 200 200 0.3 0.2 0.1 0.0

Fig. 3. Subset of adjusted LOS displacement maps overlaid onto a shaded relief map. Up) Ascending Sentine! crack dis placement maps from January 2016 and January 2017 with respect to the image acquired in October 2014. A) Perpendicular baseline (in meters) as function of cime. Black triangles and red lines represent the SAR images and the interferograms used in this srudy. B) Cumulative LOS displacements and their la uncertainties (in red) recorded ac the location of the GNSS station MAUL2 and the GNSS vecrors recorded ac MAUL2 projected into the radar LOS (in blue). Middle) Same as previously but for the descending Sentine] crack displacement maps from January 2016 and March 2017 with respect to the image acquired in October 2014. Bottom) Same as previously but for the descending ALOS2 crack 130 displacement maps from March 2016 and May 2017 with respect to the image acquired in March 2015. The satellite to ground radar line of sight LOS is shown with a white arrow. LOS displacements toward the satellite are positive. The areas shown in gray indicate area which are not coherenL

(7)

A Up dlsplacement Eastlng dlsplacement m/y m/y 0.00 0.05 0.10 0.15 0.20 -0.10 -0.05 0.00 0.05 0.10 B

I

2 :::ic: ALOS 112

""' :::ic: Asc. SENTINEL

F' :::ic: ALOS 113 ::ic GNSS

-0 2007 2009 2011 2013 2015 2017 Tlme (year)

Fig. 4. Spatial and temporal behaviors of the ground surface displacement at LdM. A) Eascing and vertical components of the mean displacement observed between November 2014 to February 2017 obtained from the combination of the ascending and the descending Sentine! lnSAR data. l..eft) Vertical component of the ground surface displacement. Right) Easting component of the ground surface displacement where negative and positive values indicate displacement to the west and to the east. respectively. Dilferent color scales are used between the two maps in order to improve the conrrast between them. B) lime series of LOS displacements and their uncertainties for ascending ALOSl lnSAR and Ascending Sentine] data at the location of the GNSS station MAU2.

LdM (more details are in Supplemental section S3). Inversions fa­ vor flat-topped sources rather than spheres or prolate ellipsoids (see Supplemental Table S1), in agreement with the estimated ra­ tio of maximum horizontal to maximum vertical displacement, equal to about 0.31 according to GNSS measurements (Dieterich and Decker, 1975; Remy et al., 2014). Therefore, we explore how a large and wide flat-topped pressurized cavity immersed in an elastic crust. can simulate a reservoir filled with a Iow-viscosity magmatic fluid. We assume that the pressure increases within this large and wide flat-topped reservoir, with the shape of an ellip­ tical truncated cone. ln a second step, and following authors like Jellinek and DePaolo (2003) or Le Mével et al. (2016) who consid­ ered a viscoelastic upper crust, we search for a mode( that would simulate the intrusion of new magma at the base of a large silicic mush reservoir itself acting as a viscoelastic boundary zone. We assume that this large reservoir is composed of Iargely crystallized rhyolite close to the solidus temperature (670 °C) or of country rock heated above the brittle-ductile transition (S00-600°C), as was deduced from geochemical and thermodynamic models for LdM (Andersen et al., 2017; Miller et al., 2017). We also assume that the timescale of surface deformation is very small compared to the temperature evolution of the system, with a diffusive time­ scale r = L 2 /k

~

1 Myr (L

~

5 km the characteristic width of the reservoir and k

~

10-6 m2/s the thermal diffusivity). Therefore, it

is reasonable to consider a steady-state temperature, and conse­ quently we define a steady-state viscosity inside a shell that

be-haves according to Maxwell viscoelasticity (Del Negro et al., 2008; Currenti and Williams, 2014; Segall, 2016 and references therein). The main difficulty then stands in inferring the geometry of this reservoir. First, we verify that our finite-element method Adeli-3D (Hassani et al., 1997), which accounts for viscoelasticity, matches classical analytical solutions (e.g. Segall, 2010, details presented in Supplemental section S4). Then, we compare the results produced by Adeli-3D with those produced by the 30-boundary-elements method MO (Cayo( and Cornet, 1998) that accounts for elasticity only. Therefore, we compare the surface displacements produced by a truncated cone embedded in an elastic domain (with MO) with those produced by this same truncated cone acting as a vis­ coelastic shell at the base of which a small pressure source is de­ fined (with Adeli-3D). Our results show that both models produce the same patterns of ground surface displacements after enough time allows for the complete transfer of pressure from the in­ ner source to the visco-elastic shell walls. This convergence, is expected from analytical models (Segall, 2010, for further details refer to Supplemental section SS). Therefore, taking these results into account, we first invert the observed ground surface displace­ ments considering that they are triggered by a pressure increase in an elliptical truncated cone embedded in an elastic medium (section 4.1), and second we use the geometrical parameters in­ ferred from that inversion to explore the influence of a viscoelastic response of this truncated cone on the temporal evolution of dis­ placement field (section 4.2).

(8)

Table 2

Parametersofthedifferentgrounddeformationmodels.X andY arethelocation ofthesourcecenterexpressedinkm(UTMWGS84zone19South).Depthisthe depthbelowthemeansurfaceelevation(i.e.,2500masl).V isthevolumeofthe ellipticaltruncatedcone.V istherateofvolumechangefromOctober,2014to July2017.P istherateofpressurechangefromOctober,2014toJuly2017.S1

isthemajoraxisandS2theminoraxis.Strikeisclockwiseanglerelativetothe

north,Dipisthedipangleoftheminoraxis,alphaisthedipangleofthesource wallwith thefollowing convention,alpha<0the sourcewallsdipoutward(i.e. aconewhichwidensdownward)andalpha>0thesourcewallsdipinward(i.e. aconewhich narrowsdownward).P andV aretherate ofpressure andof volumechangesfromOctober,2014toJuly2017,respectively.V isthevolumeof theellipticaltruncatedconeestimatedfromthe geometricalparametersinferred fromtheinversion.χ2

r isthevalueofthefitofmodeltodata.t1 istheloading

time,η:theviscosity,P thepressureinthebasalmagmaticintrusionandrmsis therootmeansquaredifferencebetweenmodeledandobserveddisplacementat MAU2.

