Modeling of the simple shear deformation of sand: effects of principal stress rotation

Similar documents
Constitutive modelling of fabric anisotropy in sand

Prediction of torsion shear tests based on results from triaxial compression tests

the tests under simple shear condition (TSS), where the radial and circumferential strain increments were kept to be zero ( r = =0). In order to obtai

Drained Against Undrained Behaviour of Sand

Micromechanics-based model for cement-treated clays

A critical state model for sands dependent on stress and density

Soil strength. the strength depends on the applied stress. water pressures are required

On seismic landslide hazard assessment: Reply. Citation Geotechnique, 2008, v. 58 n. 10, p

WEEK 10 Soil Behaviour at Medium Strains

The Role of Slope Geometry on Flowslide Occurrence

Prof. B V S Viswanadham, Department of Civil Engineering, IIT Bombay

EARTHQUAKE-INDUCED SETTLEMENT AS A RESULT OF DENSIFICATION, MEASURED IN LABORATORY TESTS

Effects of principal stress rotation on the wave seabed interactions

Influence of particle shape on small-strain damping ratio of dry sands

Stress and Strains in Soil and Rock. Hsin-yu Shan Dept. of Civil Engineering National Chiao Tung University

7. STRESS ANALYSIS AND STRESS PATHS

Evaluation of undrained response from drained triaxial shear tests: DEM simulations and Experiments

Hypoplastic Cam-clay model

THE INTERPRETATION OF LABORATORY SOIL TESTS

This is a repository copy of Experimental study of anisotropy and non-coaxiality of granular solids.

8.1. What is meant by the shear strength of soils? Solution 8.1 Shear strength of a soil is its internal resistance to shearing stresses.

INTERNATIONAL JOURNAL OF CIVIL AND STRUCTURAL ENGINEERING Volume 1, No 4, 2011

Effect of embedment depth and stress anisotropy on expansion and contraction of cylindrical cavities

Verification of the Hyperbolic Soil Model by Triaxial Test Simulations

Improvement of a hypoplastic model to predict clay behaviour under undrained conditions

3D and 2D Formulations of Incremental Stress-Strain Relations for Granular Soils

A discrete element analysis of elastic properties of granular materials

Numerical Assessment of the Influence of End Conditions on. Constitutive Behavior of Geomaterials

EFFECT OF SILT CONTENT ON THE UNDRAINED ANISOTROPIC BEHAVIOUR OF SAND IN CYCLIC LOADING

The Stress Variations of Granular Samples in Direct Shear Tests using Discrete Element Method

A simple elastoplastic model for soils and soft rocks

Ch 5 Strength and Stiffness of Sands

Cyclic Behavior of Sand and Cyclic Triaxial Tests. Hsin-yu Shan Dept. of Civil Engineering National Chiao Tung University

Small-strain constrained elastic modulus of clean quartz sand with various grain size distribution

LABORATORY MEASUREMENTS OF STIFFNESS OF SOFT CLAY USING BENDER ELEMENTS

1.8 Unconfined Compression Test

Research Article Micromechanical Formulation of the Yield Surface in the Plasticity of Granular Materials

SHEAR MODULUS AND DAMPING RATIO OF SANDS AT MEDIUM TO LARGE SHEAR STRAINS WITH CYCLIC SIMPLE SHEAR TESTS

SHEAR STRENGTH OF SOIL

EFFECT OF VARIOUS PARAMETERS ON DYNAMIC PROPERTIES OF BABOLSAR SAND BY CYCLIC SIMPLE SHEAR DEVICE

INTERPRETATION OF UNDRAINED SHEAR STRENGTH OF UNSATURATED SOILS IN TERMS OF STRESS STATE VARIABLES

A Study on Modelling the Plane Strain Behaviour of Sand and its Stability

1.5 STRESS-PATH METHOD OF SETTLEMENT CALCULATION 1.5 STRESS-PATH METHOD OF SETTLEMENT CALCULATION

Clay hypoplasticity model including stiffness anisotropy David Mašín

Cite this paper as follows:

Small-Strain Stiffness and Damping of Soils in a Direct Simple Shear Device

Module 4 Lecture 20 Pore water pressure and shear strength - 4 Topics

Ch 4a Stress, Strain and Shearing

Following are the results of four drained direct shear tests on an overconsolidated clay: Diameter of specimen 50 mm Height of specimen 25 mm

Laboratory Testing Total & Effective Stress Analysis

Soil Behaviour in Earthquake Geotechnics

3 DYNAMIC SOIL PROPERTIES

General Strength Criterion for Geomaterials

INFLUENCE OF NONASSOCIATIVITY ON THE BEARING CAPACITY

POSSIBILITY OF UNDRAINED FLOW IN SUCTION-DEVELOPED UNSATURATED SANDY SOILS IN TRIAXIAL TESTS

A kinematic hardening critical state model for anisotropic clays

Prof. B V S Viswanadham, Department of Civil Engineering, IIT Bombay

Particle flow simulation of sand under biaxial test

Constitutive Modelling of Granular Materials with Focus to Microstructure and its Effect on Strain-Localization

