International Journal of Engineering Science 84 (2014) 29–53
Contents lists available at ScienceDirect
International Journal of Engineering Science journal homepage: www.elsevier.com/locate/ijengsci
An experimental investigation of the mechanical behavior and a hyperplastic constitutive model of frozen loess Yuanming Lai a,⇑, Xiangtian Xu b, Wenbing Yu a, Jilin Qi a a State Key Laboratory of Frozen Soil Engineering, Cold and Arid Regions Environmental and Engineering Research Institute, Chinese Academy of Sciences, Lanzhou 730000, China b Institute of Transportation, Inner Mongolia University, Hohhot 010070, China
a r t i c l e
i n f o
Article history: Received 31 January 2014 Received in revised form 21 June 2014 Accepted 24 June 2014
Keywords: Frozen loess Hyperplasticity theory Stress path Thermodynamic function Constitutive model
a b s t r a c t For the engineering construction in cold regions, it is essential to understand the mechanical properties of frozen soil. In order to investigate the mechanical characteristics of frozen loess, three kinds of laboratory tests were carried out at 6 °C. These included constant-confining pressure triaxial compression tests, isotropic compression tests, and constant-slope stress path tests. The test results show that the direction of the plastic strain increment is influenced by stress path. The influence of the mechanism of the confining pressures on the strength and deformation of frozen loess is investigated in detail. An analytical method to determine pressure crushing and melting of the ice in frozen soil is proposed. In order to apply the hyperplasticity theory to modeling constitutive behavior of frozen soil, a new systematical approach to derive the yield criterion and flow rule from dissipation function is proposed based on the properties of homogenous function. From the mechanical characteristics of frozen loess, Gibbs free energy function and dissipation function of frozen loess are established by applying the hyperplasticity theory. An elasto-plastic incremental constitutive model for frozen loess is derived from the two thermodynamic functions and a method of determining the corresponding parameters is also given. Simulated results show that the constitutive model, proposed in this paper, describes well the deformation behavior of frozen loess under different stress levels and stress paths. Ó 2014 Elsevier Ltd. All rights reserved.
1. Introduction Frozen soil, including permafrost and seasonally freezing soil, accounts for about 50% of the land area on the earth (Xu, Wang, & Zhang, 2010). Loess is widely distributed in the world, occupying 1/10 of the world’s land area, in both permafrost and seasonally frozen regions. For instance, in the permafrost region of Alaska loess is widely distributed in the piedmont zone. While in the seasonally frozen region of Lanzhou city, China, loess deposits with a thickness of up to 409.9 m, which is the largest loess thickness and its geological profile is the most typical in the world. Also, there is a lot of loess or loess-like deposits in cold regions in Russia, Canada, America, China etc. With the increase in infrastructure construction in cold regions with loess, such as highways, railways, tunnels and irrigation work with frozen soil as their foundations, it is essential to study the mechanical behavior of frozen loess for engineering design, especially to establish a reasonable constitutive model for frozen loess. ⇑ Corresponding author. Tel.: +86 931 4967288; fax: +86 931 8279968. E-mail addresses:
[email protected] (Y. Lai),
[email protected] (X. Xu),
[email protected] (W. Yu),
[email protected] (J. Qi). http://dx.doi.org/10.1016/j.ijengsci.2014.06.011 0020-7225/Ó 2014 Elsevier Ltd. All rights reserved.
30
Y. Lai et al. / International Journal of Engineering Science 84 (2014) 29–53
Frozen soil consists of four phases, i.e., soil skeleton, ice, unfrozen water and air. Compared with unfrozen soil, the mechanical properties of frozen soil is much more complex. Besides the factors which are also important for unfrozen soils, such as confining pressure, water content and loading rate, temperature plays a dominant role in mechanical behavior of frozen soils. In order to understand the mechanical properties of frozen soils, a large number of experiments have been carried out (Bragg & Andersland, 1981; Chamberlain, Groves, & Perham, 1972; Fish, 1991; Lai, Long, & Chang, 2009; Zhang, Lai, Sun, & Gao, 2007). The factors influencing the strength and deformation characteristics of frozen soil are investigated using both uniaxial compression and triaxial compression with constant confining pressure conditions. Some empirical formulas are proposed. Furthermore, Azmatch, Sego, Arenson, and Biggar (2011) investigated tensile strength and stress–strain behavior of Devon silt under frozen fringe conditions. For more information on experimental studies, the readers can refer to the most up to date summarization on the state-of-the-art of frozen soils by Lai, Xu, Dong, and Li (2013). However, few previous studies were found which made an extensive investigation on mechanical properties of frozen loess. The constitutive model is a mathematical tool for describing the basic mechanical properties of materials. In the past decades, many constitutive models for unfrozen soils and other solid materials have been successfully developed by introducing plasticity theory developed for metal materials. When investigating the shear deformation of soils, Drucker, Gibson, and Henkel (1957) proposed a cap yield surface and a hardening law for the first time. According to the concept of yield locus for clay proposed by Calladine (1963) and Roscoe and Poorooshasb (1963) established an elasto-plastic constitutive model for normally consolidated soils considering critical state of soil, and later on Roscoe and Burland (1968) modified the CamClay Model. Lade and Duncan (1975, 1977) established the so-called Lade–Duncan Model, in which non-associated flow rule is adopted and plastic work is taken as the hardening parameter to describe the failure and dilatancy effects of sandy soil. Huang, Pu, and Chen (1981) established the Tsinghua elasto-plastic model for soils based on an associated flow rule. In this model, the directions of plastic strain increments under various stress states were first determined from test results, and then the yield criterion was determined according to the associated flow rule (Li, 2004). Considering the effects of tangential stress/strain rate on soils and granular materials, a class of sub loading surface models for these materials was proposed and developed (Hashiguchi, 1980; Hashiguchi, Ozaki, & Okayasu, 2005; Hashiguchi & Protasov, 2004; Hashiguchi & Tsutsumi, 2001, 2007; Tsutsumi & Hashiguchi, 2005; Tsutsumi, Hashiguchi, Okayasu, Saitoh, & Sugimoto, 2001). Cleja-Tigoiu (2000) established a constitutive equation for anisotropic finite elastoplasticity in the framework of materials with relaxed configurations and internal state variables (Cleja-Tigoiu, 1990a, 1990b). By idealizing the materials as a random array of identical, elastic, frictional spheres and assuming that the particles move with an average strain, Ragione, Prantil, and Sharma (2008) proposed a model for inelastic response of an idealized granular material. Based on the definition of critical state, classical elastoplasticity and bounding surface, Anandarajah (2008) proposed a multi-mechanism anisotropic model for granular materials. A number of elasto-plastic constitutive models were established based on the friction and compression characteristics of geotechnical materials or compressible porous materials (Cleja-Tigoiu, Cazacu, & Tigoiu, 2008; Hashiguchi, 2005; Hashiguchi, Saitoh, Okayasu, & Tsutsumi, 2002; Shen, Shao, Kondo, & Gatmiri, 2012). When the plastic strain is calculated using a single yield surface theory, it cannot explain the phenomenon that the direction of plastic strain increment varies with the stress path for some geotechnical materials. On the other hand, applying an associated flow rule in single yield surface model usually results in excessive plastic dilatancy (Chen & Saleeb, 2001). To overcome these defects, a series of multiple yield surfaces or multiple plastic potential models for geotechnical materials were proposed (Calladine, 1971; Desai, 1980; Gens & Nova, 1993; Ottosen & Ristinmaa, 2013; Peric & Ayari, 2002; Seiki, 1979; Xie & Shao, 2012; Zhu, Shao, & Mainguy, 2010). These constitutive models based on the conventional elasto-plastic theory can describe the mechanical behaviors of geotechnical materials to some extent, and some are applied in engineering practice (Chen & Saleeb, 2001). The laws of thermodynamics are universally applicable. In the study of constitutive theory for continuum or deformable solid, the early researchers realized that the thermodynamic laws should also be introduced to describe the mechanical process of continuum. Based on the irreversible thermodynamic theory, Ziegler (1977) proposed a frame for establishing constitutive models and its foundation—Ziegler’s orthogonality principle (Houlsby & Puzrin, 2000). This principle has ever been criticized on a number of grounds (Collins & Houlsby, 1997). Srinivasa (2010) also pointed out Ziegler’s orthogonality principle was based on an incorrect assumption. However, Collins and Houlsby (1997) proposed that the principle is a weak assumption and can be used to define a very wide class of constitutive equations. Furthermore, by applying convex analysis, Srinivasa (2010) showed that a same orthogonality principle can be derived from maximum rate of dissipation criterion both in rate-dependent case and rate-independent case (in this case, the first-order homogeneity of dissipation function was required). Therefore, in this study, we adopt the orthogonality principle proposed by Srinicasa to modeling the constitutive behavior of frozen loess. The approach to constitutive modeling based on thermodynamics was termed as ‘‘hyperplasticity’’ (Puzrin & Houlsby, 2001a, 2001b). In hyperplasticity, the constitutive model of a material is directly derived from two thermodynamic functions, i.e., energy potential function and dissipation function. At the same time, the elastic law, yield criterion, flow rule and hardening law corresponding to the conventional elasto-plastic theory were formulated (Collins & Houlsby, 1997; Houlsby & Puzrin, 2000) Later on, hyperplasticity was further developed and widely used in constitutive modeling of soils (Chiarelli, Shao, & Hoteit, 2003; Cleja-Tigoiu, 2003; Maugin, 1992, 1999; Rosakis, Rosakis, Ravichandran, & Hodowany, 2000; Rousselier, 2001; Vorobiev, 2008; Zhu et al., 2010). Houlsby (1981, 1982) introduced hyperplasticity to geotechnical materials and showed that the constitutive model of some geotechnical materials could be derived within the frame of irreversible thermodynamics. Collins and Houlsby (1997) further developed and improved the constitutive theory for geotechnical materials
31
Y. Lai et al. / International Journal of Engineering Science 84 (2014) 29–53
