TRANSITION CRUSTAL ZONE VS MANTLE RELAXATION: POSTSEISMIC DEFORMATIONS FOR SHALLOW NORMAL FAULTING EARTHQUAKES

Similar documents
High-Harmonic Geoid Signatures due to Glacial Isostatic Adjustment, Subduction and Seismic Deformation

Ground displacement in a fault zone in the presence of asperities

PLATE DEFORMATION - 2

Rotation of the Principal Stress Directions Due to Earthquake Faulting and Its Seismological Implications

Seismotectonics of intraplate oceanic regions. Thermal model Strength envelopes Plate forces Seismicity distributions

Slab pull, slab weakening, and their relation to deep intra-slab seismicity

Surface changes caused by erosion and sedimentation were treated by solving: (2)

Far-reaching transient motions after Mojave earthquakes require broad mantle flow beneath a strong crust

Can geodetic strain rate be useful in seismic hazard studies?

Elizabeth H. Hearn modified from W. Behr

Data Repository Hampel et al., page 1/5

Source properties of the Central Italy earthquake sequence from inversion of long-period and broad-band seismograms

United Nations Educational Scientific and Cultural Organization and International Atomic Energy Agency

A mechanical model of the San Andreas fault and SAFOD Pilot Hole stress measurements

Earthquake and Volcano Deformation

INGV. Giuseppe Pezzo. Istituto Nazionale di Geofisica e Vulcanologia, CNT, Roma. Sessione 1.1: Terremoti e le loro faglie

Predicted reversal and recovery of surface creep on the Hayward fault following the 1906 San Francisco earthquake

Effect of an outer-rise earthquake on seismic cycle of large interplate earthquakes estimated from an instability model based on friction mechanics

Basics of the modelling of the ground deformations produced by an earthquake. EO Summer School 2014 Frascati August 13 Pierre Briole

Originally published as:

The Size and Duration of the Sumatra-Andaman Earthquake from Far-Field Static Offsets

P-SV MULTIMODE SUMMATION DIFFERENTIAL SEISMOGRAMS FOR LAYERED STRUCTURES

Zoning surface rupture hazard along normal faults: insight from the 2009 M w 6.3 L Aquila, central Italy, earthquake and other global earthquakes

Crags, Cracks, and Crumples: Crustal Deformation and Mountain Building

What is the LAB Dynamically: Lithosphere and Asthenosphere Rheology from Post-loading Deformation

Co-seismic Gravity Changes Computed for a Spherical Earth Model Applicable to GRACE Data

Variations in Tremor Activity and Implications for Lower Crustal Deformation Along the Central San Andreas Fault

Unit 4 Lesson 7 Mountain Building

3 4 0B. physics COEXISTING CONTRACTION-EXTENSION CONSISTENT WITH BUOYANCY OF THE CRUST AND UPPER MANTLE IN NORTH-CENTRAL ITALY.

The influence of short wavelength variations in viscosity on subduction dynamics

Secondary Project Proposal

Rheology III. Ideal materials Laboratory tests Power-law creep The strength of the lithosphere The role of micromechanical defects in power-law creep

Elastic models of deformation in nature: why shouldn t we use the present day fault geometry?

M 7.0 earthquake recurrence on the San Andreas fault from a stress renewal model

Monitoring long-term ground movements and Deep Seated Gravitational

On the validity of time-predictable model for earthquake generation in north-east India

Graves and Pitarka Method

Global Tectonics. Kearey, Philip. Table of Contents ISBN-13: Historical perspective. 2. The interior of the Earth.

Significance of the fundamental mantle rotational relaxation mode in polar wander simulations

Chapter Review USING KEY TERMS. asthenosphere uplift continental drift. known as. tectonic plates move. object. UNDERSTANDING KEY IDEAS

Gravitational deformation after the April 6, 2009 L Aquila Earthquake detected by Cosmo-SkyMed

Depth (Km) + u ( ξ,t) u = v pl. η= Pa s. Distance from Nankai Trough (Km) u(ξ,τ) dξdτ. w(x,t) = G L (x,t τ;ξ,0) t + u(ξ,t) u(ξ,t) = v pl

Summary so far. Geological structures Earthquakes and their mechanisms Continuous versus block-like behavior Link with dynamics?

Magnitude 7.5 PALU, INDONESIA

Geologic Structures. Changes in the shape and/or orientation of rocks in response to applied stress

What Causes Rock to Deform?