Appendix A Results of Triaxial and Consolidation Tests

ON THE FACE STABILITY OF TUNNELS IN WEAK ROCKS

Three-Dimensional Failure Criteria Based on the Hoek Brown Criterion

Finite Element Method in Geotechnical Engineering

Computers and Geotechnics

Microstructural effect on the cavity expansion in a soil cylinder

Monitoring of underground construction

FROM THEORY TO PRACTICE IN GEOTECHNICAL ENGINEERING: THE MISSING LINKS. Yannis F. Dafalias

Theory of Shear Strength

GEO E1050 Finite Element Method Mohr-Coulomb and other constitutive models. Wojciech Sołowski

13 th World Conference on Earthquake Engineering Vancouver, B.C., Canada August 1-6, 2004 Paper No. 3016

ISSUES IN MATHEMATICAL MODELING OF STATIC AND DYNAMIC LIQUEFACTION AS A NON-LOCAL INSTABILITY PROBLEM. Ronaldo I. Borja Stanford University ABSTRACT

New Criterion For The Liquefaction Resistance Under Strain-Controlled Multi-Directional Cyclic Shear

Stress and fabric in granular material

Reappraisal of vertical motion effects on soil liquefaction. Citation Geotechnique, 2004, v. 54 n. 10, p

True Triaxial Tests and Strength Characteristics Study on Silty Sand Liang MA and Ping HU

PLAXIS LIQUEFACTION MODEL UBC3D-PLM

Computers and Geotechnics

both an analytical approach and the pole method, determine: (a) the direction of the

Multiaxial Constitutive and Numerical Modeling in Geo-mechanics within Critical State Theory

3-D Numerical simulation of shake-table tests on piles subjected to lateral spreading

A micromechanical approach to describe internal erosion effects in soils

walls, it was attempted to reduce the friction, while the friction angle mobilized at the interface in the vertical direction was about degrees under

EARTHQUAKE-INDUCED SETTLEMENTS IN SATURATED SANDY SOILS

Mechanics of origin of flow liquefaction instability under proportional strain triaxial compression

Discrete Element Modeling of Soils as Granular Materials

Bifurcation Analysis in Geomechanics

Module 3. DYNAMIC SOIL PROPERTIES (Lectures 10 to 16)

Effect of uniformity coefficient on G/G max and damping ratio of uniform to well graded quartz sands

CYCLIC LIQUEFACTION POTENTIAL OF LACUS- TRINE CARBONATE SILT FROM JULIAN ALPS

A Double Hyperbolic Model

Influences of material dilatancy and pore water pressure on stability factor of shallow tunnels

Changes in soil deformation and shear strength by internal erosion

General method for simulating laboratory tests with constitutive models for geomechanics

PLASTICITY FOR CRUSHABLE GRANULAR MATERIALS VIA DEM

D1. A normally consolidated clay has the following void ratio e versus effective stress σ relationship obtained in an oedometer test.

STUDY OF THE BARCELONA BASIC MODEL. INFLUENCE OF SUCTION ON SHEAR STRENGTH

Triaxial Shear Test. o The most reliable method now available for determination of shear strength parameters.

Chapter 5 Shear Strength of Soil

MODELING GEOMATERIALS ACROSS SCALES

Transcription:

