ANALYSIS OF EXPLOSIVE ACTIVITY OF TUNGURAHUA VOLCANO USING SEISMIC-ACOUSTIC DATA

Mario Calixto Ruiz

A dissertation submitted to the faculty of the University of North Carolina at Chapel Hill in partial fulfillment of the requirements for the degree of Doctor of Philosophy in the Department of Geological Sciences.

Chapel Hill 2007

Approved by, Dr. Jonathan M. Lees Dr. Allen F. Glazner Dr. Jeffrey B. Johnson Dr. Peter E. Malin Dr. Jose A. Rial

ABSTRACT

Mario Calixto Ruiz Romero

(Under the direction of Dr. Jonathan M. Lees)

“ANALYSIS OF EXPLOSIVE ACTIVITY OF TUNGURAHUA VOLCANO USING SEISMIC-ACOUSTIC DATA”

The character and mechanism of small vulcanian eruptions, often seen in andesitic volcanoes, have not been studied thoroughly. Activity of Tungurahua volcano, , is presented here using data from a temporal deployment of broad-band seismic stations and infrasonic microphones. This study focuses on three areas of volcano dynamics: a) structure of the magmatic plumbing system, b) the process of magma fragmentation inside the volcano conduit, and c) propagation of infrasound waves through a real medium. Using shear wave splitting of regional tectonic earthquakes, we found that anisotropic structures (micro-cracks alignments) are approximately perpendicular to expected tectonic orientations. The perpendicular orientation of these structures is likely related to magma intrusions in the upper part of the volcano conduit. Degassing at

Tungurahua is characterized by frequent short-duration explosions, typical of vulcanian style eruptions. Seismic and infrasound arrival times constrain the source of Tungurahua fragmentation to the shallowest section (upper 200 m) of the volcano edifice. While infrasound events cluster into families of similar waveforms, we found no temporal correlation nor relationship of event size with cluster association. The variation of amplitudes and arrival time differences between seismic and acoustic records suggests

ii that either the shear fragmentation occurs at different loci inside the conduit, or that the seismic and acoustic signals do not share the same source-time-function. In contrast to previous infrasound studies on volcanoes, we analyze the effects of wind and rough topography on the propagation of infrasound waves. Numerical simulations using Finite-

Difference Time-Domain methods confirm that wind has a significant influence on arrival times at stations around the volcano. Acoustic reduction is observed if infrasound is impeded by topographic barriers with dimensions similar to or larger than the infrasound wavelength. This research has a direct application to the important effort of volcano explosion monitoring and hazard reduction.

iii

DEDICATION

A mi mujer y mis hijos por ser el motivo y la inspiración de mi esfuerzo.

A mis padres y hermanos por todo el cariño y la confianza con que alimentaron mis sueños.

iv

ACKNOWLEDGES

I would like to thank Dr. Jonathan Lees for his help, support and teachings. I also thank Dr. Jeffrey Johnson for his constant advice and friendship.

v

TABLE OF CONTENTS

1. Introduction 1 1.1 Mechanisms of vulcanian eruptions 1 1.2 Characteristics of Tungurahua volcano 4

2. Structure of the magma plumbing system 8 2.1 Introduction 9 2.2 Seismic and volcanic activity 9 2.3 Seismic anisotropy on active volcanoes 12 2.4 Polarization of split shear-waves 15 2.5 Inversion of crack parameters 22 2.6 Temporal variation of delay times 26 2.7 Conclusions 28

3 Conduit processes 29 3.1 Introduction 29 3.2 Experiment description 31 3.3 Description of Degassing Signals 31 3.3.1 Explosion Events 32 3.3.2 Jetting tremor 33 3.3.3 Chugging signals 35 3.4 Clustering of Explosive Events 36 3.4.1 Clusters of infra-sonic signals 36 3.4.2 Clusters of seismic signals 39 3.5 Analysis of Source Locations 41 3.6 Conclusions 49

vi 4 Infrasound Propagation 51 4.1 Introduction 51 4.2 Characteristics of the explosive activity of Tungurahua 53 4.3 Finite Difference Time Domain Implementation 60 4.4 Results and discussion 63 4.4.1 Effect of topographic barriers 64 4.4.2 Effect of frequency 64 4.4.3 Wind effect on propagation with topography 64 4.4.4 Effect of source location 70 4.4.5 Effect of source’s characteristics 71 4.5 Conclusions 71

5 General conclusions 73 6 References 76

vii LIST OF TABLES

Table 2.1 TUNG1 crustal velocity model

Table 2.2 Results from cross-correlation of rotated horizontal components of seismic events containing shear waves.

Table 2.3 Results of circular statistic analysis of polarization angles of splitting shear-waves.

Table 2.4 Strike and dip solutions using inversion of polarization angles and delay- times.

Table 3.1 Description and location of instrumentation deployed on Tungurahua volcano during the 2004 field campaign.

Table 3.2 Distinctive characteristics of degassing signals at Tungurahua volcano.

Table 3.3 Number of blast events with clear arrival times of acoustic and seismic signals.

Table 3.4 Statistical estimates of explosion source depth, obtained from an analysis of delay times.

Table 4.1 Expected delay times and amplitude ratios between i and j stations (Ai/Aj exp)

viii LIST OF FIGURES

Fig. 1.1 Simplified scenario for vulcanian eruptions with a shallow fragmentation point.

Fig. 1.2 Historical description of Tungurahua April 23rd 1773 Eruption.

Fig. 1.3 Location of broad-band seismic-infrasonic stations (BB-IS) deployed on June- August 2004.

Fig. 2.1 A:Epicentral locations (circles) of tectonic events recorded by the Ecuadorean seismic network. B: Location of temporal seismic stations around Tungurahua volcano. Projection of incident rays at MAS (black dotted lines), RUN (black solid lines), and JUI (black dotted lines) are plotted. Most of the seismic rays recorded at MAS cross the central part of Tungurahua volcanic edifice.

Fig. 2.2 Cross-correlation method for shear-wave splitting

Fig. 2.3 Distribution of polarization direction of fast shear waves at stations MAS (A) and RUN (B)

Fig. 2.4 Cluster analysis of distribution of polarization angles at MAS using principal components (A) and typical dendogram signal (B).

Fig. 2.5 Residual contour plots using inversions of polarization parameters for obtaining crack parameters and delay times for MAS and RUN.

Fig. 2.6 Comparison between parameters of volcanic activity (RMS of amplitudes, maximum pressure disturbance, and amplitude) with delay times of split shear- waves.

Fig. 3.1 Monthly number of explosions recorded by the seismic network of Instituto Geofisico in log scale.

Fig. 3.2 Distribution of main types of seismic signals recorded on Tungurahua. Explosion Events (EX), Jetting tremors (RO), and Chugging (CH) events have acoustic signals.

Fig. 3.3 Waveforms and spectrograms of Explosion events (left column), Jetting tremor (center column), and Chugging events. Seismic signals are above and corresponding acoustic signals below.

Fig. 3.4 Cluster distribution of infrasonic records after applying a fuzzy logic process.

Fig. 3.5 Traditional dendogram tree showing the event clustering of acoustic signals of explosion blast (EX) events.

Fig. 3.6 Infrasonic records of each cluster of infrasound signals (A1, A2, A3, and A4).

ix Fig. 3.7 Cluster distribution of corresponding seismic records after applying a fuzzy logic process.

Fig. 3.8 Seismic records of clustered events.

Fig. 3.9 Time intervals between arrivals of acoustic pulse and first P-wave motion ∆T observed on MAS station at 3516 m (total distance) from the vent. Right panel: Distribution of ∆T times shows a poisson-like distribution.

Fig. 3.10 Schematic propagation model for EX type events in Tungurahua volcano.

Fig. 3.11 Schematic propagation model for EX type events in Tungurahua volcano.

Fig. 3.12 Distribution of residuals between the summation of observed delay times and the summation of expected times assuming a ground velocity α= 2.9 km/s for a range of U [10-300 m/s] and depths [0-400m].

Fig. 3.13 A) Boxplots of observed ∆T delay times. B) Density functions of EX source depths using a model with a vertical conduit (red line: based on ∆T at MAS; green line: based on ∆T at JUI; blue line: based on ∆T at RUN). Light gray area contains source depths between the 5th and 95th percentiles. Dark gray contains depths between the 25th and 75th percentiles.

Fig. 4.1 3-D perspective of Tungurahua volcano showing location of seismo-acoustic sensors.

Fig. 4.2 Distribution of horizontal velocities (A) and azimuthal directions (B) of Tungurahua’s ash-clouds with altitudes smaller than 7.5 km asl.

Fig. 4.3 Infrasound signals generated by a volcanic explosion at Tungurahua on July 21 (03h32 GMT), recorded on A, B, and C.

Fig. 4.4 Histograms of delay times between infrasonic arrivals at stations A-B, B-C, and A-C. Vertical dashed lines show the expected delay times using a constant infrasound speed and vent-station distances.

Fig. 4.5 Distribution of peak-to-peak amplitude ratios (Ai/Aj) of infrasonic signals generated by volcanic explosions of Tungurahua volcano.

Fig. 4.6 Left: Delay times (in %) observed on 2-Hz infrasound waves propagating with and without topography without wind (v=0). Right: Comparison between amplitudes of infrasound waves propagating with and without topography.

Fig. 4.7 Simulation of infrasound propagation from the crater floor to station A over real topographic conditions.

x Fig. 4.8 Simulated waveforms recorded in a radial profile from the crater rim (x=240 m from the source), passing through A station (x=3190 m). Source function has a frequency of 2 Hz. Waveform amplitudes are corrected by geometrical spreading attenuation.

Fig. 4.9 Contour plots showing the distribution of delay times between waveform simulations. Gray areas show the distribution delay times observed on real waveforms of explosion events. Dashed lines show the median value of delay time distributions. Dotted line mark one set of wind vectors that satisfies the observed delay times.

Fig. 4.10 Contour plots showing the distribution of amplitude ratios between 2-Hz waveform simulations at different stations. Gray areas show the distribution of amplitude ratios observed on real waveforms of explosion events. Dashed lines show the median value of amplitude ratio distributions. The set of wind vectors that better fits the observed delay times is given by dotted lines.

xi

LIST OF ABBREVIATIONS

σH maximum horizontal stress Φ polarization direction of the fast shear wave δt time difference between fast and slow shear waves

σh minimum horizontal stress α back-azimuth angles islope slope-corrected incident angle Vp, α velocity of p-waves Vs velocity of s-waves fl lower frequency cut-off values fh upper frequency cut-off values θi rotation angle snr signal-to-noise ratio κ concentration parameter of a circularly distributed data set So sample circular standard deviation R mean resultant length of a set of observations ε crack density RMS root-mean-square SWVA shear-wave velocity anisotropy

EX explosion event RO jetting tremor (roar event) CH chugging event σ correlation factors of distance matrix tc duration of compression phase tr duration of rarefaction phase ξ compression /rarefaction amplitude ratio Tac arrival time of acoustic wave Tp1 arrival time of acoustic wave ∆T delay time between arrivals of acoustic and seismic waves To origin time of explosion U rise velocity of this two-phase fluid inside the conduit z depth h elevation dh horizontal distance Π total acoustic power r distance vent-receiver ∆p pressure disturbance Km coefficient for acoustic propagation with a circular flat orifice Rb conduit radius ρair air density vej gas ejection velocities

xii θm mean circular angle ∆tA-B differences in arrival times between signals at stations A and B AA/AB ratio between amplitudes of signals at stations A and B p pressure fluctuations w velocity fluctuations ρ air density, c adiabatic speed of sound, v vectorial wind velocity, F source force acting on the medium (dipole sources), and Q mass source (monopole sources) κ adiabatic bulk modulus b mass buoyancy ∆x node spacing in x direction ∆y node spacing in x direction C Courant number

xiii

CHAPTER 1

INTRODUCTION

1.1 Mechanisms of vulcanian eruptions

In general, volcanic eruptions are controlled by physical (viscosity, temperature, density) and chemical (composition) properties of magmas. As magma rises because of buoyancy, the confining pressure drops below the supersaturation level, and its dissolved volatiles begin to exsolve and form bubbles. Generally this phenomenon occurs at a depth of a few hundred meters for H20 and larger depths for CO2 (Sparks, 1978). Once bubbles are formed, they grow by decompression and diffusion of volatiles from the melt (Dobran, 2001). Gas (H2O) solubility increases with depth exponentially, with the largest void fraction values at shallow depths

(Marchetti et al., 2004). The distribution of void fraction in the conduit is determined by the upward motion of the magma, volatile exsolution, and bubble degassing. The increase in void fraction induces a drop in density, sound velocity (Aki et al., 1978; Chouet, 1996), and bulk modulus, and a significant increase in the mixture viscosity (Jaupart and Vergniolle, 1989).

Magma viscosity, volatile content, and magma rise velocity, the parameters that determine the eruption character, are related to the chemical composition of magmas. Basic magmas, usually with high temperatures and low viscosity, erupt with Strombolian and Hawaiian styles, depending mainly on magma rise speed (Parfitt and Wilson, 1995). Coalescent bubbles can rise easily, and if they reach ascent velocity higher than the magma ascent velocity, then they can reach the upper part of the conduit, and in some cases even the surface, where they explode

(Parfitt and Wilson, 1995). In the other case, highly viscous silicic magmas do not exhibit bubble coalescence

(Sparks, 1978). Even so, they can reach void fractions exceeding the 70-83% level (Papale,

1998) at low confinement pressures. At this level, magma tends to explode inside the conduit with a fragmentation mechanism. A fragmentation wave can propagate downward forcing more closed gas vesicles to shatter. Simultaneously, a mixture of expanding gas and fragmentation products rapidly flow upward to the atmosphere (Alidivirov and Dingwell, 2000).

Short-lived (seconds to minutes) mild-size (<0.1 km3) vulcanian eruptions are very common at many strato-volcanoes clustered along subducting plate margins (Self et al., 1979). If the conduit is overpressured relative to the surrounding rock and magma porosity exceeds 60%, volatiles will escape through conduit walls. Degassing and crystallization increase the viscosity of the magma several orders of magnitude at the shallow part of the conduit, favoring the formation of a rock-plug (Melnik and Sparks, 2005). Reducing permeability at the upper part of the conduit leads to higher amounts of retained gas and increasing excess pressure. Prior to a vulcanian explosion, overpressure reaches a maximum in the uppermost few hundred meters of a volcanic conduit (Melnik and Sparks, 2005). If the caprock seal is not sufficient, gases will continue to exsolve and increase conduit pressure until an explosion occurs. Plug thickness and density, along with depth and magnitude of overpressure are controlled by magma flow rate, surrounding rock permeability and bubble connectivity threshold (Diller et al., 2006). According to Clarke (2002), vulcanian eruptions initiate when disruption of the caprock enables the expanding gases to rise and eject pyroclasts at speeds several times larger than the ascent rate of underlying magma. Two mechanisms of explosions have been envisaged: a) shear induced fragmentation emitting ash and gas through a ring-shaped opening and b) plug fragmentation that releases sufficient confining pressure to cause the partial ejection of the plug. Higher flow rates increase the magnitude and extent of conduit overpressure, decreasing the density and thickness of the confining cap, which may favor explosion over extrusion (Diller et al., 2006)

2 Mechanisms of volcanic eruptions have been studied using seismic and infrasonic instruments (Fig. 1.1). Seismic wavefields are generated inside the conduit and propagate through the ground. Infrasonic waves are generated at the vent (conduit-atmosphere interface) and propagate through the atmosphere for events with muzzle ejection velocities smaller than sound velocity in air. These infrasonic signals exhibit little atmospheric scattering, while seismic signals are more heavily altered by propagation effects (Garces et al., 1999; Johnson et al.,

2004). Strombolian activity has been extensively studied using seismic and infrasonic data

(Caplan-Auerbach and McNutt, 2003; Johnson and Aster, 2005; Ripepe et al., 2001; Vergniolle and Brandeis, 1994). A smaller number of volcanoes exhibit vulcanian eruptions, and consequently, there is little infrasonic data on vulcanian events. Infrasound records of vulcanian eruptions have been recorded on Sakurajima (Garces et al., 1999), Suwanosenjima (Iguchi and

Ishihara, 1990), Tungurahua (Johnson et al., 2003), and Fuego (Johnson et al., 2004).

Fig. 1.1. Cartoon depicting a simplified scenario for vulcanian eruptions with a shallow fragmentation locus (S), where seismic signals are originated. Infrasound is generated at the crater floor (A) and is recorded at a microphone on the volcano flank.

3 1.2 Characteristics of Tungurahua volcano

Tungurahua (1.45°S, 78.43°W, 5023 m) is an andesitic strato-volcano in the Real

Cordillera of Ecuador, one of the most active volcanoes of the Northern . Its current cone has a steep flank (30-35° slopes) and up to 3,200 m of local vertical relief. A 300 m-wide and

100-m deep crater is located on the upper part of its northwestern flank (Samaniego et al., 2003).

A high temperature pit, with a diameter no larger than 25 m (estimated from FLIR images taken by Ramón and Boker on 03/20/2003 in Samaniego et al.(2003)), can be related to the active vent.

The current volcanic cone is built over two ancient edifices that were destroyed by large avalanche blasts (Hall et al., 1999). Tungurahua volcano experienced at least 17 major eruptions in the last three millennia (Le Pennec et al., 2006). Since the Spanish conquest (1534),

Tungurahua has had five major eruptive periods (1640-1641, 1773-1777, 1886-1888, 1916-1918,

1999-present) (Egred, 2004; Hall et al., 1999). Generally, these eruptions were characterized by tephra-and-ash falls covering the volcano flanks, (especially on the western slopes), pyroclastic flows, lava flows down the north-west and south valleys, and lahars (Fig. 1.2). Lavas from the last five eruptive periods have a silica content of 57-65% (Hall et al., 1999). Vulcanian explosions and transitions between vulcanian to strombolian styles dominate the eruptive activity of

Tungurahua.