within the same frame. They presented a new method for deriving constitutive models from the Gibbs free energy function and dissipation function and theoretically showed the non-associated behavior of the single yield surface model for frictional materials. Houlsby and Puzrin (2000) suggested that the constitutive model for dissipative materials can be established from the four energy functions according to Legendre transformation. Using the multiple yield surface models in conventional plastic theory, Puzrin and Houlsby (2001a, 2001b) extended the single internal variable to multiple internal variables and internal functions. Collins (2005) put forward the concept of frozen elastic energy in soils. Based on this, the thermodynamic compatibility of constitutive models for some soils was investigated (Collins, 2003; Collins & Hilder, 2002; Collins & Kelly, 2002b). There are many advantages to establish constitutive models based on irreversible thermodynamic theory. The first is that the laws of thermodynamics are automatically satisfied. Secondly, the formulation is compact. It is necessary to establish some competitive constitutive models in a uniform theory frame, which makes comparison of different models more convenient so as to select the most promising one (Houlsby & Puzrin, 2000). The complexities mentioned above means that the constitutive models for unfrozen soils are not suitable for frozen soils. Thus a constitutive model which describes the main mechanical properties of frozen soils should be established (Lai, Yang, Chang, & Li, 2010; Lai, Zhang, & Li, 2009). In frozen soil engineering, frozen soil is affected by both mechanical and thermal loads. The deformation characteristics of frozen soils are influenced by the soil temperature change, ice–water phase change and moisture migration. It is therefore a typical hydro-thermo-mechanic coupling problem (Lai et al., 2009). This feature of frozen soil shows that the deformation process under a surcharge load is actually a complex thermodynamic process. It is necessary to establish a constitutive model within the framework of irreversible thermodynamic theory which can automatically couple the constitutive model and other differential equations, making the hydro-thermo-mechanic coupling model more consistent and systematical. In this study, the frozen Lanzhou loess is taken as study object. A series of triaxial compression tests were carried out under constant confining pressure, to study the influence of confining pressure on the strength and deformation characteristics from a new viewpoint. Isotropic compression tests and constant-slope tests were used to investigate the influence of the stress path on the deformation characteristics of frozen soil. In this paper, based on the test results and the hyperplasticity theory, specific expressions of Gibbs free energy function and dissipation function for frozen Lanzhou loess are proposed. An incremental elasto-plastic constitutive model for frozen Lanzhou loess, within the frame of irreversible thermodynamics, is established and the method to determine the model parameters is presented. In addition, to make the hyperplasticity more applicable and comprehensible, a method of deriving yield criterion and flow rules from dissipation function without Legendre transformation and convex analysis is proposed. 2. Mechanical test on frozen loess All the mechanical tests mentioned in this study were conducted under axisymmetric conditions. First of all, the mechanical symbols used in this study are defined and their transformational relations are given. They are as follows:
r1 – maximum axial stress (maximum principle stress), r3 – confining pressure, p – mean stress, q – deviator stress or equivalent stress, pa – atmospheric pressure, pa = 0.101325 MPa, ea – axial strain, e3 – lateral strain, ev – volumetric strain, ec – equivalent strain or shear strain, Under axisymmetric condition,
p¼
r1 þ 2r3 3
;
q ¼ r1 r3 ;
2 3
ev ¼ ea þ 2e3 ; ec ¼ ðea e3 Þ:
2.1. Test material and sample preparation The soil tested in this study was loess collected from Donggang town, Lanzhou City, a seasonally frozen soil region. The grain size distribution of the tested loess is shown in Table 1. To ensure the uniformity of the samples, the loess was mixed
Table 1 Grain size distribution of the loess (%). <0.005 mm
0.005–0.05 mm
0.05–0.075 mm
0.075–0.10 mm
0.10–0.25 mm
>0.25 mm
11.21
68.46
14.03
4.54
1.76
0
32
Y. Lai et al. / International Journal of Engineering Science 84 (2014) 29–53
with water at a moisture content of about 17%, and then kept for 6 h without evaporation to allow the water to be uniformly distributed in the soil. After this, the mixed loess was filled in a cylindrical mold, to make cylindrical samples of 12.5 cm (height) 6.18 cm (diameter) with the desired dry densities. The basic physical parameters of the loess are listed in Table 2. 2.2. Sample freezing and test temperature To avoid large frost heaving during freezing, the samples were put in rigid molds and were quickly frozen at 30 °C. After 48 h of freezing, the samples were taken from the molds and mounted with epoxy resin platens on both ends and covered with rubber sleeve. Finally, the samples were kept in an incubator for over 12 h until the samples reached the target testing temperature of 6.0 °C. 2.3. Test equipment The test equipment used in this study is a cryogenic triaxial apparatus modified from the MTS-810 material test machine, which can measure volumetric strain of sample in triaxial test under constant confining pressure. Fig. 1 shows a photo and schematics of the equipment. The cryogenic triaxial apparatus can be controlled by load, displacement and time, respectively. The test is automatically controlled by a computer program, and the data recorded automatically. The maximum axial load is 100 KN. The axial displacement variation range is 85 cm 85 cm. The available confining pressure range is 0 20 MPa. The liquid used in confining pressure loading is aeronautic hydraulic oil. The medium in the cooling system is alcohol. The test temperature range is 30 °C 25 °C. 2.4. Three types of mechanical tests for frozen loess In this study, three types of mechanical tests on frozen Lanzhou loess were carried out under 6.0 °C, (a) triaxial compression test under constant confining pressure (including the case of loading–unloading–reloading test), (b) isotropic compression test (including the case of loading- unloading–reloading test) and (c) constant-slope stress path test. These tests can not only be used to study the mechanical behavior of frozen soil under different loading modes (or different stress paths), but also to obtain the parameters for constitutive model and to verify the constitutive models. 2.4.1. Constant confining pressure triaxial compression–CCPTC test In these tests the confining pressure is kept constant during the test. In this study, a series of CCPTC tests with confining pressures ranging from 0.5 MPa to 17.0 MPa were carried out. Before testing, the temperature in the pressure chamber was set at 6.0 °C, and the samples were mounted into the apparatus. After the temperature in the pressure chamber was stable, the confining pressure was applied and kept constant for 5 min before the axial load was applied. The axial strain rate was 1.67 104/s. The tests were stopped when the axial train reached 20%. To obtain the elastic shear module of the samples, a series of loading–unloading–reloading tests (under CCPTC condition) were carried out with confining pressures ranging from 0.5 MPa to 9.0 MPa. The first unloading began when the axial strain reached to ea1. After unloading the deviator stress equaled to zero (r1 r3 = 0). Then another loading was applied until the axial strain reached to ea2, followed by an unloading. In this study, 7 loading–unloading cycles were carried out with a strain rate of 1.67 104/s. 2.4.2. Isotropic compression–IC test To investigate mechanical behavior of frozen soil samples under hydrostatic pressure and determine the elastic volumetric module of the samples, isotropic compression tests with loading–unloading–reloading cycles (under IC condition) on frozen Lanzhou loess were carried out. In these tests, the axial strain was controlled. The displacement of axial pressure piston was limited to zero, and the tests were controlled by confining pressure loading system. The reaction force of the axial pressure piston generated by the confining pressure was used to make the constant hydrostatic pressure in the pressure chamber. The curves of volumetric strain versus hydrostatic pressure were obtained. In the loading–unloading–reloading tests (isotropic compression), four different stress levels were used to investigate the influence of confining pressures (or stress level) on elastic volumetric module of frozen soil. In addition, in the isotropic compression tests loading rates ranged from 0.1 MPa/min to 10 MPa/min. It was found that the loading rate had little influence on the isotropic compression curve when lower than 0.1 MPa/min. Therefore, in the isotropic compression tests a loading rate of 0.1 MPa/min was used.
Table 2 Basic physical parameters of the loess. Density
Water content (%)
Specific gravity
Liquid limit (%)
Plastic limit (%)
2.01 g/cm3
16.9
2.69
27
15.4
Y. Lai et al. / International Journal of Engineering Science 84 (2014) 29–53
33
Fig. 1. Cryogenic triaxial apparatus. (a) Photo, (b) Schematics.
2.4.3. Constant slope stress path–CSSP test To investigate the influence of stress path on the deformation of the frozen loess, in addition to the constant confining pressure triaxial compression tests, four kinds of constant-slope stress path tests were carried out on the frozen loess. The constant-slope means that the slope of stress path d q/d p was kept constant during the whole test process. The constant confining pressure compression test actually refers to the constant-stress path test of dq/dp = 3. The four stress paths SP1, SP2, SP3 and SP4, with corresponding slopes 98.24, 113.68, 99.74 and 77.90, respectively, as shown in Fig. 2 were carried out. In the four CSSP tests, the confining pressures of 2 MPa, 3 MPa, 4 MPa and 5 MPa, respectively, were applied. Then the axial loading began with an axial loading rate of 0.006KN/s and the confining pressure loading began with a loading rate of 0.001 MPa/s. 2.5. Test results and analyses 2.5.1. Effects of confining pressure on the mechanical behaviors of frozen loess The CCPTC test results for frozen loess under different confining pressures are shown in Figs. 3 and 4. Fig. 3 shows the influence of confining pressure on the stress–strain curves. When confining pressure is below 2 MPa, the stress–strain curves
Fig. 2. Four constant-slope stress paths in p–q plane.
34
Y. Lai et al. / International Journal of Engineering Science 84 (2014) 29–53
Fig. 3. Stress–strain curves of frozen loess from CCPTC tests.
Fig. 4. Volumetric strain curves of frozen loess from CCPTC tests.
show softening, which decreases with the increase of confining pressure. When the confining pressure exceeds 2 MPa, the stress–strain curves show hardening. Fig. 4 shows how the confining pressure influences the volumetric strain of frozen loess during shear process. Under all confining pressures there is first of all contraction followed by dilatancy. When the confining pressure is lower than 2 MPa, the contraction stage of frozen loess is relatively small. Dilatancy of frozen loess occurs at axial strains less than 10%, and the final dilatancy is relatively large. The reason for this is that the radial confinement is low when the confining pressure is low, and the relative movement between solid particles in the sample after shear failure results in volumetric expansion. When the confining pressure is higher than 2 MPa, only small volumetric expansion occurs during the shear process of the sample. In particular when the confining pressure is more than 7 MPa, the volumetric strain is small, between 0.037% to 0.642%. In order to describe more concretely the influence of confining pressure on the mechanical behavior of frozen loess more accurately, the strength qf and the initial slope of stress–strain curve, Ei, are introduced. The strength is defined according to the standard for test methods of soils GB/T 50123–1999 (People’s Republic of China National Standard, 1999). That is to say, for the softening curves shown in Fig. 3, the peak deviation stress is taken as strength, while for hardening curves, the strength is equal to the deviation stress when axial strain is 15%. qf and Ei are the two main variables determining the shape of stress–strain curves. qf and Ei can therefore be used to determine the influence of confining pressure on the mechanical behavior of frozen loess. The strength-confining pressure relationship and initial slope-confining pressure relationship are shown in Figs. 5 and 6, respectively.
Y. Lai et al. / International Journal of Engineering Science 84 (2014) 29–53
35
Fig. 5. Relationship between strength and confining pressure.
Fig. 6. Relationship between Initial slope and confining pressure.