Acta Geotechnica (9) 4:193 1 DOI 1.17/s1144-9-94-3 RESEARCH PAPER Modeling of the simple shear deformation of sand: effects of principal stress rotation Marte Gutierrez Æ J. Wang Æ M. Yoshimine Received: 1 November 8 / Accepted: 7 May 9 / Published online: 23 July 9 Ó US Government 9 Abstract The paper presents a simple constitutive model for the behavior of sands during monotonic simple shear loading. The model is developed specifically to account for the effects of principal stress rotation on the simple shear response of sands. The main feature of the model is the incorporation of two important effects of principal stress on stress strain response: anisotropy and non-coaxiality. In particular, an anisotropic failure criterion, cross-anisotropic elasticity, and a plastic flow rule and a stress dilatancy relationship that incorporate the effects of non-coaxiality are adopted in the model. Simulations of published experimental results from direct simple shear and hollow cylindrical torsional simple shear tests on sands show the satisfactory performance of the model. It is envisioned that the model can be valuable in modeling in situ simple shear response of sands and in interpreting simple shear test results. Keywords Constitutive model Dilatancy Granular materials Non-coaxiality Simple shear M. Gutierrez (&) Division of Engineering, Colorado School of Mines, Golden, CO 841, USA e-mail: mgutierr@mines.edu J. Wang Department of Building and Construction, City University of Hong Kong, Kowloon, Hong Kong e-mail: jefwang@cityu.edu.hk M. Yoshimine Department of Civil Engineering, Tokyo Metropolitan University, Hachioji, Tokyo 192-397, Japan e-mail: yoshimine-mitsutoshi@c.metro-u.ac.jp 1 Introduction One of the most common types of in situ soil deformation is that of simple shear. Simple shear condition is predominant during earthquake shaking of level grounds when deformation propagates vertically upwards from the bedrock in the form of shear waves. Soils within localized failure zones also deform in simple shear. Simple shear deformation can be simulated in element tests using the direct simple shear (DSS) device, or the hollow cylindrical torsional (HCT) simple shear device. The main drawback of simple shear devices is the non-uniformity of the stress and strain distributions within the soil specimen. Despite this shortcoming, simple shear testing continues to have wide use in practice due to the relevance of the test to in situ conditions and the ease of carrying out the test. Another drawback in the use of simple shear testing is the difficulty of interpreting the test results. A key feature of simple shear loading is that the principal stress directions are not fixed but rotate during shearing. As a result, the orientations of the failure planes are not a priori known and depend on the degree of principal stress rotation. Simple shear devices, where the sample is constrained from deforming laterally, have only two degrees of freedom corresponding to the shear and vertical deformations. Thus, the principal stress rotation cannot be directly controlled and only limited rotation can be achieved [22]. Experimental determination of the degree of principal stress rotation requires measurement of the variation of the lateral confining stress during shearing. However, this measurement is difficult to perform and is not usually carried out in routine DSS tests. Due to anisotropy, the response of sands during loading depends on the principal direction. Experimental data show that the shear strength and deformation of sands are

194 Acta Geotechnica (9) 4:193 1 dependent on the loading direction. Due to anisotropy, soils are expected to deform during principal stress rotation even when the shear stress level or the mobilized friction angle is kept constant [1, 8, 13]. Another important effect of principal stress rotation is the non-coincidence or non-coaxiality of the principal stress and principal plastic strain increment directions. Extensive experimental data from Hollow Cylindrical Testing have shown that the principal stress direction lags behind the principal plastic strain increment direction when the principal stresses are rotated [1, 7, 8, 13]. The deviation or non-coaxiality between the principal stress and principal plastic strain increment directions is significant for loading involving pure stress rotations at low mobilized friction angles. Non-coaxiality has important implications on the post-localization response of granular soils as shown by [8, 25]. Early attempts at modeling simple shear behavior of soils were made by [6, 14, 15, 18, ]. These models did not fully account for the effects of principal stress rotation. Recently, models that consider the effects of principal stress rotation on simple shear response have been proposed, e.g. [19, 26]. However, most of these models are based on complicated theories such as hypo-plasticity, making them difficult to use in routine geotechnical practice. The objective of this paper is to present a simple constitutive model for the behavior of sands during monotonic simple shear loading that fully accounts for the effects of principal stress rotation. The main feature of the model is the incorporation of the two important effects of principal stress on stress strain response: anisotropy and non-coaxiality. In particular, an anisotropic failure criterion, cross-anisotropic elasticity, and a plastic flow rule and a stress dilatancy relationship that incorporate the effects of principal stress rotation are adopted in the model. The model uses simple equations, and the implementation does not require complicated numerical integration. The model can be easily and conveniently implemented in a spreadsheet environment, allowing for their ease of use in practice. The model is validated against published experimental DSS and HCT data on sands. 2 Model for non-coaxial simple shear deformation of sand The analytical model for the non-coaxial simple shear deformation of sand is formulated in two-dimensional plane strain conditions where the stress and strain increment tensors are defined as r ij ¼ r x r r r y ; de ij ¼ de x de de de y ð1þ Note that de = dc /2, where dc is the engineering shear strain increment. The two-dimensional plane strain assumption implies that the out-of-plane strains are zero, i.e., de zz ¼ de zx ¼¼: All stresses are assumed to be effective, and will be denoted without the symbol prime ( ). In formulating the model, the stress invariants s and t (which correspond to the mean and shear stresses, respectively, in the plane of deformation) and strain rate invariants dv and dc (which correspond to volumetric and shear strain rates, respectively) will be used. These invariants are defined as follows: s ¼ 1 2 ðr x þ r y Þ; t ¼ 1 1=2 4 ðr x r y Þ 2 þ r 2 ð2þ dv ¼ de x þ de y ; dc ¼ððde x de y Þ 2 þ 4de 2 Þ1=2 ð3þ From the Mohr circle for stress (Fig. 1a), it can be seen that s is equal to the distance to the center of circle from the origin, and t is equal to the radius of the circle. Similarly, from the Mohr circle for strain (Fig. 1b), dv/2 is equal to the distance to the center of circle from the origin, and dc/2 is equal to the radius of the circle. τ Plane of maximum stress ratio φ dss φ a d γ b o σ3 Pole for direction ( σx, σ) 2α s t σ 1 ( σy, σ) σ 1 direction σ ( dεyy, dγ ) dε 3 Pole for direction 2β dγ/2 ψ dv/2 dε 1 (, d γ ) dε 1 direction dε Zero extension direction Fig. 1 Mohr s circles: a stress, b incremental strain for simple shear conditions