Appearance of volcanic tremor in 1993 preceded the last eruptive period (Ruiz et al.,

1999) whose first explosion occurred on October 5th 1999, followed by significant activity in

November-December 1999 (Molina, 2001). On October 17th 1999, the alert level at Tungurahua

Volcano was raised to Orange, leading to an evacuation of over 26,000 people from Baños, a town 8 km from the summit, as well as other villages surrounding the volcano (Tobin and

Whiteford, 2002). Activity in November and December 1999 was dominated by ash-and-tephra falls on the West flank and mushroom shaped columns, sometimes extending higher than 7 km.

During this time period, however, no pyroclastic flows occurred. After November, volcanic

4 activity decreased and the majority of evacuees returned to Baños on January 5th, 2000. A climatic eruptive phase occurred in July –August 2006 with strong ash emissions, widespread pyroclastic flows and lava flows.

Fig. 1.2 Historical description of Tungurahua April 23rd 1773 Eruption (J. Egred, unpublished document).

Between the initial activity and the climatic phase, Tungurahua presented four additional explosive cycles with smaller levels of volcanic activity between 2000 and 2006 with more than

6,000 volcanic explosions. Small eruptive cycles were characterized by ash emissions, and vulcanian and strombolian activity. Between large activity peaks, a persistent degassing has been observed with a background in SO2 concentrations of 582 tons/day (5th percentile of monthly averages of SO2 measurements). Despite minor fumarolic activity on the northern flank, practically all degassing events occur through the crater. Three seismo-volcanic phases can be distinguished in the fourth cycle of volcanic activity: a) a vent-clearing phase with the occurrence of large explosions and increasing seismic activity [julian days 181-194]; b) a degassing phase with low seismic activity and intense explosive activity [days: 194-215], and c) a vent-sealing

5 phase, where a decrease of magma input allows formation of a magma plug, retention of gas bubbles, and occasional generation of large explosions [days: 215-226].

Besides a permanent surveillance system maintained by local institutions, a temporary seismic and infrasonic station was deployed during the highest active cycle (Oct. 1999) on a ridge just in front of the northwestern flank. This instrument recorded a series of explosion events and long-duration degassing signals (Johnson et al., 2003; Johnson et al., 1999). In addition, an electret microphone installed at the headquarters of the Tungurahua Volcano Observatory (~13 km from the vent) recorded 14 infrasonic explosion signals from August 11 through August 31,

2000.

In late June-July 2004, three temporary stations with collocated seismic and infrasonic sensors were deployed on SW, NE and SW flanks. These instruments were located at horizontal distances between 3.20 and 5.51 km from the active vent (Fig. 1.3). More than 2,000 seismic events with a volcanic origin were recorded between June 29 and August 12, 2004. Despite this activity level, approximately 25,000 people reside in high risk areas such as Baños and small villages and farms on the volcanic flanks.

Using seismic, infrasonic, geochemical data, and visual observations collected during field campaigns, I will analyze the explosive activity of Tungurahua volcano in order to characterize the different degassing types, locate the explosion sources, characterize their mechanisms, and study their temporal evolution. The present analysis contains three parts:

a) A study of the anisotropic structures beneath the present cone. Using shear-wave splitting of regional earthquakes we expect to define the orientation of stresses and micro-porous structures that define the feeding system of Tungurahua volcano.

b) Using seismic and infrasonic records, we will define the different degassing modes, and will constraint the source location of the explosion events that characterize the vulcanian eruptions of Tungurahua.

6 c) Considering that wind and strong topographic conditions are present in andesitic volcanoes, such as Tungurahua, we will discuss the effects of topography and moving atmosphere on the propagation of infrasound generated by volcanic explosions.

Conclusions of these studies could be useful for hazard assessments and monitoring activities in active volcanoes.

Fig. 1.3 Location of broad-band seismic-infrasonic stations (BB-IS) deployed on June-August 2004 on SW (MAS), NW (JUI) and NE (RUN) flanks of Tungurahua volcano. Vertical and horizontal axes are in kilometers.

7

CHAPTER 2

STRUCTURE OF THE MAGMA PLUMBING SYSTEM

In volcanic conduits, magma rises due to its lower density compared to the surrounding medium. Silicic magmas usually have central vents that are fed by many dikes with widths ranging from less than 10 cm to more than 1 m (Dobran, 2001). Porous and fractured media with prevalent orientations can also be considered magma plumbing structures. This chapter analyzes the structures that may compose the magmatic feeding system of beneath Tungurahua volcano.

Different techniques have been used to map the magma plumbing systems. Tomographic studies have been used for mapping volcanic structures. However, tomographic inversions do not have resolutions finer than 0.5 km. Using a tomographic inversion of volcano-tectonic events, a high- velocity central zone and lateral high-velocity elongations beneath the southern and northern flak were found in Tungurahua. In this chapter, we will study the seismic anisotropy beneath

Tungurahua in order to find the distribution of the probable main magmatic conduits and its correlation with present activity. This chapter is part of a paper entitled: “Seismic Anisotropy beneath Tungurahua volcano” submitted to the Journal of Volcanology and Geothermal Research by Ruiz M., Lees J., and Yang M.

2.1 Introduction

Seismic anisotropy can provide detailed information about the state of stress in volcanic regions by examining the effect of faults and cracks on shear waves as they pass through regions of intensely anisotropic media. As shear waves propagate through an effective anisotropic medium (faults, cracks, or aligned microcracks), they split into two or more components with different velocities and different polarizations (Crampin and Booth, 1985). Once split shear waves leave the anisotropic medium they retain a time separation between the faster and the slower waves.

Shear wave analyses have been applied to anisotropic structures in the whole mantle, upper mantle, and crust (Babuska and Cara, 1991; Crampin and Chastin, 2003; Crampin and

Lovell, 1991; Crampin and Peacock, 2005; Hiramatsu et al., 1998; Savage, 1999). Recently, shear-wave analysis has been used to evaluate the stress state of active volcanoes for possible midterm eruption forecasting (Gerst and Savage, 2004). In this chapper, we demonstrate the presence and describe the characteristics of local seismic anisotropy beneath Tungurahua volcano using shear-wave splitting of regional seismic events during an active volcanic cycle in the summer of 2004. Tungurahua volcano was chosen for this study because it is one of the most active andesitic volcanoes of the volcanic arc in Ecuador and is surrounded by active seismo- tectonic sources.

2.2 Seismic and volcanic activity

Rapid subduction of the Nazca plate (58 ± 2 mm/a; (Kreemer et al., 2003; Trenkamp et al., 2002)) along the Ecuador-Colombia trench is the primary driving force of Ecuadorean mainland tectonics. The oblique convergence of the Carnegie ridge with the continental crust causes the tectonic escape of the northwestern part of (North-Andean Block) toward the NNE (Ego et al., 1996; Pennington, 1981). The southern border of the North-Andean

9 Block is a set of dextral faults that crosses Ecuador from the Colombia Sub-Andean zone to the

Gulf of Guayaquil (Ego et al., 1996; Eguez et al., 2003). Besides these transcurrent faults, two sets of reverse and thrust faults align along a NNE direction including faults and folds in the interior of the Inter-Andean Valley (Ego et al., 1996; Winter, 1990) and a set of transpressional faults along the Sub-Andean zone (Eguez et al., 2003). These three fault systems constitute the source of intense seismic activity employed in the present analysis (Fig. 2.1).

Focal mechanisms of seismic events located in the central part of the Ecuadorean Arc show a general ESE-WNW trending direction of horizontal compression (Segovia, M., comm. per., 2005). The same compressive direction (N121°) is obtained from the summed moment tensor for the southern part of the North-Andean block (Ecuador) (Corredor, 2003). Using P wave-polarities, a simultaneous inversion of focal mechanisms for a unique moment tensor was performed for a tectonic swarm near Guagua volcano ( seismic swarm) (Legrand et al., 2002), indicating a major horizontal compressive stress σH, oriented N116.5°. Focal mechanism inversions of shallow events in the Sub-Andean Seismic Zone of Ecuador yield an E-

W trending sub-horizontal compression (σH = 99°) (Ego et al., 1996). All current geologic and seismic evidence suggests that regional tectonic stresses in the central part of continental Ecuador have an ESE-WNW orientation.

Stress tensors in areas close to active volcanoes may differ substantially from regional orientations. For example, the major compressive stress in Guagua Pichincha volcano, Ecuador, shows a σH orientation of N213°, approximately 97° from the σH orientation of the adjacent Quito swarm (Legrand et al., 2002). Other examples include significant variations of stress orientation between earthquakes beneath Mt. Rainier and events located in the surrounding areas

(Giampiccolo et al., 1999). Many studies of local stress regimes beneath active volcanoes (Crater

Peak, St.Helens, Etna, Usu, Unzen, and Soufriere Hills) show a ~90 rotation with respect to the maximum regional horizontal compression (Roman and Cashman, 2006). Seismic events with a

10 ~90° rotated stress tensor were observed in a swarm beneath Aoba volcano, Ambae Island

(Rouland et al., 2001). Using inversions of focal mechanisms prior to and during the 1992 eruption of Crater Peak, Alaska, Roman et al (2004) found a strong temporal correlation between a ~90° horizontal rotation of p-axes and periods of magmatic activity. A compilation of stress orientations on active volcanoes show that a ~90° rotation occurred prior, during, or after the main eruptive phase (Roman and Cashman, 2006). We present here a comparison between the regional stress orientation and polarization orientation observed at Tungurahua volcano during its eruptive activity in 2004.

Tungurahua volcano experienced at least 17 major eruptions in the last three millennia

(Le Pennec et al., 2006). Since the Spanish conquest (1534), Tungurahua has had five major erupting periods (1640, 1773-1777, 1886-1888, 1916-1918, 1999-present), including ash columns, ash falls, lahars, and occasional pyroclastic and lava flows (Egred, 2004; Hall et al.,

1999). During the last period of activity (1999-present) about 6,000 explosions were recorded by the Instituto Geofisico-Ecuador.

To augment the permanent volcano monitoring system at Tungurahua, three broad-band seismic stations (MAS, JUI, and RUN) were deployed on the volcano flanks during julian days:

184 to 226, at horizontal distances between 3.2 to 5.5 km from the active vent (Fig. 2.1-B).

Nearly 1,400 explosions, more than 600 tremor-like acoustic signals, and 145 tectonic earthquakes were recorded over 46 days of operation (Ruiz et al., 2006). Three seismo-volcanic phases can be distinguished in the fourth cycle of volcanic activity: a) a vent-clearing phase with the occurrence of large explosions and increasing seismic activity [julian days 181-194]; b) a degassing phase with low seismic activity and intense explosive activity [days: 194-215], and c) a vent-sealing phase, where a decrease of magma input allows formation of a magma plug, retention of gas bubbles, and occasional generation of large explosions [days: 215-226].

11

Fig. 2.1 A: Epicentral locations (circles) of tectonic events recorded by the Ecuadorean seismic network. Tungurahua volcano is marked with a triangle. Quito city is marked with a white square. Contour topographic lines are every 1,000 m. B: Location of temporal seismic stations around Tungurahua volcano. Projection of incident rays at MAS (black dotted lines), RUN (black solid lines), and JUI (black dotted lines) are plotted. Most of the seismic rays recorded at MAS cross the central part of Tungurahua volcanic edifice. Topographic contour lines are every 500 m. Black bar in the upper left corner shows a 5 km scale.

2.3 Seismic anisotropy on active volcanoes

The alignment of fluid-filled micro-cracks and porous space can result in splitting bi- refringence of near-vertically traveling S-waves into two approximately orthogonal polarizations

12 with different propagation velocities; a fast one polarized parallel along the dominant crack direction, and a slow one polarized perpendicular to it (Babuska and Cara, 1991; Crampin and

McGonigle, 1981; Gerst and Savage, 2004; Tang et al., 2005). Regardless of the initial S-wave polarization, the polarization direction of the fast shear wave (Φ) is parallel to the strike of the dominant crack orientations (Crampin et al., 1986), and approximately parallel to the direction of maximum horizontal compressive stresses (Crampin et al., 1986; Crampin and Chastin, 2003).

As fractures open parallel to the minimum compressional stress, the time difference (δt) between fast and slow components provides a relative measure of stress σH (Crampin, 1987; Crampin et al., 2004). Fluctuations in time-delays between the split shear-waves have been used to directly monitor stress-induced changes (Crampin, 1994). The differential time delay between the arrivals of the fast and slow shear waves (δt) is also closely related to crack density and crack aspect ratio

(Crampin, 1987; Hudson, 1981; Tang et al., 2005).

To avoid waveform distortions of incoming shear waves, or contamination with reflected

S-to-P converted phases (Nuttli, 1961), split shear-waves should be located inside a so-called

“shear-wave window cone” which is the cone of ray paths with angles of incidence less than 35-

45° from the free surface (Crampin and Chastin, 2003; Crampin and Peacock, 2005). For angles of incidence within this window, the behavior of shear waves at the surface is similar to that of the incident wave (Booth and Crampin, 1985; Crampin and Lovell, 1991). For angles of incidence outside this window, shear waves have such severe interactions with the free surface that almost all similarities with the incoming waveform are irretrievably lost (Booth and Crampin,

1985; Crampin and Lovell, 1991). The effect of the topography on the determination of incident angles is addressed below.

Stresses related to eruptive activity can influence the alignment of fluid-filled micro- cracks, and consequently cause seismic anisotropy (Gerst and Savage, 2004). For example, shear-wave splitting has been observed in volcanic regions such as the East Rift Zone of Hawaii

13 and the Phlegraean Fields (Savage et al., 1989). In the Phlegraean Fields, Savage et al. (1989) reported that the shear-wave polarizations were consistent with the inferred stress field direction.

Booth et al. (1992) also found shear-wave splitting in an experiment conducted in the Kaoiki region, Hawaii. Munson et al. (1993; 1995) found a velocity anisotropy exceeding 10% in southern Hawaii. Bianco et al. (1998; 1999) used shear wave splitting on seismic events from

1993 to 1996 to show that the fast split directions are closely parallel to the main volcanic fault system of Mt. Vesuvius. Volti and Crampin (2003) reported the presence of shear wave splitting in volcanic areas of Iceland. An increase of delay times between shear-waves was related directly to the end of the Vatnajokull eruption in October 1996 (Volti and Crampin, 2003). At Mount

Ruapehu, New Zealand, Miller and Savage (2001) found indications for a temporal change in anisotropy associated with changes in pore fluid pressures or changes in crack orientations due to a significant volcanic activity. Additional changes in the orientation of shear-wave polarizations were observed several years later on Mount Ruapehu when conditions on the volcano returned to a pre-eruption stress state (Gerst and Savage, 2004; Miller and Savage, 2001). Gerst and Savage

(2004) proposed that the source of anisotropy changes at Ruapehu was a change of local stresses caused by a magmatic injection in a dike (or a swarm of subparallel dikes) parallel to regional σH.

This injection caused a rotation of local stresses and the crack orientation to be aligned perpendicular to regional σH, and therefore parallel to minimum horizontal stress (σh) (Gerst and

Savage, 2004).

A distinctive change in fast-shear wave polarization, known as 90º-flips, has been observed in areas with high pressure fluid injections or above seismically active faults (Crampin and Peacock, 2005; Crampin et al., 2002). According to numerical models, this change occurs when the pore-fluid pressures reach levels of fracture criticality (Crampin and Peacock, 2005;

Crampin et al., 2002). Gerst and Savage (2004) do not discard the possibility that high-pressure pore fluids can generate such rotations of Φ.

14

2.4 Polarization of split shear-waves

To examine the presence of anisotropy beneath Tungurahua volcano, we measured Φ and

δt on tectonic seismic events recorded on three flanks of the volcano. Between June 29 and

August 13, 2004 (julian days: 181 and 226; hereafter dates will refer to julian days), 145 regional events were recorded at three broad band temporary stations (Fig. 2.1) (Ruiz et al., 2006). Most of these regional events are located along the Pisayambo fault zone, the most active part of the

North-Andean block boundary in the center of Ecuador (Guillier et al., 2001). Reliable hypocentral locations of 69 events were obtained by the Instituto Geofisico-Ecuador using a regional seismic network of 43 seismic stations. The velocity model used for ray-tracing

(TUNG1) is a combination of the 1-D local velocity model (<20 km) (Molina et al., 2005) and the regional ASW velocity model for Ecuador (>20 km) (Calahorrano, 2001) (Table 2.1). An independent three-dimensional P-wave velocity model confirms the presence of low velocities

(~3 km/s), especially in the north-west and southern flanks, above a basal high velocity layer near sea level (Molina et al., 2005). Back azimuth angles (α) are computed using regional epicenters determined by the Instituto Geofisico-Ecuador. Incident vectors on flat surfaces were calculated after computing the travel ray-paths between the hypocenter and the recording station with the

TUNG1 velocity model. The rough topography on the volcano has considerable gradients (Fig.

2.1B) so a slope-corrected incident angle (islope) was computed taking into account the normal vector to the surface and the incident vector at a flat surface. The normal vector to the slope around a seismic station was calculated by averaging topography values in an 800 m2 area centered on the station. This area included at least 2 wavelengths of incident shear-wave, since most tectonic events have a spectral energy peak around 5 Hz and upper layer S-wave velocity of

1.73 km/s (Table 2.1). We have assumed a Vp/Vs ratio of 1.73 (Poisson’s ratio = 0.25) in all our estimates.

