arxiv: v1 [physics.geo-ph] 4 Feb 2015

Similar documents
Proceedings, International Snow Science Workshop, Banff, 2014

A porosity-based Biot model for acoustic waves in snow

Hydrogeophysics - Seismics

Wave Propagation in Fractured Poroelastic Media

Juan E. Santos a,b,c, Gabriela B. Savioli a and Robiel Martínez Corredor c a

THE ROCK PHYSICS HANDBOOK

Lawrence Berkeley National Laboratory

Acoustics in Porous Media.

3D simulations of an injection test done into an unsaturated porous and fractured limestone

Mecanum. Acoustic Materials: Characterization. We build silence. Mecanum Inc.

7.2.1 Seismic waves. Waves in a mass- spring system

Scale-up in poroelastic systems and applications to reservoirs

Receiver. Johana Brokešová Charles University in Prague

The acoustic characterization of porous media and its standards

6298 Stress induced azimuthally anisotropic reservoir - AVO modeling

D scattering of obliquely incident Rayleigh waves by a saturated alluvial valley in a layered half-space

LINK BETWEEN ATTENUATION AND VELOCITY DISPERSION

Borehole Geophysics. Acoustic logging measurements

SUMMARY INTRODUCTION EXPERIMENTAL PROCEDURE

LECTURE 5 - Wave Equation Hrvoje Tkalčić " 2 # & 2 #

Stress-induced transverse isotropy in rocks

Crosswell tomography imaging of the permeability structure within a sandstone oil field.

16 Rainfall on a Slope

Wave Propagation in Anisotropic, Anelastic, Porous and Electromagnetic Media

A Thesis. Presented to. the Faculty of the Department of Earth and Atmospheric Sciences. University of Houston. In Partial Fulfillment

Flow interpretation implications for Poro-Elastic Modeling

19 th INTERNATIONAL CONGRESS ON ACOUSTICS MADRID, 2-7 SEPTEMBER 2007

13th International Symposium on Nondestructive Characterization of Materials (NDCM-XIII), May 2013, Le Mans, France

Physical and Biological Properties of Agricultural Products Acoustic, Electrical and Optical Properties and Biochemical Property

ABSTRACT 1. INTRODUCTION

KOZENY-CARMAN EQUATION REVISITED. Jack Dvorkin Abstract

Interface conditions for Biot s equations of poroelasticity Boris Gurevich The Geophysical Institute of Israel, P.O. Box 2286, Holon 58122, Israel

2.1 Introduction to waves

Mechanics of Earthquakes and Faulting

Sound Propagation in Porous Media

Comparative review of theoretical models for elastic wave attenuation and dispersion in partially saturated rocks

inter.noise 2000 The 29th International Congress and Exhibition on Noise Control Engineering August 2000, Nice, FRANCE

Soft Bodies. Good approximation for hard ones. approximation breaks when objects break, or deform. Generalization: soft (deformable) bodies

Modeling seismic wave propagation during fluid injection in a fractured network: Effects of pore fluid pressure on time-lapse seismic signatures

Lecture 8: Tissue Mechanics

Theoretical and Experimental Studies of Seismoelectric Conversions in Boreholes

An alternative Biot s displacement formulation for porous materials

Analytical and Numerical Investigations on the Vertical Seismic Site Response

Game Physics. Game and Media Technology Master Program - Utrecht University. Dr. Nicolas Pronost

Shear Wave Velocity Estimation Utilizing Wireline Logs for a Carbonate Reservoir, South-West Iran

Wavefields in porous media: lithology, pressure and saturation effects

Modeling Scattering from Rough Poroelastic Surfaces Using COMSOL Multiphysics

A comparison and review of theories of the acoustics of porous materials.

A Multi-Physics Study of the Wave Propagation Problem in Open Cell Polyurethane Foams

Electroseismics for CO 2 storage and hydrocarbon reservoirs. Fabio I. Zyserman

Numerical simulation of the Biot slow wave in water-saturated Nivelsteiner Sandstone

DOWN-HOLE SEISMIC SURVEY AND VERTICAL ELECTRIC SOUNDINGS RABASKA PROJECT, LÉVIS, QUÉBEC. Presented to :

On measurement of mechanical properties of sound absorbing materials

Geology 229 Engineering Geology. Lecture 5. Engineering Properties of Rocks (West, Ch. 6)

Fundamentals of Ultrasonics

Finite Element and Plate Theory Modeling of Acoustic Emission Waveforms. NASA Langley Research Center. Hampton, VA *University of Denver

SPE These in turn can be used to estimate mechanical properties.

Macroscopic theory Rock as 'elastic continuum'

Site Characterization & Hydrogeophysics

3 Mathematical modeling of the torsional dynamics of a drill string

Geology 2112 Principles and Applications of Geophysical Methods WEEK 1. Lecture Notes Week 1

Validity of the limp model for porous materials: A criterion based on the Biot theory

Semeniuk, B., Göransson, P. (2018) Modelling the Dynamic Viscous and Thermal Dissipation Mechanisms in a Fibrous Porous Material In:

Elements of Rock Mechanics

A look into Gassmann s Equation

Wave Equation. Finite difference Calculation in Tesseral. Content. 1 Introduction. Finite difference solution of Wave Equation in Tesseral

Dynamic permeability in poroelasticity

Summary PHY101 ( 2 ) T / Hanadi Al Harbi

ERTH2020 Introduction to Geophysics The Seismic Method. 1. Basic Concepts in Seismology. 1.1 Seismic Wave Types

PEAT SEISMOLOGY Lecture 9: Anisotropy, attenuation and anelasticity

ROCK PHYSICS MODELING FOR LITHOLOGY PREDICTION USING HERTZ- MINDLIN THEORY

Downloaded 08/30/13 to Redistribution subject to SEG license or copyright; see Terms of Use at

Seismic Velocity Dispersion and the Petrophysical Properties of Porous Media

PEAT SEISMOLOGY Lecture 2: Continuum mechanics

Water, Inertial Damping, and the Complex Shear Modulus

TIME-DEPENDENT BEHAVIOR OF PILE UNDER LATERAL LOAD USING THE BOUNDING SURFACE MODEL

Abstract. 1 Introduction

Propagation of seismic-induced electromagnetic waves in a semi-infinite porous medium: A Fourier transform approach

Geophysical Journal International

ScienceDirect. Model-based assessment of seismic monitoring of CO 2 in a CCS project in Alberta, Canada, including a poroelastic approach

Fluid Flow Fluid Flow and Permeability

International Snow Science Workshop Grenoble Chamonix Mont-Blanc

On the study of elastic wave scattering and Rayleigh wave velocity measurement of concrete with steel bar