Acta Geotechnica (9) 4:193 1 195 2.1 Non-coaxial stress dilatancy relationship Coaxiality is a fundamental assumption in continuum mechanics. It implies that the directions of principal plastic strain increment and principal stresses coincide. Gutierrez and Ishihara [7] showed that coaxiality does not hold for loadings involving principal stress rotation, and at the same time, conventional plasticity models expressed in terms of the usual stress and plastic strain increment invariants implicitly assume coaxiality in the plastic flow rule. They showed that one way to integrate the effects of non-coaxiality in the plastic deformation of granular soils is by modifying the expressions for energy dissipation and stress dilatancy relations. They derived the following stress dilatancy relationship that is valid for loadings involving varying degrees of principal stress rotation by accounting for the degree of non-coaxiality in the calculation of dissipated energy: dv d dc ¼ sin / c c t ð4þ s where dv d is the volumetric strain increment due to dilatancy, and / c is the friction angle at phase transformation, and c is the non-coaxiality factor equal to: c ¼ cos 2D ð5þ The non-coaxiality angle D is equal to the deviation between the principal stress and the principal strain increment directions: D ¼ja bj ð6þ The angles a and b are shown in Fig. 1 and are defined from the Mohr s circle of stress and strain increment as tan 2a ¼ 2r ; tan 2b ¼ 2de ð7þ r y r x de y de x In terms of the dilation angle w and the mobilized friction angle /, Eq.4 can also be written as sin w ¼ sin / c c sin / ð8þ sin w ¼ dv d dc ; sin / ¼ t ð9þ s In addition to the above stress dilatancy relation derived from Critical State Soil Mechanics, another widely used stress dilatancy relation is the one derived by [23]. Gutierrez and Wang [9] recently extended Rowe s stress dilatancy relationship to account for non-coaxiality. In terms w and /, the non-coaxial Rowe s stress dilatancy equation is given as sin w ¼ sin / c sin / ð1þ cð1 sin / sin / c Þ Eq. 1 was found to provide a better fit of experimental simple shear data than Eq. 8, and thus, Eq. 1 will be used below in the development of the simple shear model. In simple shear, the lateral strain is equal to zero, de x =. As a result, the volumetric strain increment is equal to the vertical strain increment due to dilatancy, i.e., dv d = de yd, and the dilatancy ratio dv d /dc is equal to de yd sin w ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi de 2 yd þ dc2 ð11þ From the above equation, the simple shear dilatancy ratio de y /dc can be written as de yd sin w ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ¼ tan w: ð12þ dc 1 sin 2 w 2.2 Total vertical strain increment In addition to the vertical strain increment from dilatancy de yd, a change in the vertical stress will also induce a vertical strain de yc due to compression of the soil sample. Since the soil sample is constrained from deforming laterally, the compression induced vertical strain increment de yc is related to the vertical stress increment dr y by the constrained modulus M: de yc ¼ dr y ð13þ M The total vertical strain increment de y is now equal to de y ¼ de yc þ de yd ¼ dr y M þðtan wþ dc : 2.3 Shear stress strain relationship ð14þ A relationship that is widely used in the modeling of the non-linear stress strain response of soils is the hyperbolic stress strain model. Here, the hyperbolic relation will be used in terms of the shear stress ratio r /r y and the shear strain c : r r y ¼ G max tanð/ dss Þ p c tanð/ dss Þ p þ G max c ð15þ where (/ dss ) p is the peak direct simple shear friction angle corresponding to the peak value of the direct simple shear stress ratio, i.e., tan(/ dss ) p = (r /r y ) p and G max is the maximum initial shear modulus. The above equation models a shear stress ratio r /r y increasing asymptotically towards the peak shear stress (/ dss ) p with increasing shear strain c. From the Mohr s circle of stress (Fig. 1), the simple shear friction angle / dss = tan -1 (r /r y ) can be related to the plane strain friction / = sin -1 (r 1 - r 3 )/(r 1? r 3 ) and the principal stress direction a as