15

Layer Layer top Vp(km/s) Vs(km/s) (km) 1-3 3 1.73 2 -2.5 4 2.31 3 1.5 5.5 3.18 465.9 3.41 5166.3 3.64 6276.7 3.87

Table 2.1 TUNG1 crustal velocity model. Depth datum is the sea level. Layer top is negative for altitudes above sea level.

From 69 regional tectonic events exhibiting split shear-waves, only 47 have an incident angle (islope) inside the shear-wave window and acceptable signal-to-noise ratio of S-waves (Table

2.2). Split shear waves of 24 of these events were recorded at MAS, 20 at RUN, and 3 at JUI, with 11 events recording simultaneously at 2 stations, and one at three stations. All selected events had signal-to-noise ratios greater than 2.1 (Table 2.2). Epicentral distances to stations

MAS and RUN have median values of 40 km and 32 km, respectively. In both cases, the distribution of epicentral distances has interquartile ranges of 2.3 km, showing that most of the events used in this anisotropy analysis come from a small epicentral area. Frequency peaks of S- waves are in the range of 2.1 – 8.7 Hz at 95% confidence (Table 2.2). Having most events within the same frequency range reduces the effect of frequency-dependent anisotropy, observed in other studies (Gerst, 2003; Marson-Pidgeon and Savage, 1997). In order to include most of the energy of S-waves, signals were band-pass filtered with variable lower (fl) and upper (fh) frequency cut- off values, chosen by examination of the spectral content of incoming S-waves. Median values for fl and fh were 3 and 9 Hz, respectively. Each signal was integrated to displacement in the time domain, and windowed to include the onset of S-waves and at least the complete first cycle of the

S-wave arrival (Fig. 2.2). Integration of filtered velocity waveforms acts as a low-pass filter that

16 lessens the effects of high frequency noise, improving sensitivity at low frequencies. Lengths of data windows used for cross-correlation were typically in the range of 0.4 to 2.4 s.

Event 210.084 JUI rotated waveforms

5 Sfast Sslow 1. 40 NS 0 .5 s 0 0 4 dis um 5 vel um/ EW 0. 0

x1 dt=80 ms 40

1.5 x2 shear wave window

02040 0 500 1000 2000 3000

time s time ms

hodogram Cross-correlation function 6 0. 2 0. .2 0 .0 0 dt (s) x1 um 2 2 0. 0. 0.4 0.6

1.0 0. 5 0.0 0.5 50 100 150

x2 um angle (degrees)

Fig. 2.2 Upper panels show the horizontal components (NS-EW) of velocity ground motion waveforms (left) of event at 210.84 (07/28/04) recorded at JUI. Ground displacement waveforms (right) rotated Φ degrees (angle with the maximum cross-correlation between horizontal components). Low-left panel shows the hodogram with the vertical component in the direction of the fast shear-wave. Lower-right panel shows the distribution of cross-correlation function for polarization angles [0-180°] and normalized delay times of splitting shear-waves. Maximum cross-correlation point is marked with a black dot.

17 Delay times and polarization angles were computed using a cross-correlation approach

(Bowman and Ando, 1987; Fukao, 1984). At each rotation angle, θi (i from 0 to 180°), the cross- correlation between the rotated horizontal components was determined, producing a grid of cross correlation coefficients and delay times (Gao et al., 1998). The θi rotation angle with the highest correlation was taken as Φ, the orientation that best resolved the shear-wave splitting (Bowman and Ando, 1987; Fukao, 1984; Gao et al., 1998). Delay times (δt) are given as time lags at the orientation angle with highest correlation (Fig. 2.2, Table 2.2). Events with incidence angle larger than the shear-wave window, low signal-to-noise ratios, low correlation coefficients, or very unstable solutions were not included (Table 2.2).

18 Event julian day lat lon depth mag code fmax islope α snr r Φ δ t 1 186.324 -1.01 -78.05 13.1 3.7 3 6.1 28.9 42.1 2.8 0.75 105 56 2 191.494 -1.04 -78.29 14.3 3.6 1 5.3 23.0 21.8 5.1 0.98 78 8 3 191.496 -1.04 -78.28 8.8 3.6 1 6.3 23.4 22.9 4.5 0.85 98 8 4 192.767 -1.06 -78.32 9.0 3.9 1 6.1 24.1 19.2 3.3 0.80 52 8 5 193.410 -1.22 -78.25 14.8 3.1 1 6.8 18.9 39.0 3.8 0.88 91 16 6 200.220 -1.08 -78.25 28.0 3.1 1 6.8 17.8 28.3 3.2 0.87 3 8 7 200.241 -1.20 -78.25 11.2 3.5 1 6.8 20.2 37.6 5.8 0.92 17 8 8 200.330 -1.91 -77.94 185.1 4.6 1 0.7 11.3 128.6 7.2 0.88 69 48 9 200.849 -1.19 -78.24 8.4 3.8 1 6.8 20.7 37.8 6.5 0.94 97 8 10 200.953 -1.21 -78.23 11.0 3.2 1 6.8 19.8 40.8 3.0 0.96 37 8 11 201.692 -1.22 -78.44 16.1 4 1 4.4 23.7 6.2 2.8 0.87 15 16 12 203.507 -2.63 -77.04 7.0 4 3 3.4 15.2 131.1 3.7 0.81 95 24 13 207.524 -2.51 -79.19 43.2 4.1 3 6.8 31.3 215.5 5.1 0.78 124 304 14 209.243 -1.22 -78.30 25.3 3.2 2 8.0 32.4 37.3 4.9 0.75 93 24 15 210.084 -0.97 -79.24 113.7 4.2 3 1.3 30.4 299.1 9.3 0.94 39 80 16 210.084 -0.97 -79.24 113.7 4.2 2 1.6 29.7 301.0 6.4 0.87 27 72 17 212.188 -1.22 -78.21 18.6 4.5 1 4.2 9.3 44.5 8.8 0.97 100 8 18 212.188 -1.22 -78.21 18.6 4.5 3 4.6 23.7 46.5 10.3 0.95 16 8 19 212.188 -1.22 -78.21 18.6 4.5 2 5.3 34.9 49.4 4.5 0.93 40 8 20 212.204 -1.21 -78.21 9.3 3.9 3 4.8 27.5 45.2 7.5 0.86 15 8 21 212.214 -1.21 -78.22 9.0 3.4 1 6.8 19.7 43.2 5.0 0.92 13 16 22 212.214 -1.21 -78.22 9.0 3.4 3 4.6 27.6 44.9 5.4 0.92 26 16 23 212.627 -1.22 -78.23 14.8 3 1 5.6 18.5 42.6 3.4 0.77 15 16 24 212.627 -1.22 -78.23 14.8 3 3 4.6 25.7 44.2 4.6 0.90 26 8 25 212.646 -1.22 -78.22 14.8 3.1 1 6.6 18.5 43.5 3.4 0.74 18 8 26 212.646 -1.22 -78.22 14.8 3.1 3 4.6 25.6 45.3 4.9 0.95 29 16 27 212.870 -1.22 -78.24 14.8 3 1 6.6 18.9 40.2 3.9 0.87 10 8 28 212.870 -1.22 -78.24 14.8 3 3 8.7 26.2 41.1 5.6 0.83 12 8 29 213.090 -1.22 -78.23 14.8 3.2 1 5.8 18.7 42.3 4.4 0.92 98 16 30 213.090 -1.22 -78.23 14.8 3.2 3 6.6 25.9 43.8 4.4 0.90 14 16 31 213.710 -1.23 -78.22 14.8 3.2 3 6.8 25.5 45.3 5.5 0.92 16 16 32 213.710 -1.23 -78.22 14.8 3.2 1 6.8 25.5 45.3 3.5 0.82 66 8 33 215.631 -1.27 -78.46 13.0 3.2 1 5.1 26.0 3.0 3.7 0.73 5 16 34 215.808 -1.50 -78.14 12.0 NA 1 6.1 18.9 93.6 6.0 0.70 1 16 35 216.537 -1.18 -78.25 10.3 3.9 1 5.6 20.8 35.6 2.6 0.74 6 8 36 216.537 -1.18 -78.25 10.3 3.9 3 4.1 29.4 35.3 3.9 0.77 21 48 37 216.560 -1.19 -78.25 14.8 2.9 3 9.5 27.5 37.2 2.1 0.80 62 64 38 216.819 -1.20 -78.27 14.8 2.9 1 5.6 19.7 34.8 2.3 0.75 20 8 39 216.819 -1.20 -78.27 14.8 2.9 3 4.1 27.8 34.3 2.3 0.87 46 16 40 217.010 -1.21 -78.29 14.5 2.9 3 4.6 27.8 34.3 4.0 0.84 30 16 41 218.133 -1.20 -78.26 9.8 3.5 1 5.1 20.8 35.7 2.9 0.90 108 8 42 218.133 -1.20 -78.26 9.8 3.5 3 4.1 29.4 35.3 2.2 0.84 72 56 43 219.317 -1.21 -78.22 16.0 3.5 1 5.8 18.1 42.5 2.9 0.90 99 8 44 219.317 -1.21 -78.22 16.0 3.5 3 4.6 25.5 44.0 4.5 0.80 178 24 45 219.897 -1.22 -78.23 22.0 3.4 3 4.6 22.7 43.0 6.0 0.80 43 32 46 222.430 -1.21 -78.22 12.8 0 3 4.6 26.7 44.0 3.9 0.83 100 80 47 223.531 -1.22 -78.24 15.2 0 1 5.8 18.6 40.9 3.6 0.91 105 24

Table 2.2 Results from cross-correlation of rotated horizontal components of seismic events containing shear waves. Numbers 1, 2, and 3 of the seventh column correspond to stations MAS, JUI, and RUN,

respectively. Back-azimuths (α), slope-corrected incidence angles (islope), and split shear-wave orientations (Φ) are given in degrees in clock-wise direction. Delay times (δt) are in milliseconds (ms). Depth is in km and spectral peak frequency in Hz. Signal-to-noise ratio (snr) is non-dimensional. Maximum correlation coefficient is given by r.

19 Assuming a von Mises distribution of polarization angles, we applied the Rayleigh test

(Davis, 2002; Mardia, 1972) to check the randomness of the distribution of Φ at each station.

Rayleigh test showed that distributions of Φ at MAS and RUN are not random at 99% confidence

(Fig. 2.3). Using circular statistics we found that shear-wave polarizations at RUN have a significant preferential trend with mean direction Φ = 47.5° and a circular standard deviation error (Gaile and Burt, 1980) So=42.8° (Fig. 2.3, Table 2.3).

MAS: 24 events RUN: 20 events B A 0 10 0 4 330 30 330 30

300 5 60 300 2 60

270 90 270 90

240 120 240 120

210 150 210 150 180 180

Fig. 2.3 Distribution of polarization direction of fast shear waves at stations MAS (A) and RUN (B). Each bean is 10º wide. Occurrence number for each bean is in italics.

20

Set # events R κ Φ (°) So (°)

MAS-A 13 0.97 16.93 16.3 13.6

MAS-B 11 0.97 16.93 91.8 13.6

RUN 20 0.76 2.45 47.5 42.8

Table 2.3 Results of circular statistic analysis of polarization angles of splitting shear-waves. R is the mean resultant length of a set of observations. κ is the concentration parameter of a circularly distributed data set. Φ is the angular average of all polarization angles. So is the sample circular standard deviation. Polarization angles at MAS were divided into two subsets (MAS-A and MAS-B).

At station MAS, polarizations exhibit a bimodal distribution. Fuzzy-logic cluster analysis (Kaufman and Rousseeuw, 1990; Pison et al., 1999) of distances between polarization angles at MAS suggests the presence of two clusters (Fig. 2.4). One cluster, MAS-A contains 13 events and has an orientation mean Φ=16.1°. Orientation of the second cluster, MAS-B, with 11 events, is approximately perpendicular to the former one with a mean value Φ=91.8°. In both cases, the circular standard deviation (Gaile and Burt, 1980) is So=13.6° (Table 2.3). Since the distance between the two means is greater than two standard deviations, we may assume that polarization angles at MAS have a bimodal distribution (Reschenhofer, 2001). The anisotropy derived orientations of MAS-A and RUN are almost perpendicular to the regional orientation of tectonic stresses in central Ecuador. Polarization angles of subset MAS-B, however, have a mean direction close to the orientation of σH found in regional studies (Corredor, 2003; Ego et al.,

1996; Legrand et al., 2002). Fast-shear wave polarization angles (Φ ) with a prevalent orientation regardless of azimuth or incidence angle are expected for shear-wave splitting originating in a single system of thin, parallel, vertical cracks (Elkibbi et al., 2005). More complex subsurface fracture configurations such as parallel non-vertically dipping crack systems or intersecting multi- crack systems can generate non- uniformly parallel Φ distributions. The shallower the crack dip,

21 the more likely fast shear-wave polarizations will deviate from the direction of crack strike

(Elkibbi et al., 2005).

A Distance between shear wave polarization angles at MAS B Distance of fast shear wave polarization angles

22

19 0 0 0 5 1

1 24 18 20 0 11 23 162 15 8 08 0 12 1310 4 6

14 06 21 Distance Component 2 10 04 2 1

9 0

7 3 9

17 1 3 4 20 7 5 15 6 8 2 21 22 24 17 19 12 18 20 14 11 23 10 13 16

60 40 20 0 20 40 60 Component 1 Seismic Events at MAS These two components explain 100 % of the point variability.

Fig. 2.4 Cluster analysis of distribution of polarization angles at MAS. A: Two clusters of polarization angles are seen in the principal-components plot. B: Dendogram of distribution of distances between polarization angles at MAS, showing two clusters

2.5 Inversion of crack parameters

Multiple observations at each station provide an opportunity to estimate properties of cracks beneath stations. In this study, we apply the inversion scheme introduced by Yang et al.

(2005) to invert the observed data (Ф and δt) for crack orientation (crack strike and dip angle) and crack density (ε). Since we intend to estimate crack properties from two separate datasets (Ф and

δt), a double-response regression problem is proposed (Draper and Smith, 1998). In general, it is unlikely that both datasets provide the same regression solution. For this reason, the inversion scheme divides the original double-response inversion problem into two connected single- response problems by taking advantage of the inherent characteristics of Ф and δt (crack orientation is more sensitive to polarization distribution; crack densities are more sensitive to delay times) (Yang et al., 2005). The best fitting fracture model relies on simultaneous

22 minimization of both Ф and δt residual functions in the model space of crack strike and dip for different crack densities. The minimization is achieved via a non-linear least-squares algorithm, whereby the goodness-of-fit of a given crack model is evaluated based on the following root mean square (RMS) estimates (Elkibbi et al., 2005):

1 2 ⎡ 1 N ⎤ 2 RMS(Φ) = ⎢ ∑()()Φ0 n − (Φt )n ⎥ ⎣ N − 2 n−1 ⎦

1 2 ⎡ 1 N ⎤ 2 RMS()δt = ⎢ ∑()()δt0 n − (δtt )n ⎥ ⎣ N − 2 n=1 ⎦

where Φ0 and Φt are observed and theoretical fast shear-wave polarization angles respectively, δt0 and δtt are observed and theoretical time delays, respectively. N is the number of observations recorded at a given seismic station. Refer to Yang (2004), Yang et al. (2005), and Elkibbi et al.,

(2005) for details.

Crack parameters were thus estimated via inversion of polarization data sets recorded at stations MAS and RUN. Two low residual areas were found on inversions of crack parameters of

MAS. The area that contains the minimum residual point (crack azimuth N47º, dip 68º, crack density of 0.014) is delimited by the azimuths N30°-N60° with dipping angles larger than 60°

(Fig. 2.5A). This range is significantly different from the orientation of maximum horizontal regional stresses (Table 2.4, Fig. 2.5A). The second low residual area corresponds to a crack system with a NW trend, nearly orthogonal to the former low residual area. Inversion of delay times also shows low-residuals in areas with large dip angles (Fig. 2.5B). Assuming that the differential shear-anisotropy is 100 times ε (Crampin, 1994), we have, in the case of MAS, a shear-wave velocity anisotropy (SWVA) of 1.4%, which lies below the fracture-criticality level

(Crampin and Peacock, 2005).

23 The best solution at RUN represents a sub-vertical NNE crack (strike=20º, dip angle 63°, crack density 0.143). The low-residual region containing this minimum point is bounded by azimuths 10-60°, and dip angles larger than 60° (Fig. 2.5C). This region is located close to the minimum residual area on inversions of delay times (Fig. 2.5D). A second small area of low residuals of polarization inversion corresponds to a NW crack with shallower inclination angle

(30-80°). In this case, the best fit crack corresponds to a shear-wave velocity anisotropy of 14%, which is above the fracture-criticality level (Crampin and Peacock, 2005).

No inversion of a crack model was attempted for JUI, because only four events have slope-corrected incident angles (islope) inside the shear-window. Large islope angles occur because most events are from Pisayambo seismic zone located northeast of JUI, with incident rays pointing to the SW; where the slope vector is pointing to the north, so that most of the events fall out of the shear-window. We thus excluded these data from our analysis.

Station N Crack Crack dip Crack density RMS RMS delay strike (°) (°) polarization (°) times (ms/km) MAS 24 47 68 0.014 40.98 3.93 MAS 24 138 57 0.014 42.70 4.04 RUN 20 20 63 0.143 32.39 17.66 RUN 20 122 52 0.143 35.92 17.89

Table 2.4 Best fitting strike and dip solutions using inversion of polarization angles and delay times. Delay times were normalized considering a 3 km-thick anisotropic layer (low velocity zone from seismic tomography).