It can be seen in Fig. 5 that the variation of the strength of frozen loess with confining pressures can be divided into three phases. Under relatively low confining pressure, i.e., from 0.5 MPa to 4 MPa, pressure melting does not occur, so the increase of confining pressure only makes the solid particles (soil particles and ice particles) more compact. This increases the sliding and interlocking frictional forces, which enhance the strength of frozen loess. In confining pressure ranges from 4 MPa to 13 MPa, the ice in the sample begins to be crushed and pressure melting occurs. The ice crushing reduces the ice bonding force between the particles, while the pressure melting of ice increases the unfrozen water film on the granular surfaces, both of which decrease the angle of internal friction of the frozen loess. At the same time, the extent of crushing and pressure melting of the ice increase with the increase in confining pressure. The strength of frozen loess decreases with the increase of confining pressure. When the confining pressure exceeds 13 MPa (a critical value in this study), the pressure melting produces enough porous water to saturate the sample. The confining pressure increment is almost totally carried by the porous water. The effective confining pressure affecting on the solid skeleton no longer increases. The studies on other types of frozen soil show similar tendencies (Chamberlain et al., 1972; Parameswaran & Jones, 1981; Xu, Lai, Dong, & Qi, 2011). From Fig. 6, it can be seen that the initial slope of stress–strain curve with varying confining pressure can also be divided into three phases. In the first phase, the increase of confining pressure makes the solid granular more compact, which results
36
Y. Lai et al. / International Journal of Engineering Science 84 (2014) 29–53
in the initial slope increasing with the increase of confining pressure. It should be noted that the confining pressure range for the initial slope (from 0.5 MPa to 13 MPa) is larger than that for strength (from 0.5 MPa to 4 MPa). This indicates that in the range from 4 MPa to 13 MPa the extent of ice crushing is greater than that of pressure melting. The reason for this is that the ice crushing decreases the strength of frozen loess other than the initial slope. The variation difference between initial slope and strength with confining pressure also exists in frozen sandy soils (Xu et al., 2011). In the second phase, the pressure melting of ice increases with the increase in confining pressure, so the initial slope decreases with the increase of confining pressure in this phase. In the third phase, it is similar to the variation of strength with confining pressure. The initial slope is not influenced obviously by confining pressure, but remains constant. It further shows that when confining pressure reaches a critical value, the pore water generated by the pressure melting of ice carries almost all the confining pressure increment, and the mechanical behavior of frozen loess is not influenced. 2.5.2. The influence of stress path on the direction of plastic strain increment of frozen loess Numerous studies show that the direction of plastic strain increment of unfrozen soil is related to not only stress state, but also stress increment direction (Hambly, 1972; Lade & Duncan, 1976; Lewin & Burland, 1970; Tatsuoka & Ishihara, 1974). This implies that the deformation behavior of unfrozen soil is influenced by the stress path. To describe this characteristic of unfrozen soils, some constitutive models for unfrozen soils, considering the influence of stress path, have been established (Seiki, 1979; Shen, 1980; Yin & Duncan, 1984). However, for frozen soils, almost no reports on the influence of stress path on the direction of plastic strain increment have been published. In this study, two directions of plastic strain increment under stress paths with slopes of dq/dp = 3.00 and dq/dp = -113.68 were investigated, as shown in Fig. 7. In order to identify how the direction of plastic strain increments vary with the change of stress path from geometry, the stress space (p, q) in Fig. 7 is chosen together with the plastic strain increment space depv ; depc , namely, the two sets of coordinates coincide. It can be seen from Fig. 7 that, after stress states A and B reach point C along stress paths AC and BC, respectively, the plastic strain increments generated by the stress increments along these two stress paths are not of the same direction. This indicates that for frozen loess, the direction of plastic strain increment is not alone determined by the stress state, and the influence of stress path should also be taken into account. 2.5.3. Variation of elastic parameters with confining pressures When calculating the elastic strain for elasto-plastic materials, the elastic parameters should be determined. In this study, the elastic volumetric module is determined using isotropic compression loading–unloading–reloading tests. The elastic shear module is determined by the loading–unloading–reloading tests under constant confining pressure compression. The ev ln (1 + p/pa) curve, obtained from isotropic compression loading–unloading–reloading tests, is shown in Fig. 8. It can be seen in Fig. 8 that an elastic after-effect exists in the unloading phase in isotropic compression loading–unloading–reloading tests on frozen loess, and that hysteresis loop of energy dissipation is formed during unloading–reloading process. The elastic strain of frozen soil accounts for small amount of the total strain (Lai et al., 2010), so for calculation convenience, the hysteresis loop of ev ln (1 + p/pa) in unloading–reloading phase is approximated to a straight line. The slope of each straight line can be calculated by the following expression:
Fig. 7. Directions of plastic strain increment under different stress paths.
Y. Lai et al. / International Journal of Engineering Science 84 (2014) 29–53
37
Fig. 8. Isotropic compression loading–unloading–reloading curve of frozen loess.
k¼
deev d lnð1 þ p=pa Þ
ð1Þ dee
According to Eq. (1), one has d lnð1 þ p=pa Þ ¼ kv . For convenience, a dimensionless parameter K = 1/k is introduced which has the following relationship with the conventional elastic volumetric module K:
K¼
dp ¼ Kðp þ pa Þ deev
It can be seen from Fig. 8 that k varies with the confining pressure at the unloading points, as does K. From the test results, the relationship between K and confining pressure r3, as shown in Fig. 9, can be fitted by:
K ¼ Ak ðr3 =pa Þ þ Bk
ð2Þ
where Ak = 10.244, Bk = 38.25. Fig. 10 shows the q ec curve obtained from loading–unloading–reloading tests under CCPTC conditions. Similar to the test curve of isotropic compression loading–unloading–reloading test, elastic after-effect also exists in unloading phase for
Fig. 9. Relationship between elastic volumetric module and confining pressure.
38
Y. Lai et al. / International Journal of Engineering Science 84 (2014) 29–53
Fig. 10. Loading–unloading–reloading curve of frozen loess under CCPTC conditions.
loading–unloading–reloading tests under CCPTC conditions, and closed hysteresis loops are also formed by unloading curve and reloading curve. It can be seen that the hysteresis loops in the q ec curve can be approximated as straight lines with slope of C. The relationship between deviator stress and elastic shear strain can be expressed by:
q ¼ Ceec
ð3Þ
In Eq. (3), C is defined as elastic shear modulus. It should be noted that the slope of hysteresis loop, under a given confining pressure, decreases with increase in strain. This is elasto-plastic coupling. The elastic shear modulus for each loading– unloading cycle is shown in Fig. 10. Since the elastic shear modulus decreases gradually with the increase in the strain, the elasto-plastic coupling of frozen loess is considered to be a kind of damage behavior. As the strain increases, the microdefects and micro-cracks in frozen loess are continually merged and expanded, resulting in the continuous deterioration of elastic properties of frozen loess. The stress concentrations at the original defects of the soil may induce the pressure melting of ice during loading process. This is the reason that the shear elastic modulus of frozen loess decreases with the increase of the strain. Since the effect of plastic strain in each cycle is far greater than that of elastoplastic coupling, the elastoplastic coupling phenomenon is ignored. The slope of the first hysteresis loop is taken as the elastic shear modulus. That is to say, the plastic strain is the main part in the total strain for each cycle, and is much larger than the elastic strain. From Fig. 10, it can be seen that the elastic strain only accounts for 4.1% of the total strain in the seventh cycle. Therefore, the contribution of elastic strain to the total strain can be almost ignored, and the effect of elastoplastic coupling phenomenon on both plastic strain and elastic strain also be ignored. Similarly, when the elastic strain accounts for a large proportion in the total strain for an elastoplastic coupling materials, the stress–strain relationship can also be simplified without considering the elastoplastic coupling effect. Furthermore, it is shown by the test results that the elastic shear modulus is influenced by confining pressure. The relationship between elastic shear modulus C and confining pressure r3, as shown in Fig. 11, can be expressed by
C ¼ AC pa ðr3 =pa ÞBC
ð4Þ
where AC = 6001.18, BC = 0.1603. 3. Hyperplasticity theory In this study, a constitutive model for frozen loess was derived by applying the hyperplasticity theory which was developed by researchers (Collins and Houlsby, 1997; Houlsby and Puzrin, 2000; Puzrin and Houlsby, 2001). This theory includes two types: single internal variable hyperplasticity (Collins & Houlsby, 1997; Houlsby & Puzrin, 2000) and multiple internal variable hyperplasticity (Puzrin & Houlsby, 2001a, 2001b), corresponding to the single yield surface model and the multiple yield surface model in classical plasticity, respectively. In the hyperplasticity, the orthogonality principle and first-order homogeneity are two important conditions that the dissipation function needs to meet besides the second law of thermodynamics. These two conditions are very important in constructing the constitutive model for rate-independent materials, since they restrict the selective range of the dissipation function, which also in turn, provides the theoretical basis for its specific form. The basis was extensively discussed by many researchers (Collins, 2005; Collins & Houlsby, 1997; Houlsby
Y. Lai et al. / International Journal of Engineering Science 84 (2014) 29–53
39
Fig. 11. Relationship between elastic shear module and confining pressure.
& Puzrin, 2000; Puglisi & Truskinovsky, 2005; Srinivasa, 2010; Ziegler, 1958, 1963). Collins and Houlsby (1997) applied the degenerate Legendre transformation of dissipation function to derive the yield criterion and flow rule. Srinivasa (2010) provided the existence proof of yield function and flow rule from positively homogenous dissipation function by using gauge function and convex analysis. Although these approaches established the relationship between dissipation function and yield criterion, it relied on the use of complicated mathematical operations which are unfamiliar to most geotechnical engineers. According to the properties of homogeneity function, a systematic method is proposed to derive the yield criterion and the flow rule from dissipation function. This method is more accessible to geotechnical specialist and more convenient to apply. 3.1. Single internal variable hyperplasticity (corresponding to the single yield surface model) In irreversible processes, internal variables are introduced and used to describe internal structure, deformation histories and dissipation behavior of materials (Ziegler, 1977). They can be scalars, vectors and tensors. More details about internal variables were introduced by Maugin (1992). The hyperplasticity using only one internal variable in thermodynamic functions to describe dissipative characteristics is named as single internal variable hyperplasticity. According to the frame of single internal variable hyperplasticity developed by Collins and Houlsby (1997) and Houlsby and Puzrin (2000), if the plastic strain tensor epij is chosen as internal variable, the Gibbs free energy G and the dissipation function D have the following forms:
G rij ; epij ¼ G1 ðrij Þ þ rij epij þ G2 epij D rij ; epij ; e_ pij ¼ hs_ i
ð5Þ ð6Þ
where rij is stress tensor, h is temperature and s_ i is the rate of entropy production per unit volume. For various materials, it should be noted that the dissipation function in Eq. (6) may be either stress-dependent or stress-independent. If the dissipation function is stress-independent, then the normal flow rule applies in the true stress space. Otherwise, the flow rule is non-associated in true stress space. The connection between first-order homogenous dissipation function, yield criterion and flow rule was detailed by Rajagopal and Srinivasa (2005) and Srinivasa (2010). By virtue of Eqs. (5) and (6), the first law of thermodynamics and the second law of thermodynamics can be written as follows:
_ rij ; ep Þ ¼ @G r_ ij þ @G e_ p ¼ r_ ij eij þ D Gð ij @ rij @ epij ij D rij ; epij ; e_ pij ¼ hs_ i P 0
ð7Þ ð8Þ
Substituting Eq. (5) into Eq. (7), gives
eij ¼
@G ; @ rij
D ¼ hs_ i ¼
eeij ¼ eij epij ¼ @G p e_ @ epij ij
@G1 @ rij
ð9Þ ð10Þ
40
Y. Lai et al. / International Journal of Engineering Science 84 (2014) 29–53
In the irreversible process, the rate of the entropy production s_ i was expressed as the sum of products of thermodynamic forces and thermodynamic flows (Li, 1986). Then, in Eq. (10), the internal variable rate e_ pij is referred to as the thermodynamic is taken as a thermodynamic force. vij is defined as a generalized stress in hyperplasticity theory (Collins & flow and vij ¼ @@G ep ij
Houlsby, 1997; Houlsby & Puzrin, 2000). For rate-independent material, much researchers assumed that the dissipation function is a first-order homogeneous function in rate of internal variable e_ pij , and gave various explanations (Collins & Houlsby, 1997; Houlsby & Puzrin, 2000; Collins, 2005; Puglisi & Truskinovsky, 2005; Puzrin and Houlsby 2001; Srinivasa, 2010). According to the first-order homogeneity of dissipation function, owing to Euler’s theorem, one has
D¼
@D p e_ @ e_ pij ij
ð11Þ
Srinivasa (2010) proposed the orthogonality principle for rate-independent materials, based on the maximum rate of dissipation criterion and first-order homogeneity of dissipation function, as following:
@D
vij ¼ _ p @ eij
ð12Þ
According to the first-order homogeneity of D, following yield criterion and flow rule in generalized stress space can be derived based on Srinivasa’s method (Srinivasa, 2010):
Y
rij ; vij ; epij ¼ 0
e_ pij ¼ k
ð13Þ
@Y @ vij
ð14Þ
where Y is the yield criterion in generalized stress space, and k is a multiplier. Furthermore, by differentiating Eq. (5) with respect to epij , one has
vij ¼ rij þ
@G2 @ epij
ð15Þ
2 qij ðepij Þ ¼ @G p was defined as back stress (Collins & Houlsby, 1997). It is the central point of yield surface in true stress space @e ij
and is used to describe kinematic hardening in conventional plasticity, while isotropic hardening is defined by a dependency of yield criterion (13) on internal variable epij . In order to derive the incremental constitutive model, the following consistency condition of yield criterion should be introduced:
_ rij ; v ; ep Þ ¼ @Y r_ ij þ @Y v_ ij þ @Y e_ p ¼ 0 Yð ij ij @ rij @ vij @ epij ij
ð16Þ
From Eq. (15), one has
v_ ij ¼ r_ ij þ
@ 2 G2 p e_ @ epij @ epkl kl
ð17Þ
The multiplier k can be determined by substituting Eqs. (14) and (17) into Eq. (16):
k ¼
@Y @ rij
þ @@Y v
ij
2 @Y @ G2 @ vkl @ ep @ ep kl ij
r_ ij
þ @@Yep
ij
ð18Þ @Y @ vij
By differentiating the second part of Eq. (9) with respect to time, the expression of the elastic strain rate is
e_ eij ¼
@ 2 G1 r_ kl @ rij @ rkl
ð19Þ
Then, from Eqs. (14), (18) and (19), the elasto-plastic incremental constitutive model can be given by:
0
_ pij
_ eij
e_ ij ¼ e þ e ¼ 2
@ G1 where C ep ijkl ¼ @ rij @ r kl
C ep ijkl
r_ kl
@ 2 G1 ¼@ @ rij @ rkl
@Y @Y @Y @Y @ vij @ rkl þ@ vij @ vkl 2G @ @Y @Y þ @Y @Y 2 @ vst @ ep @ ep @ vmn @ ep @ vmn mn st mn
@Y @Y @ vij @ rkl
1
@Y þ @@Y vij @ vkl
@ 2 G2 @Y @Y @ vst @ ep @ epmn @ vmn st
þ @@Y ep
mn
@Y
Ar_ kl
@ vmn
is the elasto-plastic compliance tensor.