196 Acta Geotechnica (9) 4:193 1 sin / sin 2a tan / dss ¼ ð16þ 1 þ sin / cos 2a Note that Eq. 16 is valid for all levels of the mobilized shear friction angle /, including the peak condition, provided the corresponding a value is used. 2.4 Principal stress rotation during simple shear loading Oda and Konishi [18], Oda [16], and Ochiai [15] developed a simple expression for the amount of principal stress rotation that occurs during simple shear loading. They arrived at an expression which reads: tanð/ dss Þ¼ r ¼ j tan a ð17þ r y This equation implies a straight-line relationship between the simple shear stress ratio on a horizontal plane and the tangent of the angle that r 1 makes with the vertical axis. The constant j can be shown to be related to the ratio of the initial horizontal to the initial vertical stress, which in most in situ cases is equal to the K o value. The above equation is compared with the simple shear experimental data of [4] in Fig. 2. As can be seen, Eq. 17 provides a good approximation of the principal stress rotation during simple shear loading of sands. The experimental results show some differences in the degree of principal stress rotation between the loose, medium dense and dense sand samples. However, the effects of density appear to be very small and will be neglected in the Princiapal stress direction α ( ) 6 5 4 3 1 Dense Medium Loose σ /σ y =κtanα..2.4.6.8 1. Shear stress ratio, σ/σy Fig. 2 Principal stress rotation during simple shear loading of Leighton-Buzzard sand (experimental data from [4]) modeling. It is also worth noting from this figure that although a full 18 principal stress rotation cannot be achieved in simple shear loading, the magnitude of principal stress is still substantial, approaching as much as 6. 2.5 Non-coaxiality angle in simple shear loading The angle of non-coaxiality D is evaluated using the plastic flow rule developed by [1] for plastic flow in the (r y - r x )/2 versus r stress rotation plane. This flow rule is described in Appendix, and gives the following expression for the non-coaxiality angle:! D ¼ a n 1 sin / 2 sin 1 sinð2a 2nÞ ð18þ sin / p where / is the mobilized friction angle, / p is the peak friction angle, and n is the principal stress increment direction defined as tan 2n ¼ 2dr dr y dr x : 2.6 Principal stress increment direction ð19þ To obtain an expression for the rotation of the principal stress increment direction n, which is required in Eq. 18, it is only necessary to note that the stress increment vector should be tangential to the stress path. The stress path is given in Eq. 17 in terms of the stress ratio r /r y and the principal stress direction a can be differentiated to obtain this tangent. First, Eq. 17 is re-written in terms of 2a: r r y ¼ j pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi! 1 þ tan 2 2a 1 tan 2a Substituting Eq. 7 and solving for r x : ðþ r x ¼ r2 þð1 jþr y ð21þ jr y The differential form of the above equation can be written as dr x ¼ 2r dr r2 dr y þð1 jþdr y jr y jr 2 y ð22þ It can be noted that when r =, dr x = (1 - j)dr y, and as a result, it can be concluded that j = (1 - K o ) when there is no shear stress applied. The stress increment dr can be determined from dc and dr y using the incremental form of Eq. 15: dr ¼ G max tan/ ðdssþp c dr y þ G maxðtan/ ðdssþp Þ 2 tan/ ðdssþp þg max c ðtan/ ðdssþp þg max c Þ 2dc ð23þ

Acta Geotechnica (9) 4:193 1 197 In case the change in effective vertical stress dr y is known or prescribed, this value together with dr x from Eq. 22 and dr from Eq. 23 can be substituted in Eq. 19 to obtain the principal stress increment direction n. For drained direct simple shear test, a constant vertical stress, dr y =, is usually used, and in this case, dr x ¼ 2r dr ; dr y ¼ ð24þ jr y The principal stress direction n (Eq. 19) can then be written as tan 2n ¼ 2dr ; dr y ¼ ð25þ dr x Combining Eqs. 22 and 24 gives: tan 2n ¼ j r y 1 ¼ j ; dr y ¼ ð26þ r tan / dss For undrained loading condition, the change in the effective vertical stress dr y can be obtained assuming that the sample volume is constant during loading, and thus, de y =. Substituting this condition in Eq. 14 gives the vertical stress increment dr y in terms of the shear strain increment during undrained shearing: dr y ¼ Mðtan wþdc : ð27þ 2.7 Anisotropic friction angle and deformation moduli Extensive experimental data show that the friction angle of sand is strongly dependent on the direction of the applied loading. Studies by [2, 13, 17, 24] showed that the friction angle of sand is highest when the major principal stress is oriented normal to the bedding plane, i.e., when a =, and decreases as a is rotated from to 9. Typical variations of / as function of a from experiments are shown in Fig. 3. The effect of the major principal stress direction on the peak friction angle is represented in the model by the following relationship: 2k sin / p ð Þ sin / p ðaþ ¼ ð28þ ð1 þ kþ ð1 kþcosð2aþ where / p (a) is the direction-dependent friction angle, / p ( ) is the friction angle for a =, and k is the ratio of the sine of the friction angles for a = 9 and a =, i.e., k = sin / p (9 )/sin / p ( ). The validity of Eq. 28 can be seen in Fig. 3 where the equation is shown to adequately represent the trend in the anisotropy of the peak friction angle as manifested by the experimental data. The best fit of Eq. 28 through the experimental data gave a value of k =.84, while the lower and upper bound curves gave values of.8 and.94, respectively. Combining Eqs. 16 sinφp(α)/sinφp( ) and 28 gives the following variation in the peak simple shear friction angle (/ dss ) p with the major principal stress direction a: 2k sin / p ð Þ sin 2a tanð/ dss Þ p ¼ ð1 þ kþ 1 k 2k sin / p ð Þ cos 2a ð29þ In similar manner, experimental data show that the elastic deformation of sand is strongly dependent on the direction of loading. The small strain deformation of sands has been widely represented by cross-anisotropic elasticity where the deformation along the bedding plane is isotropic but different from the deformation perpendicular to the bedding plane [3, 5, 11, 12, 28]. For plane strain condition, cross-anisotropic elasticity requires four independent parameters, as shown in the following stress strain relation: 8 < : 1.1 1..9 Arthur & Menzies (1972) Oda (1972).8 Tatsuoka (1987) Gutierrez (1989) Model (k=.87) Model (k=.94) Model (k=.8).7 15 3 45 6 75 9 Major principal stress direction, α ( ) Fig. 3 Variation of sin / p with major principal stress direction a dr x dr y dr 9 2 = C 11 C 12 ; ¼ 4 C 12 C 22 38 < 5 : C 33 de x de y dc 9 = ; ð3þ where C 11 = M h = constrained modulus in the horizontal direction, C 33 = M v = constrained modulus in the vertical direction, and C 33 = G vh = shear modulus in the plane of anisotropy. Due to the zero lateral strain (de x = ) condition for simple shear loading, only two of these parameters are required, and these are the vertical constrained modulus M v = M, and the shear modulus G vh = G max. It can be shown that parameter C 12 can be related to the K o ratio by the relation: C 12 M ¼ dr x ¼ K o ð31þ dr y To calculate dr x, Eq. 31 is only used for consolidation, while Eq. 22 is used for shear loading.

