H O S T E D B Y A stress–strain description of saturated sand under undrained cyclic torsional increments in volumetric strain due to dilatancy, a normalized stress–plastic shear strain relationship is employed in combination with a novel empirical Earlier experimental attempts to study the liquefaction behavior of soils date back to the 1960s when Seed and Lee (1966) conducted a series of undrained cyclic triaxial tests on The Japanese Geotechnical Society www.s journal homepage: w Soils an Soils and Foundations 2015;55(3):559–574saturated sand and reported that the onset of liquefaction was primarily governed by the relative density of the sand, the confining pressure, the stress or strain amplitude and the http://dx.doi.org/10.1016/j.sandf.2015.04.008 0038-0806/& 2015 The Japanese Geotechnical Society. Production and hosting by Elsevier B.V. All rights reserved. E-mail addresses: nalinds@uom.lk (L.I.N. De Silva), koseki@iis.u-tokyo.ac.jp (J. Koseki), chiaroga@iis.u-tokyo.ac.jp (G. Chiaro), tsato@iis.u-tokyo.ac.jp (T. Sato). Peer review under responsibility of The Japanese Geotechnical Society.1. IntroductionnCorresponding author.stress–dilatancy relationship derived for torsional shear loading. The proposed stress–dilatancy relationship accounts for the effects of over-consolidation during cyclic loading. Numerical simulations show that the proposed model can satisfactorily simulate the generation of excess pore water pressure and the stress–strain relationship of saturated Toyoura sand specimens subjected to undrained cyclic torsional shear loading. It is found that the liquefaction resistance of loose Toyoura sand specimens can be accurately predicted by the model, while the liquefaction resistance of dense Toyoura sand specimens may be slightly underestimated. (i.e., the liquefaction potential is higher). Yet, the model predictions are conservative. & 2015 The Japanese Geotechnical Society. Production and hosting by Elsevier B.V. All rights reserved. Keywords: Liquefaction behavior; Stress–dilatancy relationship; Over-consolidation; Sand; Constitutive model JEL classification: IGC; A10; D07; E08; T06shear loading Laddu Indika Nalin De Silvaa,b,n, Junichi Kosekic, Gabriele Chiaroc, Takeshi Satod aDepartment of Civil Engineering, University of Moratuwa, Sri Lanka bDepartment of Civil Engineering, University of Tokyo, Japan cInstitute of Industrial Science, University of Tokyo, Japan dIntegrated Geotechnology Institute Ltd., Japan Received 21 January 2014; received in revised form 11 December 2014; accepted 2 February 2015 Available online 16 May 2015 Abstract A constitutive model to describe the cyclic undrained behavior of saturated sand is presented. The increments in volumetric strain during undrained loading, which are equal to zero, are assumed to consist of increments due to dilatancy and increments due to consolidation/swelling. This assumption enables the proposed model to evaluate increments in volumetric strain due to dilatancy as mirror images of increments in volumetric strain due to consolidation/swelling, thus simulating the generation of excess pore water pressure (i.e., reduction in mean effective principal stress) during undrained cyclic shear loading. Based on the results of drained tests, the increments in volumetric strain due to consolidation/swelling are evaluated by assuming that the quasi-elastic bulk modulus can be expressed as a unique function of the mean effective principal stress. On the other hand, in evaluating theciencedirect.com ww.elsevier.com/locate/sandf d Foundations d Fnumber of loading cycles. Since then, extensive studies have been conducted on soil liquefaction throughout the world (Vaid and Thomas, 1995, among others) and a number of attempts have been made to define proper constitutive models to describe it (Liou et al., 1977; Liyanapathirana and Poulos, 2002, among others). Based on the results of several series of experiments on saturated Nomenclature τzθ shear stress σ 0 z; σ 0 r and σ 0 θ axial, radial and circumferential effective stress, respectively p0 mean effective stress Drini relative density measured at confining pressure of 30 kPa (τzθ=p0)max maximum shear stress ratio τzθ max peak shear stress Gzθ0 initial quasi-elastic shear modulus (¼dτzθ=dγezθ) γzθ; γ e zθ; γ p zθ total, elastic and plastic shear strain, respec- tively (engineering strain) εpvol plastic volumetric strain γref reference shear strain (¼ (τzθ=p0)/(Gzθ0=p0)) m, n, k material parameters that accounts for the stress induced anisotropy of Young's moduli, shear moduli and Poisson's ratio, respectively CE, CG factors that account for the degradation of quasi- elastic Young's and shear moduli, respectively (assumed as zero in the present study) L.I.N. De Silva et al. / Soils an560hollow cylindrical sand specimens, Towhata and Ishihara (1985a) proposed a unique correlation between the shear work and the generation of pore water pressure (PWP). Furthermore, the effects of the rotation of the principal stress axes on sand liquefaction were investigated by Towhata and Ishihara (1985b) using hollow cylindrical specimens subjected to cyclic torsional shear loading. However, compared to the large amount of experimental data existing on liquefaction and the undrained behavior of soils, very few models are available to successfully simulate the soil performance under cyclic undrained loading. Ishihara et al. (1975) proposed a model based on five postulates to trace the generation of the excess PWP of sand subjected to undrained irregular cyclic loading. This model qualitatively simulates the stress–strain relationships and the shear stress versus mean effective stress relationships. A constitutive model to simulate the cyclic undrained behavior of sand, based on the multi-spring concept, was developed by Iai et al. (1992). In this model, commonly known as the “Towhata–Iai model”, shear deformation is modeled by employing the multi- spring concept, and the generation of excess PWP is modeled using a unique correlation between the increments in excess PWP and shear work, as proposed by Towhata and Ishihara (1985a). Nishimura (2002) and Nishimura and Towhata (2004) modified the above model by expanding the multi-spring concept from two dimensions to three dimensions, while using an empirical stress– dilatancy relationship to model the generation of excess PWP bycorrelating the stress–dilatancy relationship to consolidation. Never- theless, these models do not consider the inherent anisotropy of soils. Furthermore, the steady state during liquefaction and the continuous increase in shear strain with cyclic loading cannot be properly simulated. An elasto-plastic constitutive model for sand, based on the non-linear kinematic hardening rule, was employed to inves- A Ez0/Eθ0, i.e. ratio of vertical to circumferential quasi elastic Young's moduli at isotropic stress state Y normalized shear stress (¼ (τzθ=p0)/(τzθ=p0)max) X normalized shear strain (¼γpzθ=γref ) D1 and D2 drag parameters D plastic shear moduli immediately after reversal of stress/initial plastic shear moduli (i.e., damage parameter) Dult minimum value for D S amount of hardening Sult maximum value for S OC over-consolidation ratio dεpvol=dγpzθ dilatancy ratio Rk gradient of the empirical stress–dilatancy relationship Rm the maximum value for Rk C intercept of the empirical stress–dilatancy relationship Cmin minimum value for C oundations 55 (2015) 559–574tigate the effectiveness of the cement-mixing column method and the gravel drain method as countermeasures against liquefaction by means of a