24 A B

MAS: N=24; RMS=40.98; Crack Strike=47(deg); Crack Dip=68(deg);Crack Density=0.014 MAS: N=24; RMS=3.93; Crack Strike=47(deg); Crack Dip=68(deg);Crack Density=0.014

4.5 10 10 55 4 50 30 46.96 30 45 3.5 50 50 46.96 46.96 40 3 70 70 3.93 40.98 35 p p 3.14 2.5 90 30 90 2.74 3.14 Crack Di Crack Di 25 2 -70 -70 20 1.5 -50 -50 15 67.45 46.96 1 -30 10 -30 0.5 5 -10 -10

-80 -60 -40 -20 0 20 40 60 80 -80 -60 -40 -20 0 20 40 60 80 Crack Strike Crack Strike

C D

RUN: N=20; RMS=32.39; Crack Strike=20(deg); Crack Dip=63(deg);Crack Density=0.143 RUN: N=20; RMS=17.66; Crack Strike=20(deg); Crack Dip=63(deg);Crack Density=0.143

60 10 10 50

30 38.26 30 45 50 15.8 40 50 50 32.39 17.66 35 p 38.26 40 70 p 70 13.38 15.8 30 38.26 90 0 Crack Di 74.62 30 Crack Di 25

-70 -70 20 38.26 20 -50 -50 15.8 15

10 -70 38.26 38.26 10 -30 5 -10 -10 0 -80 -60 -40 -20 0 20 40 60 80 -80 -60 -40 -20 0 20 40 60 80 Crack Strike Crack Strik e Fig. 2.5 A: Residual contour of crack plane found using inversion of two sets of polarization parameters at MAS. B: Residual contour plot of inversion of delay times. Contours are given by the root mean square of the residuals between observed data and calculated model. Thick dashed lines represent a 95% confidence. Two important zones with low residuals on polarization angles correspond with low residuals on delay times. One defines an almost vertical crack with an NE orientation (Strike=47°, Dip=68°) and another with a NW orientation and a more shallow inclination angle (Strike=138°, Dip=57°). C: At RUN two areas have low residuals on inversions of polarization angles and delay times. One area has a NNE orientation (Strike=20°, Dip=63°), and another with a NW trend (Strike=122°, Dip=52°). D: Residual contour plot of inversion of delay times at RUN.

Crack orientations found with anisotropy analysis correlate spatially with structural anomalies imaged by tomographic inversion of P-wave arrival times and geological studies (Hall et al., 1999; Molina et al., 2005). Crack orientations at MAS align with a NE-SW high velocity anomaly (Vp ~ 4.8 km/s) located in the southwestern flank between 2 and 2.5 km above sea level

(Molina et al., 2005). This high velocity zone is interpreted as a solidified lateral dike system that

25 is connected to the magmatic recharge zone at the central base of the volcano (Molina et al.,

2005). A vertically oriented, high velocity zone under the northeastern flank is in agreement with crack orientations beneath RUN, and may reflect a dike system active between 1,500-3,000 years

BP (Hall et al., 1999). In general, high velocity anomalies beneath Tungurahua defined a NNE alignment across the volcanic center, which may be remnant volcanic conduits and dike systems corresponding to the preferred orientation of magma transport and the maximum horizontal compressional stress (Molina et al., 2005). Our study appears to corroborate this finding through analysis of waveforms. An alternative explanation for these orthogonal low-residual solutions, found especially at MAS, may be the “90°-flip” mechanism, which is associated with changes induced by highly over-pressured pore fluids without a variation in the stress direction (Crampin et al., 2002; Gerst and Savage, 2004).

2.6 Temporal variation of delay times

One aspect of the measurement of δt above small earthquakes in world-wide studies is the exceptionally large scatter in time delays (Crampin and Chastin, 2003; Volti and Crampin,

2003). Crampin and McGonigle (1981) showed that delays between split shear-waves vary with the propagation direction. In order to reduce the scattering effect, only 34 events were used, all from the Pisayambo seismic zone (1.18ºS -1.23ºS and 78.20ºW-78.30ºW) with small variation of back-azimuth angles and ray-paths. A significant temporal increase of delay times (δt) is observed at station RUN since julian day 215 (Fig. 2.6), which coincides with the beginning of the vent-sealing phase. Kolmogorov-Smirnov tests show that δt at RUN before and after day 215 are different at 99% confidence. This suggests that there was an increase of local stress over this period. Delay times at MAS, however, do not show any significant change. It may be that these local stress changes are restricted to the structure below the stations, and we recognize that scatter makes this observation speculative. More data over periods of magma injection will be required to answer this question definitely.

26

Seismic E nergy and P ressure Amplitude

cleaning conduit sealing 200 degassing 4.5 md earthquake 100 0 5 0

Seismic Energy Pressure (Pa) 185 190 195 200 205 210 215 220

Shearwave splitting delays

0 legend 4.5 md earthquake MAS 08 RUN JUI 06 04 2 delay times (ms) 0

185 190 195 200 205 210 215 220 time (days)

Fig. 2.6 Upper panel: Activity of Tungurahua volcano during the activity cycle of June-August 2004. Solid line shows the seismic amplitude curve for this period, smoothed with a spline function. Seismic amplitude is a time series composed by 1-hour long root mean squared (RMS) amplitudes computed over 2-15 Hz bandpass filtered seismic signals of vertical component. Stars represent the pressure disturbance of large explosion events from Tungurahua. On day 212 a 4.5 magnitude earthquake occurred 40 km from Tungurahua. Lower panel: Temporal distribution of delay times of shear splitting waves observed at MAS (circles), JUI (crosses) and RUN (triangles).

27 2.7 Conclusions

We have demonstrated that there is measurable shear wave splitting in the proximity of

Tungurahua Volcano. Polarization orientations at stations RUN (NNE of the active crater) show a clear trend in the NNE direction. To the southwest station MAS, there is a bimodal distribution of polarization angles, with one preferred direction trending NNE and another EW. The NNE direction differs from the regional orientation of maximum horizontal stress. Iterative inversions for crack orientation and crack density at MAS and RUN produced NE and NNE crack orientation, respectively. Both crack orientations differ from the maximum regional stress σH.

Another set of low residuals has a NW trend, almost orthogonal to cracks corresponding to minimum residual regions found at MAS and RUN. Sets of orthogonal polarizations found especially at MAS, can be explained by a “90°-flip” mechanism observed in areas with highly over-pressured pore fluids. However, since there are geological structures with NNE-NE trends under the NNE and SW flanks, and the majority of split shear waves recorded at MAS cross the central part of the volcanic edifice (Fig. 2.1B), we conclude that the NNE-NE anisotropy represents the stress directions below the central part of Tungurahua volcano, and is associated with a system of cracks or dykes composing the feeding system of Tungurahua.

On a more speculative note, temporal variations of splitting time delays appear to track changes observed in eruptive style within the activity cycle of Tungurahua in the summer 2004.

Three characteristic phases were observed during this period: vent-clearing, degassing, and a vent-sealing phase. An increase of δt was observed on split shear-waves of tectonic events coming from a small epicentral area (Pisayambo seismic zone) recorded at station RUN. This increase may have been caused by a local change of stress levels beneath station RUN, coinciding with the beginning of the sealing phase.

28

CHAPTER 3

CONDUIT PROCESSES

This chapter is based on the paper: “Source constraints of Tungurahua volcano explosion events”, which was published in the Bulletin of Volcanology, volume 68, 2006, pags. 480-490.

Here, I present an analysis of the different types of infrasound signals generated by degassing events of Tungurahua volcano, especially signals from explosion events.

3.1 Introduction

Since the sizable eruptive activity in late 1999 ended, Tungurahua exhibited four eruptive cycles (Fig. 3.1). The most recent eruptive cycle started in late May 2004, reached its climax in

July, and continues to the present (December 2004) with minor bursts of activity (several explosions per day). Diverse measurements carried out by the Geophysical Institute of the

National Polytechnic School of Ecuador (IG-EPN) confirmed that Tungurahua reached a moderate activity level during the most recent period (May-December 2004) with frequent ash columns of altitudes no higher than 3 km above the crater. COSPEC measurements carried out by IG-EPN (Instituto Geofisico - OVT, 2004) indicated SO2 concentrations of 65 Tons/h

(07/08/04) and 81 Tons/h (07/28/04), showing that Tungurahua had slightly higher degassing levels than the 2003-2004 background (43.1 ± 27.9 tons/h).

EXPLOSIVE ACTIVITY OF TUNGURAHUA 4 2004 deployment 3.5 Onset climatic peak3 peak4 phase 3

peak1 peak2 2.5

2

1.5

1 log (number of explosions per month)

0.5

0 1999 2000 2001 2002 2003 2004 2005 2006 TIME

Fig. 3.1 Monthly number of explosions recorded by the seismic network of Instituto Geofisico in log scale. Peaks of activity are labeled with a month-year code. Data compiled from Instituto Geofisico data base.

In October 1999, a temporary infrasonic microphone was deployed on a ridge just in front of the northwestern flank. During a 4-day operation, this instrument recorded series of explosion pulses and an almost continuous jetting signal (Johnson et al., 2003). In addition, a low-frequency microphone installed at the headquarters of the Tungurahua Volcano Observatory

(~13 km from the vent) recorded 14 infrasonic explosion signals from August 11 through August

31, 2000. All the infrasonic signals showed a single N-type (compression-rarefaction) pulse in a low frequency band (0.5-2.1 Hz). Explosion signals have been further observed with ground- coupled waves at seismic stations, especially at station PATA (4.6 km from the vent). The time difference between P-waves and infrasonic ground-coupled waves ranges from 11 to 16 s at

PATA with a mean of 13.3 s (Ruiz et al., 2001). In this paper, we describe a temporary deployment of three seismic and infrasonic stations installed in summer 2004 to monitor explosion activity and further constrain eruption at Tungurahua.

30 3.2 Experiment Description

In late June 2004, two stations (RUN and MAS) with collocated seismic and infrasonic sensors were deployed on NE and SW flanks. Another station (JUI) was deployed on the NW flank on July 19, 2004. These instruments were located at horizontal distances between 3.20 and

5.51 km from the active vent (Fig. 1.3; Table 3.1). Power failures due to ash falls, wind and heavy rains occasionally affected the instrument operation. However, more than 2,300 seismic events were recorded between June 29 and August 12, 2004, and most of these (2,142) were related to the volcanic activity. Others correspond to tectonic activity, especially the earthquakes of the Pisayambo seismogenic zone, and the main shock and aftershock sequence of a local 4.2 tectonic earthquake (July 30, 2004).

Station Sensors Gains Lat. S Lon. W Altitude Total Distance MAS CMG40T, 802 v/m/s, 48.4 -1.4836° -78.4684° 3310 m 3516 m LD2200C mv/pa

JUI CMG40T, 800 v/m/s, 48.4 -1.4353° -78.4607° 2965 m 4168 m LD2200C, mv/pa, 80 McCH2 mv/pa*

RUN CMG3T, 1496 v/m/s, -1.4221° -78.4223° 2700 m 5889 m LD2200C, 48.4 mv/pa, McCH1 100 mv/pa*

Table 3.1 Description and location of instrumentation deployed on Tungurahua volcano during the 2004 field campaign. All stations had Reftek-130 dataloggers, Guralp broadband seismic sensors, and Larson Davis microphones. Additional microphones (McCH1 and McCH2) were installed in order to check gain values, environmental and instrumental signal distortions.

3.3 Description of Degassing Signals

Degassing events, characterized by seismo-acoustic signals, comprised 95% of the total volcanic seismic signals recorded on Tungurahua during our survey. These degassing signals are classified as: Explosion Blast (EX), Roars (RO), and Chugging (CH). Descriptions are

31 summarized in Table 3.2. Other volcanic events without acoustic (infrasonic) signals are: Long

Period (LP), Hybrid (HB), Volcano-tectonic (VT), and Tremor (TR). Figure 3.2 shows the distribution of these categories.

3.3.1 Explosion Events

At all stations, infrasonic records of EX events (Fig. 3.3A) have impulsive compression onsets, followed by rarefaction pulses, and a short duration coda with an exponential decaying envelope. These signals contain more energy in the 1-3 Hz band

Fig. 3.2 Distribution of main types of seismic signals recorded on Tungurahua. Explosion Events (EX), Jetting tremors (RO), and Chugging (CH) events have acoustic signals. These events are related to degassing processes, and account for more than 95% of total volcanic signals. Tremor (TR), Long-Period (LP), Hybrid (HY), and Volcanic-Tectonic (VT) events do not have an acoustic component.

without a very long period component, and are similar to explosion signals recorded on

(Johnson and Lees, 2000), Karymsky (Johnson and Lees, 2000), Arenal (Hagerty et al., 2000), and Fuego (Johnson et al., 2004). Amplitudes of blast signals span three orders of magnitude from 0.1 to 180 Pa (July 21, 03h32 GMT, the largest signal recorded on MAS station). Often EX events are followed by tremor-like signals (Fig. 3.3B) with irregular envelopes and high

32 frequency content (up to the signal Nyquist frequency of 62.5 Hz). Time lapses between the impulsive explosion blast and the tremor-like signal range from 5 to 100 s. Several of these high frequency signals have larger pressure amplitudes than the initial blast.

Seismic signals of explosion events (Fig. 3.3 left column) are characterized by an emergent compressive onset on vertical components (Tp1), followed by a secondary compressive phase (Tp2). Some seismic waveforms have very clear ground-coupled airwaves with high frequency content.

3.3.2 Jetting tremor

Jetting tremor waveforms (called “roars” by locals) are characterized by emergent initial arrivals with long duration codas both in seismic and infrasonic records (Fig. 3.3 central column).

No clear blasts are associated with the initiation of these signals. Acoustic records of some eruptions from Stromboli (Woulff and McGetchin, 1976) have a pattern similar to the one observed in RO events: (1) emergent onset, (2) coda punctuated by distinct peaks, and, (3) a gradual decay to background. Dominant amplitudes span the 1-4 Hz band with no significant energy above 10 Hz. In a few cases, however, jetting tremors seem to be generated by overlapping cascades of smaller explosions; but in general no single blast signals are recognized.

33

Fig. 3.3 Seismic (upper row) and infrasonic (lower row) traces and spectrograms of explosions (left column), jetting tremor (central column), and chugging event (right column) recorded at MAS station. Time scales are in seconds (explosion signal lasts 45 s, jetting tremor 80 s, and chugging event 35 s). Infrasound amplitudes in pascals and seismic amplitudes in µm/s. A bar over upper right corner of spectrograms shows the FFT window length (2 s). Left panels: Seismic and infrasonic records of an EX event (07/21/2004 03h32m) with a maximum ground motion peak-to-peak velocity is 928 µm/s. Infrasound has a maximum peak-to-peak pressure disturbance of 182 Pa. Central panels: Seismic record of jetting tremor with a maximum ground motion peak-to-peak velocity of 31 µm/s. Infrasound signal has long duration and broad frequency content. Right panels: Seismic signal of chugging event covered by seismic noise. Infrasound record shows a sequence of equally-spaced pulses with initial compressive polarization. Maximum peak-to-peak pressure is 0.09 Pa.

34 3.3.3 Chugging signals

In the course of this deployment, chugging (CH) events were recognized for the first time at Tungurahua. However, this does not mean that chugging signals were not produced in the past as acoustic monitoring at Tungurahua has been intermittent. Although relatively rare compared to other degassing signals, 16 events showed sequences of pulses at time intervals between 0.5 to

1.0 s (Fig. 3.3 right column), with a saw-tooth shape on infrasonic records, similar to CH events observed on other active volcanoes (Hagerty et al., 2000; Johnson et al., 1998; Johnson and Lees,

2000). Each pulse consists of an initial compression followed by a rarefaction, resembling a periodic sequence of small blast events. Seismic records of CHs have emergent onsets and tremor-like codas without clear distinctive pulses. Spectrograms of Tungurahua’s CH events show that most of the energy is in the 1-3 Hz band and gliding is not apparent.

Signal Type EX RO CH

Seismic signal First arrival on vertical Emergent, in few cases Emergent. No component is emergent starts with a very small explosion event as a with compressive polarity blast. Waveform trigger. Waveforms in all azimuths. Waveform envelope has spindle have tremor-like envelope has spindle shape. shape. No ground envelopes. No ground Ground coupled acoustic coupled acoustic signals coupled acoustic signal signal is noticeable. are apparent. Small is apparent. Small amplitudes. amplitudes.

Infrasonic Clear compressive pulse Emergent first arrival Sequence of signal followed by rarefaction and waves are followed by a compressive pulses with small amplitude coda. chaotic wave train with time intervals between Frequency 1-3 Hz. Some larger amplitudes. 0.5-1.0 s. No gliding. events are followed by a Frequencies 1-10 Hz. high frequency rumbling (up to 62 Hz). Large range of amplitudes (up to 180 Pa).

Table 3.2 Distinctive characteristics of degassing signals at Tungurahua volcano.

35 3.4 Clustering of Explosive Events

Quantitative cluster analysis was applied to the highest quality data in order to find populations of EX events with common characteristics. To achieve a consistent data set, we used

361 events with reliable seismic and acoustic arrival-times at MAS. Events with picking errors greater than 0.1 s were rejected. The high quality events were partitioned using fuzzy logic programs in R-software (Kaufman and Rousseeuw, 1990). This grouping method considers that dissimilarities between signals can be measured by a distance defined as the complement of the matrix of maximum cross-correlation coefficients of each pair of signals (Pison et al., 1999). The correlation matrix is used to associate a cluster score for each waveform such that the scores represent a level of association in each group (Pison et al., 1999). The new clustering matrix is then rotated by principal components. Clusters are graphically displayed on a principal component plot with ellipses containing all events of each cluster. Distance between ellipses measures the degree of dissimilarity between clusters.