Geodesy (InSAR, GPS, Gravity) and Big Earthquakes

A MODEL OF PLATE CONVERGENCE IN SOUTHWEST JAPAN, INFERRED FROM LEVELING DATA ASSOCIATED WITH THE 1946 NANKAIDO EARTHQUAKE

THERMAL THERMAL AND AND RHEOLOGIC SIGNATURES OF

Centroid moment-tensor analysis of the 2011 Tohoku earthquake. and its larger foreshocks and aftershocks

Seismic Activity near the Sunda and Andaman Trenches in the Sumatra Subduction Zone

Regional Geodesy. Shimon Wdowinski. MARGINS-RCL Workshop Lithospheric Rupture in the Gulf of California Salton Trough Region. University of Miami

2D versus 1D ground-motion modelling for the Friuli region, north-eastern Italy

Long-term Crustal Deformation in and around Japan, Simulated by a 3-D Plate Subduction Model

Materials and Methods The deformation within the process zone of a propagating fault can be modeled using an elastic approximation.

Separating Tectonic, Magmatic, Hydrological, and Landslide Signals in GPS Measurements near Lake Tahoe, Nevada-California

SUPPLEMENTARY INFORMATION

Sendai Earthquake NE Japan March 11, Some explanatory slides Bob Stern, Dave Scholl, others updated March

Northern Sicily, September 6, 2002 earthquake: investigation on peculiar macroseismic effects

THE SEISMICITY OF THE CAMPANIAN PLAIN: PRELIMINARY RESULTS

Ground surface deformation of L Aquila. earthquake revealed by InSAR time series

The March 11, 2011, Tohoku-oki earthquake (Japan): surface displacement and source modelling

THE 1976 FRIULI (NE ITALY) THRUST FAULTING EARTHQUAKE: A REAPPRAISAL 23 YEARS LATER

Coseismic and postseismic slip distribution of the 2003 Mw=6.5 Chengkung earthquake in eastern Taiwan: Elastic modeling from inversion of GPS data

GLOBAL SOURCE PARAMETERS OF FINITE FAULT MODEL FOR STRONG GROUND MOTION SIMULATIONS OR PREDICTIONS

RELOCATION OF THE MACHAZE AND LACERDA EARTHQUAKES IN MOZAMBIQUE AND THE RUPTURE PROCESS OF THE 2006 Mw7.0 MACHAZE EARTHQUAKE

Modelling coseismic displacements during the 1997 Umbria^Marche earthquake (central Italy)

NOTES AND CORRESPONDENCE Segmented Faulting Process of Chelungpu Thrust: Implication of SAR Interferograms

Kinematics of the Southern California Fault System Constrained by GPS Measurements

EARTHQUAKE LOCATIONS INDICATE PLATE BOUNDARIES EARTHQUAKE MECHANISMS SHOW MOTION

DETAILS ABOUT THE TECHNIQUE. We use a global mantle convection model (Bunge et al., 1997) in conjunction with a

Effect of the Emperor seamounts on trans-oceanic propagation of the 2006 Kuril Island earthquake tsunami

Deformation cycles of great subduction earthquakes in a viscoelastic Earth

Lecture 2: Deformation in the crust and the mantle. Read KK&V chapter 2.10

Rupture history of the 1997 Umbria-Marche (Central Italy) main shocks from the inversion of GPS, DInSAR and near field strong motion data

Supporting Online Material for

Deformation of Rocks. Orientation of Deformed Rocks

Afterslip and viscoelastic relaxation following the 1999 M 7.4 İzmit earthquake from GPS measurements

11.1 Rock Deformation

Earthquake patterns in the Flinders Ranges - Temporary network , preliminary results

2 compared with the uncertainty in relative positioning of the initial triangulation (30 { 50 mm, equivalent to an angular uncertainty of 2 p.p.m. ove

Asish Karmakar 1, Sanjay Sen 2 1 (Corresponding author, Assistant Teacher, Udairampur Pallisree Sikshayatan (H.S.), Udairampur, P.O.

4 Deforming the Earth s Crust

Centroid-moment-tensor analysis of the 2011 off the Pacific coast of Tohoku Earthquake and its larger foreshocks and aftershocks

Widespread Ground Motion Distribution Caused by Rupture Directivity during the 2015 Gorkha, Nepal Earthquake

Christine M. Puskas. Objective: A geophysical position where a background in geophysical modeling, data processing, and GPS studies can be applied.