Model form Model1 Model2 Model3

Parameters inferred from elastic inverse models

X position (km) 361.267 361.948 360.949 Y position (km) 6007.115 6006.876 6007.321 Depth (km) 5.42 3.00 5.15 S1(km) 11.52 3.00 12.58 S2(km) 8.92 2.22 8.85 Strike (◦) 23 25 24 Dip (◦) −19 −23 −23 Height (m) 1000 3000 2000 Alpha (◦) 0 −60 59 P (MPa y−1) 1.35 1.45 1.66 V (km3y−1) 0.0214 0.0190 0.0205 V (km3) 80.97 137.17 100.75 χ2 r 1.07 1.08 1.08

Parameters inferred from visco-elastic inverse models

t1(days) 1980.95 1742.5 1549.36

η(Pa s) 9.97e16 9.5e16 9.94e16

P (MPa) 36 18.7 23

rms (m) 0.04 0.02 0.01

4.1. Modelingofgrounddeformationwithanelasticrheology

Todetermine thebest fittingreservoirgeometry, we useMC3 forlinearelasticmedia(CayolandCornet,1998) andthefollowing inversionscheme.Forallmodelcases,weassumeahomogeneous crust withPoisson’sratioof0.25andashearmodulus of13GPa obtainedfromthevelocity modelfromCardona etal. (2018), as-suming a constant density ofthe upper crust of2500 kg/m3. In

agreement with the result of the decomposition, we invertonly thefirstprincipalcomponentofthedifferentdatasetinorderto search forthe geometries that best explain thedata, ratherthan invertingthedisplacementdataavailableateachepoch.Next,we usetheresultingbestgeometriesfortheellipticaltruncatedcone to compute the pressure history of the sources using the same approach as Remy et al. (2014). The inversion is performed us-ing a neighborhood search algorithm (Sambridge, 1999a, 1999b), which finds optimalvaluesof the parameters by minimizing the chi-squareandallowsfora detailedexploration oftheparameter space.Thereducedchi-square(χ2

r)oftheresidualsbetween

obser-vationsandmodeleddisplacementsusedtocomparethegoodness offitofthedatatothemodelsisgivenby:

χ2 r = 1 Nk XobsXmod σ 2

WhereN isthelengthofdataandmodelvectors,k isthenumber offree parametersofthe model, Xobs and Xmod aretheobserved

and the modeled displacement and σ is the data uncertainties. The ellipticaltruncatedconesource dependson theten parame-terslistedinTable2.AsalreadyobservedinRemyetal. (2014),the

inversion indicates that thegeographical location,the azimuthof thedisplacementsourceandthedipofitsmajoraxesarewell re-solved.Howeveratrade-off appearswhenconstrainingthedepth, thickness anddipofthe sourcewalls, leading to a continuum of acceptable modelswithadepthrangingfrom2000mto6000m. Thistrade-offleadsalsotoastronguncertaintyonthe determina-tionofthesourcevolumewhichvariesfromabout50km3 to200

km3. Nevertheless, whatever the source considered, the value of therateofvolumechangeisverysimilarandcloseto0.02km3/yr fromOctober2014toJanuary2017.Table2presentsthree charac-teristicmodelsfromthecontinuumofacceptablemodelsthatare abletomatchthedata.Thesemodelsprovideslightlylowervalues of χ2

r than the Okada-type dislocation model (see Supplemental

Section S3). Nevertheless,similarly to the Okada-typedislocation model(χ2

r of8.50),theyfailtoproperlymatchtheGNSSdata(χr2

of 6.50) essentially due to significant discrepancies between the observed andthemodeled horizontaldisplacements,inparticular atstationPUEL.Ontheotherhand, χ2

r fortheInSARdataincreases

with thetime spanbetween a givenInSAR time seriesmap and the referenceimage usedforthe leastsquare adjustment,dueto thepresenceoflocalizedresidualpatternslocatedaroundthelake. Possible explanations for thismisfit patterncan be that the ge-ometry ofthesource ismorecomplexthan thesimpletruncated coneusedhere,orthatsubsurfaceheterogeneitiesaltertheground displacement.InFig.5wecomparethedisplacementobservedon the largest period of each track andon a three-year2014–2017 periodofGNSSdatawiththemodeleddisplacementsusingoneof the best ellipticaltruncatedcones (Model 3in Table 2).In order to check if these models are able to reproduce the ground dis-placement field observed since 2007, we use them to invert the bestinterferogramsobtainedfromALOS1(SupplementalFigs. S19 andS20).Consideringthat thesemodelsstilladjustwellthedata, yielding χ2

r ofabout1,weconcludethatitisreasonabletoassume

that thesource ofgroundsurfacedisplacements hasnot changed significantlyneitherinshapenorinlocationsince2007.

4.2. Modelingofgrounddeformationwithavisco-elasticrheology

Inthissection,wemodelthetemporalevolutionoftheground displacementcausedby asourcesimulatinganintrusionofmafic magma located at the base of a viscoelastic mush reservoir im-mersed in an elastic crust. First, we use the 3D finite-element codeADELI(Hassanietal.,1997),whichhasbeenusedtosimulate a varietyofgeodynamicalcontexts, includinglong-term deforma-tionatsubduction zones(Hassani etal., 1997; Cerpa etal., 2015) and short-term interseismic deformation (Contreras et al., 2016). Thiscodeusesatime-explicitdynamicrelaxationmethod(Cundall,

1988) to solve thequasi-static equation ofmotion,andcan han-dleavarietyofvisco-elasto-plasticrheologies.Moredetailscanbe foundinChéryetal. (2001),Cerpaetal. (2015) andGerbaultetal. (2018).Themodeledmeshdomainincludesanellipticaltruncated coneembedded inasurroundingelasticcrust, witha pressurized ellipsoidalsourceatitsbasesimulatingmagmarecharge.Maxwell visco-elasticity isassumedinsidethisellipticaltruncatedcone,to simulateacrystalrichmushreservoir inagreementwithDragoni and Magnanensi (1989), who showed that this rheology repro-duces well the behavior of volcanic rocks at high temperature. Supplemental section S4 provides benchmarksof ADELIwith so-lutionsforaviscoelasticshellinan infinitemedium(Dragoniand Magnanensi,1989) andwithsolutionsaccountingforafreesurface (Segall,2010),illustratingthataresolutionof100misrequiredto generateaccuratesolutions.Wetestafull3Dmodelgeometry:the modeled domain isa cube oflateral extent80 km,large enough to minimize border effects. The mesh is composed of a total of 6,9×105 elements,withhighestresolution(100 m)betweenthe

(9)

high-A)

Observed 0.0 0.1

B)

Observed - 0.2 m Modeled -Modeled 02 0.3 0.4 -0.10 -0.05 0.00 0.05 0.10

Fig. S. Maps of the observed and modeled GNSS and lnSAR dara using one of the besr elliptical truncated cone mode!. A) Upper. middle and bonom Example of three LOS displacement maps. the mode) prediction and the residuals between observed and modeled dara. B) Observed cumulated Horizontal and venical GPS displacements ( red vectors) from 2014 to 2017 and modeled displacements (blue vectors).

