You are on page 1of 23

Results of the International Workshop on Constitutive Relations for Soils /Grenoble /6-8 September 1982

Predictions of the results of laboratory tests on a clay using a critical state model
Oxford University, UK
D.M.WOOD G.T.HOULSBY & C.P.WROTH

3.1

Cambridge University, UK

ABSTRACT: An elastic-plastic model for describing the stress-strain behaviour of a clay is presented. The model is based on the Modified Cam-Clay model, but includes an additional yield surface to model failure of overconsolidated samples on a Hvorslev surface. The model is used to predict the behaviour of a kaolinite in a number of repeated loading tests, based on the results of monotonic triaxial compression and extension tests. Special account is taken in the model of the possibility of the clay being unsaturated. 1 INTRODUCTION The authors were invited to participate in the International Workshop on Constitutive Behaviour of Soils, to be held at Grenoble in September 1982. The purpose of the Workshop was to allow a critical comparison to be made of various mathematical models of mechanical behaviour of soils. Carefully controlled laboratory tests on both reconstituted clay and sand were conducted at the University of Karlsruhe; some of the test results were -provided to the participants as "input" tests for fitting of their models and establishment of soil properties. The remaining tests ("output" tests) are those for which the participants were expected to make predictions, which had to be compared in detail with the experimental data at the Workshop. In view of the proven success of the concepts of Critical State Soil Mechanics, and in particular the family of Cam-Clay models, to describe the data from high quality tests on reconstituted clays, it was decided to accept the invitation and to use these concepts as the basis of the predictions called for by the organisers of the Workshop. A similar exercise took place at the Symposium on Plasticity and Generalised Stress-Strain Behaviour of Soils, held at McGill University, Montreal, in May 1980. The predictions made on that occasion, and a full description of the assumptions made and methods used, were given by Wroth & Houlsby (1980). In that event the Modified Cam-Clay model was used for predicting the

behaviour of two natural clays and of reconstituted kaolin. It is well recognised that one deficiency of this model is that it leads to overestimates of the strengths of heavily overconsolidated clays which are better described by the Hvorslev failure criterion. Consequently, for this Workshop, the Modified Cam-Clay Model was adapted to incorporate the Hvorslev failure criterion. In order to avoid possible confusion and misunderstanding this new model is called the Roscoe-Hvorslev model. It is fully described in a later section of this paper. A further minor variation of the model had to be introduced to allow for the consequences of the samples of clay not being fully saturated, and being tested without back pressure. Initial perusal of the data of the input tests suggested that this had occurred; subsequently, it was found that the data could only be fitted well by the Roscoe-Hvorslev model (RHM) if allowance was made for incomplete saturation. The method adopted to allow for partial saturation is described later. Because it was considered that none of the existing family of Critical State models adequately represents the behaviour of sand, no predictions for the output tests on sand have been attempted. 2 IMPORTANT FEATURES OF SOIL MODELS The role of a mathematical model for the

99

description of soil behaviour is twofold: firstly, as a simple quantitative framework against which soil behaviour may be assessed and, secondly, as a complete qualitative model for analysis and design. In its first role the model is of use both in judging the quality and consistency of data, and in predicting the character and trends of behaviour of a soil under a variety of circumstances. For this purpose the model must represent the essential features of soil behaviour in as simple a manner as possible. A simple model is also desirable for the second role; this is because, in a numerical calculation (such as one carried out by means of the finite element method), an excessively complex model may be difficult to implement and expensive to use. Very often the quality of the available data about the actual soil from the field will not warrant the additional precision of calculation achieved using a complex model. If, however, the model is to be used for mathematical calculations, it is important that it should be properly founded in the theory of continuum mechan ics and should be internally consistent. A simple model will require few parameters for the description of a soil, and each of these should have a phys ical significance so that an assessment can easily be made of its importance and the likely results of any changes in its value. The parameters should be measurable directly in a small number of simple tests Both the purpose and the limitations of the model should be well understood. The use of one model should lead to an understanding of its range of applicability and of the types of problem for which it may be useful. The trends of behaviour predicted by the model under extreme conditions should be explored and their importance assessed. Some essential features of soil behaviour under conditions of monotonic loading which should be included in a model are listed below. The importance lies not in the fact that these features should be included but that the model should reproduce well-established experimentally determined behaviour. 1. A two (or more) phase material (material properties must be expressed in terms of 2. Both non-linear response and irre coverable 3. Plastic dilation or compression 4. Failure conditions (e.g. MohrCoulomb or Hvorslev) 5. The influence of consolidation history 6. Time effects (both primary consoli-