Sequence Stratigraphy. Historical Perspective

GPS Strain & Earthquakes Unit 5: 2014 South Napa earthquake GPS strain analysis student exercise

Magnitude 7.1 PERU. There are early reports of homes and roads collapsed leaving one dead and several dozen injured.

Answers: Internal Processes and Structures (Isostasy)

Three-dimensional numerical simulation on the coseismic deformation of the 2008 M S 8.0 Wenchuan earthquake in China

to: Interseismic strain accumulation and the earthquake potential on the southern San

Tandem afterslip on connected fault planes following the 2008 Nima-Gaize (Tibet) earthquake

P.M. Siciliani 1, C. Fidani 1,2, D. Marcelli 2, M. Arcaleni 1, S. Tardioli 1 1. Osservatorio Sismico Andrea Bina, Perugia, Italy 2

Preface and Overview. Folded strata in the mountains of Italy (ca AD), Leonardo da Vinci

Chapter 16. Mountain Building. Mountain Building. Mountains and Plate Tectonics. what s the connection?

GEOLOGY Vol. II - Mantle Dynamics and Plate Kinematics - Carlo Doglioni, Roberto Sabadini

Material is perfectly elastic until it undergoes brittle fracture when applied stress reaches σ f

Lecture 9 faults, folds and mountain building

The Mechanics of Earthquakes and Faulting

Transcription:

IC/99/80 United Nations Educational Scientific and Cultural Organization and International Atomic Energy Agency THE ABDUS SALAM 1NTKRNATIONAL CENTRE FOR THEORETICAL PHYSICS TRANSITION CRUSTAL ZONE VS MANTLE RELAXATION: POSTSEISMIC DEFORMATIONS FOR SHALLOW NORMAL FAULTING EARTHQUAKES R. Riva 1 Dipartimento di Scienze della Terra, Sezione Geofisica, Universita di Milano, via L. Cicognara 7, 1-20129 Milano, Italy, A. Aoudia 2 Dipartimento di Scienze delta Terra, Universita di Trieste, via E. Weiss 1, 1-3^127, Trieste, Italy and The Abdus Salam International Centre for Theoretical Physics, SAND Group. Trieste, Italy, L.L.A. Vermeersen 3 DEOS, Faculty of Aerospace Engineering, Delft University of Technology, Kluyverweg 1, NL-2629 HS Delft, The Netherlands, R. Sabadini 4 Dipartimento di Scienze della Term, Sezione Geofisica, Universita di Milano, via L. Cicognara 7, 1-20129 Milano, Italy and G.F. Panza 5 Dipartimento di Scienze della Terra, Universita di Trieste, via /;'. Weiss 1, 1-34127, Trieste, Italy and The Abdus Salam International Centre for Theoretical Physics, SAND Group, Trieste, Italy. MIRAMARE - TRIESTE July 1999 'E-mail: nccardo@sabadini.geofisica.unimi.it 2 E-mail: aoudia@geokuno.univ.lrieste.it 3 E-mail: b.vermeerseu@lr.tudclft.nl 4 E-mail: bob@sabadini.geofisica.un.imi.it 6 E-mail: panza gcosuno.univ.trieste.it

Abstract. Following a normal mode approach for a stratified viscoelastic Earth, we investigate the efrects induced by shallow normal faulting earthquakes on surface postseismic deformation rates, when stress relaxation occurs in the transition zone of the crust or in the asthenosphere. We focus on the Central Apennines (Italy), where the 1997 Umbria-Marche earthquake sequence took place. The layered viscoelastic structure of the crust and mantle imposes a pattern and scale on the observed coseismic and postseismic deformations with a major contribution from the transition crustal zone. Vertical rates of deformation, for viscosity values ranging between 10 18 and 10 19 Pa s in the crustal transition zone, vary between 0.7 to a few millimeters per year, lasting for several years after the earthquake.