Attenuation and dispersion

Soil Damping Ratio: Theoretical Aspect and Measurement

Observation of shear-wave splitting from microseismicity induced by hydraulic fracturing: A non-vti story

Using the Timoshenko Beam Bond Model: Example Problem

1 Modeling Immiscible Fluid Flow in Porous Media

Chapter 1. Governing Equations of GFD. 1.1 Mass continuity

Acoustic landmine detection: a 3D poroelastic model

P- and S-Wave Velocity Measurements and Pressure Sensitivity Analysis of AVA Response

Using windscreens to improve the efficiency of noise barriers in wind: finite-difference time-domain simulations

Estimating Permeability from Acoustic Velocity and Formation Resistivity Factor

On propagation of Love waves in an infinite transversely isotropic poroelastic layer

Calibration of the petro-elastic model (PEM) for 4D seismic studies in multi-mineral rocks Amini, Hamed; Alvarez, Erick Raciel

Sound Propagation through Media. Nachiketa Tiwari Indian Institute of Technology Kanpur

Rock Mechanics Laboratory Tests for Petroleum Applications. Rob Marsden Reservoir Geomechanics Advisor Gatwick

SENSITIVITY ANALYSIS OF THE PETROPHYSICAL PROPERTIES VARIATIONS ON THE SEISMIC RESPONSE OF A CO2 STORAGE SITE. Juan E. Santos

Seismic measurement of the propagation speed of a fracture through a weak snowpack layer

Chapter 17: Waves II. Sound waves are one example of Longitudinal Waves. Sound waves are pressure waves: Oscillations in air pressure and air density

Transcription:

1 arxiv:152.1284v1 [physics.geo-ph] 4 Feb 215 Acoustic waves in a Biot-type porous snow model: The fast slow wave in light snow Rolf Sidler Department of Earth Sciences, Simon Fraser University, 8888 University Drive, BC V5A1S6 Burnaby, Canada ABSTRACT. Wave velocities, attenuation and reflection coefficients in snow can not be explained by the widely used elastic or viscoelastic models for wave propagation. Instead, Biot s model of wave propagation in porous materials should be used. However, the application of Biot s model is difficult due to the large property space of the underlying porous material. Here we use the properties of ice and air as well as empirical relationships to define the properties of snow as a function of porosity. This reduction allows to predict phase velocities and attenuation of the shear- and compressional-waves as functions of porosity or density. For light snow the peculiarity was found that the velocity of the compressional wave of the first kind is lower than the compressional wave of the second kind that is commonly referred to as the slow wave. The reversal of the velocities comes with an increase of attenuation for the first compressional wave. This is in line with the common observation that sound is strongly absorbed after a fresh snowfall. The evaluation of the sensitivity of the compressional waves to individual properties of the pore space shows a particular sensitivity of the slow wave to specific surface area, tortuosity and permeability. The results have important implications for the use of acoustic waves to evaluate snow water equivalent, specific surface area, permeability and tortuosity of snow. 1. INTRODUCTION Acoustic applications to snow have been investigated for many decades even leading to the recognition of wave velocity as a potential index property for snow (Oura, 1952; Smith, 1969; Yamada and others, 1974; Gubler, 1977; Johnson, 1982; Shapiro and others, 1997). Recently acoustic methods have been used to monitor and spatially locate avalanches (Surinach and others, ; van Herwijnen and Schweizer, 211; Lacroix and others, 212), to estimate the height of snow cover (Albert, 1; Albert and others, 213) and to estimate the snow water equivalent of dry snowpacks (Kinar and Pomeroy, 9). However, the application of acoustic methods has remained rather sparse despite the fact that acoustic equipment is inexpensive and can potentially even be used in the absence of active sound sources (van Herwijnen and Schweizer, 211). The sparse application might be due to the interpretation of acoustic measurements being complicated by the fact that sometimes lower wave velocities can be observed with the increasing density of the snow (Oura, 1952). This observation is at odds with elastic or visco-elastic wave propagation theory, for which higher velocities are expected for the considerably higher bulk and shear moduli of denser snow. Yet, the observed wave velocities can be explained with wave propagation theory for porous media, where a compressional wave of the second kind, also known as slow wave, is predicted (Johnson, 1982). Also the common observation of fresh snow absorbing a considerable amount of sound from the environment and the corresponding silence of winter corresponds much better with a porous model of wave propagation. The pertinent physical mechanisms being the low reflection coefficient at the snow surface and conversion of sound waves to strongly attenuated compressional waves. A porous model that can explain the presence of the slow compressional wave is the rigid-frame porous model (Zwikker and Kosten, 1947), commonly used in industry for nondestructive testing (Fellah and others, 7; Jocker and Smeulders, 9). The model assumes that the stiffness of the frame is much larger than the stiffness of the pore fluid with the result that two decoupled compressional waves are propagating in the medium. One traveling in the frame and one in the pore space. The large advantage of this model is that it is relatively straight forward to extract tortuosity of the pore space and pore fluid properties from the phase velocities of the slow wave. Albert and others (9) and Kinar and Pomeroy (9) recently applied the rigid-frame porous model to estimate the structural parameters of snow. When the stiffness of the frame and the stiffness of the pore fluid are in a comparable order of magnitude, the interaction between the porous frame and the pore space should be considered as in Biot s (1956a; 1962) theory for wave propagation in porous materials (Johnson, 1982). For snow this would correspond to light snow with low frame bulk moduli and to wet snow, where the pore fluid bulk modulus and viscosity are increased by water in the pore space. Here, we propose a porous snow model as a function of porosity and use it to estimate the wave velocities and attenuation of compressional and shear wave modes using plane wave solutions for Biot s (1956a) differential equation of wave propagation in porous materials. We compare the results to measurements from the literature and investigate the sensitivity of the fast and slow compressional wave to individual