ð20Þ
41
Y. Lai et al. / International Journal of Engineering Science 84 (2014) 29–53
3.2. Hyperplasticity of multiple internal variables (corresponding to the multiple yield surface model) For some geomaterials, single yield surface model can not reasonably simulate the dependency of plastic strain increment direction on stress path (Zheng, Shen, & Gong, 2002). In order to improve some of the defects in single yield surface model, a multiple yield surface model should be introduced to describe the effect of stress path on the mechanical characteristic of geomaterials. To build a multiple yield surface model based on hyperplasticity theory, a finite number of internal variables should be introduced to two thermodynamic functions and the number of internal variables is equal to the number of yield surfaces. For consistency, the hyperplasticity using the finite number of internal variables in thermodynamic functions is named as the hyperplasticity of multiple internal variables. This idea was first developed by Puzrin and Houlsby (2001a, 2001b). According to their approach, the Gibbs free energy function and the dissipation function can be expressed as: n n X X ð1Þ ðnÞ G rij ; aij ; . . . ; aij ¼ G1 ðrij Þ þ rij aijðkÞ þ Gk2 aðkÞ ij k¼1
ð21Þ
k¼1
n X ð1Þ ðnÞ ð1Þ ðnÞ ð1Þ ðnÞ ðkÞ Dk rij ; aij ; . . . ; aij ; a_ ij D rij ; aij ; . . . ; aij ; a_ ij ; . . . ; a_ ij ¼
ð22Þ
k¼1
The dissipation function D in Eq. (22) is divided into the sum of n sub-dissipation functions, and it is still a first-order homogeneous function in the rate of internal variables. In decoupled case, each sub-dissipation function Dk is a first-order ðkÞ homogeneous function in the corresponding rate of internal variable a_ ij . As a single internal variable hyperplasticity, the constitutive model with n yield surfaces can be deduced from thermodynamic functions (21) and (22), which automatically obeys thermodynamic principles. In the irreversible process, we assume that total plastic strain of geomaterials is produced in n kinds of mechanisms. If the inter-coupling effect is neglected, the total plastic strain can be split up into additive sub-plastic strains each involving indiP pðkÞ pðkÞ vidual mechanism, i.e. epij ¼ nk¼1 eij , where epij is the total plastic strain, and eij is the sub-plastic strain corresponding to the kth mechanism. Therefore, for constitutive modeling of geomaterials in multiple internal variable hyperplasticity, all of the sub-plastic strains
epðkÞ ij ðk ¼ 1; . . . ; nÞ are chosen as internal variables. Then, Eqs. (21) and (22) can be rewritten as:
n n X X pð1Þ pðnÞ ¼ G1 ðrij Þ þ G rij ; eij ; . . . ; eij rij eijpðkÞ þ Gk2 eijpðkÞ k¼1
ð21aÞ
k¼1
n X pð1Þ pðnÞ pð1Þ pðnÞ pðkÞ pðkÞ D rij ; eij ; . . . ; eij ; e_ ij ; . . . ; e_ ij Dk rij ; eij ; e_ ij ¼
ð22aÞ
k¼1
From Eqs. (21a) and (22a), the multiple yield surface model for geomaterials can be derived by using the approach proposed by Puzrin and Houlsby (2001a). By differentiating Eq. (21a) with respect to rij, decomposition of strain tensor can be deduced:
eij ¼
n @G @G1 X ¼ eeij þ epij ¼ þ epðkÞ @ rij @ rij k¼1 ij
ð23Þ
So that the rate of elastic strain e_ eij can be obtained from Eq. (19). The rate of plastic strain is expressed by sum of n rates of sub-plastic strains corresponding to each mechanism:
e_ pij ¼
n X
e_ ijpðkÞ
ð24Þ
k¼1 pðkÞ Eq. (24) shows that to solve for e_ pij it is first necessary to obtain e_ ij ðk ¼ 1; . . . ; nÞ. From orthogonality principle and firstðkÞ order homogeneity of dissipation function D, each thermodynamic force vij ðk ¼ 1; . . . ; nÞ corresponding to thermodynamic pðkÞ flow e_ ij can be given by
pð1Þ
vijðkÞ ¼
pðnÞ
pð1Þ
pðnÞ
@Dðrij ; eij ; . . . ; eij ; e_ ij ; . . . ; e_ ij Þ pðkÞ
@ e_ ij
pðkÞ
¼
pðkÞ
@Dk ðrij ; eij ; e_ ij Þ pðkÞ
@ e_ ij
ð25Þ
Based on Srinivasa’s approach (Srinivasa, 2010) the kth yield criterion Y(k)(k = 1, . . . ,n) and flow rule corresponding to Dk can be determined in the following form:
Y ðkÞ
pðkÞ rij ; vðkÞ ¼0 ij ; eij ðkÞ
e_ pðkÞ ¼ kðkÞ ij
ð26Þ
ðkÞ pðkÞ ij ; ij ; ij Þ ; ðkÞ @ ij
@Y ðr
v v
e
k ¼ 1; . . . ; n:
ð27Þ
42
Y. Lai et al. / International Journal of Engineering Science 84 (2014) 29–53
Where k(k) is a multiplier corresponding to Dk. Similar to Eq. (15), the kth thermodynamic force
vijðkÞ ¼ rij þ
@Gk2
eijpðkÞ
;
pðkÞ
@ eij
vðkÞ ij is written as
k ¼ 1; . . . ; n
ð28Þ
Furthermore, the consistency condition of each yield criterion is
Y_ ðkÞ
¼ rij ; vijðkÞ ; epðkÞ ij
@Y ðkÞ @Y ðkÞ @Y ðkÞ r_ ij þ ðkÞ v_ ijðkÞ þ pðkÞ e_ ijpðkÞ ¼ 0; @ rij @v @e ij
k ¼ 1; . . . ; n
ð29Þ
ij
Combining Eqs. (27)–(29), we have
ðkÞ þ @Y ðkÞ r_ ij @ vij ; @ 2 Gk2 @Y ðkÞ @Y ðkÞ pðkÞ pðkÞ þ pðkÞ ðkÞ
@Y ðkÞ @ rij
kðkÞ ¼
@Y ðkÞ ðkÞ @ vst @ est @ eij
@ eij
k ¼ 1; . . . ; n
ð30Þ
@ vij
Substituting Eqs. (24), (27) and (30) into Eq. (23), the multiple yield surface model, based on irreversible thermodynamics, can be derived as follows
_e _p _e _ e_ ij ¼ C ep ijmn rmn ¼ eij þ eij ¼ eij þ
n X
e_ ijpðkÞ
k¼1
2
@ G1 where C ep ijmn ¼ @ rij @ rmn
Pn
k¼1
0
ðkÞ
ðkÞ
ðkÞ
@Y ðkÞ @Y ðkÞ þ@Y ðkÞ @Y ðkÞ ðkÞ @ rmn ðkÞ ðkÞ @v @v @v mn ij ij @ 2 Gk ðkÞ ðkÞ @Y @Y @Y ðkÞ @Y ðkÞ 2 þ ðkÞ pðkÞ pðkÞ ðkÞ pðkÞ ðkÞ @e @v @e @v @e @v pq pq pq pq st st
ðkÞ
@ vst
pðkÞ
pðkÞ
@ est @ epq
ðkÞ
@ vpq
1
ðkÞ
@Y @Y þ @Y ðkÞ @YðkÞ ðkÞ n 2 X @ vij @ rmn @ vij @ vmn B @ G1 ¼@ @ rij @ rmn k¼1 @Y ðkÞ @ 2 Gk2 @Y ðkÞ þ @Y ðkÞ pðkÞ
@ epq
@Y ðkÞ ðkÞ @ vpq
C_ Armn
ð31Þ
is the elasto-plastic compliance tensor in the multiple yield surface model.
3.3. A systematic approach for deriving the yield criterion and flow rule from the dissipation function For the rate-independent materials, the existence proof for yield criterion and flow rule and their connections with dissipation function were proposed by Srinivasa (2010) applying the gauge function and convex analysis based on the maximum rate of dissipation criterion. However, the mathematical notions and tools involving convex analysis are usually unfamiliar to most geotechnical engineers and is inconvenient to applications. In this study, a new systematic approach, based on the properties of homogeneous functions, is proposed for solving the yield criterion and flow rule from the dissipation function. The new approach can be specified by three lemmas about homogeneous function as follows: Lemma 1. If f(x, y) is a homogeneous function of order n, then
@f @x
and
@f @y
is a homogeneous function of order n1.
@f @f Lemma 2. If f(x, y) is a first-order homogeneous function, then partial derivatives @x and @y can be expressed as the functions which depend on a single variable yx or yx.