3.4.1 Clusters of infra-sonic signals

Prior to cluster analysis, signals were filtered using a band-pass 0.5 – 10.0 Hz

Butterworth filter. Four second data windows with a 1 s pre-event signal were considered for analysis. Cross-correlation scores for the full data set are generally low, so only high correlation pairs were selected for analysis. Cluster analysis was applied to a set of 28 events with correlation factors (σ) larger than 0.96. From this analysis, 4 distinct clusters were observed (Fig.

3.4). Using a more hierarchical technique, a dendogram tree was produced for these events (Fig.

3.5). In this type of graph, adjacent events, represented as leaves, are the most similar. Clusters are grouped in branches with a length proportional to the level of dissimilarity (1-σ). Using such a method, the same four clusters of EX events were obtained. An inspection of stacked signals of these clusters shows the following characteristics:

36 Group A1 [1, 9, 10, 17, 19, 23, 27] events (Fig. 3.6A) have a short duration compression phase (tc=0.25 s), followed by long-duration rarefaction phase (tr=0.43 s) with a large ratio between amplitudes of compression and rarefaction peaks, hereafter represented by ξ (ξ=0.962 for A1 cluster).

Group A2 [2, 5, 8, 12, 13, 16, 20, 22, 24] events (Fig. 3.6B) have a very short compression time tc=0.23 s, followed by a short duration rarefaction tr=0.27 s. However, the compression /rarefaction ratio ξ=0.733 is small.

Group 3 Group 4 Group 2

Group 1

Fig. 3.4 Cluster distribution of infrasonic records after applying a fuzzy logic process. Four distinct clusters of blast-explosion events are circled and marked with distinct symbols. Distance (compliment of the cross correlation mean between events of a pair of clusters) is shown with straight lines connecting clusters. Length of distance denotes the degree of dissimilarity between clusters. Four clusters of EX events are recognized.

37

Group 4

Group 3

Group 1

Group 2

Fig. 3.5 Dendogram tree showing the event clustering of acoustic signals of explosion blast (EX) events. Length of the dendogram branchs denotes the degree of dissimilarity between events or event groups between the data set. Also four clusters of EX events are found by this method.

Group A3 [3, 7, 14, 15, 21] events (Fig. 3.6C) have a short duration compression phase tc=0.26 s, followed by a very long-duration rarefaction phase tr=0.51 s. Pressure ratio is high

ξ=0.932, as in Group 1.

Group A4 [4, 6, 11, 18, 25, 26, 28] events (Fig. 3.6D) have a low-frequency compression, and also low-frequency rarefaction (tp=0.31 s and tr=0.49 s). Pressure ratio is small (ξ=0.667).

Other than the differences in waveform signature noted above, the infrasonic signal clusters show no other distinctive amplitude or temporal variations.

38

Fig. 3.6 Infrasonic records of each cluster of infrasound signals (A1, A2, A3, and A4). Last row is the stacked signal. Amplitudes of infrasound signals are normalized according to the stacked one. Signals are aligned considering Tac at time 1 s.

3.4.2 Clusters of seismic signals

Before performing a cluster analysis, we applied a band-pass filter (0.5-10.0 Hz) and integrated the corresponding seismic signals to seismic displacement. Initial arrivals of the seismic signals were selected with 4-s windows. The seismic signals are more complicated than the infrasonic signals due to path effects, and perhaps, source complexities. Clustering of the seismic signals was weaker than the acoustic signals. Only two poorly-related groups (S1 and

39 S2) were found using the same fuzzy logic process (Fig. 3.7). Stacked signals corresponding to seismic records of each cluster are presented in Fig. 3.8. Most of the events of acoustic group A1

(6 out of 7 events), and A2 (5 out of 7) belong to S1, and most of the events of A3 (8 out of 9 events) belong to S2. Events from cluster A4 are split between three events in S1 and two in S2.

This apparent correlation between infrasonic and seismic clusters suggests that clusters are related to a source effect, due to the fact that these signals, traveling on completely different paths, show some degree of coherent clustering.

Fig. 3.7 Cluster distribution of corresponding seismic records after applying a fuzzy logic process. Two poorly constrained clusters (S1 and S2) of blast-explosion events are circled and marked with distinct symbols.

40

Fig.3. 8 Seismic records of S1 and S2 clusters. Last row is the stacked signal. Signals are aligned using cross-correlation lags with the first event. Traces have smooth arrivals of the first compression arrival (Tp1), however, stacked signal shows a clear arrival of Tp2 phase. Amplitudes of infrasound signals are normalized with respect to the stacked one.

3.5 Analysis of Source Locations

Delay times between impulsive acoustic pressure pulses and emergent seismic onsets

(∆T=Tac-Tp1), measured from 361 events with clear arrival times, were analyzed in order to constrain the explosion (EX) source locations. The distribution of these delays has a positive skew, with an apparent Poisson-type distribution (Fig. 3.9). Table 3.3 shows the median values and 5th and 95th percentile values of ∆T. We attribute the scatter in ∆T primarily to variations in source processes and locations, and not to acoustic propagation phenomena, because normal daily variations in ambient temperature were small (±5°C), and would produce travel time anomalies less than those observed. Average delay times between seismic and acoustic phases were analyzed for both night-time explosions and day-time explosions and show a difference of about

41 0.1 s for presumed cold and warm atmospheric conditions. This is significantly less than the observed scatter. Ambient winds could also affect acoustic arrival times. However consistent anomalies at all stations (with an azimuthal distribution of about 180 degrees), suggest that these delays are not caused by a uni-directional wind. Thus, the scatter of observed ∆T can be explained by these options: a) a seismo-acoustic point source located at variable depths, b) a spatially-fixed point source with variable ascent velocity, and c) an extended source that has a varying source time function. In this paper, we address the first option as outlined below. 16

o 361 EX events 14 12

median = 10.06 s Time interval(s) 10

190 200 210 220 0.0 0.2 0.4 0.6 0.8 TIME (julian day) PDF

Fig. 3.9 Left panel: Time intervals between arrivals of acoustic pulse and first P-wave motion ∆T observed on MAS station at 3516 m (total distance) from the vent. Notice the important spread occurred even on the same date, showing that there is no time migration of explosion source; and suggesting the presence of multiple sources. Right panel: Distribution of ∆T times shows a poisson-like distribution.

42 Station N readings Median (∆T) (s) ∆T 5% (s) ∆T 95% (s) MAS 361 10.062 9.359 12.602

JUI 24 11.904 10.998 16.175

RUN 164 16.854 16.293 18.943

Table 3.3 Number of blast events with clear arrival times of acoustic and seismic signals. Median values of delay times between these arrivals are shown with 5th and 95th percentiles of the ∆T distribution for each station.

We assume that EX events originate as a point source at time To inside the conduit at a depth z below the crater floor with fixed elevation hv ~ 4,750 m. At this point, we postulate that seismic waves are generated due to gas expansion and possible brittle fracture. The initiation of the EX event thus corresponds to the onset of gas and ash rising through a narrow conduit (or system of cracks) towards the vent. For the sake of a simplistic model, we fix the rise velocity

(U) of this two-phase fluid between the original source depth and the vent. Gas reaches the flaring portion of the conduit (i.e., the vent) near the crater floor at time (To+z/U), where gas can freely expand into a hemispherical half-space, initiating a rapid volumetric expansion, which perturbs the atmosphere and produces infrasound. An equal and opposite force directed towards the ground generates a secondary seismic pulse. These acoustic and secondary seismic waves propagate from the vent and arrive at the stations at times Tac and Tp2, respectively (Fig. 3.10).

The initial seismic pulse, corresponding to the event onset, arrives at time (Tp1=T0+tpi) at the station i (elevation hi). We assume a temperature of 10 °C with corresponding air sound speed vair=337.6 m/s. The following equations are used to estimate the travel times of first seismic and infrasonic waves:

43 −

SOUT−WEST vent 0 hv

−0.2 path Depth based on (tac−tp) times second arrival Z −0.4

−0.6 daci source air wave

−0.8

−1.0 path first compressional arrival Depth from rim (km) −1.2

MAS −1.4 conduit hi dhi −1.6 station

3 2 1 0 Distance (km)

Fig. 3.10 Schematic propagation model for EX type events in Tungurahua volcano. Source is located at depth z below crater. From z, gas-and-solid material ascends at velocity U through the conduit, reaching the crater floor at time To+U/z. Secondary seismic and infrasound signals travel in almost parallel paths from the vent to the station.

1 2 2 t pi = ( dhi + (hv − hi − z) ) (1) α

2 2 dhi + (hv − hi ) taci = (2) vair

And the time difference between arrivals is the following:

z ∆T = + t − t (3) i U aci pi

Where tpi is the travel time of the first seismic pulse from the source to the station i, and, taci is the travel time of acoustic wave from the crater to the station i=M,J,R (for MAS, JUI and

RUN stations). The horizontal distance from the conduit to the station i is dhi, and α is the seismic velocity of the volcano edifice.

44 Using a set of six explosions with clear seismic arrival times at all three stations (TpM, TpJ,

TpR), and a range of seismic velocities α [0.1-4 km/s], we calculated the residuals between the observed arrival times (left-hand side of Eq. 4) and predicted travel times of seismic waves (right- hand side of Eq. 4), finding that, the smaller residuals are at shallow depths (z < 1 km inside the conduit) for all range of possible values of α. Fig. 3.11 shows the residual distribution for α=2.9 km/s.

d 2 + (h − z − h )2 d 2 + (h − z − h )2 ⎡ d 2 + (h − z − h )2 ⎤ T + T − 2T = hM v M + hR v R − 2⎢ hM v M ⎥ pM pR pJ α α α ⎣⎢ ⎦⎥ (4)

Fig. 3.11 Schematic propagation model for EX type events in Tungurahua volcano. Arrival times of six 6 EX events and a seismic velocity α=2.9 km/s, were used in this calculation.

Next, we will use a seismic velocity of 2.9 km/s found by Molina (2001) for a half space velocity model for p-waves at Tungurahua. With α=2.9 m/s, variations in U and z, were further investigated by minimizing the residuals between observed and predicted delay times (∆T) at

45 three stations (Eq. (3), Fig. 3.12). Minimum residuals show a clear trade-off between U and source depth (z): (z=0.925U-4.419) in the ranges of U [10-300 m/s] and z [15-275 m].

Fig. 3.12 Distribution of residuals between the summation of observed delay times and the summation of expected times assuming a ground velocity α= 2.9 km/s for a range of U [10-300 m/s] and depths [0- 400m]. Residuals from all three stations were considered.

We note that the ascent rate of a two-phase fluid through a conduit or system of cracks is a very complex problem, which depends upon flow properties (density and viscosity) and path geometry (width, length, and wall roughness). However we propose that gas is able to accelerate at the vent as it leaves the confines of the conduit. Owing to conservation of material flux, the conduit velocity (U) is assumed to be less than the initial ejection velocity at the vent (vej).

Therefore, we calculated vej using the total acoustic power (Π) radiated by these events, based on the assumption that blast-type eruptions can be represented as an ideal monopole acoustic source

(Woulff and McGetchin, 1976; Vergniolle et al., 2004). Total acoustic power, emitted in a half sphere of radius equal to the distance between the vent and the microphone (r), is easily

46 calculated by integrating the square of pressure disturbance (∆p) records with Eq. 5 (Vergniolle et al., 2004). For a circular flat orifice (Km=1/16) with a Rb=10-m radius (estimated from oblique

3 FLIR pictures taken from Samaniego et al., 2003), air density (ρair) of 0.9 kg/m , and blast durations (Tb) of 3 s, we found with Eq. 6, that the median of gas ejection velocities of

Tungurahua’s EX events is 20 m/s, with the largest value reaching 130 m/s. The median of vej will be considered as a first order approximation for U. Gas ejection velocities vej reported on strombolian explosions (Chouet et al., 1974; Weill et al., 1992; Ripepe et al., 2001; Ripepe et al.,

2002) are in the range of [20-112] m/s. For vulcanian explosions at Soufriere Hills, jet velocities varying from 40 to 140 m/s were measured (Formenti et al., 2003).

πr 2 Tb Π = ∫ ∆p2dt (5) ρairvairTb 0

1 4 ⎛ v Π ⎞ v = ⎜ air ⎟ (6) ej ⎜ 2 ⎟ ⎝ 4πKm ρair Rb ⎠

Once U and α are fixed, we used Eq. (3) to find the best fitting source depth (z) in the conduit for all EX events with clear acoustic and seismic arrivals. Statistical estimates of z are given in Table 3.4 and Fig. 3.13. Considering the 5th and 95th percentiles of z, we found that explosion events in Tungurahua initiate at a depth between ~ 0-100 m below the crater floor.

Cluster events do not exhibit statistically significant source depths (at 5% of significance level).

These shallow source depths of EX events are comparable with depth estimations of explosions on other volcanoes, such as Arenal (Alvarado and Barquero, 1987; Hagerty et al., 2000),

Stromboli (Chouet et al., 1997; Ripepe et al., 2001; Chouet et al., 2003), and Etna (Gresta et al.,

2004), where depths of explosion sources are shallower than 260 m.

47

Station \ Depth (Z) Z5th perc. (m) Z25th perc. (m) Z50th perc. (m) Z75th perc. (m) Z95th perc. (m)

MAS 3 11 17 30 68

JUI 2 17 20 35 105

RUN 17 24 29 39 70

Table 3.4. Statistical estimates of explosion source depth z, obtained from an analysis of delay times ∆t at MAS, JUI, and RUN stations. Depth Z was calculated using a vertical propagation velocity of 20 m/s and a seismic velocity of 2.9 km/s for seismic waves.

Fig. 3.13 A) Boxplots of observed ∆T delay times at MAS, JUI, and RUN. The median value is indicated by a line inside the box. The ends of the box correspond to statistics called “hinges”, which are similar to quartiles. Whiskers extend to values 1.5 times the inter-quartile range. Box widths are proportional to the square root of the number of elements in each data set. B) Density functions of EX source depths using a model with a vertical conduit (red line: based on ∆T at MAS; green line: based on ∆T at JUI; blue line: based on ∆T at RUN). Light gray area contains source depths between the 5th and 95th percentiles. Dark gray contains depths between the 25th and 75th percentiles.

48 Using a higher ascending velocity (U=100 m/s), as might be expected for large explosions, we estimate an explosion depth range of ~50-200 m below crater. For U < 20 m/s, z values are smaller than those reported on Table 3.4.

Stacked records of seismic explosion signals (Fig. 3.8) show the secondary compression phase (Tp2) arriving 1.0 s after the onset of the first seismic wave (Tp1). We assume that infrasonic wave and the secondary seismic pulse are both generated at the vent; and that both waves have nearly coincident ray-paths. Using the difference between these two arrival times, we can estimate α with known propagation distance and estimated sound speed (Eq. 7). We note that this value is in agreement with our previous estimate of 2.9 km/s.

⎛ 1 1 ⎞ ⎜ ⎟ (7) Tac − Tp2 = ⎜ − ⎟d ⎝ vair α ⎠

3.6 Conclusions

During our seismo-acoustic deployment in 2004, Tungurahua volcano exhibited a diversity of degassing styles, characterized by single explosion blasts, long-duration roar events, and chugging signals. As compared to seismic records, infrasonic signals are considerably simpler, and thus provide a means for signal classification of explosion events. Four distinct cluster groups were recognized among the highly correlated explosion events. The groupings are based entirely on waveform similarity, and do not show any particular spatial or temporal patterns, and are not related to event magnitude either. We conclude that the waveform clustering is related to the source mechanics of the explosions at the vent.

Analysis of travel times supports a model where explosion blasts begin at shallow depths

(< 200 m below the crater floor) inside the conduit, generating a first pulse of seismic waves with compressive polarity in all azimuths. After the event initiation, a two-phase mixture rises up

49 through the conduit and arrives at the crater floor about 1 s later. Rapid gas expansion at the vent generates infrasonic compression waves and a secondary seismic pulse.

50

CHAPTER 4

INFRASOUND PROPAGATION

This chapter analyzes the propagation of infrasound signals generated by volcanic explosions in a medium characterized by complex topographic boundaries and homogeneous wind. A version of this chapter will be submitted for publication with the title “Propagation effects on Infrasound Waveforms of Volcanic Explosions” by Mario Ruiz, Jonathan M. Lees, and

Keith Wilson.

4.1 Introduction Infrasound has become one of the most widely used techniques in studies of ongoing volcanic eruptions. Different types of infrasound signals have been recognized in active volcanoes: a) Single impulsive infrasound waveforms generated by volcanic explosions when near-surface gas volumes burst and fracture the vent plug, inducing rapid degassing (Johnson et al., 1998); b) Infrasound signals with more complex high-frequency waveforms called jetting tremors produced by long-duration degassing (Ruiz et al., 2006); c) Sequences of discrete infrasonic pulses, known as chugging events generated by releasing gas through time-varying chocked vents (Lees et al., 2004). Explosions, jetting tremors and chugging events coexist in

Tungurahua, an andesitic volcano in the Ecudorean Andes (Ruiz et al., 2006).

At difference of seismic signals, which are strongly affected by scattering and path effects from the complex geological structure in the volcano edifice, infrasound signals suffer less from atmospheric complications (Garces et al., 1999; Johnson et al., 2003). Infrasound propagation through the atmosphere is dependent mainly on the ambient temperature and wind

(Evers and Haak, 2003). For an ideal gas, the velocity of sound is proportional to the square root of the absolute temperature (Ford, 1970; Ostashev, 1997). Even at distances as short as hundreds of meters, temperature structure of the atmosphere can impact the strength of a recorded pressure trace (Johnson, 2003). Wind velocities in the atmosphere that may reach values (~10 m/s) are not insignificant compared to sound speeds in a vacuum. For this reason, wind gradients might modify propagation arrivals of infrasound. In some cases, variations of delay times between the arrivals of acoustic and seismic waves are attributed to changes in atmospheric sound velocity

(Johnson et al., 1998).