(10)

A 110000 0 Elliptical -10000 truncated cone. -20000 (Modell, Model2, Model3) -30000 Ellipsoidal -40000 Z [m] source (ln red)

B

Madel 1 Madel 2 Madel 3

Fig. 6. Mesh geomecries for the finire element ADEU serup. A) The mode! is meshed with tecrahedral elements with high resolution in the source and up to the free surface close to the source. B) Meshes used in the viscoelastic models. The ellipsoidal source (in red) and the viscoelastic elliptical cruncated cone (in grey) are meshed with a resolution of 100 m.

relief volcanic edifice and the topographie slope in the area is Jess than the limit of 10° for which its effects become significant (Cayol

and Cornet, 1998), therefore we define a planar top-free surface. We use the three characteristic chamber's geometries found in the previous section (Table 2), since the cumulated surface displace­ ment does not change between an elastic model and a viscoelastic mode! (Supplemental section SS) (Dragoni and Magnanensi, 1989; Bonafede and Ferrari, 2009). Without reliable information concern­ ing the geometry of the basal magmatic intrusion, we choose to apply a pressure increase within a small ellipsoïdal source (volume of 10 km3) situated at the base of the large elliptical truncated

cones (Fig. 68).

A trapezoidal function over time is then used to model the vari­ ation of pressure inside a thin ellipsoïdal source Jocated at the base of the large viscoelastic shell (as in Newman et al., 2001:

Segall, 2010; Del Negro et al., 2008; Parks et al., 2015). Dur­ ing the first period, pressure increases linearly between 0 and t1 simulating a recharge of new magma mass at constant rate into an existing reservoir. During the second period, pressure re­ mains constant, which means that the characteristic time scale of the major processes of i) injection supply rate and ii) internai precipitation and exsolution rates are basically slower than the relaxation time of the viscoelastic shell, at the scale of the 10 yr period of measured surface deformation and over the breath

of the volcanic complex. During this second period, the pressure within the viscoelastic shell progressively evolves with time, from its base to its outer extremities (walls). Maximum ground sur­ face displacement is achieved when this pressure is totally trans­ ferred to the walls and the roof of the visco-elastic reservoir. Consequently at this moment, the surface displacement pattern coïncides with that obtained assuming a purely elastic medium, with a discrepancy Jess than 5% (Pascal et al., 2013, Supplemen­ tal section SS). This behavior illustrates that the surface displace­ ments pattern is clearly dominated by the pressure increase at the walls of the visco-elastic domain rather than by the pres­ sure inside the smaller ellipsoidal reservoir, as shown by previous authors with spherical sources (Dragoni and Magnanensi, 1989; Bonafede and Ferrari, 2009).

In order to reproduce the temporal evolution of LOS displace­ ments at MAU2, we solve the inverse problem and fit these dis­ placements with our visco-elastic mode! according to the following approach. A grid method is used to sample the log10 of the viscos­ ity (11) on the truncated cone mush reservoir, bounded by Iower and upper limits of 15 and 18, respectively (See Supplemental Fig. S21). The choice of these limits is based on different theoreti­ cal studies (Dragoni and Magnanensi, 1989; Bonafede and Ferrari, 2009), thermomechanical models Uellinek and DePaolo, 2003; Del Negro et al., 2008), and laboratory experiments of silicic volcanoes

(11)

A

2 ]: (1) 0 ..J 0 0

B

]: (1) 0 _, 0 0

C

2 ]: (1) 0 _, 0 0 Madel 1

11= 9.97e16[Pas]. 11=1980.95 [days) and P=36[MPaL rms=D.04[m)

I

a ALOS 112, ALOS 113 0GPS --Madel 500

---

---1000 1500

---

---

---2000 2500 3000 3500 Tlme [days) Madel 2

11= 9.Se+16[PasJ, t1 =1742.5 [daysJ and P=18.7[MPaJ, rms=0.02(m)

I

u ALOS 112,AlOS 113 -;-uGPS --Mode!

--

---500 1000 1500 2000 2500 3000 3500 Time (days) Madel 3

11= 9.94e+16(Pas), t1=1549.36 (days) and P=23{MPa), nns=0.01 (ml

I

q ALOS 112.ALOS 113 0GPS --Model

---500 1000 1500 2000 2500 3000 3500 Time (days)

'"i

..

36 4000 ni' Q. � f � 18.7 4000

..

Q. l!! 23 4000

Fig. 7. Results of the best fir of the remporal evolution of surface displacements at MAU2 in the last 10 yr (3650 days). The upper. middle and bonom figures represent the three dilferent elliptical truncated cane models used in this srudy, which simulate an elliptical truncated cane that responds viscoelastically to the inflation of an ellipsoidal source at its base. The upper panel of each figure includes the parameters associated to the best mode! '7: viscosity. r1: loading time. P: pressure. and rms: root mean square error.

(Newman et al., 2001, 2006). Then, a Levenberg-Marquardt nu­

merical optimization grid search method determines the optimum

loading duration of the magma injection (t0) together with the op­

timum overpressure (dP) inside the ellipsoidal source and the op­ timum viscosity (17) in the visco-elastic cane truncated (Levenberg,

1944). We limit the range of duration of this magma injection from

several months to 10 yr and the overpressure from 1 to 150 MPa.

We present on Fig. 7, the different viscoelastic models inferred from each geometry, that best match the observed 10 years-long time series of ground displacements at MAU2 (spanning period 2007-2017).

The resulting viscosities ( 1J) in the truncated cane reservoir are

very similar in ail three mode! geometries. Nevertheless, the dura­ tian of the loading time (to) and the pressure (dP) differ slightly.

(12)

The best minimum root mean square (rms) error between ob-served and modeled ground surface displacements are obtained for a viscosity η∼1017 Pa s in model 1, model 2 andmodel 3; resultinginrms valuesof0.04,0.02 and0.01m,respectively.The loadingtimevariesbetweent0∼4 yrformodel2andmodel3,to

amaximumt0∼5.4 yr,formodel1supporting theideathatthe

maficrechargecouldnothavebeenlongerthanthefirst6 yrand thattheremaining deformationwascausedbytheviscoelastic re-sponse ofthe surroundingshell. The applied overpressure inthe ellipticalsourceatthebaseofthe truncatedcones variesslightly showingthatathickerviscoelasticreservoirrequiresless overpres-sure to generate similar ground surface deformation; dP ranges from18MPa forthethickest viscoelasticsource(3 kmheight)to 36MPaforthethinnestsource(1kmheight).Theappropriateness ofthesevaluesarefurtherdiscussedinthefollowingsection5.