dation and time dependent behaviour of the soil skeleton) 7. Anisotropy 8. Certain well-established empirical relationships which should be reproduced by the model, e.g. the ratio of undrained strength cu to overburden i av may be expressed as a function of overconsolidation All the above features are included in below, except that: the model 1. Only the time effects due to primary consolidation are considered (i.e. those which arise through the flow of water within the soil skeleton), whereas creep, secondary consolidation, aging and cementation are excluded; thus the behaviour of the soil skeleton is rate and time independent ; 2. The model is isotropic, so that certain effects due to the reversal or rotation of the principal stress axes are not included. The Roscoe-Hvorslev model was chosen for use in the prediction workshop since it is a development of a well-established and understood model (the Modified Cam-Clay model) which the general require ments of simplicity and completeness, which for the economic solution of real problems. The new model is suitable for the analysis of tests on both normally consolidated and highly overconsolidated clays, but is not thought to be suitable for sands. 3 THE ROSCOE-HVORSLEV MODEL The name of the model is taken from the two separate surfaces, the Roscoe surface and the Hvorslev surface in (q,p*,V) space by Atkinson & Bransby (1978). The Roscoe-Hvorslev model (referred to hereafter as RHM) consists of the Modified Cam-Clay model (MCCM) (Roscoe & Burland 1968) with the additional feature of the Hvorslev failure criterion. The latter is case of the better-known Mohr a Coulomb failure criterion in which the cohesive component of strength is not constant, but is an exponential function of the current water content of the clay. The RHM has three separate and distinct ingredients: (i) the Modified Cam-Clay model, (ii) the Hvorslev failure criterion and (iii) the simulation of partial satura tion. It is described in this sequence in the three following sections of the paper, 3.1 The Modified Cam-Clay model The theory of perfect plasticity is founded

100

on the hypothesis that, during plastic deformation of a material, the strain increments are functions of the absolute stresses. The theory evolved from careful observations of the behaviour of ductile metals whose response to loading is typified by an initial phase of elastic, recoverable deformation until a yield condition is reached; after yield, a second phase of behaviour occurs which is a combination of elastic and plastic irrecoverable deformation. In general, the plastic deformations are greater by an order of magnitude than the accompanying elastic ones. The study of a typical stress-strain curve for a soil in which the stress is cycled one or more times, for example, a one-dimensional consolidation test on a clay, reveals qualitatively similar behaviour. However, there is one essential difference and that is that soils, as opposed to metals, experience irrecoverable volumetric as well as irrecoverable shear strains. The classical theory of plasticity developed for describing the large strain response of ductile metals, has had to be extended to account for plastic volumetric strains in order to become applicable to soils. Making use of this analogy with metal behaviour, and coupled with the experimental evidence for a critical state for deforming soil, a family of elasto-plastic stress-strain models has been developed at Cambridge. One such model, known as Modified Cam-Clay, initially proposed by Burland (1967) has been used extensively in finite element computations. A qualitative description of this model follows, and the mathematical details are provided in Appendix A. In detail, the typical behaviour of a soil specimen in a triaxial compression test can be characterised by the curve of

Fig.1(a). If the specimen is loaded along the path OA and then unloaded along ABC, the path ABC can be approximated as elastic behaviour, with the strain represented by OC being a permanent, irrecoverable plastic strain denoted by eP. On reloading, the specimen behaves essentially elastically along CDE (displaying a small amount of hysteresis) until at point E it experiences the previous maximum deviator stress; it yields and undergoes further plastic strain. Point F denotes failure (a unique condition for this particular test); failure must be distinguished from yield which is a progressive phenomenon and which may occur at any point along the primary loading curve OAEF{depending upon the exact loading sequence followed in the test. Analogous behaviour will be displayed by a soil specimen if tested in a consolidometer, as shown in Fig.1(b), which is an unconventional plot of effective pressure against specific volume (V) (V=l+e, where e is the voids ratio). If this diagram is rotated clockwise through 90 to give Fig.1(c), the more usual plot of consolidation is obtained. . It should be realised that for a specimen that is in the overconsolidated state, represented by point C in Fig.1(c), its preconsolidation pressure p^, given by point A, is the current yield stress for further consolidation; it is the pressure at which further plastic volumetric strains begin to occur. Now consider a soil specimen that has been isotropically normally consolidated (with o[ = c ? 2 = O3) to point A in Fig.2 (a) , and then allowed to swell isotropically to some state along the unloading curve ABC. If the specimen were then subjected to a variety of different effective stress paths, it is assumed that there would be a well-defined region of stress states for