Lemma 3. If f(x, y) is a first-order homogeneous function, then yx can be a function of two variables p and s by solving the equation ¼ ff10 ðy=xÞ y=x, here f1(y/x) = f(1, y/x), and f0 1(y/x) is a derivative of f1 with respect to independent variable yx. ðy=xÞ 1 Lemmas 1–3 can be extended to a homogeneous function with n independent variables. Then, in a six dimensional stress space (or generalized stress space), the yield criterion and flow rule from dissipation function can be derived by applying the approach defined by the Lemmas. First, the second-order symmetric tensors vij and a_ ij can be regarded as vectors with six dimensions. Thus, all of vectors ðv11 ; v12 ; v13 ; v22 ; v23 ; v33 Þ form a six dimensional vector space (generalized stress space). The dissipation function D is a first-order homogeneous function in ða_ 11 ; a_ 12 ; a_ 13 ; a_ 22 ; a_ 23 ; a_ 33 Þ, i.e., D ¼ Dðrij ; aij ; a_ 11 ; a_ 12 ; a_ 13 ; a_ 22 ; a_ 23 ; a_ 33 Þ, here rij and aij are regarded as given parameters. From orthogonality principle and Lemma 2, the generalized stress vij can be expressed as p s
@D a_ a_ a_ a_ a_ vij ¼ _ ¼ vij rij ; aij ; _ 12 ; _ 13 ; _ 22 ; _ 23 ; _ 33 ; i; j ¼ 1; 2; 3 @ aij a11 a11 a11 a11 a11
ð32Þ
In fact, Eq. (32) contains six equations and it shows that each component of the generalized stress is a function of five a_ 12 a_ 13 a_ 22 a_ 23 a_ 33 a_ 11 ; a_ 11 ; a_ 11 ; a_ 11 ; a_ 11 . Eliminating these five variables
independent variables, in which the set of independent variables is
between the six equations gives the yield criterion in six-dimensional generalized stress space:
Yðvij ; rij ; aij Þ ¼ 0 Eq. (33) defines a hyper-surface in generalized stress space, i.e., yield surface in generalized stress space.
ð33Þ
43
Y. Lai et al. / International Journal of Engineering Science 84 (2014) 29–53
Similarly, an extension of Lemma 3 for functions with six independent variables can define the flow rule in six-dimensional stress space. By the first-order homogeneity of D, one has
Dðrij ; aij ; a_ 11 ; a_ 12 ; a_ 13 ; a_ 22 ; a_ 23 ; a_ 33 Þ ¼ a_ 11 D1
a_ a_ a_ a_ a_ rij ; aij ; _ 12 ; _ 13 ; _ 22 ; _ 23 ; _ 33 a11 a11 a11 a11 a11
ð34Þ
a_ a_ a_ a_ a_ a_ a_ a_ a_ a_ rij ; aij ; _ 12 ; _ 13 ; _ 22 ; _ 23 ; _ 33 ¼ D rij ; aij ; 1; _ 12 ; _ 13 ; _ 22 ; _ 23 ; _ 33 . Thus, the orthogonality principle implies a11 a11 a11 a11 a11 a11 a11 a11 a11 a11 a_ 12 a_ 13 a_ 22 a_ 23 a_ 33 @D1 a_ 12 @D1 a_ 13 @D1 a_ 22 @D1 a_ 23 v11 ¼ D1 rij ; aij ; _ ; _ ; _ ; _ ; _ a11 a11 a11 a11 a11 @ða_ 12 =a_ 11 Þ a_ 11 @ða_ 13 =a_ 11 Þ a_ 11 @ða_ 22 =a_ 11 Þ a_ 11 @ða_ 23 =a_ 11 Þ a_ 11 @D1 a_ 33 ð35Þ @ða_ 33 =a_ 11 Þ a_ 11
where D1
vij ¼
@D1 ðwhere i and j cannotbeequalto1atthesametimeÞ @ða_ ij =a_ 11 Þ
ð36Þ
From the extension of Lemma 3 and Eqs. (35) and (36), five equations about the set of
a_ 12 a_ 13 a_ 22 a_ 23 a_ 33 a_ 11 ; a_ 11 ; a_ 11 ; a_ 11 ; a_ 11
are as follows
D1 ðrij ; aij ; a_ ij =a_ 11 Þ a_ 12 @D1 =@ða_ 13 =a_ 11 Þ a_ 13 @D1 =@ða_ 22 =a_ 11 Þ a_ 22 @D1 =@ða_ 23 =a_ 11 Þ a_ 23 @D1 =@ða_ 33 =a_ 11 Þ a_ 33 v11 ¼ @D1 =@ða_ 12 =a_ 11 Þ a_ 11 @D1 =@ða_ 12 =a_ 11 Þ a_ 11 @D1 =@ða_ 12 =a_ 11 Þ a_ 11 @D1 =@ða_ 12 =a_ 11 Þ a_ 11 @D1 =@ða_ 12 =a_ 11 Þ a_ 11 v12 ð37Þ D1 ðrij ; aij ; a_ ij =a_ 11 Þ @D1 =@ða_ 12 =a_ 11 Þ a_ 12 a_ 13 @D1 =@ða_ 22 =a_ 11 Þ a_ 22 @D1 =@ða_ 23 =a_ 11 Þ a_ 23 @D1 =@ða_ 33 =a_ 11 Þ a_ 33 v11 ¼ @D1 =@ða_ 13 =a_ 11 Þ @D1 =@ða_ 13 =a_ 11 Þ a_ 11 a_ 11 @D1 =@ða_ 13 =a_ 11 Þ a_ 11 @D1 =@ða_ 13 =a_ 11 Þ a_ 11 @D1 =@ða_ 13 =a_ 11 Þ a_ 11 v13 ð38Þ D1 ðrij ; aij ; a_ ij =a_ 11 Þ @D1 =@ða_ 12 =a_ 11 Þ a_ 12 @D1 =@ða_ 13 =a_ 11 Þ a_ 13 a_ 22 @D1 =@ða_ 23 =a_ 11 Þ a_ 23 @D1 =@ða_ 33 =a_ 11 Þ a_ 33 v11 ¼ @D1 =@ða_ 22 =a_ 11 Þ @D1 =@ða_ 22 =a_ 11 Þ a_ 11 @D1 =@ða_ 22 =a_ 11 Þ a_ 11 a_ 11 @D1 =@ða_ 22 =a_ 11 Þ a_ 11 @D1 =@ða_ 22 =a_ 11 Þ a_ 11 v22 ð39Þ D1 ðrij ; aij ; a_ ij =a_ 11 Þ @D1 =@ða_ 12 =a_ 11 Þ a_ 12 @D1 =@ða_ 13 =a_ 11 Þ a_ 13 @D1 =@ða_ 22 =a_ 11 Þ a_ 22 a_ 23 @D1 =@ða_ 33 =a_ 11 Þ a_ 33 v11 ¼ @D1 =@ða_ 23 =a_ 11 Þ @D1 =@ða_ 23 =a_ 11 Þ a_ 11 @D1 =@ða_ 23 =a_ 11 Þ a_ 11 @D1 =@ða_ 23 =a_ 11 Þ a_ 11 a_ 11 @D1 =@ða_ 23 =a_ 11 Þ a_ 11 v23 ð40Þ D1 ðrij ; aij ; a_ ij =a_ 11 Þ @D1 =@ða_ 12 =a_ 11 Þ a_ 12 @D1 =@ða_ 13 =a_ 11 Þ a_ 13 @D1 =@ða_ 22 =a_ 11 Þ a_ 22 @D1 =@ða_ 23 =a_ 11 Þ a_ 23 a_ 33 v11 ¼ @D1 =@ða_ 33 =a_ 11 Þ @D1 =@ða_ 33 =a_ 11 Þ a_ 11 @D1 =@ða_ 33 =a_ 11 Þ a_ 11 @D1 =@ða_ 33 =a_ 11 Þ a_ 11 @D1 =@ða_ 33 =a_ 11 Þ a_ 11 a_ 11 v33
a_ 12 a_ 13 a_ 22 a_ 23 a_ 33 a_ 11 ; a_ 11 ; a_ 11 ; a_ 11 ; a_ 11
ð41Þ
, while the right of Eqs. (37)–(41) are expresThe left sides of Eqs. (37)–(41) are expressions related to sions about components of generalized stress. Solving this set of equations gives,
a_ kl ¼ Lkl ðvij ; rij ; aij Þðwhere k and l cannot be equal to 1 at the same timeÞ a_ 11
ð42Þ
Eq. (42) defines the ratios between the components of internal variable rate, and naturally the flow rule in the generalized stress space. Example. Consider the model proposed by Collins and Hilder (2002) as an example to further illustrate the application of this new approach. The specific form of dissipation function in the model is
D¼
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 2 2 ðAe_ pv Þ þ Be_ pc
ð43Þ
where
A ¼ a1 p þ a2 q þ a3 pc B ¼ b1 p þ b2 q þ b3 pc
ð44Þ
ai and bi(i = 1,2,3) in Eq. (44) are the parameters related to materials, pc is normal consolidation pressure related to plastic volumetric strain. From orthogonality principle, the generalized mean stress vp and generalized equivalent stress vq can be obtained as follows:
@D
@D
vp ¼ _ p ; vq ¼ _ p @ ev @ ec
ð45Þ
44
Y. Lai et al. / International Journal of Engineering Science 84 (2014) 29–53
Substituting Eq. (43) into Eq. (45) gives
8 A2 e_ pv =e_ pc > ffi > v ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi > p > p p 2 < A2 þðBe_ v =e_ c Þ
ð46Þ
> B2 > ffi v ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi > > 2 : q ðAe_ p =e_ p Þ þB2 v
c
e_ p
e_ p
In Eq. (46), functions vp and vq depend on a single independent variable e_ vp . Eliminating e_ vp from these two functions gives c c the yield criterion in generalized stress space as
v2p A
2
þ
v2q B2
¼1
ð47Þ
In order to deduce the flow rule, Eq. (43) is rewritten as
D ¼ e_ pc D1 ¼ e_ pc
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ffi p p 2 Ae_ v =e_ c þ B2 ;
D1 ¼
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ffi p p 2 Ae_ v =e_ c þ B2
ð48Þ
Thus, the flow rule in generalized stress space can be obtained from Eq. (48) based on Lemma 3.