5. Discussion

5.1. Sourcecharacterization

ThisstudyconfirmsthattheupliftatLdMstartedsometime be-tween2004and2007(Fournieretal.,2010;Feigletal.,2013)and continuedatleastuntilMarch2017.Thedecompositionofboth In-SARandGNSStime seriesindicatesthat displacementshavebeen linearintime since early2012. Thebest inflatingtruncatedcone modelsembeddedinanelasticcrust,deducedfromtheinversions ofSentinel1andGNSSdatamatchwell theALOS1dataacquired from2007to2011,indicatingthatthesource’sshapeandlocation have not changed significantly since 2007. Our inversions show thatthecenterlocation,theorientation andthedipofthesource remain very similarwhatever its geometry.The center islocated belowthe southwestpartof thelake, its majoraxis iselongated N22◦E–23◦Eanditdipsdowntothewestwithanangleofabout 20◦.Furthermore,theinversionscannotprovidestrongconstraints on the depth of the source, due to the trade-off betweendepth andpressure.Nevertheless, theestimatedrate ofvolume change, ofabout0.02km3/yrfromOctober2014toJanuary2017,isclose

tothat obtainedby Le Méveletal. (2016) and isindependent of thesource type whenassuming a purely elastic crust andan in-compressiblemagma.

5.2. Visco-elasticparametersofthemushreservoir

Alargespectrum ofacceptable sourcegeometries explainsthe data, ranging froma thinsource (sill) to a large massive source of equivalent basal area of about 10×8 km2. The volume and

the thickness of this large massive source range from ∼60 to 200 km3 and from 1000 to 3000 m, respectively. Such values might seem large, butthey are compatible with the size of the magma chamber that wouldhave triggered the formation ofthe 12×8 km2-wide Bobadilla caldera ∼950 kyr ago (Hildreth et al., 2010). From the continuum of possible sources, we selected three representativegeometries simulating a mushreservoir that responds visco-elastically to a basal pressure change. Whatever the geometryconsidered, they fit the temporal evolution of sur-face displacements at MAU2 for a viscosity converging towards 1017Pa s.Ourresultsindicatethatathickermushreservoir(model 2 and 3 in Fig. 7B–C) explains better the temporal evolution of surface displacements than the thinnermodel 1 (Fig. 7A). Model 1 cannot explain the variation in slope that starts in 2014 (day

∼1980,Fig.7A),illustratingtheimportanceofthethicknessofthe mushreservoirinamplifyingsurfacedisplacements.

The presenceof alarge mush reservoirbeneath LdM hastwo mainimplications. (i) Althoughan elastic medium requiresa full 10 yr of inflation over the 10 yr of observed surface uplift, a visco-elasticmedium “overtopping” an ellipsoidal sourcerequires

a shorter period of active inflation (less than 6 years), the re-mainingfour-yearsreflectingitsdelayedvisco-elasticresponse.(ii) Althoughthecumulatedsurfacedisplacementsaresimilarbetween an elastic medium and a visco-elastic medium, the visco-elastic mush model requires less basal magma recharge, i.e. a volume change50%lowerthantheelasticmodel.

According toLeMevel’smodel,a 25kmdeepreservoirwould feeda 30–60mwideconduit withmaficmagma duringthe first 2yr (2007–2009)ataconstantincreasingpressurerate,itself fill-ing in a shallow crustal reservoir. Over the following years, the pressure remains constant inthe deepreservoir, butit continues to increase inthe shallow reservoir until hydrostaticequilibrium is reached. Thismodel explainsboth theamplitude andthe pro-gressivedecelerationofthegroundsurfacedisplacementswiththe elasticresponseofthebedrocksurroundingtheupperchamber.Le Méveletal. (2016) justifytheirassumptionofanelasticalbedrock by showingthat a viscous componentofdeformation increments thesurfaceupliftbyonly4%comparedtoapurelyelasticsolution. Buttheseauthorsassumedthattheviscoelasticshellcorresponds tothethermalvariationassociatedtoamaficintrusionintoasmall magmatic source (16 km3), thus insufficient to significantly am-plifygroundsurfacedisplacements.Thiswouldimplyacontinuous supply of magmatic material from below. Our assumptions here differ fromLe Mévelet al. (2016) because we considera thicker mushreservoirdevelopingoveratleast26kyr,hypothesizingthat itwouldbecontinuously reheatedwithmaficmagmaduringthat broader time-scale.This assumptionis coherentwiththeprocess suggested byAndersen et al. (2017) andothers pointingtowards thepresenceofalong-standingnear-solidusmagmabodybeneath theareaofactiveuplift,cf.Section5.3below.

5.3. Largenear-solidusmagmabodybeneathLdM

Thevolumeofthe“mush”reservoirinferredbytheinversionof InSARandGNSSdatamustbetakenwithcautionaswedidnot ac-countforthepossiblecontributionofadditionalviscousandbrittle deformation ofthe bedrock aroundandbelow thereservoir. This host rock could be mechanically damaged by dyking and meta-morphic reactions in a potentially hydrated environment, which wouldcontribute to furtherreducing its effectivestrength(Rubin Allan, 1995; Newman et al., 2001; Jellinek and DePaolo, 2003; Gerbault et al., 2018). The viscosity of about 1017 Pa s inferred

from ourmodeling studyis one orderof magnitudebelow com-mon values of crustal viscosity ∼1018 Pa s inferred for volcanic regions with a large amount of heated country rock (Newman et al., 2006; Parks et al., 2015; Le Mével et al., 2016), butit is higherthanthatobtainedinotherthermo-mechanicalstudiessuch as those of Del Negro et al. (2008) and Currenti and Williams (2014) that obtaineda crystalviscosityof ∼1016 Pa s. Thisvalue

of 1017 Pa s is in turn too high to support the hypothesis of a large quantity of molten magmabeneath theLdM volcanic com-plex, whichisinagreementwiththemainconclusionsoftheMT analysis carried out by Cordell et al. (2018), who suggestedthat a molten magma body belowthe uplift area was unlikely to be greater than 10–20 km3.Our modeled valuesare also consistent

withthegravity anomaly patternreaching 19mGal atthecenter ofinflation,whichwasinterpretedtowitnessalargeshallow sili-cicsystemoflowdensity,mostlycrystallizedmagma(Milleretal.,

2017). It is alsoconsistent withthe results ofisotopic geochem-ical studies that rule out the possibility of LdM rhyolites being generatedinthelowercrust,andratherfavorthehypothesisthat the last period of activity resulted from mafic magma intruding awell-developedpre-existingcrystallizedrhyoliticreservoirinthe shallowcrust(Andersenetal.,2017).