2 Sidler: Acoustic waves in snow parameters of the porous model such as specific surface area (SSA). 2. POROUS SNOW MODEL An inherent problem when working with Biot-type porous models is the number of material properties involved. There are no less than ten material properties that need to be specified and it is a laborious task to characterize all these parameters by measurements at the snowpack. Additionally, measurements may also disturb the properties of the snowpack which, in fact, also rapidly change over time. We try to address this problem by reducing snow as a porous material to porosity, as its single degree of freedom. The remaining nine properties are evaluated by using a priori information as well as empirical relationships that characterize property values as a function of porosity. The model is not restricted to porosity as input and additional information can be readily used to increase the precision of the results. Snow is especially well suited for this reduction as its solid structure, its matrix, consists of ice which has reasonably constant and homogeneous properties under the conditions where snow naturally exists. The same is true for the pore fluid, which, for the purpose of this study, is air. The density of snow has a linear relationship to porosity and is interchangeably used, as the much lower density of air can be well neglected beside the much higher density of ice. This assumption is, due to generally low water saturations, often applied even for wet snow packs. Density has become a more intuitive index value for snow than porosity and we annotate the density coequally to the porosity where applicable in all the Figures. Deriving empirical relationships as a function of porosity is based on the assumption that these properties evolve from the shape of the frame and that for naturally deposited and metamorphosed snow the pore shape statistically balances for a specific porosity. The four porous properties that are based on the frame geometry are matrix bulk modulus, shear modulus, permeability, and tortuosity. This section illustrates the choice of the empirical relationships that are used in the following section to predict phase velocities and attenuation of acoustic waves in snow. 2.1. Frame bulk modulus (K S ) The frame bulk modulus corresponds to the solid material of the frame. It is also called the undrained modulus, because in a unjacketed compressibility test the frame material is squeezed between the pore fluid which presses against the frame from outside the sample and from inside the pore space at the same time. The resulting dilatation of the sample originates from the compressibility of the frame material (Biot and Willis, 1957). In the case of snow, the porous frame material is ice, which has almost perfect elastic behavior that is not temperature dependent for single crystals in the temperature range between -3 to -4 C (Gold, 1958). For randomly oriented polycristals, typical values for Young s modulus are 9. to 9.5 GPa and.33 ±.3 for Poisson s ratio (Hobbs, 1974; Mellor, 1983; Schulson, 1999). Young s modulus is defined as the ratio between stress and strain in one specific direction in the amplitude range of stress where Hooke s law holds. In our formulation we use bulk modulus as a measure of the relative decrease of volume to uniform compression rather than Young s modulus. For isotropic homogeneous materials Young s modulus E with a known Poisson s ratio ν can be converted to bulk modulus K as (Mavko and others, 9) K = E 3(1 2ν). (1) Ignoring the weak temperature dependence of polycrystalline ice we assume a constant bulk modulus of K S = 1 GPa for our porous snow model. 2.2. Matrix bulk modulus (K m ) The matrix bulk modulus describes the rigidity of the porous frame structure. It is also called the drained modulus as it corresponds to the result of a test with a jacketed sample that has drainage for the pore fluid. When stress is applied to the sample, the water is able to escape and the observed dilatation results from the compressibility of the frame only. The matrix bulk modulus in snow corresponds roughly to the bulk modulus in an elastic or visco-elastic model of wave propagation and therefore has been intensely investigated and ample measurements are available for a large range of porosities. It is important to note that the effective elastic modulus is strongly dependent on strain rate and that measurements with larger strain lead to lower values than measurements based on smaller strain rates (Kry, 1975; Mellor, 1975; Schneebeli, 4). A similar behavior has been observed for polycrystalline ice (Gold, 1958; Sinha, 1978). Tests with lower and higher frequencies are often referred to as static and dynamic, respectively. For the former the strain rate is usually high and the stress and strain is directly applied and measured on the snow sample. The dynamic tests are mostly based on indirect measurements such as the measurements of acoustic wave velocities. A compilation of results for static measurements of Young s moduli and Poisson s ratio as a function of density was presented by Mellor (1975) and later confirmed by Shapiro and others (1997) who extended the results with a study with quasi-static measurements by Kuvaeva and others (1967). These values are shown in Figure 1 by dashed lines. The results presented by Mellor (1975) are based on static tests, that result in lower values than for dynamic conditions. Mellor (1975) assumes that the dynamic values for dense snow are higher by a factor of two and by a factor of five for snow with lower density. Equation (1) in combination with the Poisson s ratio from Equation (5) is used to convert Young s modulus into bulk modulus which is used in our current formulation. To asses the performance of a poroelastic wave propagation model Johnson (1982) used Young s moduli from dynamic tests (Smith, 1965; Yamada and others, 1974) which are indicated with black circles in Figure 1. For the purpose of this study we need a continuous function for the presented data that describes the compressibility of snow as a function of porosity. In lack of any function of this type in the literature we use the Krief equation (Garat and

Sidler: Acoustic waves in snow 3 1 4 8 9.5 8 7 5 3 Young s modulus (MPa) 1 3 1 2 1 1 1 Poisson's ratio.4.3.2.1 1-1..2.4.6.8 1. Fig. 1. Krief equation fitted to dynamic (solid line) and static (dotted line) measurements. Dynamic measurements are indicated with signs while static measurements are indicated with dashed lines. Dashed areas and lines are from Shapiro and others (1997). Dynamic values from Johnson (1982) and Smith (1969) are indicated with circles and stars, respectively and theoretical values obtained from numerical modeling of microtomography snow structures are indicated with diamonds and triangles for the publications by Schneebeli (4) and Reuter and others (213), respectively. others, 199; Mavko and others, 1998) 4 K m = K S (1 φ) (1 φ), (2) that is often used in rock physics for this purpose and introduce the parameters a and b as a K m = K S (1 φ) (b φ). (3) Using a least square inversion to fit Equation (3) to the Young s moduli from the literature, we get a = 3.85 and b = 7.76 for the dynamic values from Johnson (1982) and a = 7.6 and b = 2.1 for the statically evaluated Young s moduli presented by Shapiro and others (1997). The resulting curves are shown in Figure 1 and meet the overall trend, even though the form of the equation was originally intended for sediments. This might be due to comparable behavior of structural stability with decreasing solid fraction. The data records for dynamically measured Young s moduli in snow are relatively sparse for low density snow, where measurements are more difficult. The light snow is very fragile and its structure can be easily destroyed during the measurements. Schneebeli (4) therefore proposed a micro tomographic analysis of the snow structure with consequent numerical modeling of the rigidity of an alike ice frame. Theses results coincide with the dynamic measurements presented in Johnson (1982) and the method presents an efficient and accurate solution for dynamic measurements of light snow. We therefore added additional measurements of this type from Reuter and others (213) to Figure 1. These measurements fit well to the adjusted Krief equation particularly for light snow, where no other data is available....1.2.3.4.5.6.7.8 Fig. 2. Estimated variation of Poisson s ratio with porosity. Regions in dashed and dotted lines refer to Bader (1952) and Roch (1948), respectively, as shown in (Shapiro and others, 1997), circles correspond to values presented by Smith (1969). The solid line corresponds to a rough linear fit for a porosity dependent Poisson s ratio. 2.3. Shear modulus (µ S ) For homogeneous isotropic linear elastic materials with a known bulk modulus, the shear modulus can be obtained using the relationship for the bulk modulus and Poisson s ratio ν as (Mavko and others, 9) µ S = 3 2 K m(1 2ν). (4) 1 + ν If a material is compressed in one direction, the Poisson effect describes the materials tendency to expand in the two other directions. This effect is quantified by Poisson s ratio which is the negative ratio of transverse to axial strain. An estimate of snow shear modulus as a function of porosity depends, beside the porosity dependence of the matrix bulk modulus shown in Section 2.2, strongly on the variation of Poisson s ratio with porosity. Fresh snow with high porosity collapses back upon itself and does not expand much in the directions normal to the compression. Such a behavior results in a Poisson s ratio close to zero for very porous snow. With decreasing porosity the Poisson s ratio of snow increases toward a Poisson s ratio more typical for incompressible materials. Not all studies of Poisson s ration in snow are equally well suited for the application to wave propagation as snow has a strong viscous behavior under large strains that is not relevant under the small strains involved by propagating waves (Mellor, 1975). The former is generally referred to as static testing, while dynamic testing refers to the conditions relevant for wave propagation. Mellor (1975) presents a summary mainly based on findings from Roch (1948) and Bader (1952), that shows a general increase along with increasing density. Also Smith (1969) s results fit well into this trend when the one value for extremely dense snow is neglected. The Poisson s ratio of this three studies are shown with circles, dashed and pointed lines, respectively in Figure 2.