(a) Behaviour in shear

(b)

(c)

Behaviour in consolidation

Fig.l. Typical behaviour of soil specimen in consolidation and shear tests

101

which the specimen would remain elastic (The stress used for representing the data of tests are the mean normal stress p' = ^(cr^+a^+ap and deviator stress q = c'-a'. The corresponding strain variables are the volumetric strain v l+e2+e3 and the deviatoric 2 e-).) This region is bounded by a (e 1 3 yield curve. If the stresses applied to the specimen take state outside the current yield curve, it will yield and experience both plastic volumetric and plastic shear Consider a second specimen that has been normally consolidated to G and then allowed to swell to some point on the unloading curve GHI. Associated with this specimen is a larger yield curve, Fig.2(b), but one that has in practice approximately the same shape as that for the first specimen. The sizes of the yield curves are dictated by the points A and G, which lie on the normal isotropic consolidation curve. The choice of yield curve appropriate for any specimen depends on the maximum consolidation pressure (i.e. the pre-consolidation

Direction of plastic increment

P',v
(b)

Fig.2. Yield curves for specimens with consolidation

The shape of the yield curve, Fig.3(a), is assumed to be elliptical; this choice is based on considerations of energy pated plastically within the specimen (Roscoe & Burland 1968; Houlsby 1981). One semi-axis of the ellipse BA is fixed by the consolidation history relevant to the specimen The other semi-axis BX , is given by the assumption that the point X, is on f the line of critical states q Mp of failure states of normally consolidated specimens The critical state line is assumed to be parallel to the normal consolidation line in the logarithmic plot of Fig.3(b).

Failure line (Critical states)

nv

Failure line (Critical states) Normal Consolidation


*

Yield surface

v=r

Gradient-X

Rc
(a)

p ' = I
(b)

Fig.3. Details of yield surface for Modified Cam-Clay model


X

102

As shown in Fig.3(b), both consolidation and swelling lines are considered as 1 straight in jlnV-Hnp space. This represents a very minor alteration of the model from previously published descriptions of Modified Cam-Clay, for which the lines were taken as straight in V - np' space. Consequences of the changes are: 1. The numerical values of the parameters A* and K* (see below) are reduced by a factor approximately equal to V from the original parameters A and k; 2. The bulk modulus becomes truly proportional to pressure rather than depending volume (see Appendix A); also upon 3. Evidence in support of this change is given by Butterfield (1979), and a further advantage is a slight simplification of the mathematics of the model. For a specimen undergoing small the new version of the model is identical to the old one The complete description of the model requires five parameters to specify the shape and size of the yield locus for a soil specimen at a given pressure and volume, as well as the properties of the material. The parameters 1. X*, the gradient of the consolidation line in nV-np' space. This is related directly to the conventional compression index Cc' for a remoulded material by A * = CJ/(2.303V) ; c 2. K* the gradient of the swelling line, similarly related to the swelling index * Cs' by k / 3. M, the gradient of the critical state i This is related to the line in q P angle of shearing resistance < p ' cv of a normally consolidated clay by sin ( J ) ' M 6 sin d > ' / (3 cv cv) 7 4. G, the elastic shear modulus, taken to be constant;
5. r.

incremental bulk modulus, derived from the local gradient of the swelling line, can be written K
P'/K
*

. (1)

The by assuming a constant shear modulus, G, since any assumption of variation of G with pressure can result in a model which is thermodynamically unacceptable. Nov*, suppose the state of stress experienced by tHe specimen is at point J in Fig.2(b) close to the relevant yield curve. If the stress increment JKL is applied to the specimen, the increment JK will cause increments only, whereas the increment KL will cause both and plastic strain increments. As the specimen yields at K, the yield curve is expanded as the specimen undergoes consolidation. The plastic volumetric strain increment that occurs from K to L is given by a hardening law,
P

(A

K*)

P C' P '

. . .(2)

The last parameter T is required solely to locate the critical state line or the isotropic normal consolidation line on the in Fig.3(b); it is the voids critical state line for which p' has unit value. It is in some ways analogous to the liquid limit (see Schofield & Wroth 1968) and increases markedly with the plasticity index of the clay being considered. The elastic behaviour of the specimen for stress states within the current yield locus is assumed to be isotropic, but with the bulk modulus directly dependent upon the current mean stress p'. This latter feature is a consequence of the volumetric with the swelling and recompression lines AX]^ and GX2 in Fig.3(b) Simple calculation shows that for a specimen experiencing a pressure p' the