Introduction The calculation of postseismic deformation requires the assumption of a layered viscoelastic Earth model. Nearly all viscoelastic relaxation studies [e.g. Ma and Kuznir, 1995; Pollitz et ai, 1998] assume a three-layer model schematizing the brittle upper crust (UC), the lower crust (LC) and the mantle. Seismological and geological observations from regions of active faulting suggest that the seismogenic UC is likely to be separated from the LC by a transition zone (TZ) where both brittle and ductile processes can occur [e.g. Sibson, 1982; Meissner and Strehlau, 1982; Chen and Molnar, 1983; Sibson. 1989; Scholz, 1990]. These observations are more frequent in regions of active continental extension [Jackson and White, 1989] and suggest that the thickness of such crustal layering imposes a pattern and a scale on the observed coseismic and postseismic deformation. We investigate the effects of the stress relaxation in the TZ of the crust and its interplay with the mantle on the surface vertical deformation that follows shallow normal faulting earthquakes. The modeling approach we use, based on normal mode theory, is the fully analytical one described by Vermeersen and Sabadini [1997] and Sabadini and Vermeersen [1997], and it allows us to deal with some of the complexities of the real Earth, such as crust and mantle viscoelastic layering, sphericity and self-gravitation. However, other possible aspects like lateral variations, non-linear rheologies and non-hydrostatic pre-stresses that might occur in the region are not considered. Since a large amount of data may be brought to bear on this issue, we focus on the part of Central Italy, undergoing continental extension, that experienced the recent moderate Umbria-Marche 1997 normal faulting earthquake sequence [e.g. Amato et ai, 1998; Ekstrom et al, 1998; Meghraoui et ai., 1999; Hunstad et at, 1999]. The recent CROP03 seismic reflection profile [Pialli et al., 1998], aimed to the study of the deep crust in the Northern Apennines, imaged the top of the LC using regional high-amplitude reflectors, e.g. K-Horizon: thermal and phase boundary at 450 C [Cameli ei al. 7 1998].

Below the Umbria-Marche extensional belt, the top of the LC is about 20 km deep, with a Moho around 30 km [Coli, 1998], Cattaneo et al. [1999] show an abrupt cutoff of the 1997 aftershock sequence at about 9 km of depth. Thus the seismic activity is largely restricted to the upper one third of the deforming continental crust and leaves room for a 10 km thick TZ (Fig. 1). We assume that the TZ acts as a low-viscosity layer which decouples the UC from the LC. Accordingly, in our modeling we use a five layer Earth model (Fig. 1). This 5-layer model assures that saturated continuum limits are obtained for the deformation [Sabadini and Vermeersen, 1997], The relevant stratification is consistent with the range of regional velocity models estimated through the analysis of surface-wave dispersion [Calcagnile and Panza, 1981; Du et al., 1998] and regional gravity and heat flow anomalies [Delia Vedova et a/., 1991; Marson et al, 1995]. Modeling and Results We consider the strongest earthquake of the Umbria-Marche sequence that took place on the 26.09.1997 at 9:40 (Ms = 6.0). We use the moment solution of Sarao et al. [1998] retrieved by broad-band waveform inversion, in good agreement with the CMT [Ekstrom et al, 1998], Since in this first approach to the problem we are not interested in the details of the deformation pattern, we do not consider all the complexities of the earthquake. We focus on the global characteristics of crustal deformation and on a first order estimate of the expected vertical velocities. To illustrate the sensitivity of the results on the depth of the source, we neglect the flniteness of the source in the vertical direction, and besides the source depth of 7 km [Sarao et al, 1998] we consider another case where the source is embedded at a depth of 3 km, to simulate the effects of a fault nearly reaching the surface. We thus distribute the seismic moment not over a plane but along a line of dislocations, whose length along strike is 10 km; in order to resolve the deformation produced by this fault, we summed 4500 normal modes, which ensures convergence of the solution. We limit our description to the vertical deformation, most