4 Sidler: Acoustic waves in snow Shear modulus (Pa) 1 1 1 9 1 8 1 7 1 6 1 5 1 4 dynamic static 5 3.3.4.5.6.7.8 Fig. 3. Shear moduli predicted using Equation (5) for Poisson s ratio and Equation (3) for the bulk modulus. The solid and dotted line correspond to a dynamic and a static fit of Equation (3) and diamonds and stars to shear moduli presented by Johnson (1982) and (Smith, 1969), respectively. For simplicity and in view of the given uncertainties we define a linear relationship ν =.38.36 φ, (5) to characterize the general trend of the Poisson s ratio as a function of porosity. This function is shown as a solid line in Figure 2. In Figure 3 shear moduli resulting from Equations (3), (4) and (5) are compared to shear moduli from the dynamic measurements of Johnson (1982) and Smith (1969) which are indicated with diamonds and stars, respectively. The dashed line represents the prediction of shear moduli that would result from the static fit of the matrix bulk modulus (Eq. 3). To obtain predictions of shear moduli that better fit the measurements with the static fit of the matrix bulk modulus, the Poisson s ratios would have to be significantly lower which does not accord well with the values presented in Figure 2. More data is available for shear wave velocities than for dynamic shear bulk moduli. In Figure 4 the shear wave velocities resulting from shear moduli based on Equations (3), (4) and (5) are therefore compared to measured shear wave velocities from Yamada and others (1974) and Johnson (1982). For comparison also the predicted shear velocities based on the static fit of the matrix bulk modulus, and from assuming a constant Poisson s ratio of ν =.3 are shown with dotted and dashed lines, respectively. 2.4. Density (ρ S ) and (φ S ) In dry snow the density of the air filling the pore space is orders of magnitudes smaller than the density of the ice matrix and for our purposes can be neglected when converting porosity into density. The relationships and ρ S = (1 φ S ) 916.7 kg m 3, (6) φ = 1 ρ S 916.7 kg m 3, (7) 1 2 3 1 1 1 8 7 5 3 dynamic static ν =.3.2.3.4.5.6.7.8 Fig. 4. Predicted shear velocities at 1kHz based on different assumptions for Poisson s ratio and matrix bulk modulus. The solid and dotted lines correspond to a variable Poisson s ratio as shown in Figure 2 and matrix bulk moduli estimates based on dynamic or static measurements, respectively. The dashed line corresponds to a constant Poisson s ratio of ν =.3 and matrix bulk moduli estimates based on dynamic measurements. Squares and crosses correspond to shear wave velocity measurements by Johnson (1982) and Yamada and others (1974), respectively. can be used to derive the snow density ρ S from the porosity φ and vice versa based on the assumption that the density of ice is constant for temperatures where snow exists (Hobbs, 1974). Densities and porosities for typical snow types are shown in Table 1 (Paterson, 1994). 2.5. Permeability (κ) Permeability is an important property for fluid flow in porous materials. The most common approach to express permeability as a function of porosity is arguably the Kozeny-Carman equation that is based on geometrical considerations, but is basically an empirical relationship (see Appendix C). A comprehensive study of permeability dependence on porosity has been presented by Shimizu (197) and it has been shown that the Kozeny-Carman equation can also be used to express permeability as a function of porosity in snow (Calonne and others, 212). However, due to the large variety of snow types Table 1. Snow densities and porosities of the most prominent snow types as listed in Paterson (1994) New snow 5 7.92.95 Damp new snow.78.89 Settled snow 3.67.78 Depth hoar 3.67.89 Wind packed snow 35.56.62 Firn 83.9.56 Very wet snow and firn 7 8.12.23 Glacier ice 83 917.9 ρ s φ