198 Acta Geotechnica (9) 4:193 1 3 Comparisons with experimental data To illustrate its validity, the model described above is used to simulate the DSS test results obtained by [4] on Leighton-Buzzard sand, and the HCT results obtained by [27] on Kartal sand. Comparisons of model simulations and experimental results are shown in Figs. 4 and 5. The experimental data for Leighton-Buzzard sand are for drained tests, while the experimental data for Kartal sand are for undrained tests. Three sets of test results on Leighton-Buzzard sand with void ratios at the end of consolidation of e o =.739,.645 and.539 are used in the validation of the model. These void ratios correspond, respectively, to loose, medium dense and dense relative densities. All tests were done at constant r y value of 1 kpa. For Kartal sand, the test results are for vertical consolidation stresses of r yo = 4, 6 and 8 kpa. The void ratios at the end of consolidation e o for the three tests are very similar ranging from.87 to.878. Table 1 lists the model parameters used in the simulations of the experimental results. Three of the parameters, namely the peak friction angle / p, the initial shear modulus G max, and the bulk modulus M are dependent on the initial void ratio e o and the initial vertical stress r yo. For Leighton-Buzzard sand, the principal stress rotation parameter j appears to be independent of initial void ratio e o, while for Kartal sand j varies with the initial vertical stress r yo. The dependency of these parameters on e o and r yo are given by the expressions in Table 1. The dependency of G max on e o and r yo follows the empirical relation by [21]. The other parameters, namely the friction angle at phase transformation / c and anisotropy parameter k are relatively constant and independent of e o and r yo. As can be seen in Figs. 4 and 5, the simulations of the simple shear response are very close to the experimental data. There is a good agreement between the stress strain response, volumetric response (in drained tests), effective stress path (in undrained tests), and the directions of the Shear stress ratio, σ /σ y Volumetric strain, v.7.6.5.4.3.2.1 Model Experiment...2.4.6..15.1.5. -.5 -.1.8 1. loose (e o =.739) medium dense(e o =.645) dense (e o =.529) -.15..2.4.6 Shear stress ratio, σ/σy Volumetric strain, v.6.4.2...1.2.3..15.1.5. -.5 -.1..1.2.3 Shear stress ratio, σ/σy Volumetric strain, v.8.6.4.2...1.2.3.4.3..1. -.1..1.2.3 Direction α, β and ξ (deg.) 8 6 4 α ξ Model β α β Experiment ξ..2.4.6 Direction α, β and ξ (deg.) 8 6 4..1.2.3 Direction α, β and ξ (deg.) 8 6 4..1.2.3 Fig. 4 Comparison of experimental data and model predictions. Top row shear strain versus shear stress, middle row shear strain versus volumetric strain, and bottom row shear strain versus directions a, b and n (experimental data from [4])