Nearly all andesitic stratovolcanoes possess steep slopes and are surrounded by abrupt topographic variations. During periods of activity, explosion sources are typically localized along the shallow upper few hundred meters of the volcanic conduits (Garces et al., 2000; Ripepe et al.,

2001; Ruiz et al., 2006) and infrasound waves propagate from the crater floor (Ruiz et al., 2006).

Infrasound stations are usually located on volcano flanks or farther a field. Ray-paths from the volcanic source to the recording stations on the flanks may be obstructed by crater walls, hills or other intervening topographic structures.

Acoustic signal fluctuations can further be generated by the interaction between sound and topographic variations (Sack, 1995). Wind interacting with topographic disturbances can generate scattering by turbulence. Wind must speed up at increases in elevation in order to conserve mass. There are also typically turbulent vortices generated in the lee of the mountain.

All of these phenomena can affect infrasound propagation in volcanic regions. Commonly, investigations of volcano infrasound neglect the effect of wind and topography; where, instead a constant air wave velocity and straight propagation paths are assumed (Garces et al., 2000;

Hagerty et al., 2000; Ripepe et al., 2001; Ruiz et al., 2006).

In this chapter, we introduce a 2-D approach to simulate the effects of wind and topography on infrasound propagation using a Finite Difference - Time Domain (FDTD)

52 technique (Ostashev et al., 2005; Wilson and Liu, 2004). We compare the FDTD calculations to infrasound waves from explosion events recorded on different flanks with variable topography, in a medium that includes wind. Effects caused by atmospheric heterogeneities, turbulence generated by wind-topography interaction, and 3-D propagation models are beyond the scope of this paper and they will be likely considered in future research.

4.2 Characteristics of the explosive activity of Tungurahua

Tungurahua volcano has steep flanks (30-35º) and up to 3,200 m of local vertical relief

(Fig. 4.1). A 300 m- wide and 100 m-deep crater vent is located on the upper part of its northwestern flank (Samaniego et al., 2003). The only source of the explosive activity during the last eruptive period is the crater at the NW flank.

Most explosions recorded at Tungurahua generate ash-laden steam plumes with altitudes between 200 m to 8 km above the crater (G. Ruiz, personal communication). During the seismo- acoustic field experiment of June-August 2004 only 10 ash-clouds were clearly observed due to rainy and cloudy conditions. However, using observations of 404 small ash-clouds (altitudes less than 2.5 km above the crater) recorded between 1999 to 2005 (Ruiz et al., 2005), horizontal velocities of ash plumes were found to be approximately 7.0 ± 4.8 m/s at 95% confidence (Fig.

4.2A). We assume that horizontal displacements of small ash plumes are driven by winds at those altitude levels. These plumes have a preferential propagation direction from East to West

(mean θm= 266°, circular sample standard deviation s0=57°) (Davis, 2002; Gaile and Burt, 1980)

(Fig. 4.2B). Mean and standard deviations of wind velocity vectors projected in the SW, NW, and NNE directions are 5.5 ± 3.8 m/s, 3.3 ± 2.3 m/s, and -3.6 ± 2.4 m/s.

53

Fig. 4.1 Location of temporal seismo-acoustic stations (A, B, and C) on Tungurahua’s flanks. A 300 m- wide single crater is located in the upper part of the NW flank. Meteorological station El Tablon is located 8 km in NW direction.

In 2005, the Instituto Geofisico installed a digital weather station on El Tablon, a hill in front of the volcano cone, 8.5 km NW of the crater vent. During July and September 2006, winds with a mean velocity of 6.3 ± 5.5 m/s at 95% confidence (Fig. 4.2C) and a mean direction to the

WNW (θm = 299º, s0=46º) were recorded (Fig. 4.2D). In general, ash-cloud displacements and meteorological data show that, at Tungurahua, winds have a station-ward orientation in the NW and SW flanks (A and B stations) and a vent-ward orientation in the case of the vent-C direction

(NE flank).

54 horizontal ash−cloud velocity ash−cloud orientation

150 B A N 100 W 50 100 E 0 Frequency 5 mean=266 deg

S 0

0 5 10 15 velocity m/s

peak wind velocity at Tablon wind orientation D C N 300

100 200

200 W E

Frequency mean=299 deg 100

S 0

0 5 10 15 velocity m/s

Fig. 4.2 Distribution of horizontal velocities (A) and azimuthal directions (B) of Tungurahua’s ash-clouds with altitudes smaller than 7.5 km asl between 1996 and 2005. Data base from Geophysical Institute- EPN, compiled by G. Ruiz. Distribution of maximum velocity values recorded at Tablon weather station from July to September 2006 (C). Rose diagram shows the wind orientations at Tablon (D).

During the fourth cycle of activity (June-August 2004), three portable broad band seismic stations and collocated Larson-Davis free-field precision microphones were installed on the SW,

NW, and NNE flanks. These stations, originally called MAS, JUI and RUN, will be hereafter named A, B, and C, according to their horizontal distances from the single active crater of 3.2,

3.9, and 5.4 km, respectively (Fig. 4.1). To insure that pressure amplitudes were properly represented, infrasonic recordings were calibrated and the instrumental response of each microphone was removed by deconvolution. Empirical determination of microphone responses

(0.2-to 20 Hz frequency band) was estimated using a low frequency piston-phone device

(Timothy Marston, Pennsylvania State University, personal communication).

55 At all stations, infrasound signals generated by Tungurahua’s explosions have impulsive, compression onsets, followed by a more gradual rarefaction, and a short duration coda (Fig. 4.3).

The maximum pressure during this deployment was 129 Pa at A after instrument calibration.

Station A also registered 39 events with pressures larger than 10 Pa during the 44 day deployment. Most of the infrasound energy was restricted to the 1-3 Hz frequency band (Ruiz et al., 2006). Using time differences between seismic and acoustic arrivals at these three stations,

Ruiz et al. (2006) found that explosion events initiate at different sources inside the shallow part of the volcanic conduit (depth < 200 m below the crater floor). Once the mixture of gas and ash reaches the crater floor, it freely expands into the atmosphere and generates infrasound compression waves.

56 EX Jul 21 03h32

94.4 Pa 0 8

0 122.2 Pa 4 A 0 3.52

40 B 0 4.28 pressure Pa 0 8 expected arrival times total distance from vent km Vair=331 m/s 39.8 Pa 0

4 C 0 6.01

0 5 10 15

time s

Fig. 4.3 Infrasound signals generated by a volcanic explosion at Tungurahua on July 21 (03h32 GMT), recorded on A, B, and C. Signal amplitudes are scaled to pascals (left axis). Source-receiver distance is given on the right hand side axis. Dashed line shows the expected arrival times for sound velocity at 0º C. Maximum value of pressure disturbance at each station is given in pascals.

In the absence of wind and refraction effects, and under ideal gas conditions, delays of infrasound arrivals at different stations should depend only on source-receiver distances (Fig.

4.3). However, relative delay times at all three stations (∆tA-B, ∆tB-C, and ∆tA-C) were systematically larger than expected for propagating sound waves at average temperatures (0 °c at

El Tablon weather station) (Fig. 4.4, Table 4.1). Infrasound amplitude ratios at different stations

(AA/AB, AB/AC, and AA/AC) deviate from expected values based on a simple 1/r geometrical spreading amplitude reduction (Table 4.1 and Figs. 4.3 and 4.5). The anomalous amplitude ratios and arrival times suggest that topography and/or wind play a role in infrasound wave perturbation near volcanoes.

57

DELAY TIME INFRASONIC ARRIVALS

25 MAS−JUI JUI−RUN MAS−RUN median=2.50s median=5.06s median=7.58s 20 15 occurrence 10 5 0

2468 delay times s

Fig. 4.4 Histograms of delay times between infrasonic arrivals at stations A-B, B-C, and A-C. Vertical dashed lines show the expected delay times using a constant infrasound speed and vent-station distances. Dotted lines show the expected variation on arrival times for ±50 m offsets on source location.

58 Amplitude ratios of infrasound 3.5

expected ratios +/− 50m 3.0 2.5 2.0 amplitude ratios 1.5 1.0

MAS−JUI JUI−RUN MAS−RUN

Fig. 4.5 Distribution of peak-to-peak amplitude ratios (Ai/Aj) of infrasonic signals generated by volcanic explosions of Tungurahua volcano. Boxes are limited by the first and third quartile. Lines inside boxes mark the median amplitude ratio. Whiskers go out up to 1.75 times the interquartile range. A, B, and C are seismo-acoustic stations located at 3.52, 4.27, and 6.01 km from the crater. Expected amplitude ratios (cross marks) were computed using vent-station distance and 1/r attenuation.

∆t exp. ∆t obs. ∆t obs. ∆t obs. Ai/Aj Ai/Aj Ai/Aj Ai/Aj 5th % s 50th 95th exp. 5th % 50th % 95th % % s % s A-B 1.85±0.3 2.40 2.50 2.65 1.17±0.03 0.67 0.93 1.10 B-C 5.37±0.3 4.94 5.06 5.18 1.36±0.03 1.44 1.94 3.04 A- C 7.22±0.3 7.43 7.58 7.78 1.67±0.04 1.39 1.92 2.73

Table 4.1 Expected delay times (∆t exp.) and amplitude ratios between i and j stations (Ai/Aj exp) considering only source-receiver distances and statistics of delay times and amplitudes ratios observed at stations A, B, and C. Deviation on expected delay times and amplitude ratios are computed considered a deviation of ±50 m in source-receiver distances.

59 4.3 Finite Difference Time Domain Implementation

FDTD numerical simulations in 2 and 3 dimensions have been recently implemented with success in studies of sound propagation (Blumrich and Heimann, 2002; Ostashev et al., 2005;

Salomons et al., 2002; Symons et al., 2004). Wilson et al. (2004) showed that FDTD simulations are capable of reproducing acoustic propagation with reflections from hills and valleys, diffractions over hills, and wind motion. Commonly, these methods demand extensive computing-time. Three-dimensional FDTD simulations of infrasound propagation in a moving atmosphere with 5x107 nodes required 6 hours of CPU time on a parallel processing cluster with

48 processors (Symons et al., 2004). To expedite simulations with various scenarios, we limited our experiments to 2D.

In the FDTD simulation used here, we consider that acoustic propagation is governed by a set of two, coupled, first-order partial differential equations (Eq.1 and Eq. 2) obtained from a more complete set of 5 fundamental equations of fluid dynamics (Ostashev et al., 2005).

Reduction to the simpler set of equations involved the following assumptions: (1) The fluid dynamic equations were linearized with respect to pressure (p) and velocity (w) fluctuations related to the propagating sound wave; (2) the ambient medium is adiabatic; (3) the fluid is incompressible (divergence free); and (4) the effect of the ambient medium pressure gradient on the sound field is ignored (Ostashev et al., 2005; Wilson and Liu, 2004). The two resulting, coupled, first order differential equations for the acoustic pressure and the acoustic particle velocity are:

∂p = −()vr ⋅∇ p − ρc 2∇ ⋅ wr + ρc 2Q (1) ∂t r ∂wr ∇p F = −()wr ⋅∇ vr − (vr ⋅∇)wr − + (2) ∂t ρ ρ

60 Where ρ is the air density, c is the adiabatic speed of sound, v is the vectorial wind velocity, F is the source force acting on the medium (dipole sources), and Q is the mass source

(monopole sources). This set of equations can be used for wave propagation in non-moving, homogeneous uniformly moving, vertically stratified, or turbulent medium, offering the same or wider applicability than other equations used for analytical and numerical studies of sound propagation (Ostashev et al., 2005).

In order to solve equations (1) and (2), we use finite differences of pressure (p) and velocity (w) fluctuations on a spatially staggered grid. Parameters, such as p, adiabatic bulk modulus κ=ρc2, Q, and the mass buoyancy b=1/ρ are stored at integer nodes i∆x and j∆y, where i and j are integer indices. ∆x and ∆y are the spatial grid intervals in the horizontal and vertical orientations. In the staggered grid approach, quantities wx and vx are defined at the points half a grid interval off the reference grid, say (i±1/2)∆x, instead of i∆x. The y-components wy and vy are stored at staggered points (j±1/2)∆y.

We used a spatially scaled model with 16,000 and 9,000 m in x and y-directions, respectively. 2D profiles were computed using a 20 m-resolution digital elevation map of

Tungurahua area. These profiles contained the vent and each of the infrasound stations (A, B, and C). In these models, the node spacing is 13.3 m in horizontal (∆x) and 12 m in the vertical

(∆y) directions, giving a total of 9x105 nodes for the two-dimensional grid model. Topography was interpolated at grid nodes using splines. The time step was set to 10.8 ms, to satisfy the

Courant condition C<1/√2 (Kreyszig, 1993), that is required for numerical stability of the 2- dimensional FDTD calculation. The Courant number (C) for the downwind direction is defined by Eq.3.

(c + v)∆t C = (3) 1 (∆x) −2 + (∆y) −2

61 During extreme wind conditions (v=14 m/s), the Courant number 0.42, still satisfied the stability condition.

A 2,000-m thick rigid layer, representing the high-velocity geological basement, is placed at 0 m on the y-axis of the simulation model. In order to provide a realistic representation of the earth’s surface (volcano-clastic deposits with a vegetative covering), a porous layer is placed over the rigid bottom layer. Large impedance contrasts across rigid layers represent reflective surfaces, while porous bodies allow sound energy to penetrate and dissipate by viscous attenuation and thermal conduction. Highly porous surfaces, such as surfaces covered by snow, forest, and grass, absorb a large fraction of incident acoustic energy (Wilson and Liu, 2004). On the other hand, surfaces with low porosity reflect most of the incident energy. At stratovolcanoes like Tungurahua, the porous material represents the low-velocity volcanic edifice above the basement rocks (Molina et al., 2005). The porous volcanic surface is characterized by typical parameters for soil with vegetative covering (Wilson and Liu, 2004): static flow resistivity =

2x105 Pa-s m-2, tortuosity parameter = 1.4, and porosity = 0.5. Absorbing layers are placed at the sides and upper edges of the grid to suppress spurious reflections.

Implementation of accurate propagation models of infrasound in the atmosphere requires detailed knowledge of atmospheric state variables and their functional dependence (Drob et al.,

2003). In simulations discussed here, a uniform atmosphere with horizontal wind vectors was assumed. An adiabatic speed of sound C0=331.6 m/s was estimated using an ambient temperature of 0º C derived from temperature records saved at seismic stations, the average of ambient temperature measured at the El Tablon meteorological station and a temperature gradient of -

6.5º/km (Ruiz et al., 2005). In the far-field, infrasound propagation requires knowledge of background pressure and temperature gradients since these waves penetrate the upper atmosphere

(Drob et al., 2003; Wilson and Liu, 2004). In the near-field analysis presented here, these effects may be ignored.

62 For the acoustic simulation, a single-point pressure source is located inside the crater at

4,700 m above sea level (2,700 m in the FDTD model), few meters above the porous earth surface

(Fig. 4.6). The crater is surrounded by steep inner walls (>45°) with more than 100 m of topographic relief. For a source function we used a Ricker wavelet, a simple, short-duration, one- dimensional wavelet point source. The source function has characteristic frequencies fi of 1, 2, and 3 Hz, the dominant frequency range of infrasonic explosion signals. Since winds at

Tungurahua generally flow from East to West, simulations on stations A and B used wind vectors ranging from -2 to 14 m/s in station-ward direction (from vent to station). In the case of station

C, simulations used wind vectors in the station-ward direction. The vertical component of wind flow is ignored because it is small relative to the horizontal component in most outdoor settings

(Li and Wang, 1997).

4.4 Results and discussion

4.4.1 Effect of topographic barriers.- To quantify the effects of topographic variations on infrasonic waves generated by Tungurahua explosions, travel times and amplitudes from 2-Hz waveforms with and without topography are compared (Fig. 4.6). Using cross-correlation of waveforms, we found the largest arrival time perturbation due to topography at the closest station to the vent (A ~2%). Anomalies were smaller at more distant stations B (0.9%) and C (0.8%).

Topographic effects on amplitude attenuation, however, are stronger at A, where using topography waveforms are about 4 times smaller than those without topography. At B, ratio

Atopo/Aflat is 0.29, and at A is 0.42. These tests suggest that topographic effects on travel times and amplitudes are stronger at shorter propagation distances, where the influence of topographic structures near the source is larger.

63 Travel time effect Amplitude effect 1.8 v=0 m/s v=0 m/s 0.36 1.7

1.6 0.34 1.5 1/3

1.4 0.32

1.3

0.3 1.2 travel time anomaly % Amplitude ratio Atopo/Aflat 1.1 0.28 1

0.9 0.26 MAS JUI RUN

0.8 1/4 3 4 5 6 3 4 5 6 distance source−receiver km distance source−receiver km

Fig. 4.6 Left: Delay times (in %) observed on 2-Hz infrasound waves propagating with and without topography without wind (v=0). Right: Comparison between amplitudes of infrasound waves propagating with and without topography. Ratios between the amplitudes are given as fractions at the right axis.