two-dimensional liquefaction analysis (Oka et al., 1992). Later, Oka et al. (1999) further modified this model by introducing a stress–dilatancy relationship that accounts for the damage to plastic stiffness at large levels of shear strain. In addition, several other constitutive models, based on the critical state framework, are proposed in the literature. Jefferies (1993) proposed a strain-hardening model, which utilizes the state parameter, to explain the behavior of very loose to very dense sand. A unified generalized plasticity model, based on the non- linear critical state line, was proposed by Ling and Yang (2006). It should be noted that all the above-described models are based on either the critical state soil mechanics approach (e.g., Oka et al., 1992) or the energetic approach (Iai et al., 1992; Nishimura, 2002). In the current study a different and original approach is attempted by extending empirical relationships that are found to be reasonably consistent with the experimental observations. The undrained cyclic behavior of sand is simulated based on the response whereby the same sand is shown during drained cyclic loading. In fact, after appropriate normalization, the stress–strain relationship is found to be unique for drained and undrained conditions. Moreover, the generation of PWP during undrained loading can be described based on the volumetric strain response of sand during drained loading. This is done by improving the model proposed by De Silva and Koseki (2012) that can accurately simulate the drained cyclic behavior of sand, i.e., the stress–strain relationship and the volumetric strain response. However, no attempt has been made so far to utilize the above- mentioned approach in simulating the cyclic undrained behavior (i. e., liquefaction) of soil. The attempt is made in this paper, where a cyclic constitutive model is presented to describe the undrained cyclic behavior of sand. In the model, a simulation of the plastic volumetric strain due to dilatancy (dεdvol) is combined with the consolidation/swelling behavior of sand to simulate the generation of excess PWP (i.e., a reduction in mean effective principal stress) and stress–shear strain relationships. 2. Test material and procedures In order to support the modeling work, a series of drained and undrained cyclic torsional shear loading tests were conducted on saturated Toyoura sand specimens (D50¼0.162 mm, emax¼0.966, emin¼0.600, coefficient of uniformity Uc¼1.46). Hollow cylind- rical specimens, having dimensions of 20 cm in outer diameter, 12 cm in inner diameter and 30 cm in height, were prepared at initial relative densities (Drini) of 21%, 56% and 75%, as measured at a confining pressure of 30 kPa. A modified air pluviation technique, in which the sand pluviation was completed in a radial Tokyo, was used for the testing program. A recently developed local deformation measurement technique was employed in the evaluation of quasi-elastic deformation properties and volumetric strain during isotropic loading for drained specimens in the current study. Refer to De Silva et al. (2005) for details on the torsional shear apparatus and local deformation measurement system used. In order to investigate the stress–dilatancy relationship of Toyoura sand, a series of drained cyclic torsional shear tests was conducted on loose and dense specimens, with Drini¼56% and 75%, respectively, while keeping the mean effective principal stress (p0) constant at 100 kPa. Details of the stress paths employed in the drained cyclic tests are presented in De Silva and Koseki (2012). In addition to the above, a series of constant stress amplitude undrained cyclic torsional shear tests was conducted on Toyoura sand specimens, while keeping the specimen height constant, for a comparison with the model predictions. The stress paths of the cyclic undrained tests are shown in Table 1. 3. Framework for modeling of liquefaction behavior Changes in p0 during undrained loading cause the consoli- dation/swelling of a specimen, while changes in shear stress (τ) (1) IC (σ΄z ¼ σ΄r ¼ σ΄θ ¼ 50-100 kPa) (2) UTS (τzθ ¼ 0-22-22-0 kPa, until liquefaction at p΄0¼100 kPa) SAT 31 79.4 L.I.N. De Silva et al. / Soils and Foundations 55 (2015) 559–574 561SAT 32 75.2 SAT 34 50.3 SAT 33 21.3 ΄direction while slowly moving the nozzle of the pluviator in alternate clockwise/anticlockwise directions, was employed in the current study to minimize the degree of anisotropy of the horizontal bedding plane of the hollow cylindrical specimens (see details in De Silva et al. (2006)). A high-capacity medium-sized hollow cylinder apparatus, developed at the Institute of Industrial Science, University of Table 1 Stress paths and test conditions of liquefaction tests. Test Dr (%) SAT 38 75.7 SAT 28 74.6Dr: relative density (%) measured at an isotropic stress state of σc¼30 kPa, IC: iso mean effective stress at start of cyclic undrained loading.(1) IC (σ΄z ¼ σ΄r ¼ σ΄θ ¼ 50-100 kPa) (2) UTS (τzθ ¼ 0-30-30-0 kPa, until liquefaction at p΄0¼100 kPa) (1) IC (σ΄z ¼ σ΄r ¼ σ΄θ ¼ 50-100 kPa) (2) UTS (τzθ ¼ 0-40-40-0 kPa, until liquefaction at p΄0¼100 kPa) (1) IC (σ΄z ¼ σ΄r ¼ σ΄θ ¼ 50-100 kPa) (2) UTS (τzθ ¼ 0-60-60-0 kPa, until liquefaction at p΄0¼100 kPa) (1) IC (σ΄z ¼ σ΄r ¼ σ΄θ ¼ 50-100 kPa) (2) UTS (τzθ ¼ 0-30-30-0 kPa, until liquefaction at p΄0¼100 kPa) (1) IC (σ΄z ¼ σ΄r ¼ σ΄θ ¼ 50-100 kPa) (2) UTS (τzθ ¼ 0-30 kPa, flow failure at p΄0¼100 kPa) 0cause dilation. Therefore, the increments in volumetric strain (dεvol) during undrained loading, which are equal to zero, are assumed to consist of volumetric strain components due to both dilatancy (dεdvol) and consolidation/swelling (dε c vol), as expressed in Eq. (1). dεvol ¼ dεcvolþdεdvol ¼ 0: ð1Þ Stress pathstropic consolidation. UTS: undrained cyclic torsional shear loading, p0¼Initial d FL.I.N. De Silva et al. / Soils an5624. Evaluation of dεcvol Fig. 1(a) shows volumetric strain (εcvol) versus p 0 during the swelling of isotropically consolidated Toyoura sand specimens. It can be clearly seen that specimens of similar density have similar swelling curves. The εcvol values reported in Fig. 1(a) are evaluated by employing the local deformation measurement (De Silva et al., 2005) and assuming the isotropy of the horizontal bedding plane Fig. 1. Comparison of swelling curves for (a) loose and dense sand Toyoura sand specimens, (b) 10 isotropic cycles between p0 ¼100 and 400 kPa of a typical loose specimen and (c) 1st and 10th cycles of a typical loose specimen.(i.e., radial and circumferential strains, εr and εθ, respectively are equal). Local deformation measurement transducers (LDTs) are mounted on metal hinges, which are glued onto the membrane. However, excessive hinge deformation may take place when the confining stress becomes less than 50 kPa, causing an error when evaluating εcvol, as shown in the upper left corner of Fig. 1(a). In Fig. 1(a), it can also be observed that the swelling curves of Toyoura sand, subjected to isotropic unloading and reloading cycles in the range of p0 from 100 to 400 kPa, can be expressed by Eq. (2). dεcvol ¼ dp0 Ko p0 p 0 0  mk ð2Þ where Ko is the bulk modulus at the reference mean effective stress (p 0 0) and mk is a material parameter. As shown in Fig. 1(a), it is found that the value for Ko (evaluated at p 0 0¼100 kPa) is 58 MPa for dense Toyoura sand specimens (Dr¼75–80%) and 50 MPa for loose Toyoura sand specimens (Dr¼54–57%). The value for mk, for both dense and loose specimens, is taken as 0.9. The effect of large isotropic loading/unloading cycles (IC cycles) on the swelling curve is shown in Fig. 1(b). Swelling curves for the first and tenth cycles are compared in Fig. 1(c). No significant effect on the swelling curve due to the application of isotropic cycles could be observed after applying 10 IC cycles (disregarding the effects of creep during the application of 10 cycles). Therefore, a unique swelling curve is employed in the current study to model εcvol. 