Acta Geotechnica (9) 4:193 1 199 8 6 4 Experiment Model 2 4 6 8 1 8 6 4 3 σ yo =4 kpa (%) 4 6 8 Effective vertical stress, σ y (kpa) 8 6 4 σ yo =6 kpa 5 1 15 (%) 8 6 4 5 4 6 8 1 Effective vertical stress, σ y (kpa) 6 4 σ yo =8 kpa 5 1 (%) 6 4 5 15 4 6 8 1 Effective vertical stress, σ y (kpa) σ1-direction, α (deg) 25 15 1 5 2 4 6 8 1 (%) σ1-direction, α (deg.) 4 3 1 2 4 6 8 1 (%) σ1-direction, α (deg) 4 3 1 2 4 6 8 (%) 1 Fig. 5 Comparison of experimental data and model predictions. Top row shear strain versus shear stress, middle row effective stress path, and bottom row shear strain versus principal stress direction a (experimental data from [27]) Table 1 Model parameters used for simulation of simple shear behavior of Leighton-Buzzard sand and Kartal sand Parameter Leighton-Buzzard sand Kartal sand Peak friction angle / p for a = ( ) 74.6e o - 7.7 82.1(r yo /p a ) - 47.7 Friction angle at phase transformation / c ( ) 33.5 36.5 Shear modulus G max 9F(e o, r yo ) 11F(e o, r yo ) Constrained modulus M 92.7(r yo /p a ) - 8.8 Stress rotation parameter j.55 2.95-2.73(r yo /p a ) Anisotropy parameter k.85.85 pffiffiffiffiffiffiffiffiffiffiffiffiffi Fðe o ; r yo Þ¼p a r yo =p a ð2:97 e o Þ 2 =ð1 þ e o Þ, where e o = initial void ratio, r yo = initial vertical stress, p a = atmospheric pressure major principal stress a. For the Leighton-Buzzard, the predicted and measured principal strain increment direction b and the principal stress increment direction n are also in good agreement. 4 Conclusions A key feature of the response of soils during simple shear loading is the rotation of the principal stresses. Early experimental studies by [15, 18, 22] showed that the principal stress rotation is limited (typically to less than 6 ), and that it occurs mainly during small strain deformation. Consequently, the effects of principal stress rotation have been routinely neglected in the formulation of the simple shear response of soils. However, the recent emphasis on accurate representation of the pre-failure deformation of soils has made it essential to account for the effects of principal stress rotation. To meet this need, the paper presented a constitutive model that specifically

Acta Geotechnica (9) 4:193 1 accounts for the effects of principal stress rotation on the simple shear behavior of sands. The model incorporated two important effects of principal stress on stress strain response, namely anisotropy and non-coaxiality. The model used an anisotropic failure criterion, cross-anisotropic elasticity, and a plastic flow rule and a stress dilatancy relationship that incorporate the effects of noncoaxiality. The model was validated against published experimental results from direct simple shear and hollow cylindrical torsional simple shear tests on sands. Appendix: Non-coaxiality angle due to principal stress rotation The angle of non-coaxiality angle D required to determine the non-coaxiality parameter c in the stress dilatancy relationships given in Eqs. 9 and 1 is evaluated using the flow rule developed by [1] for plastic flow in the (r y - r x )/2 versus r stress rotation plane. This flow rule is shown in Fig. 6. In this figure, the strain increment components (de y - de x ) and 2de have been superimposed on the stress plane. Point A is the current stress point. Neglecting the effect of anisotropy, the failure surface is circular and centered at the origin of the (r y - r x )/2 versus r axes. The distances OA and OB are equal to the current shear stress t and the radius of the circular failure surface, respectively. These distances can be related to the current mean stress s, the mobilized friction angle / and the peak friction angle / p as follows: OA ¼ s sin /; OB ¼ s sin / p ð32þ On the (r y - r x )/2 versus r stress plane, a stress vector makes an angle equal to 2a from the (r y - r x )/2 axis (Eq. 7). Similarly, a stress increment vector makes an angle equal to 2n from the (r y - r x )/2 axis (Eq. 19). On the (de y - de x ) versus 2de strain increment plane, a strain σ 2d ε Strain increment direction increment vector has a length equal to the plastic shear strain increment dc and makes an angle equal to 2b from the (de y - de x ) axis. This plastic strain increment direction is evaluated as the normal to the failure surface at the conjugate point A which is the intersection of the failure surface and the stress increment vector extended from the current stress point. This flow rule is based on the experimental observations that plastic flow on the (r y - r x )/2 versus r stress plane is dependent on the stress increment direction [1]. This flow rule should be contrasted with conventional plasticity formulations where the plastic strain increment direction is evaluated at the current stress point independent of the stress increment direction. Details including the experimental verification of the flow rule are given in [1]. From triangle OAB in Fig. 6, the following angles can be obtained: \OAB ¼ p ð2n 2aÞ \BOA ¼ 2D 2D ¼ p \OAB \ABO Using the law of sines: OA sin \ABO ¼ OB sin \OAB \ABO ¼ sin 1 OA sin \OAB OB Substituting Eqs. 32 and 33 in Eq. 37:! \ABO ¼ sin 1 sin / sinðp þ 2a 2nÞ sin / p Substituting Eqs. 33 and 38 in Eq. 34 gives:! 2D ¼ 2a 2n sin 1 sin / sinðp þ 2a 2nÞ sin / p ð33þ ð34þ ð35þ ð36þ ð37þ ð38þ ð39þ Simplifying the above equation results in the expression for D given in Eq. 18. Conjugate stress point B Failure surface References O 2β 2 2α A Stress increment 2ξ Current stress point ( σy σx)/2 ( dεy dεx) Fig. 6 Non-coaxial flow rule in the (r y - r x )/2 versus r stress rotation plane [1] 1. Arthur JRF, Rodriguez JDC, Chua KS, Dunstan T (198) Principal stress rotation: a missing parameter. J Geotech Eng Div ASCE 16(4):419 433 2. Arthur JRF, Menzies B (1972) Inherent anisotropy in a sand. Geotechnique 22:115 128 3. Chaudhary SK, Kuwano J, Hayano Y (4) Measurement of quasi-elastic stiffness parameters of dense Toyoura sand in hollow cylindrical apparatus and triaxial apparatus with bender elements. Geotech Test J 27(1):1 13 4. Cole ERL (1967) Soils in the simple shear apparatus. Ph.D. thesis. Cambridge University