e_ pv vp B2 ¼ e_ pc vq A2
ð49Þ
Eqs. (47) and (49) are the yield criterion and flow rule derived by applying the new approach proposed by this paper, respectively. The results are similar to these of previous method. 4. Constitutive model of frozen loess The hyperplasticity theory (in Section 3) indicates that constitutive model of material can be derived from the energy potential function and the dissipation function. In this section, the thermodynamic potential function for frozen loess based on experimental results is derived. In this study, the constitutive model of frozen loess is restricted to small deformation hypothesis and isothermal conditions. 4.1. Gibbs free energy function and dissipation function of frozen loess The experimental results show that the direction of plastic strain increment of frozen loess is influenced by the stress path. Thus, this should be characterized by multiple internal variable hyperplasticity. Both shear mechanism and compressive mechanism have effect on plastic deformation of soils (Xie & Shao, 2006). The total plastic strain epij can be considered as pð1Þ the sum of two sub-plastic strains. One is related to shear mechanism eij while the other is related to compressive mech pð2Þ pð1Þ pð2Þ , namely epij ¼ eij þ eij . As a geotechnical material, frozen loess is also governed by the two mechanisms. It is anism eij assumed that the volumetric expansion of frozen loess is produced by shear mechanism and the volumetric contraction produced by compressive mechanism. According to multiple internal variable hyperplasticity, the Gibbs free energy function of frozen loess can be expressed as: pð1Þ
pð2Þ
Gðrij ; eij ; eij Þ ¼ G1 ðrij Þ þ
2 X
2 X pðkÞ Gk2 eij
k¼1
k¼1
rij eijpðkÞ þ
ð50Þ
Under the axisymmetric condition, Eq. (50) can be simplified by
1 pð1Þ pð2Þ pð2Þ pð1Þ pð2Þ pð2Þ pð1Þ pð1Þ pð2Þ ¼ G1 ðp; qÞ þ pepð1Þ þ G22 epð2Þ G p; q; epð1Þ v ; ec ; ev ; ec v þ qec þ pev þ qec þ G2 ev ; ec v ; ec pð1Þ
Here ev
¼e
pð1Þ a
þ 2e
pð1Þ 3
pð1Þ
ec
¼ 23
e
pð1Þ a
e
pð1Þ 3
ð51Þ
are plastic volumetric strain and plastic shear strain under shear mecha pð1Þ pð2Þ pð2Þ pð2Þ pð2Þ pð2Þ pð2Þ nism, respectively. It should be noted that the value of ev is negative. ev ¼ ea þ 2e3 and ec ¼ 23 ea e3 are plasand
pð2Þ
tic volumetric strain and plastic shear strain under compressive mechanism, respectively. The value of ev is positive. For frozen loess, G1(p, q) is corresponding to the elastic law, which can be determined from constant-confining pressure loading–unloading–reloading test and isotropic compression loading–unloading–reloading test. From Eqs. (1) and (3), the expression of G1(p, q) is as follows:
G1 ðp; qÞ ¼
ðpa þ pÞ lnð1 þ p=pa Þ p q2 þ K 2C
ð52Þ
Y. Lai et al. / International Journal of Engineering Science 84 (2014) 29–53
45
Considering the connection between thermodynamic functions and the formulations in conventional plasticity, pð1Þ pð2Þ pð2Þ and G22 ev ; ec for frozen loess can be determined from experimental results under the triaxial compression epð1Þ v ; ec
G12
test of constant-confining pressure and isotropic compression test.
G12
G22
pð1Þ epð1Þ ¼ ðf3t þ 1Þepð1Þ v ; ec v
! pð1Þ 2 2lec þ m pa mpa pð1Þ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi p ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi p arctan þ ln l epð1Þ m e þ n c c 2l l 4nl m2 4nl m2 pð2Þ
pð2Þ epð2Þ ¼ v ; ec
1 pa ev f3t pð2Þ 2 T 1 T 2 ev
ð53Þ
!
epð2Þ v
ð54Þ
where f3t is isotropic tensile strength, T1 and T2 are material parameters relating to volumetric hardening and l, m, n are material parameters depending on confining pressures. G12 and G22 are so called stored plastic work under shear mechanism and compressive mechanism, respectively. Similar to the stored plastic work, the dissipation function D for the frozen loess can be divided into two parts D1 and D2. 1 D and D2 are related to shear mechanism and compression mechanism, respectively. According to the experimental results of the triaxial compression tests with constant confining pressures and the isotropic compression test, the dissipation function D of frozen loess can be expressed as:
8 D ¼ D1 þ D2 > > sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiqffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi > > pð1Þ > p a ec > pð1Þ pð1Þ < D1 ¼ 2 e_ v e_ c pð1Þ 2 pð1Þ lðec Þ þmec þn > > rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi > 2 2 > pð2Þ > pð2Þ > : D2 ¼ 12 f3t þ pa ev pð2Þ e_ pð2Þ þ M2 e_ c v
ð55Þ
T 1 T 2 ev
where M is material parameter related to confining pressure. 4.2. Constitutive model based on thermodynamic functions When the Gibbs free energy function G and the dissipation function D are determined, the elasto-plastic incremental constitutive model for frozen loess can be derived. In the following, the parameters used in this model will be discussed in detail. 4.2.1. Elastic law Using the theories mentioned in Section 3, the rate of elastic strain can be determined by Eq. (52):
8 2 < e_ ev ¼ @ G21 p_ ¼ 1 p_ Kðpþp Þ @p a
ð56Þ
: e_ e ¼ @ 2 G1 q_ ¼ 1 q_ c C @q2
where the elastic parameters K and C are determined from constant-confining pressure triaxial compression, loading– unloading–reloading tests and isotropic compression loading-unloading–reloading tests. 4.2.2. Yield criterion, flow rule and hardening law According to the discussion in Section 3, the yield criterion, the flow rule and the hardening law of frozen loess can be determined by the functions of G12 ; D1 ; G22 and D2, respectively. First of all, the yield criterion, the flow rule and the hardening law of the frozen loess for shear mechanism are determined using G12 and D1, respectively. From the expression of D1 in Eq. (55), one has
8 sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffirffiffiffiffiffiffiffiffiffi pð1Þ > pa ec e_ pð1Þ > ð1Þ c @D1 > v ¼ ¼ > p pð1Þ > pð1Þ 2 pð1Þ @ e_ v e_ pð1Þ < lðec Þ þmec þn v sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffirffiffiffiffiffiffiffiffiffi > pð1Þ pð1Þ > 1 p a ec e_ v ð1Þ > > ¼ > vq ¼ @@D pð1Þ 2 pð1Þ : e_ pð1Þ e_ pð1Þ lðec Þ þmec þn c c
ð57Þ
ð1Þ
ð1Þ
The yield criterion, the flow rule in the generalized stress space (vp vq plane) are given by pð1Þ pa ec vpð1Þ vð1Þ 2 q ¼ pð1Þ pð1Þ l ec þ mec þ n
ð58Þ
vð1Þ e_ pð1Þ q v ¼ e_ pð1Þ vð1Þ p c
ð59Þ
46
Y. Lai et al. / International Journal of Engineering Science 84 (2014) 29–53
By Eq. (53), there is
8 @G12 ð1Þ > > ¼ ðf3t þ 1Þ < qp ¼ @ epð1Þ v @G12 ð1Þ > > ¼ : qq ¼ @ epð1Þ c
pð1Þ pa ec pð1Þ 2 pð1Þ
lðec
Þ
þmec
ð60Þ þn
Substituting Eq. (60) into Eq. (58), the yield criterion in true stress space (p–q plane) can be written as pð1Þ
pa ec Y ð1Þ ðp; q; epð1Þ ¼0 2 c Þ ¼ qðp þ f3t þ 1Þ ðp þ f3t Þ pð1Þ pð1Þ l ec þ mec þ n
ð61Þ
Combining Eqs. (58), (59) and (61) gives the flow rule in p–q plane by
e_ pð1Þ q v ¼ ðp þ f3t þ 1Þðp þ f3t Þ e_ pð1Þ c
ð62Þ
From Eq. (61), the hardening law under shear mechanism can be obtained as follows
h1
epð1Þ c epð1Þ ¼ 2 c pð1Þ pð1Þ l ec þ mec þ n
ð63Þ
Similarly, the yield criterion, the flow rule and the hardening law of the frozen loess under compressive mechanism can be pð2Þ pð2Þ determined using G22 and D2, respectively. Differentiating D2 with respect to e_ v and e_ c gives:
8 pð2Þ pa ev ð2Þ > @D2 1 1 > qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ffi v ¼ ¼ f þ > 3t p pð2Þ pð2Þ 2 > @ e_ v T 1 T 2 ev > pð2Þ pð2Þ 2 < 1þM2 ðe_ c =e_ v Þ pð2Þ > pa ev ð2Þ > @D2 1 M2 > qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi v ¼ ¼ f þ > 3t q pð2Þ pð2Þ 2 > @ e_ c T 1 T 2 ev : pð2Þ pð2Þ 2 ðe_ v =e_ c Þ þM2
ð64Þ
ð2Þ
ð2Þ
From Eq. (64), the yield criterion and the flow rule in generalized stress space (vp vq plane) are given by
ð2Þ p
v
2
þ
vð2Þ q
2
M2
pð2Þ
1 pa ev ¼ f3t þ pð2Þ 4 T 1 T 2 ev
!2 ð65Þ
ð2Þ M 2 vp e_ pð2Þ v ¼ pð2Þ ð2Þ e_ c vq
ð66Þ pð2Þ
Differentiating Eq. (54) with respect to ev
8 @G22 ð2Þ > > ¼ 12 < qp ¼ @ epð2Þ T v
> > : qqð2Þ ¼
@G22 pð2Þ
@ ec
pð2Þ pa ev pð2Þ 1 T 2 ev
and
epð2Þ gives: c
f3t ð67Þ
¼0
Substituting Eq. (67) into Eq. (65), the yield criterion in true stress space (p–q plane) can be written as
Y ð2Þ ðp; q; epð2Þ v Þ¼pþ
q2 2
M ðp þ f3t Þ
pð2Þ
pa ev
pð2Þ
T 1 T 2 ev
¼0
ð68Þ
Combining Eqs. (65), (66) and (68) yields the flow rule in true stress space (p–q plane) under compressive mechanism as follows 2 e_ pð2Þ M2 ðp þ f3t Þ q2 v ¼ pð2Þ 2ðp þ f3t Þq e_ c
ð69Þ
From Eq. (68), the hardening law under compressive mechanism can be obtained as follows:
h2 epð2Þ ¼ v
epð2Þ v pð2Þ T 1 T 2 ev
ð70Þ
4.2.3. The elasto-plastic incremental constitutive model for frozen loess From the thermodynamic functions G and D, the elastic law, yield criterion, flow rule, and hardening law (both of shear mechanism and compressive mechanism) for frozen loess, respectively, can be obtained. Then the elasto-plastic incremental
Y. Lai et al. / International Journal of Engineering Science 84 (2014) 29–53
47
constitutive model can be derived from these formulations. The rate of strain (strain increment for the rate-independent case) is:
(
e_ v e_ c
)
( ¼
)
e_ ev e_ ec
(
e_ pð1Þ v þ e_ pð1Þ c
)
(
e_ pð2Þ v þ e_ pð2Þ c
) ð71Þ
Eq. (56) implies
(
e_ ev e_ ec
)
" ¼
1 Kðpþpa Þ
0
#
p_ 1 q_
0
ð72Þ
C
From Eqs. (61)–(63), the sub-plastic strain rate is given by
(
where L ¼
2
)
e_ pð1Þ v e_ pð1Þ c
¼4 nlS2 2
ðlS2 þmSþnÞ
q2 ðpþf3t þ1Þðpþf3t Þ3 pa L
q ðpþf3t Þ2 pa L
q ðpþf3t Þ2 pa L
pþf3t þ1 ðpþf3t Þpa L
3
p_ 5 q_
ð73Þ
, and S is a function of p and q
ðp þ f3t Þpa m S¼ þ 2lqðp þ f3t þ 1Þ 2l
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 2 m ðp þ f3t Þpa n 2l 2lqðp þ f3t þ 1Þ l
From Eqs. (68)–(70), the following formulae can be obtained.