relevant to normal faults, evaluated at the surface along a line perpendicular to the strike of the fault, crossing its center. In the whole set of panels, the fault is dipping to the left, and the distance is measured from the intersection of the upward prolongation of the fault with the Earth surface. In Fig. 2, the depth is fixed at 7 km. Fig. 2a portrays the coseistnic and postseismic vertical deformation, measured at the surface. The cosetsmic component (solid line) shows the deformation pattern characteristic of normal faulting, with an uplifted, localized footwall and a broad subsidence in the hanging wall. The postseismic component (dashed line), that includes the coseismic one plus the extra deformation resulting from the relaxation in the TZ, is responsible for a broadening of the subsidence, that affects now also the footwall. The largest coseismic displacement is 8 cm at 10 km from the fault. Due to the relaxation in the TZ, the maximum displacement increases to 13 cm, and a broad area, that remained at rest during the coseismic deformation, is now subject to subsidence, at distances of tens of kilometers. This effect is due to the decoupling of the uppermost part of the crust with respect to the lower layers. After stress relaxation in the TZ, all the deformation is elastically supported by the UC, which is now partially free to bend being decoupled from the LC. The occurrence of normal faulting and the extension at the bottom of the UC, below the neutral plane, is responsible for the downflexure of this layer, visible in the broad subsidence of Fig. 2a. Fig. 2b provides another perspective to this physics, with the rates of vertical deformation, expected at different times after the earthquake. The wide region of negative velocity marks the broad downflexure of the UC. The two regions of positive velocity represent the peripheral response of the layer to the subsidence, and disappear a very long time after the earthquake, eventually reversing to negative values. Another interesting characteristic of Fig. 2b is the short wavelength feature at the center of the subsiding area, that follows the pattern of the coseismic displacement. We can recognize the relative maximum corresponding to the footwall, and the largest subsidence of the hanging wall, within a generalized subsidence. The

highest velocity is -0.7 mm/yr, in the hanging wall, immediately after the earthquake, at t = 0. Due to the relatively high viscosity of the TZ (10 19 Pa s) this velocity shows a minor reduction after 10 yr, and after 50 yr there is still a visible signal of 0.1 mm/yr. With a single viscoelastic layer, the rate of the vertical deformation scales inversely with the viscosity; if, for the TZ, we consider a viscosity of 10 18 Pa s, the highest velocity increases by one order of magnitude to -7 mm/yr. Since the viscosity of the TZ is largely uncertain, probably ranging between 10 18 and 10 19 Pa s, it is likely that the largest subsidence rates in the highly deforming area of the hanging wall vary in the range that we have estimated. Comparison between the ongoing GPS campaigns and model predictions will be crucial to estimate the effective viscosity of the TZ. In Fig. 2c we describe the effects of the inclusion of the stress relaxation in the mantle, where the viscosity is lowered to 10 19 Pa s, with the same sampling time of Fig. 2b. Both features of normal faulting being maintained, the striking differences with respect to Fig. 2b are: (1) the disappearance of the global subsidence responsible for the high negative velocity, (2) the persistence of the 0.4 rnm/yr difference between the footwall and the fast subsiding part of the hanging wall. The reason for the disappearance of the downflexure of the UC stands on the reduced effectiveness of the TZ, a large amount of deformation being absorbed by the relaxation in the mantle. The bending of the UC is thus substantially reduced with respect to Fig. 2b. Fig. 2d deals with stress relaxation limited to the mantle, with sampling times increased with respect to the previous cases. The characteristic deformation pattern due to normal faulting is broadened, since stress relaxation involves a larger portion of the Earth. The global subsidence disappears because the whole crust is now coupled. The amplitude of the deformation rates in the earthquake area is substantially reduced, since the whole planet is now involved in the deformation, as indicated by the slow decay with increasing distance from the epicentral area and longer relaxation times with respect to Figs.2b and c. Figs. 2b and d show that for shallow normal faulting, the TZ plays a major role in comparison with the

stress relaxation in the mantle. In Fig. 3 the source is embedded at a depth of 3 km, in the top half part of the UC. In comparison with Fig. 2, the pattern of both coseismic and postseismic displacement in Fig. 3a becomes sharper, and the largest coseismic subsidence in the hanging wait is increased by a factor of two. The uplift in the footwall, with respect to Fig. 2, is subject to a larger increase by a factor of three. This increase in the amplitude of the displacement and sharpening of its pattern is attributable to the decrease of the wavelength, caused by the shallower source depth in comparison with Fig. 2. The most interesting difference with respect to Fig. 2 stands on the upward migration of the postseismic displacement pattern with respect to the coseismic one. An extensional source located above the neutral plane of the UC causes a bending moment opposite to the one induced by the deeper source of Fig. 2, and the Earth surface is thus subject to a general uplift, rather than subsidence. From Fig. 2a and 3a we thus obtain that, due to the location of the source beneath or above the neutral plane, the postseismic curves lie completely above or below the coseismic one. The findings differ from those by Ma and Kuznir [1995] where, due to the finiteness of the source in the vertical direction, the fault cuts the whole UC through its neutral plane, which causes the crossing of the curves depicting the postseismic and coseismic vertical displacement. The difference between the highest uplift in the footwall and subsidence in the hanging wall is doubled with respect to Fig. 2. This doubling impacts also the difference in the rates of vertical displacement between these two regions in Fig. 3b, now 0.6 mm/yr in comparison with the 0.4 mm/yr of Fig. 2b. The global subsidence, caused by the deep source, disappears in the velocity pattern of Fig. 3b, in agreement with the previous observations on the flexural properties of the UC. This tendency of a general upwarping is visible also in Fig. 3c, where relaxation involves both the TZ and the mantle. In comparison with Fig. 2c, we notice also here the increase in the difference in the velocity between the footwall and the hanging wall, with an increase in the uplift rates to 0.4 mm/yr, in comparison to the 0.1 mm/yr, t =0 yr, of Fig. 2c. If