Sidler: Acoustic waves in snow 5 and corresponding differences in permeability it is desirable to have an additional variable that can express the dependency of permeability to the snow type. Xu and Yu (8) present an approach based on stochastics to express permeability as a function of porosity based on the measurements of the minimal and maximal characteristic pore sizes. To avoid the introduction of additional snow properties in form of the minimal and maximal characteristic pore sizes, which are not well known and correspondingly difficult to characterize, we stick to the Kozeny-Carman equation in the form of Equation (1) where the specific surface area is introduced to account for the snow type. The specific surface area is also an important property for the specification of snow albedo and conversely it is possible to estimate the specific surface area from snow albedo measurements (Gallet and others, 9). To retain the dependency of our model to porosity only, Equation (11) is used to relate the specific surface area to porosity. The Kozeny-Carman relation is given by κ = C r2 φ 3 (1 φ) 2, (8) where κ is the permeability, φ the porosity, and the empirical constant C for snow characterizes the geometry of the channels in the model (Mavko and others, 1998). For sediments the constant is C s =.3 (Mavko and Nur, 1997; Carcione and Picotti, 6), but one order of magnitude larger for snow C B1 =.22 (Calonne and others, 212; Bear, 1972). The grain diameter r is the equivalent sphere radius of ice grains in snow and can be computed from the specific surface area (SSA) as r V A = 3 SSA, (9) where SSA is defined as surface area per volume (m 1 ). Substituting Equation 9 into Equation 8 leads to κ = C SSA φ 3 (SSA) 2 (1 φ) 2, (1) where C SSA = 9 C B1 =.2. Due to the compaction and metamorphosis processes inherent to snow it is reasonable to assume that SSA by itself can be seen as a function of porosity or density (e.g., Legagneux and others, 2; Herbert and others, 5). Domine and others (7) use SSA = 38.2 log(ρ S ) 26., (11) to relate specific surface area to snow density ρ S. The study is based on 345 specific surface area measurements in Alpine, maritime, tundra and taiga type environments and suggests an additional adjustment by classification to fresh, recent, and aged snow as well as classification to grain shapes. The proposed relationship yields negative values for the specific surface area for porosities smaller than φ =.44 which corresponds to a densities larger than ρ s = 512 kg/m 3. Here, the relationship is used only for porosities larger than φ =.65 which corresponds to a densities smaller than ρ s = 315 kg/m 3 and a constant value of 15 m 2 /kg for specific surface area is used to estimate permeability for lower porosities. Tortuosity 2.8 2.6 2.4 2.2 2. 1.8 1.6 1.4 1.2 5 3 r =.5 r =.67 1..3.4.5.6.7.8.9 Fig. 5. Tortuosity as a function of porosity according to Berrymann s (198) equation for shape factors r =.5 (solid line) and r =.67 (dashed line). 2.6. Tortuosity (T ) Based on geometrical considerations the tortuosity T of the pore space can be estimated as ( ) T = 1 r 1 1 φ, (12) where r is the so called shape factor (Berryman, 198). For a packing of sphere the shape factor is.5 and Equation (12) reduces to ( ) T = 1 2 1 + 1 φ. (13) The shape factor has a prominent role when a rigid frame model is used to predict wave propagation in a porous material. The rigid frame model assumes independent wave propagation in the pore space and in the rigid frame (e.g., Zwikker and Kosten, 1947). This assumption simplifies Biot s equations and is assumed to be valid if the bulk modulus of the matrix is significantly larger than the bulk modulus of the pore fluid. The slow wave velocity V, for example, is given in the rigid frame model as (Carcione, 7) Kf V = ρ f T, (14) where K f and ρ f are the bulk modulus and density of the pore fluid, respectively. By assuming constant compressibility and density for the pore fluid the slow wave velocity is a function of tortuosity only. When the snow water equivalent is estimated from the slow wave based on a rigid frame model, Equation (12) is used to connect the estimated tortuosity to the snow porosity and the shape factor r is a calibration constant for the specific snow type. Kinar and Pomeroy (9) use a shape factor between.5 to.67 for this purpose. Figure 5 shows tortuosity as a function of porosity based on Equation (12) for the shape factors r =.5 and r =.67. 2.7. Pore fluid (K f, ρ f, η f ) The viscosity of the pore fluid is the actual sink of energy in a Biot-type porous material if the bulk and shear moduli are real valued and do not contain additional attenuation

6 Sidler: Acoustic waves in snow 4 Table 2. Pore fluid properties of air (Lide, 5). density, ρ a 1.3 kg/m 3 viscosity, η a 1.7 1 5 Pa s bulk modulus, K a 1.42 1 5 Pa mechanisms. The viscous dissipation is caused by the relative fluid motion toward the frame of the porous material. This fluid motion originates in the pressure differences between the high and through of a wave that propagates through the material. The flow of the pore fluid leads to conversion of kinetic energy into heat and expresses itself as attenuation and dispersion of the propagating wave. The extent of the flow and subsequent attenuation is controlled by the geometry of the pore space and the characteristics of the pore fluid. In dry snow the pore space is filled with air which can be treated as an ideal gas at the conditions where snow exists and Sutherland s formula can be used to derive the viscosity of an ideal gas as a function of temperature ( ) 3/2 T η = η + C T, (15) T + C T where T is temperature in Kelvin and C = 12 K is the Sutherland s constant for air with the reference temperature T = 291.15 K and reference viscosity η = 1.827 1 5 Pa s (Smits and Dussauge, 6). Given that the viscosity of air of -4 Celsius and Celsius changes only from η = 1.8 1 5 kg/(m s) to η = 1.7 1 5 kg/(m s) we assume the viscosity to be a constant value for our purposes. The bulk modulus of air being the pore fluid can be computed as K g = γp, (16) where γ = 1.43 Pa is the adiabatic index and p is pressure. For the standard atmosphere atm = 113.25 hpa the bulk modulus K f is 1.42 1 5 Pa. We assume the air bulk modulus to be a constant value and do not use an adjustment on elevation as this bulk modulus is orders of magnitude smaller than the other bulk moduli in the system and moderate variations have only small effects on the result. The density of air as the pore fluid can be calculated using the ideal gas law p ρ = R, (17) gast where T is the temperature and R gas = 287.58 J/(kg K) is the specific gas constant for dry air. For a temperature of -2 Celsius in a standard atmosphere the density is 1.3 kg/m 3. Again, we take this as a constant value for our model. Table 2 shows the properties of the pore fluid used for the porosity snow model in the case where the pore space is exclusively filled with air. 3. PLANE WAVE SOLUTIONS In this section we use the reduction of the porous snow model presented above to predict wave velocities and wave attenuation for snow as a function of porosity. For this purpose we 25 15 5 5 3.3.4.5.6.7.8.9 Fig. 6. Predicted velocities for compressional waves of the first (solid line) and second kind (dashed line) as a function of porosity based on empirical relationships for frame bulk and shear modulus, tortuosity and permeability. The dashed lines identify measurements of first compressional waves compiled by (Sommerfeld, 1982) and the diamonds and squares represent wave velocity measurements compiled by Johnson (1982) for compressional waves of the first and second kind, respectively. evaluate a solution for Biot s (1956b) differential equations for the porosity range in question. A convenient way to get closed form solutions is to assume plane waves solutions and substitute those into the differential equations. The complex plane wave modulus is then obtained by solving the resulting dispersion relation. As in the poroelastic case the dispersion relation is a quadratic equation, there are two roots that correspond to the compressional waves of the first and second kind (e.g., Johnson, 1982; Pride, 5; Carcione, 7). The detailed calculations are summarized in Appendix A. The phase velocity V and attenuation in terms of the dimensionless quality factor Q can then be obtained from the complex plane wave modulus as (O connell and Budiansky, 1978) [ V (ω) = Re(V 1] 1 c(ω)), (18) Q p(ω) 1 = 2 Im(Vc(ω)) Re(V c(ω)). (19) Wave velocities for both kind of compressional waves as a function of porosity are shown in Figure 6 and are compared to measurements presented by Johnson (1982) and Sommerfeld (1982). The variations of the properties of the porous material, as shown in Section 2, especially the strong decrease of bulk modulus with increasing porosity, lead to the peculiarity that the compressional wave of the first kind becomes slower than the compressional wave of the second kind for very light and highly porous snow. It was described by Oura (1952) that in this kind of snow, waves propagate mainly in the air and the icy structure only interferes with the propagation (Carcione, 7). 3.1. Sensitivity of wave velocities The predicted velocities shown in Figure 6 fit the overall trend of the measured velocities from the literature relatively P 1 P 2