Applied (total) stress path

(b) Fig.4. Stress paths for conventional undrained triaxial compression test
i

103

which is derived from the normal consoli dation line, An(V/V ) X* n(p '/p'). . (3) (The dot notation of plasticity theory has been adopted to indicate a small (timeindependent) increment.) The details of this derivation are not given here but can be found in Schofield & Wroth (1968). The important point is that no additional soil parameter is required. of the The flow rule governing the plastic strain increments is given by the condition of normality. This condition stipulates that if the associated strain increments vP and e^ are plotted on the same axes as the stresses in Fig.2(b), the vector KM of plastic strain increment is normal to the yield curve at K. The gradient of the curve at K is known from its elliptical shape, so that the ratio can be calculated and, since vP is already established (from the hardening law), eP can be evaluated. Consider now- a specimen undergoing a conventional undrained triaxial compression , with initial state as shown by point A in Fig.4. The condition of the test is such that the total volume must stay constant, i.e. that the path in Fig.4(a) coincides with a line of constant voids ratio, that is, constant volume. The initial response of the specimen must be elastic as the point A within the current yield curve, Fig.4(b). There can be no change of p' during mation as the point A in Fig.4(a) is con strained to lie on both the elastic swel ling line and the constant voids ratio stress path line. When the reaches the point B, the specimen yields. Beyond this point, both plastic and elastic volumetric strains occur, of equal magnitude but opposite sign so as to maintain the total volume constant. The ratio of the plastic volume change to the plastic shear strain is calculated as , and the test proceeds to failure at point D. The effective stress path ABCD is as shown in Fig.4(b). The total stress path for a triaxial compression test (with constant cell pressure) is given by AE. The pore pressure at any stage of the test is simply the difference in mean stress between the two curves, as shown in Fig.4(b). This model successfully reproduces the major deformation characteristics of soft clay, and is expressed in terms of effective stress allowing prediction to be made of pore pressures in undrained tests.

3.2 Hvorslev's failure criterion

The out by Hvorslev in the mid-1930's consisted of stress con shear tests on two recon trolled stituted clays. His results and findings were published in 1937 in his doctoral in German, and published in English many years later (Hvorslev 1969). His failure criterion has been confirmed since for strain controlled triaxial tests on clays by other workers, e.g. presented by Parry (1960), as interpreted by Schofield & Wroth (1968). The criterion will be discussed in this paper in terms of and is presented diagrammatically in Fig.5. In Figs 5(a) and 5(b), the conditions at failure are plotted in terms of the varip'f and Vf. Hvorslev showed that < J f for a set of specimens which failed at the same value of water content (i.e. the same volume for a saturated clay) the locus of failure points was a straight line such as A1B1C1 or A 2 B 2 C 2 in Fig.5(a). Further, he showed that the end points of lines, C , C2 lay on a straight line passing through the origin, which is now recognised as the Critical State Line. i i He defined the "equivalent pressure p'e as the pressure on the normal consolidation line corresponding to the current water content of the specimen in question (point Ej in Fig.5(b) for states B. and C ). By normalising the results by means of the equivalent pressure he pro duced the simple unique failure criterion illustrated in Fig.5(c). This failure criterion is a and subtle version of the widely-used Mohr Coulomb failure criterion. The difference is that the cohesive component of strength is seen to be a simple explicit function (exponential) of the current water content and not a constant as usually assumed. The elliptical yield loci of the MCCM are associated with swelling curves as shown in 5(d) and 5(e). For a set of specimens which fail so that their speci along a swelling curve volumes Vf such as DFG in Fig.5(e) the locus of their failure states (qf,p'f) will be the curved line shown in Fig.5(d) which lies between the critical state line and the elliptical yield locus. The surface in (q ,p',V) space containing the states A,B,C,D,F,G is known as the Hvorslev surface, and that formed by the right-hand part of the elliptical yield locus GHI is termed the

104

ql
f

Hvorslev surface

Roscoe surface

(a)

(d)

(b)

(e)

/ /

(C)

e
Fig.5. Hvorslev's failure criterion

(f)

/ c

The gently curving line DFG has a relatively complex algebraic form, and as a first attempt to incorporate the Hvorslev criterion into the family of Cam-Clay models, it has been assumed for simplicity to be straight. Furthermore,

for convenience, normalisation has been done by means of the isotropic preconsoli dation pressure p' c rather than by the equivalent (one-dimensional) pressure p'e' which affects the arithmetic but not the principle involved. The result of these