(
)
e_ pð2Þ v e_ pð2Þ c
2
M 2 ðpþf3t Þ2 q2
M2 pa Vðpþf3t Þ
h
M 2 ðp2 þpf3t Þþq2 1þT 2 M 2 pa ðpþf3t Þ
i2
3
2q M2 pa Vðpþf3t Þ
6 M2 p Vðpþf3t Þ2 ¼4 a 2q
4q2 2
2
M pa V ½M ðpþf3t Þ
2
q2
7 p_ 5 q_
ð74Þ
where V ¼ . T1 Substituting Eqs. (72)–(74) into Eq. (71) yields
(
e_ v e_ c
)
where C ep ¼
¼
C 11 C
C 11 C 21 C 12 C
p_ C 22 q_ C 12
ð75Þ
is the elastoplastic flexibility matrix. Its elements are given by
21 22 8 2 ðpþf3t Þ2 q2 q2 1 > C 11 ¼ Kðpþp þ ðpþf þ1Þðpþf þM > 3 Þ > M2 pa Vðpþf3t Þ2 a 3t 3t Þ pa L > < 2q C 12 ¼ C 21 ¼ ðpþfqÞ2 p L þ M2 p Vðpþf 3t Þ a 3t a > > > pþf3t þ1 4q2 > : C 22 ¼ C1 þ ðpþf þ 3t Þpa L M 2 pa V ½M2 ðpþf3t Þ2 q2
4.3. Determination of material parameters There are nine material parameters in the constitutive model of the frozen loess, namely, K, C, l, m, n, M, T1, T2 and f3t. According to previous investigations on frozen loess (Peng, 1998), the value of f3t is 1.721 MPa. The parameters K, C, l, m, n and M, depend on the confining pressures, and characterize the pressure crushing and melting phenomena of frozen loess. All of these parameters can be determined by triaxial compression tests under constant-confining pressures and isotropic compression tests. An exception is f3t which is obtained from isotropic tensile tests. In Section 2, the elastic parameters of K and C were determined using constant-confining pressure triaxial compression loading–unloading–reloading tests and isotropic compression loading–unloading–reloading tests, respectively. Using the elastic shear module C, the plastic shear strain in constant-confining pressure triaxial compression test can be separated from the total shear strain. And Eq. (61) implies pð1Þ
ec qðp þ f3t þ 1Þ ¼ 2 ðp þ f3t Þpa pð1Þ pð1Þ l ec þ mec þ n
ð76Þ
The values of l, m and n can be obtained by fitting experimental data to Eq. (76). The dependency of l, m and n on confining pressure is determined from the experimental results of constant-confining pressure triaxial compression test. Their expressions are given by
48
Y. Lai et al. / International Journal of Engineering Science 84 (2014) 29–53
Fig. 12. Comparisons between simulated results and test results under constant-confining triaxial compression conditions.
Y. Lai et al. / International Journal of Engineering Science 84 (2014) 29–53
Fig. 12 (continued)
49
50
Y. Lai et al. / International Journal of Engineering Science 84 (2014) 29–53
8 8 0:29464r3 =pa > < l ¼ 6:5741 10 þ 0:00381 e 0:067r3 =pa m ¼ 0:01292 e þ 3 105 r3 =pa þ 0:01426 > : n ¼ 0:01274 þ 0:03603 e0:04657r3 =pa
ð77Þ
Similarly, when the elastic volumetric module K is determined, the plastic volumetric strain of the frozen loess in isotropic compression tests can be separated from the total volumetric strain. The parameters T1 and T2 can be defined by fitting the experimental data to following equation
p0 =pa ¼
epð2Þ v pð2Þ T 1 T 2 ev
ð78Þ
where p0 is hydrostatic pressure in isotropic compressive condition. From the experimental results of isotropic compression tests and Eq. (78), T1 = 23.345 and T2 = 1.904 can be determined. In order to determine the parameter M, Eq. (68) can be rewritten as pð2Þ
p q2 ev þ 2 ¼ pa M pa ðp þ f3t Þ T 1 T 2 epð2Þ v
ð79Þ
Substituting the given values of T1, T2 into Eq. (79), the relationship between M and confining pressure is defined by:
M ¼ e0:02172r3 =pa þ
r3 =pa 0:32778r3 =pa þ 63:31
ð80Þ
4.4. Model verification In order to check the validity of the proposed constitutive model for frozen loess, a calculation program based on the proposed model Eq. (75) in this paper was formulated. Using the parameters given in Section 4.3, the simulations of constantconfining pressure triaxial compression tests and constant-slope stress path tests on frozen loess have been performed. All of the simulated results are shown in Figs. 12 and 13, respectively. Fig. 12 shows a comparison between test results and simulated results under the constant-confining triaxial compression condition. Fig. 13 shows the comparison of stress–strain curves between test results and simulated results under the constant-slope stress path condition. From Fig. 12, it can be seen that the proposed model, based on the hyperplasticity theory, has the capability of describing the mechanical behavior of frozen loess under various confining pressures. This indicates that the proposed model can describe the influences of confining pressure (or stress level) on the mechanical behavior of frozen soil, especially the pressure crushing and melting phenomena under high confining pressures. Furthermore, the volumetric contraction and dilation are also well simulated. Using the multiple internal variable hyperplasticity (corresponding to the multiple yield surfaces in conventional plasticity) to construct the constitutive model, the simulated results, shown in Fig. 13, illustrate that the proposed model can also predict the stress–strain curves under constant-slope stress path condition ideally. Therefore, the proposed model can be used to investigate the effects of stress paths on the mechanical behaviors of frozen loess.
Fig. 13. Comparisons between simulated results and test results under constant-slope stress path conditions.
Y. Lai et al. / International Journal of Engineering Science 84 (2014) 29–53
51
5. Conclusions In this paper, a series of tests, such as the constant-confining pressure triaxial compression tests (including loading– unloading–reloading), the isotropic compression tests (including loading–unloading–reloading) and the constant-slope stress path tests, were carried out on frozen loess to investigate the mechanical behavior under various loading conditions. Based on the test results, the mechanical characteristics for frozen loess are analyzed and summarized as follows: The mechanical behavior of frozen loess is influenced by the confining pressures. When the confining pressures are low (r3 < 2 MPa), the stress–strain curves show softening, which decreases with the increase in confining pressure. When the confining pressures are more than 2 MPa, the stress–strain curves show hardening. In addition, when the confining pressures are lower than 2 MPa, frozen loess experience volumetric expansion during the shear process, and an increase in confining pressure greatly reduces the dilatancy. The mechanical behavior of frozen soil under various confining pressures is investigated through the influence of confining pressures on the strength of frozen soil and its initial slope of stress–strain curve. It was found that the strength and the initial slope of stress–strain curve for frozen loess versus confining pressures can be divided into three phases. In the first phase, both of them increase with the increase in confining pressure followed by a decrease in the second phase. In the third phase when a large enough confining pressure is applied, the changes were relatively small. However, the influences of confining pressure on the strength and the initial slope of stress–strain curve are not the same. In the first stage, the influence of the confining pressure range on the strength is much smaller than that corresponding to initial slope. This enables the identification of the stress level when ice crushing happens in frozen soil using macroscopic tests. Using constant-confining pressure triaxial compression tests and constant-slope stress path tests, it is found that the plastic strain incremental direction of frozen loess is influenced by stress path. This characteristic of frozen loess cannot be described by a single yield surface or a single plastic potential. In order to construct a constitutive model that can appropriately describe the mechanical characteristics of frozen loess, following conclusions based on the hyperplasticity theory are made: Based on the properties of a homogeneous function, an approach to deriving the yield criterion and flow rule directly from orthogonality principle and first-order homogeneity of dissipation function is proposed, in which the mathematical tools with regard to Legendre transformation and convex analysis need not be introduced. A dissipation function and Gibbs free energy function with two internal variables are constructed. According to these two thermodynamic functions, an elasto-plastic incremental constitutive model for frozen loess is set up and a method of determining the material parameters in the model is also given. The stress–strain curves for frozen loess under constant-confining pressure triaxial compression and isotropic-slope stress path conditions are simulated by the constitutive model proposed in this study. The comparison of the simulated results with test results shows that the model can describe well the mechanical characteristics of frozen loess, such as ice crushing, pressure melting, dilatancy and the dependency of plastic strain direction on stress path.
Acknowledgements This research was supported by National key Basic Research Program of China (973 Program No. 2012CB026102), National Natural Science Foundation of China (41230630, 41301072), the Western Project Program of the Chinese Academy of Sciences (KZCX2-XB3-19), the Program of Higher-level talents of Inner Mongolia University (30105-125146), the foundation of State Key Laboratory of Frozen Soil Engineering (SKLFSE-ZY-03), and the Open Project Program of the State Key Laboratory of Frozen Soil Engineering (SKLFSE201208). The authors also thank Mr. Geoffrey Gay, a geotechnical engineer retired from Stuttgart University, for his kind help in revising the paper. References Anandarajah, A. (2008). Multi-mechanism anisotropic model for granular materials. International Journal of Plasticity, 24(5), 804–846. Azmatch, T. F., Sego, D. C., Arenson, L. U., & Biggar, K. W. (2011). Tensile strength and stress–strain behaviour of Devon silt under frozen fringe conditions. Cold Regions Science and Technology, 68(1-2), 85–90. Bragg, R. A., & Andersland, O. B. (1981). Strain rate, temperature, and specimen size effects on compression and tensile properties of frozen soil. Engineering Geology, 18(12), 35–46. Calladine, C. R. (1963). Correspondence on the yielding of clay. Geotechnique, 13(3), 250–255. Calladine, C. R. (1971). A microstructural view of the mechanical properties of saturated clay. Geotechnique, 21(4), 391–415. Chamberlain, E., Groves, C., & Perham, R. (1972). The mechanical behavior of frozen earth materials under high pressure triaxial test conditions. Geotechnique, 22(3), 469–483. Chen, W. F., & Saleeb, A. F. (2001). Constitutive equations for engineering materials. Wuhan: Huazhong University of Science and Technology Press [translated by Yu, T.Q., Wang, X.W., Liu, Z.H.]. Chiarelli, A. S., Shao, J. F., & Hoteit, N. (2003). Modeling of elastoplastic damage behavior of a claystone. International Journal of Plasticity, 19(1), 23–45. Cleja-Tigoiu, S. (1990a). Large elasto-plastic deformations of materials with relaxed configurations-I. Constitutive assumptions. International Journal of Engineering Science, 28(3), 171–180.