5. Modeling of monotonic stress–shear strain relationship It is a well-known fact that τzθ=p0 versus plastic shear strain (γpzθ) during drained or undrained monotonic shear is of a non- linear shape (e.g., Koseki et al., 1998; among others). A typical relationship for a cyclic undrained test, conducted on a Toyoura sand specimen of Drini¼75.7%, is shown in Fig. 2(a). γpzθ is evaluated by deducting the elastic strain (γezθ) components from the total shear strain (γzθ). The γ e zθ component is evaluated by employing the quasi-elastic constitutive model (IIS model) pro- posed by HongNam and Koseki (2005, 2008). The parameters employed in the IIS model for a dense specimen with Drini of about 75% (defined at a confining stress of 30 kPa) are listed in Table 2. It should be mentioned that, when the soil reaches the full liquefaction state (i.e., p0E0), the τzθ=p0 values become extremely sensitive to very small changes in p0. Hence, highly scattered data can be observed. Koseki et al. (2005) investigated the liquefaction properties of sand under low confining stress levels and proposed a simplified procedure to estimate the liquefaction resistance by introducing the concept of the apparent increase in effective mean principal stress (Δp0), due to particle interlocking, as well as parameter Δτ to correct τ due to possible measurement errors. Therefore, a modified stress ratio was proposed by Koseki et al. (2005): ðτzθΔτzθÞ=ðp0 þΔp0Þ. Chiaro et al. (2013) showed that the values for Δτzθ and Δp0 may slightly vary during cyclic oundations 55 (2015) 559–574loading. However, for simplicity, the values for Δτzθ and Δp0 were assumed to be constant, as originally suggested by Koseki et al. (2005) in the present study. The modified stress ratio is employed in Fig. 2(a) with Δτzθ¼1.3 kPa and Δp0 ¼0.2 kPa. Typical evaluations of Δτzθ and Δp0 are shown in Fig. 2(b). It can be clearly seen that ðτzθΔτzθÞ=ðp0 þΔp0Þ versus γpzθ is of a non- linear shape with hysteresis, which is similar to τzθ=p0 versus γ p zθ of a drained cyclic test. In order to quantitatively investigate the above, comparisons of τzθ=p0 versus γ p zθ for the virgin loading of drained and undrained specimens of similar densities are shown in Fig. 3. It can be clearly seen that, for similar densities, τzθ=p0 versus γ p zθ for the undrained test is very similar to that for the drained test. This observation suggests that it is possible to evaluate dεdvol (increment in volumetric strain due to dilatancy) during cyclic undrained loading by combining the simulation of the stress– shear strain relationship of a drained test with an appropriate stress–dilatancy relationship. Normalized stress–strain relationships during the virgin loading (backbone curves) of undrained tests with different densities are compared in Fig. 4. The peak shear stress ratio ((τzθ=p0)max) and the initial quasi-elastic shear modulus (Gzθ0) at p 0 0¼100 kPa for a specimen of Drini¼75% were taken as -1.6 -1.2 -0.8 -0.4 0.0 0.4 0.8 1.2 -0.8 -0.6 -0.4 -0.2 0.0 0.2 0.4 0.6 0.8 Δτ Δ (τ zθ -Δ τ z θ)/ (p '+ Δ p' ) γpzθ (%) -0.8 -0.4 0.0 0.4 0.8 1.2 1.6 2.0 -0.5 0.0 0.5 1.0 1.5 2.0 2.5 3.0 Δτ = 1.3 kPa τ z θ (k Pa ) p' (kPa) Δp' = 0.2 kPa Fig. 2. (a) ðτzθΔτzθÞ=ðp0 þΔp0Þ versus γpzθ relationship of undrained test and (b) evaluation of Δτzθ and Δp0. Table 2 Model parameters. Test Quasi-elastic model parametersa ram L.I.N. De Silva et al. / Soils and Foundations 55 (2015) 559–574 563SAT 38 Ezo¼215 MPa, σ΄o¼100 kPa, νzθ0¼0.18, m¼0.5, n¼0.5, k¼0.3, CE¼CG¼0.0, a¼0.7SAT 28 SAT 31 SAT 32 SAT 34 Ezo¼190 MPa, σ΄o¼100 kPa, νzθ0¼0.18, m¼0.5, n¼0.5, k¼0.3, CE¼CG¼0.0, a¼0.7 SAT 33 Ezo¼145 MPa, σ΄o¼100 kPa, νzθ0¼0.18, m¼0.5, n¼0.5, k¼0.3, CE¼CG¼0.0, a¼0.7 aDetails of the formulation of the quasi-elastic model (IIS Model) and its pa bDrag and hardening parameters are the same as those employed in the simulati loading in De Silva and Koseki (2012). cDamage parameter Dult, for drained cyclic loading, is taken as 0.20.85 and 100 MPa, respectively, based on the results of a drained test on a specimen of similar density. For Drini¼50%, the (τzθ=p0)max and Gzθ0 values are 0.78 and 60 MPa, respec- tively, and for Drini¼21%, the (τzθ=p0)max and Gzθ0 values are 0.60 and 42.9 MPa, respectively. Reference shear strain γref is taken as the ratio of (τzθ=p0)max to Gzθ0=p 0 0. Then, the Generalized Hyperbolic Equation (GHE), as proposed by Tatsuoka and Shibuya (1991) (Eq. (3)), was employed to simulate the backbone curve shown in Fig. 4. Y ¼ X 1 C1ðXÞ þ jXj C2ðXÞ ð3Þ where C1ðXÞ ¼ C1ð0ÞþC1ð1Þ 2 þ C1ð0ÞC1ð1Þ 2 cos π α0 X  mt þ1 ! ð4aÞ C2ðXÞ ¼ C2ð0ÞþC2ð1Þ 2 þ C2ð0ÞC2ð1Þ 2 cos π β0 X  nt þ1 0 B@ 1 CA ð4bÞ and X and Y are normalized plastic shear strain and shear stress parameters, respectively. Normalized stress and strain para- meters, as defined below, are selected by following the Drag parametersb Hardening parameterb, Sult Damage parameter c, Dult D1¼0.15 1.15 0.6 D2¼12 D1¼0.01 D2¼3.13 not required (specimen fails by flow failure) eters are presented in HongNam and Koseki (2005, 2008) on of the stress–shear strain relationship during drained cyclic torsional shear Tatsuoka et al. (1997) reported that the stress–strain relationships d F0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 Drained backbone (OC = 1) γpzθ (%) Drained loading SAT35(Drini = 79.3 %, OC = 1) Undrained loading SAT32(Drini = 75.2 %, OC = 1) Toyoura sand (saturated) (σ'z = σ ' r = σ ' θ)initial = 100 kPa τ z θ/p ' Undrained backbone Fig. 3. Drained and undrained τzθ=p0 versus γ p relationships during virgin L.I.N. De Silva et al. / Soils an564procedure proposed by De Silva and Koseki (2012): X ¼ γ p zθ γref and Y ¼ τzθ=p 0 ðτzθ=p0Þmax ð5Þ where γpzθ ¼ γzθγezθ and γref ¼ ðτzθ=p0Þmax=ðGzθ0=p0Þ. The GHE has 8 parameters, i.e., C1(0), C1(1), C2(0), C2(1), α0, β0, mt and nt, which can be determined from a single monotonic drained torsional shear test. Refer to Tatsuoka and Shibuya (1991) for the procedure for evaluating these parameters. It can be seen from Fig. 4 that the normalized stress–shear strain relationships of undrained tests with different densities are similar. Hence, they can be modeled by employing a single set of GHE parameters obtained either by a normally con- solidated drained test or an undrained shearing test. In order to obtain a better fitting to the experiment data, GHE parameters of undrained tests are selected by slightly modifying those of drained tests (Fig. 4). -5 0 5 10 15 20 25 30 35 40 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 Simulation using GHE parameters that are slightly modified from SAT30 SAT32 (Drini = 75.2 %) SAT28 (Drini = 74.6 %) SAT34 (Drini = 50.3 %) SAT33 (Drini = 21.3 %) Toyoura sand(saturated) p'o= 100 kPa γref = (τz /p')max/ (Gz /p' )Y= ( τ zθ /p ')/ ( τ z θ /p ') m ax X = γpzθ/γref Simulation using GHE parameters of a drained test (SAT30, OC =1) GHE parameters (modified from SAT 30) C1(0) = 35, C1(∞) = 0.3, α = 0.000146, m = 0.25 C2(0) = 0.078, C2(∞) = 1, β = 0.178, n = 0.25 Fig. 4. Comparison of the backbone curves of undrained tests with different densities. zθ loading.of soils are significantly influenced by cyclic strain hardening, damage due to straining etc., which