4.4.2 Effect of frequency.- The amplitude reduction provided by a barrier is a function of frequency, effective barrier height, and the angle through which the sound wave must turn (Ford,

1970). Considering that explosion events have peaks frequencies in the 1-3 Hz range, we explore the effect of the frequency in infrasound attenuation caused by topographic barriers, keeping constant topographic and atmospheric conditions. Numerical simulations at different frequencies show that if we increase the frequency from 1 to 3 Hz, waveform amplitudes at station decrease about 3 times. This amplitude reduction is related to the ratio between the wavelength and topographic barrier dimensions. Shorter wavelengths have more difficulties to bend around barriers than low frequencies (larger wavelengths) making topographic events stronger for high frequency signals.

4.4.3 Wind effect on propagation with topographic boundaries.- Fig. 4.7 shows snapshots of infrasonic wave propagation in a homogeneous, moving medium with a topographic

64 profile containing the crater and station A. Propagation simulations show that infrasound generated inside the crater is followed by reflections on the inner crater walls and the crater floor

(Fig. 4.7A). The propagation front subsequently rises with a small portion of the energy propagating down-slope (Fig. 4.7B and Fig. 4.7C). Diffracted infrasound waves propagate down the volcano flanks (Fig. 4.7B-D). Figure 4.7 also shows that infrasound propagates with higher velocity in the direction of station A (right to left) due to wind advection.

The effect caused by wind and topography on infrasound amplitudes can be seen using synthetic waveforms at a string of equally-spaced receiver points on the SW flank, that passes through station A (Fig. 4.8). Amplitude decay is twice larger than expected 1/r attenuation at all points, except the first one (240 m from vent) that is considered as reference. For 2-Hz waveforms the amplitude decay at points close to the topographic barrier is larger than at points far from the vent, confirming that topographic effects on amplitudes are stronger at shorter propagation distances.

65

Fig. 4.7 Simulation of infrasound propagation from the crater floor to station A over real topographic conditions. Predominant wind direction is E-W (right to left). At D, infrasound passes through station A, located at -3200 m in horizontal distance from the vent. Horizontal and vertical axes have a 1:10 scale.

66

Synthetic waveforms v=10 m/s amplitude reduction 0.6 1

0.4 0.9 0.2

x=240 m 0 0.8 975 m −0.2 1710 m 0.7 −0.4 2450 m

3190 m normalized pressure −0.6 normalized pressure 0.6 3930 m −0.8 0.5 −1

−1.2 0.4 0 5 10 15 20 0 1000 2000 3000 4000 time s distance m

Fig. 4.8 Simulated waveforms recorded in a radial profile from the crater rim (x=240 m from the source), passing through A station (x=3190 m). Source function has a frequency of 2 Hz. Waveform amplitudes are corrected by geometrical spreading attenuation. A significant amplitude decay is observed at stations located in the upper flank at distances of 975, 1710 and 2450 m from the vent, where amplitudes are less than a half of expected by geometrical spreading.

In order to compare the delay times of synthetic waveforms for different wind conditions, we plot the delay times between each pair of stations for realistic wind velocity ranges (Fig. 4.9).

Contour lines show the distribution of 81 computed delay times of synthetic waveforms. Large delay times are observed for the following conditions: high wind velocities in the vent-A direction and small velocities in the vent-B direction (Fig. 4.9A), high wind velocities in the vent-

A and C-vent directions (Fig. 4.9B), and high velocities in the vent-B and C-vent directions (Fig.

4.9C). In FDTD simulations, station-ward winds decrease infrasound travel times recorded at A

67 and B (station-ward direction) by 3-4 % when a 14 m/s downwind velocity is introduced. With vent-ward wind, we observed a 4 % increase in travel time during windy conditions (v=14 m/s) at

C compared to static conditions.

Delay times obtained from FDTD simulations were compared with observed delay times of infrasound arrivals at the three stations (shaded areas in figure 4.9). Synthetic and observed delay times agree at high wind speeds in the direction of A, low speeds in the direction of B, and high wind velocities in the opposite direction toward C. Infrasound signals with 9 m/s wind velocity towards A, 4 m/s in the direction of B, and -7 m/s towards C, is one of the solutions that satisfies these conditions (Fig. 4.9). In the case of A and B, these vectors are inside the range of wind vectors inferred from observations of ash-clouds and meteorological data from El Tambo

(8.5 km from the vent). In the case of C, there is a small discrepancy between inferred wind vectors and those estimated from ash-cloud data. This discrepancy may arise because C is furthermost from El Tablon, and ash-clouds preferentially move in the opposite direction of C.

Amplitudes of 2-Hz infrasound waves at station C propagating through an atmosphere with 14 m/s upwind velocity are 33% smaller than amplitudes at still atmosphere. Downwind (14 m/s) causes an amplitude increase of 33% and 42% at stations A and B, respectively. Amplitude ratios between simulated waveforms generated at each pair of stations with different wind velocities are shown in Fig. 4.10. Infrasound amplitude ratios AA/AB, AB/AC, and AA/AC have the following ranges for realistic wind velocities: 1.42-0.69; 2.5-1.04; 2.8-1.1, respectively. The large spread of amplitude ratios of infrasound signals recorded at Tungurahua (Table 4.1) can not be modeled with realistic wind velocities (-2 to 14 m/s). Observed amplitude ratios between stations A and B can be modeled if wind velocity is larger in the direction of B than in the direction than A. At the other cases, wind speeds larger than 14 m/s are required to simulate the complete range of observed amplitudes ratios. Wind velocity values that satisfy the delay time distributions, do not adequately fit the distribution of amplitude ratios between real waveforms recorded at A and B stations (Fig. 4.10A). The large dispersion of amplitude ratios observed in

68 real pressure signals can be related to a variety of propagation phenomena including refraction by wind and temperature gradients and scattering by turbulence. Amplitude distortions by wind and topography have implications on energy balances of volcanic explosions where the contribution of acoustic energy is generally underestimated.

A B

+ + + + + + + + + + + + + + + + + + 5th perc 50th=2.50s 95th perc

+ + + + + + + + + + + + + + + + + + 5th perc 50th=7.58s 95th perc

0 + + + + + + + + + 0 + + + + + + + + + 1 1

+ + + + + + + + + + + + + + + + + + - A m / s - A m / s

+ + + + + + + + + + + + + + + + + +

+ + + + + + + + + + + + + + + + + +

+ + + + + + + + + + + + + + + + + + wind speed VENT wind speed VENT

+ + + + + + + + + + + + + + + + + + 05 05

+ + + + + + + + + + + + + + + + + +

0510 0510

wind speed VENT - B m / s wind speed C - VENT m/s

C

+ + + + + + + + +

+ + + + + + + + +

0 + + + + + + + + + 1 5th perc 50th=5.06s 95th perc

+ + + + + + + + +

+ + + + + + + + +

+ + + + + + + + +

+ + + + + + + + + wind speed VENT - B m / s

+ + + + + + + + + 05

+ + + + + + + + +

0510 wind speed C - VENT m/s Fig. 4.9 Contour plots showing the distribution of delay times between waveform simulations. Gray areas show the distribution delay times observed on real waveforms of explosion events. Dashed lines show the median value of delay time distributions. Dotted line mark one set of wind vectors that satisfies the observed delay times.

69

AB

o o o o o o o o o o o o o o o o o o

o o o o o o o o o o o o o o o o o o

5th perc 0 o o o o o o o o o 0 o o o o o o 50th=1.92o o o m/s m/s 1 1

o o o o o o o o o o o o o o o o o o MAS MAS

o o o o o o o o o o o o o o o o o o

o o o o o o o o o o o o o o o o o o

o o o o o o o o o o o o o o o o o o wind speed VENT wind speed VENT 95th perc 50th=0.93 o o o o o o o o o o o o o o o o o o 05 05

o o o o o o o o o o o o o o o o o o

0510 0510

wind speed VENT JUI m/s wind speed RUN VENT m/s

C

o o o o o o o o o

o o o o o o o o o 5th perc 50th=1.94

0 o o o o o o o o o 1

o o o o o o o o o

o o o o o o o o o

o o o o o o o o o

o o o o o o o o o wind speed VENTJUI m/s

o o o o o o o o o 05

o o o o o o o o o

0510 wind speed RUNVENT m/s

Fig. 4.10 Contour plots showing the distribution of amplitude ratios between 2-Hz waveform simulations at different stations. Gray areas show the distribution of amplitude ratios observed on real waveforms of explosion events. Dashed lines show the median value of amplitude ratio distributions. The set of wind vectors that better fits the observed delay times is given by dotted lines.

4.4.4 Effect of source location.- Small amplitude changes are observed if the location of the infrasound source is moved 50 m from the center of the crater to one of the their edges.

Perturbations of waveform arrival times are in the same range of the anomalies caused by wind.

If we know the predominant wind conditions, numerical modeling propagation may be useful for

70 locating volcanic explosion sources in volcanoes with multiple vents (Johnson, 2005; McGreger and Lees, 2004).

Effect of source’s characteristics.- In FDTD simulations we considered a single point source. However, actual source conditions at exploding volcanoes are likely more complex.

Dowlin (1998) presents different source conditions that produce complex amplitude patterns. For example, two monopoles located inside the crater at distances less than a half wavelength may have an increased power output if their signals have the same phase; on the other hand, the power output is decreased if these sources are out of phase; signals from a set of aligned monopoles may have strong directionality on propagation patterns depending of monopole spacing, frequency, and distance to receiver; expanding and contracting cylinders can also generate infrasound with amplitude amplification nodes in the propagation pattern.

4.5 Conclusions

Numerical simulations of infrasound propagation of explosion signals of Tungurahua volcano show that topography and wind have a significant effect on acoustic amplitudes and travel times. Although, delay time anomalies of about 1-2% are directly related to topographic structures near the infrasound source, these structures can act as barriers for infrasound propagation. Amplitude reduction cause by topography depends of signal frequeny. Explosion signals with 3-Hz frequency have an amplitude reduction three times larger than 1-Hz signals.

For explosions with a frequency peak of 2Hz, an amplitude reduction to 1/2 -1/3 is caused only by topographic structures close to the vent. Wind influences travel times by producing early arrivals if infrasound propagates downwind versus delayed arrivals for upwind propagation.

Station-ward winds are able to reduce arrival times by about 3% of the travel time. On the other hand, vent-ward winds can delay the arrival times by about 4%. Comparing delay times at different stations of synthetic and observed waveforms, wind vectors that fit observations of ash- cloud and meteorological data are vent-A: 9 m/s; vent-B: 4 m/s; and C-vent: 7 m/s.

71 Comparing amplitudes at stations located on the crater rim versus on the upper volcano flanks, a 2-fold attenuation is observed above geometrical spreading in 2Hz waveforms.

Infrasound propagating in a downwind direction has amplitudes 40% than it is expected in a still atmosphere. On the other hand, upwind infrasound amplitudes are 30% smaller. If we neglect the amplitude reduction caused by topography or wind, we could underestimate the eruption strength. Large amplitude ratio spread observed in infrasound signals of explosion events can not be fully modeled using realistic wind velocities.

Delay time anomalies caused by wind and topography can be applied to location of explosion sources, especially in volcanoes with multiple vents. In the same way, a more precise estimation of energy contribution of infrasound can be obtained if amplitude distortions caused by wind and topographic barriers are considered. This study has proved that numerical simulations of infrasound propagation in moving medium with topographic boundaries can be achieved and it can be a very valuable tool in volcano research and monitoring.

72

CHAPTER 5

GENERAL CONCLUSIONS

Tungurahua is one of the most active volcanoes in the Ecuadorean Andes. Historic activity, especially during the last eruptive period presented a predominantly vulcanian style, characterized by short-duration emission of ash columns, ballistic blocks and rocks, and infrasound shocks. Dynamic processes of eruptive activity of Tungurahua volcano were studies in three areas: a) structure of plumbing magma system beneath this volcano; b) conditions for magma fragmentation inside the volcanic conduit; and c) propagation of infrasound through a realistic medium.

Using regional tectonic earthquakes recorded in broad band seismic stations deployed on the volcano flanks, we found evidence of seismic anisotropy beneath this volcano. Polarization directions (Φ) and delay times (δt) of split shear-waves were found using a method based on the cross-correlation of displacement waveforms of shear-waves at all possible rotation angles. Non- linear inversions of Φ and δt of events recorded at stations MAS and RUN with signal-to-noise ratio >2.1 and cross-correlation coefficient >0.7 establishes two sets of anisotropy sources. The anisotropic source with the lowest inversion residuals has a NE orientation and is parallel to elongated high velocity zones found by tomographic inversion. The second set of cracks has an orientation nearly perpendicular to the first set. Both sets of cracks exhibit significant offset relative to the regional tectonic stresses, suggesting that local stresses were modified by magmatic injections. During the deployment (June 29-August 13 2004), Tungurahua volcano displayed three characteristic activity phases: vent-clearing, degassing, and vent-sealing. An increase of δt was observed at the beginning of the sealing phase. Low δt values were found during the vent- clearing and degassing phases. This increase on delay times could be related to an increase in stresses beneath RUN, the closest station to the epicentral area.

During the fourth eruptive cycle of Tungurahua volcano (June-August 2004), more than

2,000 degassing events were recorded. The events are classified by waveform character and include: explosion events (the vast majority, spanning three orders of pressure amplitudes at 3.5 km from the vent, 0.1-180 Pa), jetting tremors, and sequences of repetitive infrasonic pulses, called chugging events. Travel time analysis of seismic first arrivals and infrasonic waves indicates that blast explosions start with a seismic event at a shallow depth (< 200 m), followed

~1 s later by an out-flux of gas, ash and solid material through the vent. Cluster analysis of infrasonic signals from explosion events was used to isolate four groups of similar waveforms without apparent correlation to event size, location or time. The clustering is thus associated with different source mechanisms coexisting during all the eruptive cycles. Lack of specific threshold of explosion events suggests that shear-induced fragmentation around the conduit is a more suitable mechanism for vulcanian eruptions than plug fragmentation models.

Strong topography and winds perturb infrasound generated by volcanic explosions. We study the propagation of infrasound waves by applying a Finite-Difference Time-Domain algorithm in a moving medium (presence of wind). Local topography near the rim of the crater, obtained from digital terrain maps, obstructs direct propagation from explosion source to receiver.

Local topography affects infrasound amplitude depending on the frequency content of propagating signals. For 2-Hz infrasound waves, numerical simulations show that, although the effect of near-source topographic barriers in delay times is relatively small (1-2% of travel times), its influence on amplitude attenuation is significant, with an amplitude reduction at the volcano flanks 2-3 times larger than the expected by the 1/r geometrical spreading. This amplitude

74 attenuation is more noticeable at points close to the topographic barrier (crater rim). Wind speed mainly affects travel times: downwind traveling waves arrive early and upwind waves are delayed up to 4% of travel times with realistic wind velocities. Comparing delay times of synthetic and observed infrasound waveforms at different stations, we show that wind vectors of

9 m/s, 4 m/s and -7 m/s in the direction of A, B, and C, explain travel time anomalies observed in real data. The large spread observed on the distribution of amplitude ratios of infrasound waveforms can not be totally explained by numerical simulations, suggesting that wind anomalies not included in this study (atmospheric refractions, eddies, or scattering) or source complexities may further affect infrasound amplitudes.

75

REFERENCES

Aki, K. et al., 1978. Seismic properties of a shallow magma reservoir in the Kilauea Iki by active and passive experiments. Journal of Geophysical Research, 83: 2273–2282

Alidivirov, M. and Dingwell, D.B., 2000. Three fragmentation mechanisms for highly viscous magma under rapid decompression. J. Vol. Geotherm. Res., 100: 413-421

Babuska, V. and Cara, M., 1991. Seismic Anisotropy in the Earth. Modern Approaches in Geophysics, 10. Kluwer Academic, Dordrecht, 217 pp.

Bianco, F., Castellano, M., Milano, G., Ventura, G. and Vilardo, G., 1998. The Somma-Vesuvius stress field induced by regional tectonics: evidences from seismological and mesostructural data. J. Volcanol. Geotherm. Res., 82(1): 199-218

Bianco, F. et al., 1999. The seismic crises at Mt. Vesuvius during 1995 and 1996 Physics and chemistry of the earth. Part A, Solid earth and geodesy, 24(11-12): 977-983

Blumrich, R. and Heimann, D., 2002. A linearized Eulerian sound propagation model for studies of complex meteorological effects J. Acoust. Soc. Am.: 446-455

Booth, D.C. and Crampin, S., 1985. Shear-wave polarizations on a curved wavefront at an isotropic. free-surface. Geophys. J. Astr. Soc., 83: 31-45

Booth, D.C., Wyss, M., and Gillard, D., 1992. Shear-wave polarization alignments recorded above the Kaoiki fault zone, Hawaii. Geophysical Research Letters, 19(11): 1141-1144

Bowman, J.R. and Ando, M., 1987. Shear-wave splitting in the upper-mantle wedge above the Tonga subduction zone. Geophys. J. R. Astr. Soc., 88: 25-41

Calahorrano, A., 2001. Estudio del Origen del Enjambre Sismico de la Zona Norte de la Ciudad de Quito, durante 1998-1999, Escuela Politecnica Nacional, Quito, 188 pp.