Acta Geotechnica (9) 4:193 1 1 5. Fioravante V () Anisotropy of small strain stiffness of Ticino and Kenya sands from seismic wave propagation measured in triaxial testing. Soils Found 4(4):129 142 6. Frydman S, Talesnick M (1991) Simple shear of isotropic elastoplastic soil. Int J Num Anal Meth Geomech 15(4):251 27 7. Gutierrez M, Ishihara K () Non-coaxiality and energy dissipation in granular materials. Soils Found 4(2):49 59 8. Gutierrez M, Vardoulakis I (6) Energy dissipation and postbifurcation behavior of granular soils. Int J Num Anal Meth Geomech 31(3):435 455 9. Gutierrez M, Wang J (8) Non-coaxial version of Rowe s stress dilatancy relation. Granural Matter 11:129 137 1. Gutierrez M, Ishihara K, Towhata I (1993) Flow theory for sand during rotation of principal stress direction. Soils Found 31(4):121 132 11. Jung Y-H, Chung CK, Finno RJ (4) Development of nonlinear cross-anisotropic model for the pre-failure deformation of geomaterials. Comp Geotech 31:89 12 12. Kuwano R, Jardine RJ (2) On the applicability of crossanisotropic elasticity to granular materials at very small strains. Geotechnique 52:727 749 13. Miura K, Miura S, Toki S (1986) Deformation behavior of anisotropic sand under principal axes rotation. Soils Found 26(1):36 52 14. Moroto N (1987) On deformation of granular material in simple shear. Soils Found 27(1):77 85 15. Ochiai H (1975) The behaviour of sands in direct shear tests. J Jpn Soc Soil Mech Found Eng 15(4):93 1 16. Oda M (1975) On the relation s/r n = j tan w in the simple shear test. Soils Found 15(4):35 41 17. Oda M (1981) Anisotropic strength of cohesionless sands. J Geotech Eng ASCE 17(GT9):1219 1 18. Oda M, Konishi J (1974) Rotation of principal stresses in granular material during simple shear. Soils Found 14(4):39 53 19. Osinov VA, Wu W (6) Simple shear in sand with an anisotropic hypoplastic model. Geomech Geoeng 1(1):43 5. Prevost J, Høeg K (1975) Reanalysis of simple shear soil testing. Can Geotech J 13:418 429 21. Richart FE, Hall JR, Woods RD (197) Vibrations of soils and foundations. Int Ser Theor Appl Mech. Prentice-Hall, Englewood Cliffs, NJ 22. Roscoe KH (197) The influence of strains in soil mechanics 1th Rankine Lecture. Geotechnique (2):129 17 23. Rowe PW (1962) The stress dilatancy relation for static equilibrium of an assembly of particles in contact. Proc R Soc Lond Ser A Math Phys Sci 269(1339):5 527 24. Tatsuoka F (1987) Discussion on the strength and dilatancy of sands. Geotechnique 37(2):219 226 25. Vardoulakis I, Georgopoulos IG (5) The stress dilatancy hypothesis revisited: shear-banding related instabilities. Soils Found 45(2):61 76 26. Yang Y, Yu HS (6) Numerical simulations of simple shear with non-coaxial soil models. Int J Num Anal Meth Geomech 3(1):1 19 27. Yoshimine M, Özay R, Sezen A, Ansal A (1999) Undrained plane strain shear tests on saturated sand using a hollow cylinder torsional shear apparatus. Soils Found 39(2):131 136 28. Zeng X, Ni B (1999) Stress-induced anisotropic G max of sands and its measurement. J Geotech Geoenviron Eng ASCE 125(9): 741 749