are caused by the rearrange- ment of particles during cyclic loading, and proposed additional rules to model these features to account for behaviors under more general stress conditions. In this regard, a conceptual approach was implemented for dense Toyoura sand under a plane strain condition by applying a horizontal shift to the basic skeleton curves, i.e., dragging the basic skeleton curve along the X-axis (strain parameter axis) in the opposite direction to its loading direction, while applying the (extended) Masing's rule (refer to Masuda et al. (1999) and Tatsuoka et al. (2003) for further details). It was assumed that the amount of drag β, applied to one basic skeleton curve in one loading direction, is a function of the plastic shear strain accumulated in the opposite loading direction (Masuda et al., 1999; Balakrishnaiyer and Koseki, 2000; Tatsuoka et al., 2003; HongNam and Koseki, 2008). The same approach was employed to model the cyclic stress–strain relationships of Toyoura sand under cyclic torsional shear loading (HongNam and Koseki, 2008; De Silva and Koseki, 2012). The dragged backbone curve can be written as follows: Y ¼ ðXβÞ 1 C1ðXβÞ þ jXβj C2ðXβÞ ð6Þ where β denotes the amount of drag, which can be evaluated by a drag function as shown below. β ¼ X 0 1 D1 þ X0D2 ð7Þ where D1 is a fitting parameter (i.e., initial gradient of the drag function) and D2 is the maximum amount of drag. D1 and D2 can be experimentally determined. X0 is the accumulated normalized plastic shear strain in one direction (positive or negative direction). HongNam and Koseki (2008) employed D1¼0.45 and D2¼3.13 for dense Toyoura sand (Drini¼71%) subjected to cyclic torsional shear loading starting from an isotropic stress state (σ΄z ¼ σ΄r ¼ σ΄θ ¼ 100 kPa). However, it was observed by De Silva and Koseki (2012) that the application of drag alone is not sufficient for simulating the cyclic stress–shear strain relationship close to the peak stress of the material. 6.1. Modification of extended Masing's rule In view of the limitation of Masing's rule with drag in simulating the stress–shear strain relationship close to the peak stress of the material, De Silva and Koseki (2012) proposed two conceptual modification factors, which take into account6. Modeling of cyclic stress–shear strain relationship by using extended Masing's rule Subsequent unloading/reloading cycles are modeled by employing the procedure proposed by De Silva and Koseki (2012), as briefed below. oundations 55 (2015) 559–574the hardening behavior during cyclic loading (reduction in damping ratio with constant stress amplitude cyclic loading) to dilative at the phase transformation state (usually when τzθ=p'E0.5 in the case of Toyoura sand). In addition, De Silva and Koseki (2012) proposed a conceptual equation for hardening parameter S by assuming that S can be expressed as a hyperbolic function of the total normalized plastic strain up to the current turning point, as follows: S¼ 1þ PΔX  Upto current turing point D2 D1 þ PΔX  Upto current turing point Sult1ð Þ ð11Þ where Sult is the maximum value for S after applying an infinite number of cycles, and D1 and D2 are the same drag parameters Toyoura sand with Drini¼50% (De Silva and Koseki, 2012). d Foundations 55 (2015) 559–574 565and damage to plastic stiffness (dτzθ=dγ p zθ) at large stress levels, while maintaining continuity in the simulation. The two parameters in the GHE, C1(X¼0) and C2(X¼1), represent the initial plastic stiffness and the peak strength, respectively. Therefore, the damage occurring to the plastic stiffness can be obtained by multiplying C1(X) by damage factor D, and the hardening can be obtained by multiplying C2(X) by hardening factor S. Note that in the approach used by Masuda et al. (1999), Balakrishnaiyer and Koseki (2000), and Tatsuoka et al. (2003), unique backbone curves were used to model subsequent cyclic branches by employing extended Masing's rules. Therefore, the hysteresis curve, starting from an arbitrary point A, can be obtained by employing the extended Masing's rules with damage and hardening by using Eq. (8). Y ¼ YAþ XXA 1 C1 X  XA np   D þ jXXAj npC2 X  XAnp   S ð8Þ In order to maintain continuity, the dragged backbone curve, which is in the same direction as the curve to be modeled, should be modified, as shown in Eq. (9), to determine the proportional parameter np by using the extended Masing's rules. Refer to Masuda et al. (1999) and Tatsuoka et al. (2003) for details on the extended Masing's rules and sub-rules. Y ¼ ðXβÞ 1 C1ðXβÞD þ jXβj C2ðXβÞS ð9Þ Note that parameters D and S were employed in Eqs. (8) and (9) in such a way that D and S are constant for a given hysteresis curve, but they change from curve to curve. An evaluation of D and S for a particular hysteresis curve is made based on a few empirical equations, as follows. The plastic shear modulus (D) can be expressed as an “S curve”, as proposed by De Silva and Koseki (2012) and shown in Eq. (10). D¼Dultþ 1þexp γnð Þ 1þexp Δγpzθ p γn   1Dultð Þ ð10Þ where jΔγpzθjp is the total plastic shear strain (%) accumu- lated between the current and the previous turning points, and Dult is the minimum value for D, which would be applied to evaluate the minimum value for the plastic shear modulus. γn corresponds to the value for γpzθ (in %) at which the volumetric behavior of the material changes from contractive to dilative, and is taken as 0.8. De Silva and Koseki (2012) proposed Dult¼0.2 for drained cyclic torsional shear loading, although experimental evidence suggests that Dult may lie between 0.2 and 0.6. In the current study, for an improved simulation of the experimental results, Dult¼0.6 is used for the case of cyclic undrained loading. Yet, it should be noted that the D value is assumed to be equal to L.I.N. De Silva et al. / Soils an1.0 (i.e., no damage to the plastic shear modulus) until the volumetric behavior of the material changes from contractiveThe same values are employed for the respective densities in the simulation of undrained behavior. The concepts of drag, hard- ening and damage during drained cyclic torsional shear loading are illustrated in Fig. 5. 7. Evaluation of dεdvol In order to evaluate dεdvol, it is necessary to combine dγ p zθ with an appropriate stress–dilatancy relationship. For the above purpose, the stress–dilatancy relationship, proposed by De Silva and Koseki (2012), is modified as follows. During undrained cyclic loading, the mean effective stress (p0) mainly decreases with the number of cycles. It is assumed in this study that the above reduction in p0 is associated with -2.0 -1.5 -1.0 -0.5 0.0 0.5 1.0 1.5 2.0 2.5 3.0 -80 -60 -40 -20 0 20 40 60 80 τ γ (%) used in Eq. (7). As suggested by De Silva and Koseki (2012), Sult¼1.15 was employed in the current study. Note that the values for D1 and D2 differ for the cases with and without damage and hardening. HongNam and Koseki (2008) proposed D1¼0.45 and D2¼3.13 for Toyoura sand (Drini¼71%) subjected to cyclic torsional shear loading starting from an isotropic stress state (σ΄z ¼ σ΄r ¼ σ΄θ ¼ 100 kPa). After introducing the damage and hardening factors, D1¼0.15 and D2¼12 were found to be appropriate for dense Toyoura sand with Drini¼75%, and D1¼0.01 and D2¼3.13 was suggested forFig. 5. Illustration of the concepts of drag, hardening and damage using a typical drained cyclic torsional shear test (De Silva and Koseki, 2012). over-consolidation and cyclic mobility. Firstly, the soil under- going a decrease in p0 is subjected to over-consolidation until the stress state exceeds the phase transformation stress state (Ishihara and Li, 1972) for the first time (i.e., the first instance where the volumetric strain increment changes from contractive to dilative behavior, i.e., dεdvolo0). Then, the soil will enter the stage of cyclic mobility. 