Jellinek and DePaolo (2003) pointed out that the accumula-tion of large quantities of magma for many years in silicic

(13)

vol-canoes is favored by warm-wall rocks with lower effective vis-cosity compared to conventional bedrocks. Conjointly for silicic volcanoes, Cooper and Kent (2014) suggested for Mount Hood, USA,that nearlyall ofthedurationofmagmastorage(>88% and perhaps 99%) occurs at near-solidus conditions, i.e. under rela-tivelycool(<750◦),viscousanduneruptibleconditions.They sug-gestedthateruptionsinthesesystemsoccur viaremobilizationof a cold andrigidstate that reactsimmediately tomafic recharge, as had also been suggested for other silicic volcanoes such as Taupo volcano, Long Valley, and others (Charlier et al., 2004; Hildreth,2004; HildrethandWilson,2007; CooperandKent,2014; DegruyterandHuber,2014).Thisconceptofmagmastorageisalso similartothatproposedintherecentstudybySzymanowskietal. (2017) from the analysis of zircon andtitanite crystals collected inthe Kneeling Nun Tuff (New Mexico). These authors conclude that magma has been mainly storedat cold temperatures rang-ingfrom680to730◦.Therefore,despitethedetectionbyMTofa largepartially moltenmagma bodyinthe northernzone ofLdM, whichcould forecast an imminenteruption, the smallvolume of highmeltconcentrationsimmediatelybelowLdMprobablyreflects that the mafic recharge is not able to promote an eruption. A largenear-solidusmagmabodybeneathLdMmaintainedina non-eruptible state, mightstill be reactivatedbefore an eruption, but notnecessarily“enough”.

A significant influx of magma over a long period of time is still required to allow for the development of a long-lived large near-solidusmagma body (Annen, 2009; Galetto etal., 2017), or elsediffusivecooling wouldfreezeit.AtLdM,theabundanceand thefrequency ofpost-glacialrhyolitic eruptionsare characterized bya prolonged historyofexplosive silicic eruptionssince 1.5Ma (forinstancetheBobadilla calderacollapse∼950kaago,Hildreth etal., 2010). The development of this highcrystal rich reservoir withsuccessivetransientmaficrechargesmightbelinkedwiththe

>60 m magnitudeof surface deformation accumulatedsince 9.4 thousandyearsinthesouthern-eastzoneofLdM,andthatwas in-terpretedastheevidencefortheincremental growthofamagma reservoirjustunderneath (Singeret al.,2014,2018). Another hy-pothesisisthattheLdMcomplexisfedfromalargemagmatic sys-temat>8kmofdepthtothenorththat couldprovidesufficient heatandmelttomaintaintheshallowersystem(Reyes-Wagneret al.,2017; Cordelletal., 2018).Finally,severalstudiesconvergeto the same interpretation that the past 50 rhyolitic eruptions that encircled the lake in the postglacial periodwere derived from a single andlarge shallowsilicic systemoflow density.This silicic magmaticsystemtogether withvolatileminglingwithin itwould furthermoreinhibitthedirect ascentofmafic magma tothe sur-face(Hildrethetal., 2010; Singer etal.,2014; Milleretal.,2017; Andersenetal.,2017).

5.4.Implicationonthesourceoverpressure

The three representative geometries of viscoelastic truncated conesthatwetestedmatchtheobservedevolutionofground sur-face displacements with a maximum residual of 4 cm between models and observations. The models require a basal overpres-suresource ranging between 18 MPa for the largest viscoelastic reservoir(3 kmthick,model 2)and36MPaforthe thinnest vis-coelasticreservoir (1 kmthick, model 1). Thispressure/thickness relationship can be justified with a simplified 1D analogy: con-sidering a similar total surface uplift or deformation rate (dv) and a similar viscosity (η) for both geometrical domains, but a different height H1, H2 over which this deformation occurs,

one can write dv=HP1=HP2 (invoking the

stan-dardstress-strain-rateviscousconstitutiverelationship).Thisleads

PH1=PH2,showingthat thethickertheviscoelastic

do-main,thesmallerthe overpressurerequiredto producethesame

uplift rate. Such overpressures remain realistic since 18 MPa is of the order of the tensile strength of crustal rocks and a com-mon value assumed for overpressurised magmatic crustal reser-voirs. 36 MPa is also likely possible in contexts of rather dry andcompetentbedrockenvironments(JellinekandDePaolo,2003; Gerbaultetal.,2018).

Finally,ourstudyconfirmsthatviscoelasticityisakeyfactorto promotelong-termsurfacedisplacementsoverdecades.Ourmodel 3 inparticular, which best fitsthe data between2007and 2017 at LdM, predicts that a four-year continuous injection of mafic magmaatthebaseofamushreservoircanpromote3.8m of cu-mulativedisplacementsatstationMAU2within50yr(Fig.8).This increase reflectsthe slowtransference ofoverpressurebetweena pressurizedbasalsourceandthewallsofthelargemushreservoir, duetoitsretardedviscoelasticresponse.Whileanelasticmodelis a goodindicatorof thepressureatthewallsofa largesource (if relatedtoasmallinjectionofmagmaoverashortperiodoftime) itcannotreflecttheactivationofalargemagmachamberforsilicic volcanoesdisplayinglong-termdeformationsuch asLdM.Insuch casesviscoelasticshells shouldratherbe considered.Future mea-surementswillhelpclarifythegeometryofthislargevisco-elastic shell,andtheassociatedinflatingoverpressure,furtherhelpingto predicttheevolutionofthisvolcanichazard.

6. Conclusion

Ourstudyquantifies forthe firsttime thecharacteristic prop-erties ofa mushreservoir promotinglarge surface displacements above a silicic volcanic system such asLdM, Long Valley, Campi Flegrei, etc. Such large mushreservoirs maybe associated to re-currentmaficrechargesthatthesevolcaniccomplexes experience, butdonotnecessarilyindicateimminenteruption.

From the analysisof a10 years-long combinedtime series of InSAR and GNSS data we first confirm the spatial and temporal evolution of the regional uplift (up to 2 m of accumulated dis-placement) observedattheLdMvolcaniccomplex.Asan alterna-tive to previously proposed models explaining such surface dis-placements,we tested3Dvisco-elasticmodels simulatingamafic magma recharge atthe base of a crystal rich reservoir behaving visco-elastically. The tested geometries of visco-elastic reservoirs match the observed evolution of ground surface displacements over time, with a viscosity approaching 1017 Pa s. The resulting

overpressure ranges between 18 and 36 MPa depending on the reservoir’s dimension. Our models support a scenario in which a basal mafic intrusion first inflated during the first 4 or 6 yr of active deformation, then triggering a continuous slow transfer ofpressure betweenthisbasal loadtothe viscoelasticreservoir’s walls, promotingan increase ingroundsurface displacements for uptoanother50yr.

Ourbest fitting dimensions, viscosity andoverpressure of the visco-elasticmushreservoirareconsistentwithprevious interpre-tationsofalargelong-lived,near-solidusmagmabodyunderneath LdM.Thisraisestheissueofhowsuchmagmabodiesare thermo-mechanically maintained over time, by other mechanisms than standard thermaldiffusion/dykingprocesses (metamorphism, per-colation through porous media etc.), andcalls for improving our understandingofthemulti-scalephysicsofvolcanicplumbing sys-tems,alsowiththehelp ofhighquality geophysicalandgeodetic campaigns.