Sidler: Acoustic waves in snow 7 well. Nevertheless, the measurements still show some variation towards the predicted values. Such behavior is generally expected as the relationships where chosen to express an average snowpack and an individual snowpack can deflect from such conditions, for example due to unusual weather conditions. Here we asses if this variation can be explained with the presented model and if the snow properties can be kept in a reasonable range to do so. For this reason, the reduction model in Section 2 is evaluated with an additional variation of individual properties with the objective to describe and quantify the key properties that lead to the variation found in the measurements. Please note that the analysis presented here is particularly designed for snow and its conclusions do therefore not necessarily apply in general for a Biot-type porous material. The bulk modulus of the solid material, for example, is known to have a strong influence on the compressional wave velocity, but in the particular case of snow it is a good assumption to keep this property constant. The compressional wave velocity of the first kind which has a strong variation with porosity is sensitive almost exclusively to the frame bulk modulus and the shear modulus. In Figure 7 the predicted wave velocity is compared to wave velocity measurements from Smith (1969) and Johnson (1982). Even though it is well known that static tests for the bulk modulus lead to measurements that are not well suited to predict wave propagation in snow we would like to accentuate this point by showing that static values lead to a significant underestimation of acoustic velocities in the model presented here. The dashed line in Figure 7 corresponds to the predicted wave velocities that will result if the relationship for the frame-bulk-modulus was fitted to static values. The fit is shown in Figure 1a) with a dashed line too. It is evident that especially for snow with higher porosity the statically measured bulk moduli result in wave velocities that are too low. The variation of the compressional wave velocity with respect to the bulk and shear modulus is illustrated in Figure 7b). The spread of the measured velocities about the predicted velocities can neither be explained with a variation of 25 % in the bulk modulus nor with a variation of 25 % in the shear modulus. Given that the scales of bulk and shear moduli in Figures 1 and 3 are logarithmic it is not a large surprise that these relatively small changes are not sufficient to explain the variation. On the contrary it is notable that a combination of 25 % in both moduli actually suffices to explain the observed variation. The wave velocities of the compressional wave of the second kind as a function of porosity are shown in Figure 8 and compared to measurements from Oura (1952) and Johnson (1982). Sensitivity tests on the underlying model show that the compressional wave velocity of the second kind depends almost exclusively on variations in permeability and tortuosity. Frame bulk modulus and shear modulus have almost no influence on the velocity and variations are therefore not shown here. Tortuosity has a stronger lever on the wave velocity than permeability and 3 % variation in tortuosity shows larger velocity changes than 5 % change in permeability. The compressional wave of the second kind generally shows only little variation with porosity and is more sensitive to a) b) 25 15 5 dynamic static 5 3.3.4.5.6.7.8.9 25 15 5 5 3 25 % variation in K m 25 % variation in µ s 25 % variation in both.3.4.5.6.7.8.9 Fig. 7. Predicted phase velocities for the compressional wave of the first kind at 1 khz. a) predictions based on dynamic (solid line) and static (dashed line) measurements of matrix bulk modulus compared to measured wave velocities and b) sensitivity toward variations in shear and matrix bulk modulus. Diamonds represent wave velocity data from Johnson (1982). Crosses correspond to measurements from Smith (1969) the geometrical structure of the pore space than towards the properties of the porous frame. 3.2. Dispersion and attenuation The measurement of acoustic velocities, where only the information about the first arrival of an acoustic wave is exploited, is a straight forward and robust method to obtain information about a specific material, for example the snow density. Yet for other applications this information content is often not sufficient or ambiguous. One way of obtaining more information is to analyze the dispersion and attenuation of a propagating wave packet (e.g., Norris, 1989). Dispersion manifests itself in seismograms as a change in waveform and is usually quantified by the concept of phase velocity. Even though a wave with a single frequency does not exist, the analyzed waveforms often have a reasonable small bandwidth and can be compared to the phase velocity of its center frequency or Fourier analyzed to reveal the individual velocities of its contained frequencies. Attenuation leads to a decay in amplitude and is generally analyzed in the frequency domain.