7.1. Stress–dilatancy relationship during virgin loading and before exceeding the phase transformation stress state It can be observed in Fig. 6(a) and (b) that the stress–dilatancy relationship during cyclic loading before exceeding the phase transformation stress state is different from that after exceeding the phase transformation stress state (refer to the explanations given within Fig. 6(a) and (b)). In addition, it can be observed from Fig. 6(a) that the effects of over-consolidation significantly alter the stress–dilatancy relationship during virgin loading (refer to the line shown as “start of loading” in Fig. 6(a)). Since sand will be subjected to over-consolidation/swelling during undrained cyclic loading, the stress–dilatancy relationships during different stages are addressed separately in the current study. The stress–dilatancy relationships, employed in the evalua- tion of dεdvol, consist of four equations, namely, Eqs. (A), (B), (C) and (D) in Fig. 7(a). The stress–dilatancy model is summarized in Table 3. According to the proposed model, the stress–dilatancy relationship during the virgin loading of normally consolidated Toyoura sand is given by Eq. (12) (refer to Eq. (A) of Fig. 7(a)) with Rk¼1.3 and C¼0.6. τzθ p0 ¼ Rk  dεdvol dγpzθ   7C for dγpzθ40 and dγ p zθo0; respectively ð12Þ Since Toyoura sand was subjected to undrained cyclic torsional shear loading from a normally consolidated stress state, dεdvol during virgin loading was evaluated in the current study by applying the stress–dilatancy relationship for virgin loading, as given by Eq. (A) in Fig. 7(a). If the loading direction after virgin loading is reversed before exceeding the phase transformation stress state, a different linear relationship is employed, by referring to Fig. 6, to account for the -0.4 -0.2 0.0 0.2 0.4 0.6 0.8 τ z θ/p ' -0.4 Rk= 1.5, C= -0.50C= -0.18 Cyclic loadings after exceeding -0.4 -0.2 0.0 0.2 0.4 0.6 0.8 D C BA L.I.N. De Silva et al. / Soils and Foundations 55 (2015) 559–574566-1.5 -1.0 -0.5 0.0 0.5 1.0 1.5 -0.6 phase transformation stress state -dεdvol/dγ p zθ the phase transformation stress state for the first time (-0.50 ≥ τz /p' ≥ 0.50) Fig. 6. Stress–dilatancy relationships during cyclic torsional shear loading (a) before exceeding the phase transformation stress state for the first time and (b) bilinear stress–dilatancy relationship after exceeding the phase transformation-1.5 -1.0 -0.5 0.0 0.5 1.0 1.5 -0.8 -0.6 -dεdvol/dγ p zθ -0.2 0.0 0.2 0.4 0.6 0.8 phase transformation stress state Subsequent unloading Subsequent unloading Rk= 0.33, τ z θ/p ' Subsequent loading Rk= 1.5, C= 0.50 Subsequent unloading Rk= 0.33, C= 0.18 Start of loadingstress state. Adopted from De Silva et al. (2014).-0.8 -0.4 0.0 0.4 0.8 1.2 -0.8 -0.6 -0.8 -0.4 0.0 0.4 0.8 1.2 -0.8 -0.6 -0.4 -0.2 0.0 0.2 0.4 0.6 0.8 Fig. 7. Proposed stress–dilatancy relationships for (a) different stress states during undrained cyclic torsional shear loading of dense sand and (b) the stress state within the over-consolidation boundary surface. Adopted from De Silva et al. (2014). .. (A den ely  Rk ) and cu d Fohardening behavior until the stress state exceeds the phase transformation stress state for the first time (refer to steep dashed lines in Fig. 6(a) and (b)). Refer to the stress–dilatancy relationship denoted by Eq. (B) in Fig. 7(a) for further details. Eq. (B), as shown in Fig. 7(a), is obtained by employing Rk¼2.2 and C¼0.5 in Eq. (12) for both dense and loose Toyoura sand (Drini¼75% or Drini¼50%). 7.2. Effects of over-consolidation on stress–dilatancy relationship The over-consolidation ratio (OC) continuously changes (i. e., increases) when the cyclic loading continues within the Table 3 Stress–dilatancy relationships to evaluate dεdvol during undrained cyclic loading Simulation case Stress–dilatancy relationships Virgin loadinga Subsequent loading Case 1 (Bilinear) Eq. (A) with Rk ¼ 1:3; C¼ 0:6 Eq. (A) with Rk ¼ 1:5;C¼ 0:5 and Eq Case 2 (Modified bilinear) Eq. (A) with Rk ¼ 1:3; C¼ 0:6 If τzθ=p0maxr C Eq. (B) with Rk ¼ 2:2;C¼ 0:5for both (Drini 50%) Toyoura sand, respectiv Case 3 (Modified bilinearþEffects of OC) Eq. (A) with Rk ¼ 1:3; C¼ 0:6 If τzθ=p0maxr Cτzθ=p04 C  lnðOCÞ τzθ=p0rC Same as Case 2.1 Eq. (C) with (Drini¼75% respectively OC: over-consolidation ratio. aVirgin loading always starts from a normally consolidated stress state in the specimen. L.I.N. De Silva et al. / Soils anlimits of the phase transformation stress state (i.e., |τzθ=p0| oC). Hence, the effects of the change in OC on the stress– dilatancy relationship, within the above stress range, are taken into account by applying Eq. (13) (Oka et al., 1999) (refer to Eq. (C) in Fig. 7(a)). Note that the model by Oka et al. (1999) was formulated using the stress and strain invariants, while maintaining the objectivity. For simplicity, on the other hand, the stress–dilatancy relationship formulated in this study is described one-dimensionally using specific stress components. As will be described at the end of Section 9, an attempt is under way to extend the current stress–strain description into a generalized three-dimensional modeling. For dτzθ40 and dτzθo0:  dε p vol dγpzθ   ¼Dk τzθ p0  τzθp0 =lnðOCÞ   Rk 0 @ 1 A ð13Þ where Dk ¼  τzθp0 =ðC  lnðOCÞÞ 1:5 The Dk value changes with OC. It is assumed that Eq. (13) continues until Dk¼1 with Rk¼2.2 and C¼0.50 for both dense and loose specimens and then follows Eq. (13) with Dk¼1 (i.e., changes in Dk due to changes in the stress state arenot considered after Dk41). Refer to the stress–dilatancy relationship denoted by Eq. (C) in Fig. 7(a) for details. The stress state at which Dk¼1 is defined as the over-consolidation boundary surface (Oka et al., 1999). When Dk¼1 C¼ τzθ p0 =lnðOCÞ ð14Þ Then, Eq. (13) can be rewritten as follows:  dε p vol dγpzθ   ¼ τzθ p0 C Rk ! ð15Þ Eq. (15) corresponds to Eq. (B) in Fig. 7(a). After that, the ) with Rk ¼ 0:33; C¼ 0:18 for immediately after stress reversal If τzθ=p0max4C se (Drini 75%) and loose (Case 2.1) Eq. (D) with Rmax ¼ 1:5; Cmin ¼ 0:36and Eq. (A) with Rk ¼ 0:33;C¼ 0:18 for immediately after stress reversal Eq. (D) with Rmax ¼ 1:5; Cmin ¼ 0:36and Eq. (A) with Rk ¼ 0:33;C¼ 0:18for immediately after stress reversallnðOCÞ  ¼ 2:2; C¼ 0:5 for both dense loose (Drini¼50%) specimens, rrent study τzθ=p0max: the maximum value of τzθ=p0 currently applied to the undations 55 (2015) 559–574 567stress–dilatancy relationship follows Eq. (15) until the stress state exceeds the phase transformation stress state for the first time and then follows the modified bilinear stress–dilatancy relationship (Eq. (D) shown in Fig. 7(a)). In short, if stress reversal occurs before exceeding the phase transformation stress state, the stress–dilatancy relationship follows a combination of Eqs. (B) and (C) until the stress state exceeds the phase transformation stress state for the first time. The combination of Eqs. (B) and (C), for dτzθ40 with different values for OC, is illustrated in Fig. 7(b). Rearranging the terms in Eq. (14), we get the over- consolidation boundary surfaces for positive and negative shear stress increments, as shown in Eqs. (16a) and (16b). For dτzθ40: τzθ ¼ C  p0  lnðOCÞ ð16aÞ Similarly, for dτzθo0: τzθ ¼ C  p0  lnðOCÞ ð16bÞ Fig. 8 shows a typical over-consolidation boundary surface for a normally consolidated specimen subjected to cyclic undrained torsional shear loading starting from p 0 0¼100 kPa. minimum value for C after application of a large number of d Fconstant stress amplitude cyclic loadings (i.e., Cmin¼0.36 adopted from De Silva et al. (2014)) and D is the sameThe phase transformation stress state for the above specimen is also indicated in Fig. 8. 