52
Y. Lai et al. / International Journal of Engineering Science 84 (2014) 29–53
Cleja-Tigoiu, S. (2003). Consequences of the dissipative restrictions in finite anisotropic elasto-plasticity. International Journal of Plasticity, 19(11), 1917–1964. Cleja-Tigoiu, S., Cazacu, O., & Tigoiu, V. (2008). Dynamic expansion of a spherical cavity within a rate-dependent compressible porous material. International Journal of Plasticity, 24(5), 775–803. Cleja-Tigoiu, S. (1990b). Large elasto-plastic deformations of materials with relaxed configurations-II. Role of the complementary plastic factor. International Journal of Engineering Science, 28(4), 273–284. Cleja-Tigoiu, S. (2000). Nonlinear elasto-plastic deformations of transversely isotropic material and plastic spin. International Journal of Engineering Science, 38(7), 737–763. Collins, I. F. (2003). A systematic procedure for construction critical models in three dimensions. International Journal of Solids and Structures, 40(17), 4379–4397. Collins, I. F. (2005). Elastic/Plastic models for soils and sands. International Journal of Mechanical Sciences, 47(4-5), 493–508. Collins, I. F., & Hilder, T. (2002). A theoretical framework for constructing elastic/plastic constitutive models of triaxial tests. Numer. International Journal for Numerical and Analytical Methods in Geomechanics, 26(13), 1313–1347. Collins, I. F., & Houlsby, G. T. (1997). Application of thermomechanical principles to the modeling of geotechnical materials. Proceeding of the Royal Society AMathematical Physical and Engineering Sciences, 453(1964), 1975–2001. Collins, I. F., & Kelly, P. A. (2002b). A thermomechanical analysis of a family of soil models. Geotechnique, 52(7), 507–518. Desai, C. S. (1980). A general basis for yield, failure and potential functions in plasticity. International Journal for Numerical and Analytical Methods in Geomechanics, 4(4), 361–375. Drucker, D. C., Gibson, R. E., & Henkel, D. J. (1957). Soil mechanics and work-hardening theories of plasticity. Transactions of the American Society of Civil Engineers, 122(1), 338–346. Fish, A.M. (1991). Strength of frozen soil under a combined stress state. In Proceedings of the sixth international symposium on ground freezing, Beijing (Vol. 1, pp. 135–145). Gens, A., & Nova, R. (1993). Conceptual bases for a constitutive model for bonded soils and weak rocks. In Proceedings of the symposium on geotechnical engineering of hard soils soft rocks, Balkema (pp. 485–493). Hambly, E. C. (1972). Plane strain behaviour of remoulded normally consolidated kaolin. Geotechnique, 22(2), 301–317. Hashiguchi, K. (1980). Constitutive equations of elastoplastic materials with elastic–plastic transition. Journal of Applied Mechanics, 47(2), 266–272. Hashiguchi, K. (2005). Generalized plastic flow rule. International Journal of Plasticity, 21(2), 321–351. Hashiguchi, K., Ozaki, S., & Okayasu, T. (2005). Unconventional friction theory based on the subloading surface concept. International Journal of Solids and Structures, 42(5-6), 1705–1727. Hashiguchi, K., & Protasov, A. (2004). Localized necking analysis by the subloading surface model with tangential-strain rate and anisotropy. International Journal of Plasticity, 20(10), 1909–1930. Hashiguchi, K., Saitoh, K., Okayasu, T., & Tsutsumi, S. (2002). Evaluation of typical conventional and unconventional plasticity models for prediction of softening behavior of soils. Geotechnique, 52(8), 561–578. Hashiguchi, K., & Tsutsumi, S. (2001). Elastoplastic constitutive equation with tangential stress rate effect. International Journal of Plasticity, 17(1), 117–145. Hashiguchi, K., & Tsutsumi, S. (2007). Gradient plasticity with the tangential subloading surface model and the prediction of shear band thickness of granular materials. International Journal of Plasticity, 23(5), 767–797. Houlsby, G.T. (1981). A study of plasticity theories and their applicability to soils (Ph.D. thesis). University of Cambridge. Houlsby, G.T. (1982). A derivation of the small-strain incremental theory of plasticity form thermodynamics. In Proceedings of IUTAM symposium on deformation and failure of granular materials, Delft (pp. 109–118). Houlsby, G. T., & Puzrin, A. M. (2000). A thermomechanical framework for constitutive models for rate-independent dissipative materials. International Journal of Plasticity, 16(9), 1017–1047. Huang, W. X., Pu, J. L., & Chen, Y. J. (1981). Hardening rule and yield function for soils. Chinese Journal of Geotechnical Engineering, 3(3), 19–26. Lade, P. V., & Duncan, J. M. (1976). Stress-path dependent behavior of cohesionless soil. Proceedings of ASCE, 102(GT1), 51–68. Lade, P. V., & Duncan, J. M. (1975). Elasto-plastic stress–strain theory for cohesionless soil. Proceedings of ASCE, 101(GT10), 1037–1053. Lade, P. V., & Duncan, J. M. (1977). Elasto-plastic stress–strain theory for cohesionless soils with curved yield surface. International Journal of Solids and Structures, 13(11), 1019–1035. Lai, Y. M., Long, J., & Chang, X. X. (2009). Yield criterion and elasto-plastic damage constitutive model for frozen sandy soil. International Journal of Plasticity, 25(6), 1177–1205. Lai, Y. M., Xu, X. T., Dong, Y. H., & Li, S. Y. (2013). Present situation and prospect of mechanical research on frozen soils in China. Cold Regions Science and Technology, 87(MAR), 6–18. Lai, Y. M., Yang, Y. G., Chang, X. X., & Li, S. Y. (2010). Strength criterion and elastoplastic constitutive model of frozen silt in generalized plastic mechanics. International Journal of Plasticity, 26(10). 1641–1484. Lai, Y. M., Zhang, M. Y., & Li, S. Y. (2009). Theory and application of cold regions engineering. Beijing: Science Press. Lewin, P. I., & Burland, J. B. (1970). Stress probe experiments on saturated normally consolidated clay. Geotechnique, 20(1), 38–56. Li, G. X. (2004). Advanced soil mechanics. Beijing: Tsinghua University Press. Li, R. S. (1986). The irreversible thermodynamics and the dissipative structure. Beijing: Tsinghua University Press. Maugin, G. A. (1992). The thermomechanics of plasticity and fracture. Cambridge: Cambridge University Press. Maugin, G. A. (1999). The thermomechanics of nonlinear irreversible behaviours. Singapore: World Scientific. Ottosen, N. S., & Ristinmaa, M. (2013). Thermodynamically based fictitious crack/interface model for general normal and shear loading. International Journal of Solids and Structures, 50(22-23), 3555–3561. Parameswaran, V. R., & Jones, S. J. (1981). Triaxial testing of frozen sand. Journal of Glaciology, 27(95), 147–155. Peng, W. W. (1998). Tensile strength of frozen loess varying with strain rate and temperature. Chinese Journal of Geotechnical Engineering, 20(3), 31–33. People’s Republic of China National Standard GB/T 50123-1999 (1999). Standard for soil test method. Beijing: China Planning Press. Peric, D., & Ayari, M. A. (2002). On the analytical solutions for the three-invariant Cam-clay model. International Journal of Plasticity, 18(8), 1061–1082. Puglisi, G., & Truskinovsky, L. (2005). Thermodynamics of rate-independent plasticity. Journal of The Mechanics and Physics of Solids, 53(3), 655–679. Puzrin, A. M., & Houlsby, G. T. (2001a). Fundamentals of kinematic hardening hyperplasticity. International Journal of Solids and Structures, 38(21), 3771–3794. Puzrin, A. M., & Houlsby, G. T. (2001b). A thermomechanical framework for rate-independent dissipative materials with internal functions. International Journal of Plasticity, 17(8), 1147–1165. Ragione, L. L., Prantil, V. C., & Sharma, I. (2008). A simplified model for inelastic behavior of an idealized granular material. International Journal of Plasticity, 24(1), 168–189. Rajagopal, K. R., & Srinivasa, A. R. (2005). On the nature of constraints for continua undergoing dissipative processes. Proceeding of the Royal Society A: Mathematical, Physical & Engineering Science, 461(2061), 2785–2795. Rosakis, P., Rosakis, A. J., Ravichandran, G., & Hodowany, J. (2000). A thermodynamical internal variable model for the partition of plastic into heat and stored energy in metals. Journal of the Mechanics and Physics of Solids, 48(3), 581–607. Roscoe, K. H., & Burland, J. B. (1968). On the generalized stress–strain behaviour of wet clay. Engineering plasticity. Cambridge: Cambridge University Press. Roscoe, K. H., & Poorooshasb, H. B. (1963). A theoretical and experimental study of strains in triaxial compression tests on normally consolidated clays. Geotechnique, 13(1), 12–28. Rousselier, G. (2001). Dissipation in porous metal plasticity and ductile fracture. Journal of the Mechanics and Physics of Solids, 49(8), 1727–1746.
Y. Lai et al. / International Journal of Engineering Science 84 (2014) 29–53
53
Seiki, O. (1979). A mechanical model for the stress–strain behaviour of normally consolidated cohesive soil. Soils and Foundations, 19(3), 29–44. Shen, W. Q., Shao, J. F., Kondo, D., & Gatmiri, B. (2012). A micro–macro model for clayey rocks with a plastic compressible porous matrix. International Journal of Plasticity, 36(SEP), 64–85. Shen, Z. J. (1980). The rational form of stress–strain relationship of soils based on elasto-plasticity theory. Chinese Journal of Geotechnical Engineering, 2(2), 11–19. Srinivasa, A. R. (2010). Application of the maximum rate of dissipation criterion to dilatant, pressure dependent plasticity models. International Journal of Engineering Science, 48(11), 1590–1603. Tatsuoka, F., & Ishihara, K. (1974). Yielding of sand in triaxial compression. Soils and Foundations, 14(2), 63–76. Tsutsumi, S., & Hashiguchi, K. (2005). General non-proportional loading behavior of soils. International Journal of Plasticity, 21(10), 1941–1969. Tsutsumi, S., Hashiguchi, K., Okayasu, T., Saitoh, K., & Sugimoto, M. (2001). Mechanical response of subloading surface model with tangential plasticity. Journal of Applied Mechanics, 4(8), 375–382. Vorobiev, O. (2008). Generic strength model for dry jointed rock masses. International Journal of Plasticity, 24(12), 2221–2247. Xie, S. Y., & Shao, J. F. (2006). Elastoplastic deformation of a porous rock and water interaction. International Journal of Plasticity, 22(12), 2195–2225. Xie, S. Y., & Shao, J. F. (2012). Experimental investigation and poroplastic modelling of saturated porous geomaterials. International Journal of Plasticity, 39(DEC), 27–45. Xu, X. T., Lai, Y. M., Dong, Y. H., & Qi, J. L. (2011). Laboratory investigation on strength and deformation characteristics of ice-saturated frozen sandy soil. Cold Regions Science and Technology, 69(1), 98–104. Xu, X. Z., Wang, J. C., & Zhang, L. X. (2010). Physics of frozen soils. Beijing: Science Press. Yin, Z. Z., & Duncan, J. M. (1984). A stress–strain relationship for dilative and non-dilative soils. Chinese Journal of Geotechnical Engineering, 6(4), 24–40. Zhang, S. J., Lai, Y. M., Sun, Z. Z., & Gao, Z. H. (2007). Volumetric strain and strength behavior of frozen soils under confinement. Cold Regions Science and Technology, 47(3), 263–270. Zheng, Y. R., Shen, Z. J., & Gong, X. N. (2002). Generalized plastic mechanics. Beijing: China Architecture and Building Press. Zhu, Q. Z., Shao, J. F., & Mainguy, M. (2010). A micromechanics–based elastoplastic damage model for granular materials at low confining pressure. International Journal of Plasticity, 26(4), 586–602. Ziegler, H. (1958). An attempt to generalize Onsager’s principle and its significance for rheological problems. Zeitschrift fur Angewandte Mathematik und Physik, 9(b), 748–763. Ziegler, H. (1963). Some extremum principles in irreversible thermodynamics with application to continuum mechanics. Progress in Solid Mechanics, 4(91), 92–193. Ziegler, H. (1977). An introduction to thermomechanics. Amsterdam: North-Holland.