105

assumptions is the reduced form of Fig.5(f). Because Hvorslev's original tests were stress-controlled he studied conditions at failure and he could not study post-peak behaviour. This is now possible with strain-controlled triaxial tests. Suppose a drained test with mean normal effective 1 stress p kept constant is conducted on an overconsolidated specimen with initial state at L in Fig.6. The response of the specimen is expected to be of the form given in the diagram with a quasi-elastic stage LM, failure observed at M, and subsequent work-softening as the specimen dilates, sucks in water and approaches a critical state condition at N. The post-peak behaviour MN causes real difficulties both experimentally and numerically. The specimen is almost certain to behave non-homogeneously and to develop a

narrow rupture zone undergoing intense shear (and softening) while the surrounding clay marginally unloads elastically. This form of behaviour has been discussed more fully by Schofield & Wroth (1968 8.5). On the macroscopic scale the dilation and increase in water content (or specific volume) of the saturated specimen which occurs in the rupture zone will not be observed at the boundaries of the specimen. In order to represent this phenomenon in a realistic and simple way, the microscopic softening is neglected. The consequence is that the specimen is simulated as having an elastic, perfectly plastic response PQR, as illustrated in Fig.7(b). This means that the Hvorslev failure surface is being treated as a yield surface with a nonassociated flow rule. In reality, this simplification means that the post-peak drop in strength is

106

associated

Associated

/ ap x

/ X

/ c

(a)
Fig.7. Assumed Roscoe-Hvorslev yield surface being ignored. Should the introduction of the Hvorslev failure surface prove to be a useful development in modelling soil behaviour, but this deficiency turns out to be important, then it would be possible to adopt a different flow-rule which allows dilation with a consequent amount of softening as the specimen approaches the critical state. The introduction of the Hvorslev surface (taken as straight in the section shown in Fig.6(f)) requires only one additional parameter for complete specification of the model. This is because the critical state line forms the outer edge of the surface? in Fig.7(a) the point X which is the crown of the ellipse forming the MCC yield surface establishes the position of the end of the straight line forming the section of the Hvorslev surface. The additional parameter chosen is the positive (dimensionless) constant a, where the abscissa of point Y in Fig.7(a) is given by -ap' x ,or by -ap'c/2. 3.3 Incomplete saturation Many natural clays are not fully saturated having gas present in the pore fluid. Even if the gas is mostly dissolved in the pore water, the pore water will be relatively compressible, and it cannot be assumed to be incompressible in comparison with the soil skeleton (as is usual for a saturated soil). For such a case, an undrained test when no pore fluid is allowed to cross the boundary of the specimen is not synonymous with a constant volume test. It is difficult in preparing specimens of reconstituted clay to ensure full saturation, and the only guarantee of achieving it is to apply sufficient back pressure,

(b)

until the pore pressure parameter B equals unity. It was evident from careful scrutiny of the input test data that some of the specimens of clay were not fully saturated. In order to model these data, a simple method had to be devised to simulate the effects of incomplete saturation. The problem only arises in simulating the behaviour of soils in conventional undrained tests, because the increased compressibility of the pore fluid will have no effect on a fully-drained test. . In the undrained tests (without back pressure) each specimen was assumed at first to behave as fully drained. Specimens subjected to triaxial compression, or which yield at low overconsolidation ratios in triaxial extension, undergo compressive volumetric strains. At a certain compressive strain, vsat* it was assumed that all gas bubbles within the specimen were compressed to zero volume and the sample was then assumed to behave in the usual undrained manner undergoing zero volumetric strain. In the absence of back pressure it would be expected that the bubbles might reappear when the pore pressure falls to a low value. This value was arbitrarily set as atmospheric pressure, so that whenever the lateral total stress becomes equal to the cell pressure it is again assumed that the sample behaves as fully drained, i.e. it is assumed that the sample can develop only positive, never negative, pore pressures with respect to atmospheric. The above procedure attributes zero stiffness to any bubbles in the pore water, treating them as infinitely compressible until a sudden point at which they are reduced to zero volume, and the pore fluid can be taken as incompressible.