7.3. Stress–dilatancy relationship after exceeding the phase transformation stress state If the loading direction during virgin loading is reversed after exceeding the phase transformation stress state, the stress–dilatancy relationship will follow the modified bilinear stress–dilatancy relationship, as given below, for subsequent cyclic loadings (refer to Eq. (D) in Fig. 7(a)). τzθ p0 ¼ ðRmax  DÞ  dεdvol dγpzθ   7 Cmin D ð17Þ where Rmax is the maximum value for Rk in Eq. (12) (i.e., Rmax¼1.5 adopted from De Silva et al. (2014)), Cmin is the 0 20 40 60 80 100 120 -40 -30 -20 -10 0 10 20 30 40 τ z θ (k Pa ) p' (kPa) OC boundary surface τz = × × Phase transformation line (τzθ/p' = C ) p'0 Fig. 8. Typical over-consolidation boundary surface and phase transformation stress state for a normally consolidated specimen with no initial shear. L.I.N. De Silva et al. / Soils an568damage parameter as in Eq. (10). The following boundary conditions were specified for the RmaxD and Cmin/D values by referring to the experimental data (based on De Silva et al. (2014)). Note that, RmaxD should not be less than 1.0. The maximum value for Cmin/D should not be larger than 0.60. Therefore, the RmaxD value in Eq. (2) varies between 1.5 and 1.0, while the Cmin/D value varies between 0.36 and 0.60 depending on the accumulated plastic strain between the current and the previous turning points (i.e., damage parameter D). The application of different stress–dilatancy equations for different stress states during a typical cyclic undrained test is illustrated in Fig. 9. The four-phase stress–dilatancy model, employed in the current study, is summarized in Table 3. By combining Eq. (1) with (2), the following equation can be derived to evaluate the change in mean effective stress (dp0) during undrained cyclic loading: dp0 ¼ Ko p0 p0o  mk ðdεdvolÞ ð18Þ8.1. Dense sand behavior The experimentally obtained effective stress path (τzθ versus p0) and stress–shear strain relationships (τzθ versus γzθ) of a dense Toyoura sand specimen, subjected to undrained cyclic torsional shear loading with stress amplitude (τzθ=p 0 o) equal to 0.22, are shown in Fig. 10(a) and (a1), respectively. In order to show the improvement in the simulation results with the introduction of the modifications to the stress–dilatancy relationship, a simulation ofwhere dεdvol is the volumetric strain increment due to dilatancy, which is evaluated by combining the simulation of stress–shear strain relationship during drained cyclic torsional loading with the proposed stress–dilatancy model. Then, by the numerical integration of Eq. (18), the generation of excess PWP with the shear stress level (or p0 versus τzθ) can be established. In addition, the stress–strain relationship (i.e., τzθ versus γ p zθ) can also be obtained. 8. Simulation of liquefaction behavior 0 20 40 60 80 100 120 -40 -30 -20 -10 0 10 20 30 40 D C B B τ z θ (k Pa ) p' (kPa) A Stress state exceeds the phase transformation state for the first time application of Eqs. A, B, C, D for different stress-dilatancy regions Fig. 9. Application of different stress–dilatancy relationships for different stress states during undrained cyclic torsional shear loading. oundations 55 (2015) 559–574the stress paths was carried out in three steps, as denoted by (b), (c) and (d) in Fig. 10. The corresponding stress–strain relation- ships are denoted by (b1), (c1) and (d1), respectively. First, a unique bilinear stress–dilatancy relationship (refer to De Silva et al. (2014) for details), without considering any effects of over-consolidation or change in the stress–dilatancy relationship with cyclic loading, is employed (refer to Case 1 in Table 3). The simulated stress path and the stress–strain relationship that corresponds to each of the above stress– dilatancy relationships are shown in Fig. 10(b) and (b1), respectively. It can be seen that the reduction in the change in p0, due to the effects of over-consolidation (see Fig. 10(a)), cannot be accurately simulated by employing the above stress– dilatancy relationship, as shown in Fig. 10(b). In addition, dense Toyoura sand shows a continuous increase in shear strain after the onset of cyclic mobility, as shown in Fig. 10 (a1), which is similar to that of a loose sand specimen, and ends up yielding the same curve (closed loop) for subsequent cycles, as shown in Fig. 10(b1). -20 0 20 40 60 80 100 120 -30 -20 -10 0 10 20 30 -6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6 -30 -20 -10 0 10 20 30 -20 0 20 40 60 80 100 120 -30 -20 -10 0 10 20 30 -6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6 -30 -20 -10 0 10 20 30 -20 0 20 40 60 80 100 120 -30 -20 -10 0 10 20 30 -6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6 -30 -20 -10 0 10 20 30 -20 0 20 40 60 80 100 120 -30 -20 -10 0 10 20 30 -6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6 -30 -20 -10 0 10 20 30 Fig. 10. Simulations of stress paths and stress–strain relationships using different stress–dilatancy relationships (τzθ=p0 ¼0.22, Drini¼75.7%). L.I.N. De Silva et al. / Soils and Foundations 55 (2015) 559–574 569 Fig. 11. Simulations of stress paths and stress–strain relationships using different stress–dilatancy relationships (τzθ=p0 ¼0.40, Drini¼79.4%). L.I.N. De Silva et al. / Soils and Foundations 55 (2015) 559–574570 Consequently, the variations in the stress–dilatancy relation- ship during cyclic loading are taken into account in the simulation by employing the modified bilinear stress–dilatancy relationship (refer to De Silva et al. (2014)) without consider- ing any effects due to over-consolidation (refer to Case 2 in Table 3). It can be seen from Fig. 10(c) that the stress path after the onset of cyclic mobility is improved and, in addition, the stress–strain relationship is improved to some extent showing a continuous increase in shear strain with cyclic loading after the onset of cyclic mobility, as shown in Fig. 10 (c1). However, when the stress path enters the steady state, the stress–strain relationship becomes a closed loop. Furthermore, since the effects of over-consolidation are not taken into account in the above simulation, the reduction in the change in p0 due to the effects of over-consolidation, as shown in Fig. 10(a), cannot be accurately simulated. Finally, the further modified stress–dilatancy model, sum- marized in Case 3 (Table 3), which considers the effects of over-consolidation and changes in the stress–dilatancy rela- tionship during cyclic loading, was employed in the simulation of the stress path and stress–strain relationships during undrained cyclic loading, as shown in Fig. 10(d) and (d1), respectively. It can be observed that the simulations for both the stress path and the stress–strain relationship are certainly -20 0 20 40 60 80 100 120 -40 -30 -20 -10 0 10 20 30 40 Toyoura sand (saturated) Test SAT34 Drini = 50.3 % τ z θ (k Pa ) p' (kPa) Exper ime nt OC boundary surface phase transformation line -10 -8 -6 -4 -2 0 2 4 6 8 10 -40 -30 -20 -10 0 10 20 30 40 Toyoura sand (saturated) Test SAT34 Drini = 50.3 % τ z θ (k Pa ) γzθ (%) Experiment -20 0 20 40 60 80 100 120 -40 -30 -20 -10 0 10 20 30 40 Test SAT34 D rini = 50.3 % Bilinear stress-dilatancy model τ z θ (k