relaxation is limited to the mantle (Fig. 3d) the same observations made about Fig. 2d can be done, except that now the uplift of the footwall is enhanced with respect to the subsidence in the hanging wall. The rates of the vertical deformation cannot be compared with observations yet, surveying being under way. Our results can thus be compared only with the observed coseismic displacement, but with caution, since for our purposes we are not interested in modeling the details of the source [Hunstad et al, 1999] but rather in the physics of the expected process of postseismic deformation. If we consider the results for the shallow source, we obtain a coseismic displacement, from the footwall to the hanging wall of 20 cm, comparable with the one observed from SAR interferometry [Stramondo et al., 1999], Our results, reproducing the coseismic signal, provide a first estimate of the expected rates of vertical postseismic deformation in the Umbria-Marche epicentral area. Conclusions For shallow normal faulting earthquakes the stress relaxation in the TZ plays a major role in the pattern of the postseismic deformation, in comparison with the relaxation in the mantle. This is due to the proximity of the source to the rheological discontinuity between the elastic UC and soft TZ, which decouples the UC in such a way that its bending properties become crucial to determine the deformation pattern. An appropriate test area for studying the effects of the TZ is the central Apennines, Italy, where a clear cutoff in the depth distribution of earthquakes indicates that the seismogenic layer is limited to the first ten kilometers of the crust, and a well-developed TZ lies below this depth. Depending on the effective viscosity of this layer, postseismic vertical velocities ranging between 0.7 to a few millimeters per year are expected for several years after the earthquake, for viscosities ranging between 10 18 and 10 19 Pa s. Acknowledgments. This work is supported by the Active deformation project - AC-DEF - (Italian Ministry of Research).

References Amato, A., and 18 co-authors, The Umbria-Marche, Central Italy earthquake sequence: a first look to main shocks and aftershocks, Geophys. Res. Lett., 25, 2861-2864, 1998. Calcagnile, G. and G. F. Panza, The main characteristics of the lithosphere-asthenosphere system in Italy and surrounding regions, Pure Appl. Geophys., 119, 865-879, 1981. Cameli, G. M., I. Dini, and D. Liotta, Brittle/Ductile boundary from seismic reflection lines of Southern Tuscany (Northern Apennines, Italy), Mem. Soc. Geol. It, 52, 153-162, 1998. Cattaneo, M.. and 10 co-authors, The 1997 Umbria-Marche (Italy) earthquake sequence: analysis of the data recorded by the local and temporary networks, J. Seismology, submitted, 1999. Chen, W. P. ; and P. Molnar, Focal depths of intracontinental and intraplate earthquakes and their implications for the thermal and mechanical properties of the lithosphere, J. Geophys. Res., 88, 4183-4214, 1983. Coli, M., Considerations on the crustal modelling along the CROP-3 Profile, Mem. Soc. Geol. It, 52, 401-412, 1998. Delia Vedova, B., I. Marson, G. F. Panza, and P. Suhadolc, Upper mantle properties of the Tuscan-Tyrrhenian area: a key for understanding the recent tectonic evolution of the Italian region, Tectonophys., 195, 311-318, 1991. Du, Z. J., A. Michelini, and G. F. Panza, EurlD: a regionalised 3-D seismological model of Europe. Phys. Earth Planet Inter., 105, 31-62, 1998. Ekstrom, G., A. Morelli, and A. M. Dziewonski, Moment tensor analysis of the Central Italy earthquake sequence of September-October 1997, Geophys. Res. Lett, 25, 1971-1974, 1998. Hunstad; I., M. Anzidei, M. Cocco, P. Baldi, A. Galvani, and A. Pesci, Modelling coseismic displacements during the 1997 Umbria-Marche earthquake (Central Italy), Geophys. J. Int., in press, 1999.