107

4 SELECTION OF MATERIAL PARAMETERS The Roscoe-Hvorslev model used in this prediction exercise involves six material , each of which has a clear physical meaning. In this section each of physical these parameters is defined, significance noted and the means for determining it 4.1 The Critical State stress ratio M The concept of Critical States is central to the model. The critical state is approached asymptotically by a normally consolidated or lightly overconsolidated sample as it is sheared monotonically, and represents a state in which the soil can be sheared indefinitely with no further change of effective stresses or volume. It is found empirically that, at the critical state, the stress ratio q/p' is essentially independent of p', and this critical state stress ratio is denoted by M (capital y) . The value of M is sometimes found to differ between triaxial compression and extension tests, and this effect could be incorporated within the model by adopting a non-circular generalisation of the yield locus in the octahedral plane. For simplicity a circular generalisation (i.e. equal M in compression and extension) is adopted here. The value of M is most easily determined by plotting the stress paths of shear tests on normally consolidated samples in p'-q space and determining the stress ratio at the end of the test (i.e. at very large strain). Adopting this procedure for tests 1.1 and 1.2 yields a value* of 0.74 in compression and 0.78 in extension. The value of 0.74 has been used for all predictions, which is a fairly typical value for M, which usually falls in the range 0.7-1.2. The value of M is directly related to an angle of friction at constant volume (p'cv* e.g. compression M 6 sin < b 'cv 3 sin ( f ) 'cv . . .(4)

ship between AnV and An p' is approximately linear for any constant stress ratio path, and that the gradient is independent of the stress ratio. The gradient of this line is termed X* (cf. the use of X for the gradient in V An p ) The value of X* may be obtained from that part of a consolidation test which exceeds the initial preconsolidation pressure. The most readily available tests are usually one-dimensional oedometer tests, in which V and av are measured. For a normally consolidated material it is found that Ko (the coefficient of earth pressure at rest) is reasonably constant, so that the gradient in J i n V - An av' space is the same as that in An V - An p' space r . ' space and Plots of test 2 in An V An c i test 3 in An V An p space were therefore used to determine X*, with a value of 0.1225 being chosen. This value is reason able for a clay, with values in the range 0.01-0.2 being common. The lower values apply to less compressible, siltier materials. 4.3 The gradient of swelling lines k * The gradient of swelling and reconsolidation lines is also found to be approxi mately constant in An V An p samples subjected to constant stress ratio * paths, and this gradient is denoted by K (cf. the use of K for the slope in V - An p' space). Since K0 is not constant for overconsoli dated samples, but varies with OCR, it follows that p' is not a constant multiple i * of av during rebound and that K cannot be obtained directly from the slope of a An V - An av' plot from an oedometer test. The variation of K0 with OCR is not usually known for any particular clay, so it is necessary to use the gradient of the An V - An av line to determine k*. The data from test 3 are typical, showing some hys , and a value of k* of 0.0165 was chosen. The ratio of K*/ X * is therefore 0.135. This value is occasionally as high as 0.3 and may be as small as 0.05 for clays. The elastic bulk modulus may be calculated as proportional to pressure and equal
to P'/K*.

Thus, M of 0.74 corresponds to < f > ' cv of 19.2. 4.2 The gradient of normal consolidation lines X* It is observed empirically that for a normally consolidated clay the relation-

4.4 The shear modulus G It is well known that the shear modulus of a soil varies with the mean effective stress, and also with the preconsolidation

108

. Introducing a variation of G with pressure results, however, in non conservative elasticity if the bulk modu lus does not depend on shear stress. Simi larly, the introduction of G dependent on pc' means that the model becomes one of coupled elastic-plastic behaviour, which is considerably more complicated than uncoupled plasticity. A constant shear modulus is therefore used in the model for simplicity. The value of the shear modulus was few chosen from the gradient of the loading increments on overconsolidated samples. It is obtained by plotting the tests in e-q space with the gradient being equal to 3 G. A value of 7000 kN/m was chosen. This may be compared with the correlation G 75 cu suggested by Wroth et al. (1979); taking a typical test (input 2 test 1.1) with cu 90 kN/m would indi For the cate G of the order of 6750 kN/m tests at lower pressures, a lower value of G might be more appropriate, and for very small values of shear strain a higher value of G should be used. 4.5 Critical specific volume at unit
r

Substituting the values of p' , q, pc' and M for the failure points of tests 1.4.2 and 1.4.4 results in a value of a = 0.141, which compares well with data for other clays. 5 EVIDENCE OF INCOMPLETE SATURATION

One parameter is needed to locate the iso tropic normal consolidation line, or the i this is critical state line in V-p usually chosen to be the volume on the critical state line corresponding 2 to unit pressure 1 kN/m , and denoted by r (capital gamma). It is only required if absolute values of volume (or water content) are needed. In this exercise the volumetric of strains (related to the volume) are asked for, so that the value of r is irrelevant and it has not been determined. 4.6 Hvorslev surface intercept a Those samples with an OCR greater than 2 or so (the exact value will depend on the type of test chosen to bring the sample to failure) will be expected to fail in a brittle manner on the Hvorslev surface Since the surface is chosen to pass through the point (pc'/2, Mp^/2) in p'-q space it c is only necessary to determine one para meter in order to locate the surface This is conveniently taken as a (see Fig.7(a)) where the surface is described by: M + .(5) q (P' + ap A/ 2) . 1+a