Acknowledgements

ASARdatawereprovidedbytheEuropean SpaceAgency(ESA) throughtheprojectENVISAT-AO#857andCategory1#2899. PAL-SARdatafromtheALOSsatellitemissionoperated bytheJapanese

(14)

Fig. 8. Predictedspatialandtemporalevolutionofthegroundsurfacedisplacementfrommodel3. Up)ModeledtemporalevolutionofthesurfacedisplacementatMAU2for thenext50years. Bottom)Modeledaccumulatedgroundsurfacedisplacementsfrom2007to2017.

Aerospace ExplorationAgency (JAXA)were used underthe terms and conditions of the fourth ALOS 2 Research Announcement (project 1142).We thankthe European SpaceAgency for provid-ingusSENTINELandENVISATdataandOVDAS- SERNAGEOMINfor providing the GNSS data. C. Novoa acknowledges CONICYT-Becas Chile for her PhD scholarship (folio: 72170643). A. Tassara ac-knowledgessupportfromFONDECYT(project1151175).F.Delgado

acknowledgesCentreNationald’EtudesSpatiales forapostdoctoral fellowship.WethankR.HassaniforhisusefulhelpwiththeAdeli software. The editor Tamsin Mather and the two anonymous re-viewers providedusefulandconstructivesuggestions that greatly improved this study. Several figures were made using the GMT software(Wessel andSmith,1998).The GMSHGnusoftwarewas usedtobuildthemodels;meshes(gmsh.info).

(15)

Appendix A. Supplementarymaterial

Supplementarymaterialrelatedtothisarticlecanbefound on-lineathttps://doi.org/10.1016/j.epsl.2019.06.005.

References

Andersen, N.L., Singer, B., Jicha, B., Hildreth, E., Fierstein, J., Rogers, N., 2012. Evo-lution of rhyolite at Laguna del Maule, a rapidly inflating volcanic field in the Southern Andes. In: Proceedings of the Fall Meeting of Am. Geophys. Un. San Francisco, USA.

Andersen, N.L., Singer, B.S., Jicha, B.R., Beard, B.L., Johnson, C.M., Licciardi, J.M., 2017. Pleistocene to Holocene growth of a large upper crustal rhyolitic magma reservoir beneath the active laguna del Maule Volcanic field, Central Chile. J. Petrol. 58 (1), 85–114.

Annen, C., 2009. From plutons to magma chambers: thermal constraints on the accumulation of eruptible silicic magma in the upper crust. Earth Planet. Sci. Lett. 284 (3–4), 409–416.

Bonafede, M., Ferrari, C., 2009. Analytical models of deformation and residual gravity changes due to a Mogi source in a viscoelastic medium. Tectonophysics 471 (1–2), 4–13.

Cardona,C., Tassara,A., Gil Cruz,F.,Lara,L.E., Morales,S., Kohler,P.,Franco, L., 2018.CrustalseismicityassociatedtorapidupliftatLagunadelMaulevolcanic complex,SouthernVolcanicZoneofAndes.J.Volcanol.Geotherm.Res.https:// doi .org /10 .1016 /j .jvolgeores .2018 .01.009.

Cayol, V., Cornet, F.H., 1998. Effects of topography on the interpretation of the defor-mation field prominent volcanoes – application to Etna. Geophys. Res. Lett. 25, 1979–1982.

Cerpa, N.G., Araya, R., Gerbault, M., Hassani, R., 2015. Relationship between slab dip and topography segmentation in an oblique subduction zone: insights from numerical modeling. Geophys. Res. Lett. 42 (14), 5786–5795.

Charlier, B.L.A., Wilson, C.J.N., Lowenstern, J.B., Blake, S., Van Calsteren, P.W., David-son, J.P., 2004. Magma generation at a large, hyperactive silicic volcano (Taupo, New Zealand) revealed by U–Th and U–Pb systematics in zircons. J. Petrol. 46 (1), 3–32.

Chéry, J., Zoback, M.D., Hassani, R., 2001. An integrated mechanical model of the San Andreas fault in central and northern California. J. Geophys. Res., Solid Earth 106 (B10), 22051–22066.

Contreras, M., Tassara, A., Gerbault, M., Araya, R., Bataille, K., 2016. Interseismic de-formation at subduction zones investigated by 2D numerical modeling: case study before the 2010 Maule earthquake. Andean Geol. 43 (3), 247–262. Cooper, K.M., Kent, A.J., 2014. Rapid remobilization of magmatic crystals kept in cold

storage. Nature 506 (7489), 480.

Cordell, D., Unsworth, M., Diaz, D., 2018. Imaging the Laguna del Maule Volcanic Field, central Chile using magnetotellurics: evidence for crustal melt regions laterally-offset from surface vents and lava flows. Earth Planet. Sci. Lett. 488, 168–180.

Cundall, P.A., 1988. A microcomputer program for modeling large-strain plas-ticity problems. In: Numerical Methods in Geomechanics (Innsbruck 1988), pp. 2101–2108.

Currenti, G., Williams, C.A., 2014. Numerical modeling of deformation and stress fields around a magma chamber: constraints on failure conditions and rheology. Phys. Earth Planet. Inter. 226, 14–27.

Degruyter, W., Huber, C., 2014. A model for eruption frequency of upper crustal silicic magma chambers. Earth Planet. Sci. Lett. 403, 117–130.

DelNegro, C.,Currenti,G.,Scandura,D., 2008.Temperature-dependent viscoelas-ticmodelingofgrounddeformation: applicationtoEtnavolcano duringthe 1993–1997inflationperiod.Phys.EarthPlanet.Inter.https://doi .org /10 .1016 /j . pepi .2008 .1010 .1019.

Dieterich, J.H., Decker, R.W., 1975. Finite element modelling of surface deformation associated with volcanism. J. Geophys. Res. 80 (29), 4094–4102.

Dragoni, M., Magnanensi, C., 1989. Displacement and stress produced by a pressur-ized, spherical magma chamber surrounded by a viscoelastic shell. Phys. Earth Planet. Inter. 56, 316–328.

Feigl,K.,LeMével,H.,TabrezAli,S.,Cordova,L.,Andersen,N.,DeMets,C.,Singer, B.,2013.RapidupliftinLagunadelMaulevolcanicfieldoftheAndean South-ernVolcaniczone(Chile)2007–2012.Geophys.J.Int.https://doi .org /10 .1093 /gji / ggt438.

Fournier, T.J., Pritchard, M.E., Riddick, S.N., 2010. Duration, magnitude, and frequency of subaerial volcano deformation events: new results from Latin America using InSAR and a global synthesis. Geochem. Geophys. Geosyst. 11 (1), Q01003. Galetto, F., Acocella, V., Caricchi, L., 2017. Caldera resurgence driven by magma

vis-cosity contrasts. Nat. Commun. 8 (1), 1750.