Jackson, J. A., and N. X White, Normal faulting in the upper continental crust: observations from regions of active extension, J. Struct. Geol., 11, 15-36, 1989. Ma, X. Q., and N. J. Kuznir, Coseismic and postseismic subsurface displacements and strains for a dip-slip normal fault in a three layer elastic-gravitational medium, J. Geophys. Res., 100, 12,813-12,828, 1995. Marson, L, G. F. Panza, and P. Suhadolc, Crust and upper mantle models along the active Tyrrhenian rim, Terra Nova, 7, 348-357, 1995. Meghraoui, M., V. Bosi, and T. Camelbeeck, Fault fragment control in the Umbria-Marche, Central Italy, earthquake sequence, Geophys. Res. Lett., 26, 1069-1072, 1999. Meissner, R., and J. Strchlau, Limits of stress in continental crust and their relation to the depth-frequency relation of shallow earthquakes, Tectonics, 1, 73-89, 1982. Pialli, G., M. Barchi, and G. Minelli (eds), B.esults of of the CROP03 Deep Seismic Reflection Profile, Mem. Soc. Geol. It, Vol. LII t 657 pp., 1998. Pollitz. F. F., R. Burgraann, and P. Segall, Joint estimation of afterslip rate and postseismic relaxation following the 1989 Loma Prieta earthquake, J. Geophys. Res., 103, 26,975-26,992, 1998. Sabadini, R., and L. L. A., Vermeersen, Influence of lithospheric and mantle stratification on global post-seismic deformation, Geophys. Res. Lett., 2/ h 2075-2078, 1997. Sarao A., F. Romanelli, G. Costa and G. F. Panza, Source inversion and macroseismic modelling of the Umbria-Marche main events, paper presented at XXIII EGS General Assembly, Nice, France, 1998. Schol/., C. H. The Mechanics of Earthquakes and Faulting, Cambridge Univ. Press, 439 pp., 1990. Sibson, R. H., Fault zone models, heat flow, and the depth distribution of earthquakes in the continental crust of the United States, Bull. Seismol. Soc. Am., 72, 151-163, 1982. Sibson, R. H.. Earthquake faulting as a structural process, J. Struct. Geol, 11, 1-14, 1989.

Stramondo, S., and 12 co-authors, The September 26, 1997, Colfiorito, Italy, earthquakes: Modeled coseismic surface displacement from SAR interferometry and GPS, Geophys. Res. Lett., 26, 883-886, 1999. Vermeersen, L. L. A., and R. Sabadini, A new class of stratified viscoelastic models by analytical techniques, Geophys. J. Int., 129, 531-570, 1997. 10

FIGURE CAPTIONS Figure 1. Earth model and viscosity parameters used in this study. Figure 2. Coseismic and postseismic vertical surface displacements and rates calculated for a 7 km deep normal faulting line-source. The profiles are across the center of the fault (x = 0 km) and perpendicular to its strike. The fault is dipping to the left and is buried in the upper layer of Fig. 1. (a) Coseismic (solid line) and postseismic (dashed line) displacements, (b) Postseismic rates due to relaxation in the TZ expected at different times after the earthquake, with t = 0, 10, 20, 50 yr, solid, dashed, dash-dotted and dotted respectively, (c) same as in (b) including the effects of relaxation in the mantle, (d) Postseismic rates due to relaxation in the mantle expected at t= 0, 100, 500, 1000 yr, solid, dashed, dash-dotted and dotted respectively after the earthquake. Figure 3. Coseismic and postseismic vertical surface displacements and rates calculated for a 3 km deep normal faulting line-source. Terminology as in Fig. 2 11

^^_ ^ ^^ Surface Elastic upper crust (UC) Transition zone: v = 10 (TZ) Pa s 10 km. 10 1 ' Pas 20 km 2 1 Lower crust (LC): v = 10 Pa s Asthenosphere: v= 10 Pa s Inviscid core I 10 1 ' Pas 30 km 2891 km 6371 km Fig-1 12

Rate of vertical displacement (mm/yr) Vertical displacement (mm) oo

Rate of vertical displacement (mm/yr) Vertical displacement (mm) o b