At the start of the exercise of fitting the input data to the Roscoe-Hvorslev model, the event was to plot the results of the consolidation tests in the conventional manner, and the results of the triaxial tests in terms of the vari ables p', q, V and e. The results of input test 1.1, which is an undrained compression test on an isotropically normally consolidated sample, are plotted as individual crosses in Fig.8. The effective stress path is evident in Fig.8(a). All evidence from other undrained tests on saturated samples of clay is that the effective stress path curves to the left immediately at the start of the test at point A. If, however, the sample behaves initially as though it were drained, it would have an effective stress path (in a conventional test with constant cell pressure) of gradient 3. It is considered to be most significant that the path is initially at gradient 3 from A to B, and at B it undergoes a marked change to the curved path BCD, as positive excess pore-pressures develop. It was on the of this evidence that it was believed that the test could not have been conducted with back pressure, and that the sample in question was not fully saturated Similar from the other input tests was observed and two other examples are shown in Figs 9 and 10. Consequently enquiries were made at that time (March 1982) and it was confirmed that back pres sures had not been used in the series of input tests The single set of basic soil properties assumed to apply to all the tests was selected as described in section 4 and Table 1. Selected values of soil proper for Roscoe-Hvorslev model. M Critical state stress ratio 0.74 X* Gradient of normal consoli dation line 0.1225 k* Gradient of swelling line 0.0165 7000 kN/m G Shear modulus r Critical volume Undetermined at unit pressure a Hvorslev surface intercept 0.141

109

q kN/m

q kN/m

300

300

200

200

100
B

100

100

200
(a)

300
p'kN/m
2

<00

500

10
(b)

207,

Fig.8. Input test 1.1 listed in Table 1. For each of the input triaxial tests the value of the partial saturation parameter v s a t was chosen on a trial and error basis to provide the best fit to the data. These values are included in Table 2. Table 2. Initial states of input tests Test
2

1.1 1.2 1.4.1 1.4.2 1.4.3 1.4.4

pQ' kN/m 400 400 301 68 101 22 2 pc' kN/m 400 400 600 600 200 200 .012 .03 .0025 .0055 .0017 .006

For input test 1.1 (Fig.8) the value of v was sat 0.012, i.e. 1.2% compressive volumetric strain was allowed to occur before the sample was assumed to become effectively saturated and thereafter behave in an undrained (constant volume) manner. In Fig.8 the continuous line represents the results computed for the test, which provide a very close fit to the experimental data points. The part AB of the effective stress path, in Fig.8(a), of gradient 3 is the drained response of the sample (which yields at A and is always behaving plastically); at B a total compressive volumetric strain of 1.2% has

q kN/m

q kN/m

H
+ ++ + + + + +
+ +

100

200
p'kN/m (a)

300

400

10

207

(b)

Fig.9. Input test 1.4.1

110

occurred whereupon the sample is modelled by the undrained path BCD, asymptotically reaching the critical state D after infinite shear strain. Note that in Fig.8(b) the computed result shows a small increase in stiffness at B when the abrupt change from drained to undrained behaviour is assumed to occur. For a fuller discussion of the problem of mnsaturation, see Appendix B. Input test 1.4.1 is an undrained test on a sample with an initial isotropic overconsolidation ratio of 2, shown in Fig.9. The value of vsa-j- adopted was 0.25%, which corresponds to the volumetric strain experienced by the sample during the initial drained phase from E to F. This phase is entirely within the yield surface so there is zero plastic volumetric strain, and the total volumetric strain consists solely of the component. i i i i At point F the sample becomes and behaves in an undrained manner. The stress state is still within the yield locus, so that the sample is assumed to behave in an (isotropic) elastic manner with the consequence that p' remains constant and the effective stress path FG is parallel to the q-axis. At point G, the stress state reaches the Roscoe yield surface, the sample yields and approaches the critical state at H. In this case the general character of the stress path is well represented, but the stress-strain curve EFGH in Fig.9(b) is a poor match to the experimental data. Input test 1.4.2 was on a heavily overconsolidated sample with OCR =8.8. The value of v sa t was also very small (0.55%).
q kN/m
2