P a) p' (kPa) Simulation Phase transformation l ine -10 -8 -6 -4 -2 0 2 4 6 8 10 -40 -30 -20 -10 0 10 20 30 40 Toyoura sand (saturated) Test SAT34 Drini = 50.3 % Bilinear stress-dilatancy model τ z θ (k P a) γpzθ (%) Simulation Toyoura sand (saturated) L.I.N. De Silva et al. / Soils and Foundations 55 (2015) 559–574 571-20 0 20 40 60 80 100 120 -40 -30 -20 -10 0 10 20 30 40 Test SAT34 Drini = 50.3 % Modified bilinear stress-dilatancy model τ z θ (k P a) p' (kPa) Simulation Toyoura sand(saturated)Fig. 12. Simulations of stress paths and stress–strain relationships using-10 -8 -6 -4 -2 0 2 4 6 8 10 -40 -30 -20 -10 0 10 20 30 40 Toyoura sand (saturated) Test SAT34 Drini = 50.3 % Modified bilinear stress-dilatancy model γp (%) τ z θ (k P a) Simulationzθ different stress–dilatancy relationships (τzθ=p0 ¼0.30, Drini¼50.3%). obtained from the experimental data is compared with that obtained from the simulation results by employing different stress–dilatancy relationships. However, it can be seen that the simulation was significantly improved when the modified bilinear stress–dilatancy relationship was employed while considering the effects of over-consolidation. It can be seen that the liquefaction resistance of dense Toyoura sand speci- mens is underestimated (i.e., in the simulation liquefaction occurs faster) by the simulated results. On the other hand, the experimentally evaluated liquefaction resistance curves of the loose Toyoura sand specimens (obtained using data from this study and Chiaro et al., 2012) are similar to those obtained from the simulation results after employing the modified bilinear stress–dilatancy relationship with the effects of over- consolidation (refer to Fig. 14(b)). 9. Conclusions The following main conclusions can be derived from the above study: 15a) -1 0 1 2 3 4 5 6 7 8 9 10 0 5 10 γpzθ (%) τ z θ (k P Experiment Fig. 13. Comparison of (a) stress paths and (b) stress–strain relationships (Drini¼21.3%). d Fimproved after introducing the effects of over-consolidation into the modified bilinear stress–dilatancy relationship. A continuous increase in strain can be observed until the stress path enters the steady state. It should be noted that the simulation started producing the same curve when jγpzθj was close to about 4%, while the experimental data shows a continuous increase in shear strain with cyclic loading. This strain level (jγpzθj¼4%) is close to the strain level at which the peak stress state of dense Toyoura sand is mobilized in drained shearing. Further modification of the stress–strain relationship, which considers the strain- softening behavior observed during cyclic torsional shear loading (Kiyota et al., 2008), would be necessary to address the above issue for dense sand. However, this was not attempted in the current study. Since the simulation of the stress–strain relationship of dense Toyoura sand subjected to cyclic undrained loading using the bilinear stress–dilatancy model (refer to Fig. 10(b1)) gives the same curve whenγpzθ 3%, liquefaction resistance is defined in the current study as the number of cycles required to yield a double amplitude shear strain of 6%. A comparison of the experimental stress paths and the stress–strain relationships of dense Toyoura sand specimens, subjected to undrained cyclic torsional shear loading, with their simulation results for stress amplitudes (τzθ=p 0 o) equal to 0.40, is shown in Fig. 11. The same stress–dilatancy relation- ships as employed in Fig. 10(b)–(d) are utilized in Fig. 11. 8.2. Loose sand behavior The experimental effective stress paths and stress–strain relationships of a loose Toyoura sand specimen (Drini¼50%) with stress amplitudes (τzθ=p 0 o) equal to 0.30 are shown in Fig. 12(a) and (a1), respectively. It can be seen in Fig. 12(c) and (c1) that the simulations of the stress path and the stress– strain relationship, respectively, become consistent with the corresponding experimental data when the modified bilinear stress–dilatancy relationship is employed (compare Fig. 12(a) and (a1) with Fig. 12(c) and (c1), respectively). The simulation continues until the double amplitude of shear strain becomes about 18% and then gives the same curve for further cycles. Fig. 13(a) compares the experimentally obtained stress paths of a very loose specimen (Drini¼21.3%) with its simulation, while Fig. 13(b) compares the experimentally and numerically obtained stress–strain relationships. Note that the specimen shows flow failure and the simulation shows a similar tendency. Therefore, only the stress–dilatancy relationship during virgin loading (refer to Table 3) is required for the simulation. 8.3. Liquefaction resistance curve Fig. 14(a) and (b) shows the liquefaction resistance curves for dense and loose Toyoura sand specimens, respectively. In the current study, liquefaction resistance is defined as the L.I.N. De Silva et al. / Soils an572number of cycles required to yield a double amplitude shear strain of 6%. In each figure, the liquefaction resistance curve0 20 40 60 80 100 120 0 5 10 Bilinear stress-dilatancy model τ z θ ( kP p' (kPa) Experiment 15 20 25 a) Toyoura sand (saturated) Test SAT33 Drini = 21.3 % Bilinear stress-dilatancy model Simulation20 25 Toyoura sand (saturated) Test SAT33 Drini = 21.3 % Simulationounda(1)tions 55 (2015) 559–574A unique swelling curve, that does not change with the number of cycles, has been proposed to evaluate the (2) (3) (4) Fig. and d Fo0.20 τ z0.25 0.30 Simulations Bilinear Modified bilinear Modified bilinear + OC Experiment θ/p ' o0.1 0.2 0.3 2 4 6 8 10 20 40 60 80 Toyoura sand Drini = 75 % no.of cycles τ no.of cycles required to have DA = 6 % 0.350.4 0.5 0.6 Simulations Bilinear Modified bilinear Modified bilinear + OC Experiment zθ /p ' o L.I.N. De Silva et al. / Soils anincrements in volume change due to consolidation/swelling during undrained cyclic loading. The normalized stress–shear strain relationships of undrained tests with different densities were found to be similar. Hence, they can be modeled by employing a single set of GHE parameters obtained either by normally consolidated drained tests or undrained shearing tests. τzθ=p0 versus γ p zθ of an undrained specimen is very similar to that of a normally consolidated drained specimen of similar density. Hence, similar GHE para- meters can be used for a drained normally consolidated specimen to evaluate the increments in volume change due to the dilatancy of an undrained specimen. In addition, drag parameters, damage parameters and hard- ening parameters were also similar for both drained and undrained loadings. The stress path during undrained cyclic loading is divided into four sections, namely (1) virgin loading, (2) stress path within the limits of phase transformation stress state, (3) stress path within the limits of over-consolidation bound- ary surfaces and (4) stress path after exceeding the phase transformation stress state for the first time. Different stress–dilatancy relationships are proposed for each section of the stress path to evaluate the increments in volume change due to dilatancy. De c C De G De S c S 0.10 0.15 0.6 0.81 2 4 6 8 10 20 40 Toyoura sand Drini = 50 % no.of cycles no.of cycles required to have DA = 6 % After Chiaro et al.