8 Sidler: Acoustic waves in snow 35 3 25 15 35 3 25 15 5 3 % variation in tortuosity 5 % variation in permeability.55.6.65.7.75.8.85.9.95 Fig. 8. Predicted phase velocities for the compressional wave of the second kind at 5 Hz. The dashed and dotted lines correspond to phase velocities for 3 %variation in tortuosity and 5 %variation in permeability, respectively. Squares represent wave velocity measurements cited by Johnson (1982). Crosses correspond to measurements from Oura (1952). Please note that an increasing tortuosity decreases the velocity of the second compressional wave while an increase of permeability increases the velocity. Attenuation may also influence the waveform if some parts of the frequency spectrum are attenuated more than others. One of the main differences between elastic, viscoelastic and poroelastic wave propagation theory is that in the later two body waves are subject to dispersion and attenuation even in homogeneous materials. The associated attenuation and dispersion characteristics can be exploited to obtain additional information about the material which is especially appealing for poroelastic theory as it allows for conclusion on hydrological properties of the material (e.g., Pride, 5). In practical application it is often a particular challenge to obtain accurate measurements for attenuation and dispersion as geometrical spreading and material heterogeneity have similar effects on the seismograms and therefore have to be well known to be separated in an analysis. Here we use the reduction model presented in Section 2 and plane wave theory to solve Biot s (1957) differential equations for wave propagation in porous materials in order to characterize phase velocity and attenuation for snow. Using this approach the phase velocity and attenuation depends on porosity as well as on frequency. Figures 9, 1, 11 and 12 therefore represent perpendicular cuts through the corresponding surfaces. In Figure 9 the attenuation for the compressional wave of the first kind is shown as a function of porosity for three different frequencies. There are two porosity ranges that have orders of magnitude difference in attenuation levels. Light snow with a porosity φ.8 is considerably more attenuating for compressional waves of the first kind than denser snow. This separation corresponds roughly to a separation between freshly fallen and aged snow (Judson and Doesken, ) A similar difference between light and dense snow can be seen in the dispersion of the compressional wave of the first kind. The phase velocities for snow with the porosities φ =.9 5 a) b) Attenuation (1/Q) Attenuation (1/Q).5.4.3.2.1 35 Hz 1 khz 1 khz 3 25 15..55.6.65.7.75.8.85.9.95.6.5.4.3.2.1 35 3 25 Hz 1 khz 1 khz..55.6.65.7.75.8 Fig. 9. Predicted attenuation for the compressional wave of the first kind as a function of porosity. Figure b) shows a fragment of a) for porosities below φ =.8. Light snow is orders of magnitudes more attenuating than denser snow., φ =.85, φ =.8, φ =.7, and φ =.6 are shown in Figure 1. The corresponding densities are 138 kg/m 3, 183 kg/m 3, 275 kg/m 3, and 367 kg/m 3, respectively. It is evident that for dense snow the dispersion is almost negligible with only a few meters per second difference for the frequency range between 1 Hz and 1MHz. This result is in accordance with Johnson (1982) who found small changes in phase velocity for similar snow densities. For light snow, however, the compressional wave of the first kind shows significant dispersion especially in the range below 1 khz. The significant difference of absorption levels of sound between fresh and aged snow has been widely observed and found its way even into literature and lyrics (e.g., Aiken, 195). The phenomena is sometimes associated with the reflection coefficient of the snow that transmits a large amount of the energy of the propagating waves into the snowpack (Watson, 1948; Nicolas and others, 1985). The reflection coefficient is, however, only a part of the mechanism of sound absorption. Moreover, the reflection coefficient by itself does not explain why there is a significant difference in terms of absorption between fresh and aged snow. The model presented here shows that the gradual change in porous properties with increasing porosity can lead to a dramatic change in acoustic 5

Sidler: Acoustic waves in snow 9 1 8 φ =.9 φ =.85 φ =.8 φ =.7 φ =.6 a) Attenuation (1/Q).6.5.4.3.2 φ =.7 φ =.6 φ =.5 φ =.4.1 1 1 1 1 2 1 3 1 4 1 5 1 6 Frequency (Hz) Fig. 1. Phase velocities for the compressional wave of the first kind. The dispersion in dense snow is insignificant while light snow features a large frequency dependence. properties in the special case when the compressional wave of the first kind becomes actually slower than the compressional wave of the second kind which is generally referred to as the slow wave (see Figure 6). In this case the attenuation level for the compressional wave of the first kind becomes considerably higher and waves propagating in the snowpack rapidly decay. Homogeneous Biot-type porous materials are known to have a characteristic peak of attenuation (Geertsma and Smit, 1961; Carcione and Picotti, 6). In Figure 11 these attenuation peaks are shown for snow of different densities. The Figure is split into two subfigures to account for the significant difference of attenuation levels for light and dense snow that was already apparent in Figure 9. For dense snow the peak attenuation shifts towards higher frequencies and the attenuation level decreases with density. The same is true for light snow but with considerably higher attenuation levels and and shift of the attenuation peak to higher frequencies above the porosity level where the order of the compressional wave changes. However, the attenuation peaks shown above are valid only when the material is strictly homogeneous. It has been shown that the poroelastic relaxation mechanisms are considerably different for heterogeneous materials and a stochastic distribution of porosity values would rather lead to a constant Q characteristic for the attenuation (Pride and others, 4; Sidler and others, 213). The phase velocities and attenuation of the compressional wave of the second kind as a function of porosity is shown in Figure 12. In contrast to the compressional wave of the first kind there is no distinctive difference for dense and light snow. The sharp bend in phase velocity and attenuation is due to the relationship between porosity and the specific surface area that has a singularity at ρ s = 512 kg/m 3 and results in negative values for denser snow (Domine and others, 7). A constant value of SSA = 15 m 1 was therefore chosen for snow with φ <.62 which corresponds to a density of ρ s > 343 kg/m 3. b) Attenuation (1/Q). 1 1 1 1 2 1 3 1 4 1 5 1 6 Frequency (Hz).1.8.6.4.2 φ =.82 φ =.85 φ =.87 φ =.9. 1 1 1 1 2 1 3 1 4 1 5 1 6 Frequency (Hz) Fig. 11. Frequency dependent attenuation for for the compressional wave of the first kind in (a) medium to dense and (b) light snow. The peak of the attenuation shifts toward higher frequencies for denser snow. Please note that the amplitude of the attenuation is orders of magnitude larger for light snow with a porosity φ.8. 3.3. Specific surface area The specific surface area (SSA) is a simple measure for the exposed surface per volume of snow. It is important as the exposed ice surface controls chemical reactions and physical processes (Warren, 1982; Daly and Wania, 4; Gallet and others, 9). The structure of the pore space generally increases in complexity with increasing specific surface area and is related to fractal geometry. Direct measurements of the specific surface area are hampered by the problem of measuring the surface of fractal bodies that increases with the precision of the measurements (Mandelbrot, 1967; Arakawa and others, 9; Zermatten and others, 214). Therefore indirect measurements are often used to quantify the specific surface area (Matzl and Schneebeli, 6; Legagneux and others, 2). An other indirect way to quantify specific surface area is to measure permeability and porosity of snow and relate the two to specific surface area (Calonne and others, 212). Even though a decrease of permeability with increasing specific surface area is not mandatory for any random composition of the pore space, it can be shown that this is the case for cylindrical pores or a packing of spheres (Appendix C). The success of the Kozeny-Carman equation shows that such a