Caplan-Auerbach, J. and McNutt, S.R., 2003. New insights into the 1999 eruption of Shishaldin volcano, Alaska, based on acoustic data. Bull. Volcanol., 65(6): 405-417

Chouet, B.A., 1996. Long-period volcano seismicity; its source and use in eruption forecasting. Nature, 380(6572): 309-316

Clarke, A.B., Voight, B., Neri, A. and Macedonio, G., 2002. Transient dynamics of vulcanian explosions and column collapse. Nature, 415: 897-901

Corredor, F., 2003. Seismic strain rates and distributed continental deformation in the northern Andes and three-dimensional seismotectonics of northwestern South America. Tectonophysics, 372: 147-166

Crampin, S., 1987. Geological and industrial implications of extensive dila-tancy anisotropy. Nature, 328: 491–496

Crampin, S., 1994. The fracture criticality of crustal rocks. Geophys. J. Int., 118: 428-438

Crampin, S. and Booth, D.C., 1985. Shear-wave polarizations near the North Anatolian Fault - II. Interpretation in termsof crack- induced anisotropy. Geophys. J. R. Astron. Soc., 83: 75- 92

Crampin, S. et al., 1986. Shear-wave polarizations in the Peter the First Range indicating crack- induced anisotropy in a thrust-fault regime. Geophys. J. Roy. Astr. Soc., 84: 401-412

Crampin, S. and Chastin, S., 2003. A review of shear-wave splitting in the crack-critical crust. Geophys. J. Int., 155: 221-240

Crampin, S. and Lovell, J.H., 1991. A decade of shear wave splitting in the Earth’s crust: what does it mean? what use can we make of it? and what should we do next? Geophys. J. Int., 107: 387-407

Crampin, S. and McGonigle, R., 1981. The variation of delays in stress-induced anisotropic polarization anomalies. Geophysical Journal of the Royal Astronomical Society, 64(1): 115-131

Crampin, S. and Peacock, S., 2005. A review of shear-wave splitting in the compliant crack- critical anisotropic Earth. Wave Motion, 41: 59-77

Crampin, S., Peacock, S., Gao, Y. and Chastin, S., 2004. The scatter of timedelays in shear-wave splitting above small earthquakes. Geophys. J. Int., 156: 39–44

Crampin, S., Volti, T., Chastin, S., Gudmundsson, A. and Stefánsson, R., 2002. Indication of high pore-fluid pressures ina seismically-active fault zone. Geophys. J. Int., 151: F1-F5

Davis, J.C., 2002. Statistics and data analysis in geology. Wiley, New York Chichester, xvi, 638 pp.

Diller, K., Clarke, A.B., Voight, B. and Neri, A., 2006. Mechanisms of conduit plug formation: Implications for vulcanian explosions. Geophysical Research Letters, 33: L20302

Dobran, F., 2001. Volcanic processes : mechanisms in material transport. Kluwer Academic/Plenum Publishers, New York, xxviii, 590 pp.

Dowling, A.P., 1998. Seady-State Radiation from Sources. In: M. Crocker (Editor), Handbook of Acoustics. Wiley-IEEE, pp. 99-117.

Draper, N.R. and Smith, H., 1998. Applied regression analysis. Wiley Interscience Publication, New York, 505-566 pp.

77

Drob, D.P., Picone, J.M. and Garcés, M.A., 2003. Global morphology of infrasound propagation. J. Geophys. Res., 108(D21)

Ego, F., Sébrier, M., Lavenu, A., Yepes, H. and Egüez, A., 1996. Quaternary state of stress in the Northern Andes and the restraining bend model for the Ecuadorian Andes. Tectonophysics, 259(1-3): 101-116

Egred, J., 2004. Historia de las erupciones del volcán Tungurahua. Escuela Politécnica Nacional, Instituto Geofísico.

Eguez, A. et al., 2003. Database and Map of Quaternary faults and folds of Ecuador and its off- shore regions. 03-289, USGS.

Elkibbi, M., Yang, M. and Rial, J.A., 2005. Crack-induced anisotropy models in The Geysers geothermal field. Geophysical Journal International, 162(3): 1036-1048

Evers, L.G. and Haak, H.W., 2003. Tracing a meteoric trajectory with infrasound. Geophysical Research Letters, 30(24, 2246): 1-4

Ford, R.D., 1970. Introduction to Acoustics. Elsevier, New York, 154 pp.

Fukao, Y., 1984. Evidence from core-reflected shear waves for anisotropy in the Earth's mantle. Nature, 309: 695 - 698

Gaile, G.L. and Burt, J.E., 1980. Directional Statistics. . Concepts and Techniques in Modern Geography, 25. University of East Anglia, Norwich, 39 pp.

Gao, Y. et al., 1998. Temporal changes in shear-wave splitting at an isolated swarm of small earthquakes in 1992 near Dongfang, Hainan Island, southern China. Geophysical Journal International, 135: 1-102

Garces, M.A. et al., 1999. Infrasonic precursors to a vulcanian eruption at Sakurajima Volcano, Japan. Geophysical Research Letters, 26(16): 2537-2540

Garces, M.A., McNutt, S.R., Hansen, R.A. and Eichelberger, J.C., 2000. Application of wave- theoretical seismoacoustic models to the interpretation of explosion and eruption tremor signals radiated by Pavlof Volcano, Alaska. Journal of Geophysical Research, 105(2): 3039-3058

Gerst, A., 2003. Temporal Changes in Seismic Anisotropy as a new Eruption Forecasting Tool?, Victoria University of Wellington, Wellington.

Gerst, A. and Savage, M.K., 2004. Seismic Anisotropy Beneath Ruapehu Volcano: A Possible Eruption Forecasting Tool. Science, 306: 1543-1547

Giampiccolo, E., Masumeci, C., Malone, S.D., Gresta, G. and Privitera, E., 1999. Seismicity and Stress-Tensor Inversion in the Central Washington Cascade Mountains. Bull. Seis. Soc. Am., 89(3): 811-821

78 Guillier, B. et al., 2001. Seismological evidence on the geometry of the orogenic system in central-northern Ecuador (South America). Geophysical Research Letters, 28(19): 3749- 3752

Hagerty, M.T., Schwartz, S.Y., Garces, M.A. and Protti, M., 2000. Analysis of seismic and acoustic observations at Arenal Volcano, Costa Rica, 1995-1997. Journal of Volcanology and Geothermal Research, 101(1-2): 27-65

Hall, M.L., Robin, C., Beate, B., Mothes, P. and Monzier, M., 1999. Tungurahua Volcano, Ecuador; structure, eruptive history and hazards. Journal of Volcanology and Geothermal Research, 91(1): 1-21

Hiramatsu, Y., Ando, M., Tsukuda, T. and Ooida, T., 1998. Three-dimensional image of the anisotropic bodies beneath central Honshu, Japan. Geophys. J. Int., 135: 801-816

Hudson, J.A., 1981. Wave speeds and attenuation of elastic waves in material containing cracks. Geophysical Journal of the Royal Astronomical Society, 64: 133-150

Iguchi, M. and Ishihara, K., 1990. Comparison of earthquakes and air-shocks accompanied with explosion eruptions at Sakurajima and Suwanosejima Volcanoes. Disas. Prev. Res. Inst., Kyoto Univ. , 33(B-1): 1-12

Jaupart, C. and Vergniolle, S., 1989. The generation and collapse of a foam layer at the roof of a basaltic magma chamber. Journal of Fluid Mechanics, 203: 347-380

Johnson, J.B., 2003. Generation and propagation of infrasonic airwaves from volcanic explosions. Journal of Volcanology and Geothermal Research, 121(1-2): 1-14

Johnson, J.B., 2005. Source location variability and volcanic vent mapping with a small-aperture infrasound array at Stromboli Volcano, Italy. Bulletin of Volcanology, 67: 1-14

Johnson, J.B. and Aster, R.C., 2005. Relative partitioning of acoustic and seismic energy during Strombolian eruptions. J. Volcanol. Geotherm. Res., 148: 334-354

Johnson, J.B., Aster, R.C. and Kyle, P.R., 2004. Volcanic eruptions observed with infrasound. Geophys. Res. Lett., 31(L14604), doi:10.1029/2004GL020020.

Johnson, J.B. et al., 2003. Interpretation and utility of infrasonic records from erupting volcanoes. Journal of Volcanology and Geothermal Research, 121(1-2): 15-63

Johnson, J.B., Lees, J.M. and Gordeev, E., 1998. Degassing explosions at Karymsky Volcano, Kamchatka. Geophysical Research Letters, 25(21): 3999-4042

Johnson, J.B., Ruiz, M.C. and McChesney, P., 1999. Low frequency acoustic monitoring and wind-filtering at active volcanoes. Eos (Transactions, American Geophysical Union), 80(46): F1148

Kaufman, L. and Rousseeuw, P.J., 1990. Finding Groups in Data. John Wiley & Sons, New York, 342 pp.

79 Kreemer, C., Holt, W.E. and Haines, J.A., 2003. An integrated global model of present-day plate motions and plate boundary deformations. Geophys. J. Int., 154: 8-34

Kreyszig, E., 1993. Advanced engineering mathematics. Wiley, New York.

Le Pennec, J.L., Jaya, D. and van der Plicht, J., 2006. Recurrence pattern of Late Holocene eruptions at Tungurahua volcano (Ecuador) revealed by radiocarbon dating of charcoal in deposits, Cities on Volcanoes IV. IAVCEI, pp. 64.

Lees, J.M., Gordeev, E.I. and Ripepe, M., 2004. Explosions and Periodic Tremor at Karymsky Volcano, Kamchatka, Russia. Geophysical Journal International, 0(0), doi:10.1111/j.1365-246X.2004.02239.x.

Legrand, D. , Calahorrano, A., Guillier, B., Rivera, L. , Ruiz, M., Villagomez, D., Yepes, H., 2002. Stress tensor analysis of the 1998-1999 tectonic swarm of northern Quito related to the volcanic swarm of Guagua Pichincha volcano, Ecuador. Tectonophysics, 344(1): 15- 36

Li, K.M. and Wang, Q., 1997. Analytical Solutions for Outdoor Sound Propagation in the Presence of Wind. The Journal of the Acoustical Society of America, 102(4): 2040-2049

Marchetti, E., Ichihara, M. and Ripepe, M., 2004. Propagation of acoustic waves in a viscoelastic two-phase system; influence of gas bubble concentration. Journal of Volcanology and Geothermal Research, 137(1-3): 93-108

Mardia, K.V., 1972. Statistics of Direction Data. Academic press, New York.

Marson-Pidgeon, K. and Savage, M.K., 1997. Frequency-dependent anisotropy in Wellington, New Zealand. Geophysical Research Letters, 24(24): 3297-3300

McGreger, A.D. and Lees, J.M., 2004. Vent Discrimination At Stromboli Volcano, Italy. Journal of Volcanology and Geothermal Research, 137(1-3): 169-185, doi:10.1016/j.jvolgeores.2004.05.007.

Melnik, O.E. and Sparks, R.S.J., 2005. Controls on conduit magma flow dynamics during lava dome building eruptions. Journal of Geophysical Research, 110(B2): B02209

Miller, V. and Savage, M., 2001. Changes in Seismic Anisotropy After Volcanic Eruptions: Evidence from Mount Ruapehu. Science, 293(5538): 2231 - 2233

Molina, I., 2001. Actividad Pre-eruptiva del volcan Tungurahua, Universidad de Caldas, Manizales-Colombia, 201 pp.

Molina, I., Kumagai, H., Le Pennec, J.L. and Hall, M., 2005. Three-dimensional P-wave velocity structure of Tungurahua Volcano, Ecuador. Journal of Volcanology and Geothermal Research, 147(1-2): 144-156

Munson, C., Thurber, C. and Li, Y., 1993. Observations of shear wave splitting of the southeast flank of Mauna Loa volcano, Hawaii. Geophys. Res. Lett., 20: 1139-1142

80 Munson, C., Thurber, C., Li, Y. and Okubo, P.G., 1995. Crustal shear wave splitting in southern Hawaii: Spatial and temporal analysis. J. Geophys. Res., 100(B10): 20,367-20,377

Nuttli, O., 1961. The effect of the earth’s surface on the S-wave particle motion. Bull. Seism. Soc. Am., 51: 237-246

Ostashev, V.E., 1997. Acoustics in Moving Inhomogeneous Media. E&FN SPON, London.

Ostashev, V.E. et al., 2005. Equations for finite-difference, time-domain simulation of sound propagation in moving inhomogeneous media and numerical implementation. J. Acoust. Soc. Am. , 117(2): 503-517

Papale, P., 1998. Volcanic Conduit Dynamics. In: A. Freundt and M. Rosi (Editors), From Magma to Tephra. Elsevier, pp. 55-89.

Parfitt, E.A. and Wilson, L., 1995. Explosive volcanic eruptions; IX, The transition between Hawaiian-style lava fountaining and Strombolian explosive activity. Geophysical Journal International, 121(1): 226-232

Pennington, W.D., 1981. Subduction of the Eastern Panama Basin and Seismotectonics of Norhwestern South America. Journal of Gephysical Research, 86(B11): 10753-10770

Pison, G., Struyf, A. and Rousseeuw, P.J., 1999. Displaying a clustering with CLUSPLOT. Comput. Stat. Data Anal., 30(4): 381-392, doi:10.1016/S0167-9473(98)00102-9.

Reschenhofer, E., 2001. The Bimodality Principle. Journal of Statistics Education, 9(1)

Ripepe, M., Ciliberto, S. and Della-Schiava, M., 2001. Time Constrained for Modelling Source Dynamics of Volcanic explosions at Stromboli. J. Geophys. Res., 160: 8713-8727

Roman, D.C. and Cashman, K.V., 2006. The origin of volcano-tectonic earthquake swarms. Geology, 34(6): 457-460

Roman, D.C., Moran, S.C., Power, J.A. and Cashman, K.V., 2004. Temporal and Spatial Variation of Local Stress Fields before and after the 1992 Eruption of Crater Peak Vent, Mount Spurr Volcano, Alaska. Bulletin of the Seismological Society of America, 94(6): 2366-2379

Rouland, D. et al., 2001. The December 1994 seismic swarm near Aoba (AMBAE) volcano, Vanuatu, and its relationship with the volcano processes. Tectonophysics, 338: 23-44

Ruiz, G., Le-Pennec, J.L., Palacios, E., Hall, M. and Yepes, H., 2005. Calculo indirecto del volumen del material emitido por el Volcan Tungurahua. Investigaciones en Geociencias Instituto Geofisico - Escuela Politecnica Nacional

Ruiz, M., Hall, M., Samaniego, P., Ruiz, G. and Villagomez, D., 1999. Seismic activity in Tungurahua volcano: correlation between tremor and precipitation rates, 4th International Symposium on Andean Geodynamics, Gottingen-Germany.

Ruiz, M.C., Lees, J.M. and Johnson, J.B., 2006. Source constraints of Tungurahua volcano explosion events. Bulletin of Volcanology, 68: 480-490

81

Sack, R.A., West, M., 1995. A parabolic equation for sound propagation in two dimensions over any smooth terrain profile: The generalised terrain parabolic equation (GT-PE). Applied Acoustics, 45(2): 113-129

Salomons, E.M., Blumrich, R. and Heimann, D., 2002. Eulerian time-domain model for sound propagation over a finite-impedance ground surface.Comparison with frequency-domain models. Acta Acustica, 88: 483-492

Samaniego, P. et al., 2003. Los peligros volcánicos asociados con el Tungurahua. Los peligros volcánicos en Ecuador, 1. Corporación Editora Nacional, Quito, 100 pp.

Savage, M.K., 1999. Seismic Anisotropy and Mantle Deformation: What have we learned from Shear Wave Splitting. Rev. Geophys., 37(1): 65-106

Savage, M.K., Shih, X.R., Meyer, R.P. and Aster, R.C., 1989. Shear-wave anisotropy of active tectonic regions via automated S-wave polarization analysis. Tectonophysics, 165(1-4): 279-292

Self, S., Wilson, L. and Nairn, I.A., 1979. Vulcanian eruption mechanisms. Nature, 277: 440-443

Sparks, R.S.J., 1978. The dynamics of bubble formation and growth in magmas: A review and analysis. J. Volcanol. Geotherm. Res., 3: 1-37

Symons, N.P. et al., 2004. 3D staggered-grid finite-difference simulation of sound refraction and scattering in moving media, 11th International Symposium on Long Range Sound Propagation, Fairlee, VT.

Tang, C., Rial, J.A. and Lees, J.M., 2005. Shear-wave splitting: A diagnostic tool to monitor fluid pressure in geothermal fields. Geophysical Research Letters, 32(21): L21317, doi:10.1029/2005GL023551.

Tobin, G.A. and Whiteford, L.M., 2002. Community Resilience and Volcano Hazard: The Eruption of Tungurahua and Evacuation of the Faldas in Ecuador. Disasters, 26: 28-48

Trenkamp, R., Kellogg, J.N., Freymueller, J.T. and Mora, H., 2002. Wide plate margin deformation, southern Central America and northwestern South America, CASA GPS observations. Journal of South American Earth Sciences, 15: 157-172

Vergniolle, S. and Brandeis, G., 1994. Origin of the sound generated by strombolian explosions. Geophysical Research Letters, 21(18): 1959-1962

Volti, T. and Crampin, S., 2003. A four-year study of shear-wave splitting in Iceland: 2. Temporal changes beforeearthquakes and volcanic eruptions, in New insights into structural interpretation and modelling. In: D.A. Nieuwland (Editor), Geol. Soc. Lond., Spec. Publ. Geol. Soc. Lond., pp. 135-149.

Wilson, D.K. and Liu, L., 2004. Finite-Difference, Time-Domain Simulation of Sound Propagation in a Dynamic Atmosphere. TR-04-12, U.S. Army Engineer Research and Development Center / Cold Regions Research and Engineering Laboratory.

82 Winter, T., 1990. Mécanismes des déformations récentes dans les Andes Equatoriennes, Université de Paris Sud-Orsay, Paris.

Yang, M., 2004. Inversion of Shear-Wave Splitting Data in Geothermal Reservoirs, University of North Carolina, Chapel Hill, 92 pp.

Yang, M., Elkibbi, M. and Rial, J.A., 2005. An inversion scheme to model subsurface fracture systems using shear wave splitting polarization and delay time observations simultaneously. Geophys. J. Int., 160: 936-947,

83