(2012) 14. Liquefaction resistance curves: (a) dense Toyoura sand, Drini¼75% (b) Loose Toyoura sand, Drini¼50%.1 De Sr troceedings of the International Symposium on Geomechanics and eotechnics of Particulate Media, Ube, Yamaguchi, Japan, pp. 29–33. ilva, L.I.N., Koseki, J., Sato, T., Wang, L., 2005. High capacity hollow ylinder apparatus with local strain measurements. Proceedings of the econd Japan–U.S. Workshop on Testing, Modeling and Simulation, vol. 56. Geotechnical Special Publication, ASCE16–28. ilva, L.I.N., Koseki, J., Wahyudi, S., Sato, T., 2014. Stress–dilatancyte PSilva, L.I.N., Koseki, J., 2012. Modelling of sand behavior in drained yclic shear. In: Miura (Ed.), Advances in Transportation Geotechnics II. RC Press, pp. 686–691. Silva, L.I.N., Koseki, J., Sato, T., 2006. Effects of different pluviation chniques on deformation property of hollow cylinder sand specimens. In:References Balakrishnaiyer, K., Koseki, J., 2000. Modelling of stress–strain relationships of a reconstituted gravel subjected to large cyclic loading. In: Elnashai, A. S., Antoniou, S. (Eds.), Implications of Recent Earthquakes on Seismic Risk: Series on Innovation in Structures and Construction, vol. 2; 2000, pp. 105–114. Chiaro, G., Kiyota, T., Koseki, J., 2013. Strain localization characteristics of loose saturated Toyoura sand in undrained cyclic torsional shear tests with initial static shear. Soils Found. 53 (1), 23–34. Chiaro, G., Koseki, J., Sato, T., 2012. Effects of initial static shear on liquefaction and large deformation properties of loose saturated Toyoura sand in undrained cyclic torsional shear tests. Soils Found. 52 (3), 498–510.Acknowledgments Our special appreciation goes to Prof. R. Uzuoka of the University of Tokushima, Japan, for his valuable suggestions in developing the cyclic elasto-plastic model. In addition, the authors sincerely acknowledge the Ministry of Education, Culture, Sports, Science, and Technology (Grants-in-Aid for Scientific Research #21360221), Japan for providing the financial assistance for the current research.(5) The simulations of stress paths and stress–strain relation- ships of Toyoura sand subjected to undrained cyclic torsional shear loadings were significantly improved after employing the proposed four-phased stress–dilatancy relationship. (6) Complete liquefaction and steady state can be accurately simulated by the proposed model; however, further modifica- tions are required to address the strain-softening behavior at large strain levels of dense sand. In order to develop a generalized three-dimensional model based on the proposed stress–strain description, extended research has been initiated as preliminarily described by Namikawa et al. (2011) on the results from a numerical simulation of drained cyclic loading behavior. To maintain the objectivity, in terms of the independence from the coordinate systems, the three-dimensional model adopts the concept of an infinite number of nesting surfaces (Mroz et al., 1978). Its application to the simulation of undrained cyclic loading behavior, as is the case with the current study, is under way. Details will be reported elsewhere. undations 55 (2015) 559–574 573elationships of sand in the simulation of volumetric behavior during cyclic orsional shear loadings. Soils Found. 54 (4), 845–858. HongNam, N., Koseki, J., 2005. Quasi-elastic deformation properties of Toyoura sand in cyclic triaxial and torsional loadings. Soils Found. 45 (5), 19–38. HongNam, N., Koseki, J., 2008. Deformation characteristics of dry Toyoura sand in large cyclic torsional loading and their modeling. In: Burns, Mayne, Santamarina (Eds.), Proceedings of the Fourth International Symposium on Deformation Characteristics of Geomaterials, vol. 2, IS Atlanta, USA, pp. 801–807. Iai, S., Matsunaga, Y., Kameoka, T., 1992. Strain space plasticity model for cyclic mobility. Soils Found. 32 (2), 1–15. Ishihara, K., Li, S., 1972. Liquefaction of saturated sand in triaxial torsion shear test. Soils Found. 12 (2), 19–39. Ishihara, K., Tatsuoka, F., Yasuda, S., 1975. Undrained deformation and liquefaction of sand under cyclic stresses. Soils Found. 15 (1), 29–44. Jefferie, M.G., 1993. Nor-Sand: a simple critical state model for sand. Géotechnique 43 (1), 91–103. Kiyota, T., Sato, T., Koseki, J., Abadimarand, M., 2008. Behavior of liquefied sand under extremely large strain levels in cyclic torsional shear tests. Soils Found. 48 (5), 727–739. Koseki, J., Hamaya, S., Tatsuoka, F., Maeshiro, N., 1998. Elastoplastic deformation characteristics of Toyoura sand during liquefaction. Geotech- nical Engineering and Soil Dynamics III, vol. 75. Geotechnical Special Namikawa, T., Koseki, J., De Silva, L.I.N., 2011. Three dimensional modeling of stress-strain relationship of sand subject to large cyclic loading. In: Proceedings of the 5th International Conference on Earthquake Geotechni- cal Engineering. Santiago, Chile, paper-ID: TMSNA, CD-ROM, pp. 1–10. Nishimura, S., 2002. Development of Three Dimensional Stress–Strain Model of Sand Undergoing Cyclic Undrained Loading and Stress–Axes Rotation (M.Eng. thesis). Dept. of Civil Engineering, The University of Tokyo, Japan. Nishimura, S., Towhata, I., 2004. A three-dimensional stress–strain model of sand undergoing cyclic rotation of principal stress axes. Soils Found. 44 (2), 103–116. Oka, F., Yashima, A., Kato, M., Sekiguchi, K., 1992. A constitutive model for sand based on the non-linear kinematic hardening rule and its application. In: Proceedings of 10th World Conference on Earthquake Engineering, pp. 2529–2534. Oka, F., Yashima, A., Tateishi, Y., Taguchi, Y., Yamashita, S., 1999. A cyclic elasto-plastic constitutive model for sand considering a plastic-strain dependence of the shear modulus. Geotechnique 49 (5), 661–680. Seed, H.B., Lee, K.L., 1966. Liquefaction of saturated sands during cyclic loading. Proc. ASCE SM6, 105–134. Tatsuoka, F., Shibuya, S., 1991. Modelling of non-linear stress–train relations of soils and rocks – Part 2: new equation, Seisan-kenkyu. J. IIS Univ. Tokyo 43 (10), 435–437. Tatsuoka, F., Jardine, R.J., Lo Presti, D., Di Benedetto, H., Kodaka, T., 1997. Characterising the pre-failure deformation properties of geomaterials, Theme Lecture for the Plenary Session No. 1, Proceedings of XIV IC on stress–strain relations of sand in cyclic plane strain loading. J. Geotech. L.I.N. De Silva et al. / Soils and Foundations 55 (2015) 559–574574generalized plasticity. J. Eng. Mech. 132 (12), 1380–1391. Liou, C.P., Richart, F.E., Streeter, V.L., 1977. Numerical model for liquefac- tion. J. Geotech. Eng. Div. ASCE, 103; 589–606. Liyanapathirana, D.S., Poulos, H.G., 2002. A numerical model for dynamic soil liquefaction analysis. Soil Dyn. Earthq. Eng. 22 (9–12), 1007–1015. Masuda, T., Tatsuoka, F., Yamada, S., Sato, T., 1999. Stress-strain behavior of sand in plane strain compression, extension and cyclic loading tests. Soils Found. 39 (5), 31–45. Mroz, Z., Norris, V.A., Zienkiewicz, O.C., 1978. An anisotropic hardening model for soils and its application to cyclic loading. Int. J. Numer. Anal. Methods Geomech. 2, 203–221.Geoenviron. Eng. ASCE, 129; 450–467. Towhata, I., Ishihara, K., 1985a. Shear work and pore water pressure in undrained shear. Soils Found. 25 (3), 73–85. Towhata, I., Ishihara, K., 1985b. Undrained strength of sand undergoing cyclic rotation of principal stress axes. Soils Found. 25 (2), 135–147. Vaid, Y., Thomas, J., 1995. Liquefaction and postliquefaction behavior of sand. J. Geotech. Eng. 121 (2), 163–173.Ling, H.I., Yang, S., 2006. Unified sand model based on the critical state and SMFE, vol. 4, Hamburg, pp. 2129–2164. Tatsuoka, F., Masuda, T., Siddiquee, M.S.A., Koseki, J., 2003. Modeling thePublication, ASCE385–397. Koseki, J., Yoshida, T., Sato, T., 2005. Liquefaction properties of Toyoura sand in cyclic torsional shear tests under low confining stress. Soils Found. 45 (5), 103–113.