Gerbault, M., Hassani, R., Novoa Lizama, C., Souche, A., 2018. Three-dimensional failure patterns around an inflating magmatic chamber. Geochem. Geophys. Geosyst. 19 (3), 749–771.

Hassani, R., Jongmans, D., Chery, J., 1997. Study of plate deformation and stress in subduction processes using two-dimensional numerical models. J. Geophys. Res. 102 (B8), 17951–17965.

Hildreth, W., 2004. Volcanological perspectives on Long Valley, Mammoth Moun-tain, and Mono Craters: several contiguous but discrete systems. J. Volcanol. Geotherm. Res. 136 (3–4), 169–198.

Hildreth, W., Wilson, C.J., 2007. Compositional zoning of the Bishop Tuff. J. Petrol. 48 (5), 951–999.

Hildreth, W., Godoy, E., Fierstein, J., Singer, B., 2010. Laguna del Maule Volcanic field: eruptive history of a Quaternary basalt-to-rhyolite distributed volcanic field on the Andean range crest in central Chile. Serv. Nac. Geol. Min., Bol. 63. 142 pp. Jellinek, M.A., DePaolo, D.J., 2003. A model for the origin of large silicic magma

chambers: precursors of caldera-forming eruption. Bull. Volcanol. 65, 363–381. Le Mével, H., Feigl, K., Cordova, L., DeMets, C., Lundgren, P., 2015. Evolution of unrest

at Laguna del Maule volcanic field (Chile) from InSAR and GPS measurements, 2003 to 2014. Geophys. Res. Lett. 42 (16), 6590–6598.

Le Mével, H., Gregg, P.M., Feigl, K., 2016. Magma injection into a long-lived reser-voir to explain geodetically measured uplift: application to the 2007–2014 un-rest episode at Laguna del Maule volcanic field, Chile. J. Geophys. Res., Solid Earth 121 (8), 6092–6108.

Levenberg, K., 1944. A method for the solution of certain non-linear problems in least squares. Q. Appl. Math. 2 (2), 164–168.

Mastelark, T., 2007. Magma intrusion and deformation predictions: sensitivities to the Mogi assumptions. J. Geophys. Res., Solid Earth 112 (B6).

Miller, C.A., Williams-Jones, G., Fournier, D., Witter, J., 2017. 3D gravity inversion and thermodynamic modelling reveal properties of shallow silicic magma reservoir beneath Laguna del Maule, Chile. Earth Planet. Sci. Lett. 459, 14–27.

Newman, A.V., Dixon, T.H., Ofoegbu, G.I., Dixon, J.E., 2001. Geodetic and seismic constraints on recent activity at Long Valley Caldera, California: evidence for viscoelastic rheology. J. Volcanol. Geotherm. Res. 105 (3), 183–206.

Newman, A.V., Dixon, T.H., Gourmelen, N., 2006. A four-dimensional viscoelastic de-formation model for Long Valley Caldera, California, between 1995 and 2000. J. Volcanol. Geotherm. Res. 150, 244–269.

Okada, Y., 1985. Surface deformation due to shear and tensile faults in a half-space. Bull. Seismol. Soc. Am. 75, 1135–1154.

Parks, M.M., Moore, J.D., Papanikolaou, X., Biggs, J., Mather, T.A., Pyle, D.M., et al., 2015. From quiescence to unrest: 20 years of satellite geodetic measurements at Santorini volcano, Greece. J. Geophys. Res., Solid Earth 120 (2), 1309–1328. Pascal, K., Neuberg, J., Rivalta, E., 2013. On precisely modelling surface

deforma-tion due to interacting magma chambers and dykes. Geophys. J. Int. 196 (1), 253–278.

Perfettini, H., Avouac, J.P., Tavera, H., Kositsky, A., Nocquet, J.M., Bondoux, F., Chlieh, M., Sladen, A., Audin, L., Farber, D., Soler, P., 2010. Seismic and aseismic slip on the Central Peru megathrust. Nature 465 (7294), 78–81.

Pritchard, M.E., Mather, T.A., McNutt, S.R., Delgado, F.J., Reath, K., 2019. Thoughts on the criteria to determine the origin of volcanic unrest as magmatic or non-magmatic. Philos. Trans. R. Soc. Lond. A 377 (2139), 20180008.

Remy, D., Froger, J.L., Perfettini, H., Bonvalot, S., Gabalda, G., Albino, F., Cayol, V., Legrand, D., Saint Blanquat, M.D., 2014. Persistent uplift of the Lazufre volcanic complex (Central Andes): new insights from PCAIM inversion of InSAR time se-ries and GPS data. Geochem. Geophys. Geosyst. 15 (9), 3591–3611.

Rubin Allan, M., 1995. Getting granite dikes out of the source region. J. Geophys. Res., Solid Earth 100 (B4), 5911–5929.

Sambridge, M., 1999a. Geophysical inversion with a neighbourhood algorithm, I: ap-praising the ensemble. Geophys. J. Int. 138, 727–746.

Sambridge, M., 1999b. Geophysical inversion with a neighbourhood algorithm, II: searching a parameter space. Geophys. J. Int. 138, 479–494.

Segall, P., 2010. Earthquake and Volcano Deformation. Princeton University Press. Segall, P., 2016. Repressurization following eruption from a magma chamber with a

viscoelastic aureole. J. Geophys. Res., Solid Earth 121 (12), 8501–8522. Segall, P., 2019. Magma chambers: what we can, and cannot, learn from volcano

geodesy. Philos. Trans. R. Soc. A 377, 20180158.

Singer, B.S., Andersen, N.L., Le Mével, H., Feigl, K.L., DeMets, C., Tikoff, B., Thurber, C., Jicha, B., Cardona, C., Cordova, L., Gil, F., Unsworth, M., Williams-Jone, G., Hildreth, W., Fierstein, J., Vaszquez, J., 2014. Dynamics of a large, restless, rhy-olitic magma system Laguna del Maule, southern Andes, Chile. GSA Today 24 (2), 4–10.

Singer, B.S., Le Mével, H.l.n., Licciardi, J.M., Cordova, L., Tikoff, B., Garibaldi, N., An-dersen, N.L., Diefenbach, A.K., Feigl, K.L., 2018. Geomorphic expression of rapid Holocene silicic magma reservoir growth beneath Laguna del Maule, Chile. Sci. Adv. 4 (6).

Szymanowski, D., Wotzlaw, J.r.-F., Ellis, B.S., Bachmann, O., Guillong, M., von Quadt, A., 2017. Protracted near-solidus storage and pre-eruptive rejuvenation of large magma reservoirs. Nat. Geosci. 10, 777.