1 Sidler: Acoustic waves in snow a) b) Attenuation (1/Q) 35 3 25 15 35 3 25 15 Hz 5 1 khz 1 khz.55.6.65.7.75.8.85.9.95 2.5 2. 1.5 1..5 35 3 25 15..55.6.65.7.75.8.85.9.95 5 Hz 1 khz 1 khz Fig. 12. a) Phase velocity and b) attenuation for the compressional wave of the second kind for three different frequencies. The signs denote velocity measurements by Oura (1952) and Johnson (1982). relationship can also be seen in naturally deposited materials (Bear, 1972; Mavko and others, 9). As acoustic waves in porous materials are sensitive to the permeability, they can potentially be used to obtain information on specific surface area in snow. Figure 13 shows the variation of the compressional wave velocity of the second kind due to a change in specific surface area. Fixed values for the specific surface area are plotted with dashed and dotted lines for SSA = 15 m 2 /kg and SSA = 9 m 2 /kg, respectively. The solid line represents the resulting velocities when Equation (11) is used to obtain specific surface area as a function of porosity. The range of the variation is significant and larger for denser snow than for light snow where the pore space is large enough that the permeability is only little affected by tiny structures that can highly increase the surface area of the pore space. Moreover, the attenuation levels of both compressional wave also significantly increase with increasing specific surface area. In Figure 14 attenuation for both compressional waves is shown for a constant specific surface area of SSA = 15 m 2 /kg, and SSA = 9 m 2 /kg, as well as for specific surface area being a function of porosity. When comparing Figure 14 to Figure 1 and Figure 12b it becomes obvious that the effects of the 5 35 3 25 35 3 25 15 15 SSA (φ) SSA = 15 m 2 /kg 5 SSA = 9 m 2 /kg.55.6.65.7.75.8.85.9.95 Fig. 13. Predicted phase velocities for the compressional wave of the second kind at 5 Hz for a specific surface area as a function of porosity (solid line) and constant specific surface are values of SSA = 15 m 2 /kg (dashed line) and SSA = 9 m 2 /kg (dotted line). Squares and crosses correspond to measurements of Johnson (1982) and Oura (1952), respectively. specific surface has a similar effect on attenuation as a change of frequency. This is an effect of scale. A higher specific surface area goes along with lower permeability and therefore has a similar effect as if the wavelength would be increased (Equation A1). 4. CONCLUSIONS A model to predict phase velocities and attenuation of acoustic waves as a function of snow porosity is presented. The model is based on Biot s (1956a) model of wave propagation in porous materials and uses empirical relationships to assess tortuosity, permeability, bulk, and shear moduli as a function of porosity. The properties of the ice frame of the snow and air as the pore fluid are assumed to be constant. The model is not restricted to porosity as a single degree of freedom and additional information on specific surface area or any other of the properties characterizing a Biot-type porous medium can be readily incorporated. For light snow with a porosity φ.8 the particularity is found that the velocities of the compressional wave of the first kind is slower than the phase velocities of the compressional wave of the second kind which is commonly referred to as the slow wave. Such a reversal of the velocities of the compressional waves has, to our knowledge, not been reported before and is due to the week structure of the porous frame in fresh and light snow. The wave velocity reversal is a relatively sharp boundary for the attenuation level of the compressional wave of the first kind and is orders of magnitudes larger for highly porous snow. This finding is in accordance with the well known observation that freshly fallen snow absorbs most of the ambient noise, while after a relatively short time this absorbing behavior vanishes. An analysis of the sensitivity of the compressional waves reveals that the compressional wave of the first-kind is sensitive mainly to matrix and shear bulk modulus. A variation of 25 5

Sidler: Acoustic waves in snow 11 a) b) Attenuation (1/Q) Attenuation (1/Q).35.3.25.2.15.1.5 5 SSA (φ) SSA = 15 m 2 /kg SSA = 9 m 2 /kg 3..3.4.5.6.7.8.9 2. 1.5 1..5. 5 SSA (φ) SSA = 15 m 2 /kg SSA = 9 m 2 /kg 3.3.4.5.6.7.8.9 Fig. 14. Predicted attenuation at 5Hz for the compressional wave of a) the first and b) second kind as a function of porosity. The dashed and dotted lines correspond to permeability estimates based on the fixed values of SSA = 15 m 2 /kg and SSA = 9 m 2 /kg for the specific surface area, respectively. The solid line corresponds to an empirical relationship presented by Domine and others (7) and a constant value of SSA = 15 m 2 /kg for densities above 315 kg/m 3. % in both, shear and matrix bulk modulus can characterize the variability in the measured velocities. The compressional wave of the second kind shows continuous characteristics over the porosity range with considerably higher attenuation levels than the compressional wave of the first kind. The attenuation decreases with increasing porosity and frequency dependence is considerably more distinct than for the compressional wave of the first kind. The velocity depends strongly on tortuosity, permeability, and the related specific surface area. The variation of measured wave velocities for the compressional wave of the second kind can be obtained by altering the tortuosity by 3 % or by altering the permeability by 5 %. Comparing the predictions of our porous snow model to velocity measurements from the literature it becomes clear that statically measured matrix bulk moduli are generally too low to predict the actually measured velocities. Dynamic measurements or indirect measurements based on micro tomographic imaging of the ice frame and numerical modeling of the compressibility of an alike ice structure provide more coherent results. This model is a viable prerequisite for numerical modeling of acoustic wave propagation in snow, which allows, for example, to assess the design of acoustic experiments to probe for snow properties or to assess the role of acoustic wave propagation in artificial or skier triggered snow avalanche releases. 5. ACKNOWLEDGEMENTS This research was founded by a fellowship of the Swiss National Science Foundation. APPENDIX A. PLANE WAVE VELOCITIES IN A POROUS MATERIAL To derive the plane wave velocities and attenuation of compressional waves for Biot-type materials, we closely follow Carcione (7, pp. 273). Biot s (1956a; 1956b) equations of motion are given by j σ (S) ij = ρ 11 2 ttu (S) i φ j p f = ρ 12 2 ttu (S) i + ρ 12 2 ttu (f) i + ρ 22 2 ttu (f) i + b(u (S) i b(u (S) i u (f) i ), (A1) u (f) i ), (A2) where σ is the stress tensor, ρ ij are mass coefficients which take into account that the relative fluid flow through the pores is not uniform, b = φ 2 η/κ is the friction coefficient with η being the viscosity, φ the porosity, and κ the permeability, u is the displacement vector with the superscripts (S) and (f) referring to the solid matrix and the fluid, respectively. The differential operator is used with the Einstein summation convention for notational brevity. Adding Equations (A1) and (A2) using the concept of total stress σ ij = σ (S) ij + σ (f) δ ij, (A3) where δ is the Kronecker matrix, and the stress component of the fluid is leads to where σ (f) = φp f, (A4) j σ ij = ρ 2 ttu (S) i + ρ f 2 ttw i, (A5) j p f = ρ f 2 ttu (S) i + Y t w i, (A6) w = φ(u (f) u (S) ), (A7) is the relative displacement between the solid frame and the pore fluid and the relations ρ 11 + ρ 12 = (1 φ)ρ s, ρ 22 + ρ 12 = φρ f, (A8) (A9) are used. Statistical isotropy of the kinetic energy is used to relate the density coefficients to macroscopic properties of the porous material (Biot, 1956a). The low-frequency viscodynamic operator Y therefore is given by Y (t) = m t δ(t) + η κ δ(t), (A1)