Townsend, M., Huber, C., Degruyter, W., Bachmann, O., 2019. Magma chamber growth during inter-caldera periods: insights from thermo-mechanical model-ing with applications to Laguna del Maule, Campi Flegrei, Santorini, and ASO. Geochem. Geophys. Geosyst.

Wagner,V.R.,Diaz,D.,Darcy,C.,Unsworth,M.,2017.Regionalelectricalstructureof theAndeansubductionzoneincentralChile(35◦–36◦S)usingmagnetotellurics. EarthPlanetsSpace.https://doi .org /10 .1186 /s40623 -017 -0726 -z.

Wessel, P., Smith, W.H., 1998. New, improved version of Generic Mapping Tools re-leased. Eos Trans. AGU 79 (47), 579.

(16)

Supplementary Material for: Viscoelastic

relaxation: a mechanism to explain the decennial

large surface displacements at Laguna del Maule

silicic volcanic complex

Novoa C., Remy D., Gerbault M., Baez J.C, Tassara A., Cordova L., Cardona C., Granger M, Bonvalot S., Delgado F.

May 29, 2019

The Supplement contains:

1 • Supplementary Text: S1 to S5 2 • Supplementary Figures: S1 to S21 3 • Supplementary Tables: S1 to S2 4

Introduction

5

Section S1 presents in details GNSS and InSAR processing used in this study. 6

Section S2.1 presents a principal component analysis based on singular value 7

decomposition and Section S2.2 details the strategy adopted to determine 8

the easting and vertical displacement maps from ascending and descending 9

Sentinel-1 data. Section S3 presents modeled ground surface displacement 10

using classical analytical models. Section S4 presents a benchmark between 11

the analytical solutions of Dragoni et al. (1989) and the numerical solutions 12

obtained with the 3D finite-element code ADELI, for a viscoelastic rheology 13

in an infinite medium. S4 presents another comparison between the numer-14

ical solutions of ADELI and those by Segall (2010), accounting for a free 15

surface. Section S5 presents the comparison of surface displacement patterns 16

between a pressurized elliptical source immersed in an elastic medium and 17

a small pressurized source surrounded by a viscoelastic shell. In this section 18

we show that the viscoelastic model reproduces the surface displacements 19

produced by the elastic model after a period of time. 20

(17)

S1

GNSS and InSAR processing

21

S1.1 GNSS data 22

Ground deformation at LdM was characterized from daily solutions at five 23

continuous GNSS stations installed by the Observatorio Volcanol´ogico de los

24

Andes del Sur (OVDAS) covering a period of about five years (May 2012 to 25

April 2017). The GNSS data were provided by OVDAS and processed using 26

the GIPSY software from JPL. We obtained the precise ground displace-27

ment time series in three dimensions at GNSS stations after correcting for 28

plate motion and long spatial wavelength post-seismic surface deformation 29

associated with the 2010 Maule earthquake derived from the data recorded 30

at MAUL station, about 39 km away from the center of the LdM volcanic 31

complex (Li et al., 2018). The 1σ accuracy of the daily station positions is 32

estimated at 3 mm and 5 mm for the horizontal and the vertical components, 33

respectively. 34

S1.2 InSAR data 35

SAR imagery over LdM was acquired by two satellite missions: C-band 36

Interferometric Wide Swath mode images from the European Space Agency 37

(ESA) Sentinel-1A/B and L-band Stripmap mode images from ALOS 1 and 38

2 operated by the Japan Aerospace Exploration Agency (JAXA). The C-39

band SAR dataset (Table 1) includes 33 acquisitions from ascending orbits (6 40

October 2014 to January 2017), and 32 acquisitions from descending orbits 41

(23 October 2014 to 23 March 2017). The L-band SAR dataset is composed 42

of 17 PALSAR images from ascending orbits (26 January 2007 to 8 January 43

2011) and 5 PALSAR-2 images from descending orbits (12 March 2015 to 4 44

May 2017). 45

We used the DIAPASON software (Diapason, 2006) to generate the dif-46

ferential interferograms, using the two-pass method described by Massonnet 47

and Feigl (1995). For Sentinel data, we used the 3 arcsec Shuttle Radar 48

Topographic Mission digital elevation model (SRTM) to remove the topo-49

graphic contribution and the 1 arcsec SRTM for the other SAR data. We 50

used the 1 arcsec SRTM to provide a geographic framework (UTM WGS84) 51

for the interferograms. The criteria used for selecting data were, to have a 52

perpendicular baseline lower than 1500 m, and that for the ALOS-2 and Sen-53

tinel 1 satellites orbital with separations not exceeding 200 m, no selection 54

was taken into account. 55

Next, a weighted power spectral density filter (Goldstein et al., 1988) 56

was used to filter the interferograms which were later unwrapped using an 57

implementation of the Statistical cost, Network- flow Algorithm for Phase 58

Unwrapping -SNAPHU (Chen and Zebker, 2002). The resulting interfero-59

grams indicate large seasonal variation in coherence with a strong decrease 60

Figure

Fig. 1. Reference map of  the  scucly area  in  souchem Chile corresponding  co the black box  in  the righc-hand  inset map
Fig. 2. Ground surface displacement ac the  five  GNSS stations operating ac LdM. The red circles and associated name show the location of the GNSS stations
Fig. 3. Subset of  adjusted LOS displacement maps overlaid onto a shaded relief map. Up) Ascending Sentine! crack dis placement maps from January 2016 and January 2017  with respect to the image acquired in October 2014
Fig. 4. Spatial and temporal behaviors of the ground surface displacement at LdM. A) Eascing and vertical components of the mean displacement observed between November  2014 to February 2017 obtained from the combination of the ascending and the descending
+7

Références

Documents relatifs

As an application of that abstract result, the approximate controllability and exponential stabilizability results for the interaction between a single mode Jeffreys fluid

Supplemental section S4 provides benchmarks of ADELI with so- lutions for a viscoelastic shell in an infinite medium (Dragoni and Magnanensi, 1989) and with solutions accounting for

Using the M-integral approach, we develop a method for sepa- rating mixed-mode fractures in viscoelastic media during the creep crack growth process.. The M-integral approach

Mechanical response under a moving load of a viscoelastic four-layer structure with a bonded or a slip boundary between layers 2 and 3: influence of the velocity on the normal

In Section 4 we discuss generalizations of our compressible heat-conducting model for Maxwell fluids to (i) models with an equivalent Lagrangian descrip- tion, in the spirit of

Time-series of GPS data (red) versus three models—orange: model of viscoelastic relaxation only with Maxwell rheologies (viscosities equal to the steady state viscosity of our

On the other hand, in Section 4, we prove that under some restrictions on the initial data and if the interior source dominates the boundary damping then the L p -norm of the

Truong, Existence and asymptotic expansion for a viscoelastic problem with a mixed nonhomogeneous condition, Nonlinear Analysis, Theory, Methods &amp; Applications, Series A: Theory