The results are displayed in Fig.10; IJ represents the drained phase, JK the undrained phase, and K repre sents the state at which the stresses reach the Hvorslev failure surface There after the sample is assumed to behave perfectly plastically K to L. It should be remembered that the Hvorslev surface has been selected to go through the failure (i.e. the peak) point, so that the computed result for K in Fig.10(a) does not repreof the experimental sent a match but a data. 5.1 Summary of input tests Each test was analysed as fully saturated throughout, and also as partially saturated by an amount to give the best The optimum values of v s a t are listed in Table 3. Table 3. Optimum values of parameter repre senting partial saturation in input tests Po 2 Pc' 2 kN/m kN/m 1.1 1.2 1.3 1.4.1 1.4.2 1.4.3 1.4.4
* *

Optimum vSat 0.012 0.03 0.01

emax

400 400 400 300 68 101 22

400 400 400

0. 16 -0.1 3 cycles 0 .007 ^ -0.0064 600* 0.0025 0.16 600 0.0055 0.16 200** 0.0017 0. 16 200 0.006 0.16 175

with pc' Better fit with pc'


2

q kN/m

100

100
L
+ + + +

50

50

i 50
(a)

I
p' kN/m
2

100

10
(b)

207.

Fig.10. Input test 1.4.2

111
X

6 OUTPUT TESTS Having established the importance and influence of the incomplete saturation of the clay samples of the input tests, a major problem arises in making predictions for the output tests as the degree of partial saturation of each sample (which will almost certainly vary, as was the case of the input tests) is unknown and unspecified It was decided to make predictions for the two following cases for each output , (a) full saturation, and (b) incomplete saturation, with a value of v s a ^ = 0.01. This figure was chosen as a round number close to the average of the optimum values found for the input tests. The predictions are given in the form of plots as specified in the following Figs 11-24. 6.1 Summary of output tests Each was modelled as fully saturated throughout and also as partially saturated with v s a t = 0.01 (i.e. experiencing a total compressive volumetric strain of 1% before becoming fully saturated). The test conditions are listed in Table 4. No predictions are made for output test 2.0 since the specifying the test were inadequate. It is thought that since this output test is referred to as corres ponding to input test 3.0, it involved an initial stress condition of approximately . However, unlike input test 3.0, which is essentially a consolidation test at rate ratio (and hence would be expected, and is observed, to follow a constant stress ratio path), this
\

xoct k N / m

\ 001

ei 0015

Fig.11. Predicted results

test 1.1 fully saturated

test is a shear test at constant volume. None of the input tests provides adequate information about the shape of the yield locus very close to the origin. Tests in

p'kN/m

Table 4. Conditions of output tests Test kN/m


2

kN/m

Selected test conditions

1.1

400 400 400 400 300

1.2.1 1.2.2
1.2.3 1.3

25 cycles of strain ex from 0.01 to - 0.01 400 64 cycles of shear stress 2 q from +59 to - 64 kN/m 400 64 cycles of shear stress 2 q from +80 to - 90 kN/m 400 64 cycles of shear stress 2 q from +100 to -122 kN/m 600 Extension test to 0.15 e 400
l

Notes: (1) For an undrained test 3 T (2) q oct


7i

001

0005

0-01

0015

Fig.12.

test 1.1 fully saturated

112

Toct kN/m

0-04

eI E

002

H0-01

ei 0015
10
100

-0-02-

-0-04

Fig.13. Predicted results

test 1.1 unsaturated

Fig.15. Predicted results: test 1.2.1 fully saturated

' kN/m

004

eI E

0-02

200-

10

100

log N

100-

-0-02

1 0-005

I
0-01

001

1 ei 0-015

-0-04

Fig.14. Predicted results

test 1.1 unsaturated

Fig.16

test 1.2.1 unsaturated

113

eI E

I E

004

0-04

002

002

log N

10
-002

100

log N

-002

-004 Fig.17. Predicted results: test 1.2.2 fully saturated

-00/. Fig.19. Predicted results: test 1.2.3 fully saturated

eI E

I E

0-04

0-02

log N

log N

-002

-0-02

-0-04 Fig.18. Predicted results: test 1.2.2 unsaturated

-004 Fig.20. Predicted results: test 1.2.3 unsaturated

114

Toct

2 kN/m

Ioct kN/m

ei

Fig.21. Predicted results

test 1.3 fully saturated

Fig.23. Predicted results

test 1.3 unsaturated

p'kN/m

p'kN/m

400

400

200

200

0-05

0-1

ei 0-15

H 005

+
0-1

ei 0-15

Fig.22. Predicted results: test 1.3 fully saturated

Fig.24. Predicted results: test 1.3 unsaturated

115

this region of stress space are difficult to carry out and little information about the general behaviour of clays in this region is available. It is certainly not sensible to extrapolate the elliptical yield locus (princi of critical) pally valid on the "wet" or the Hvorslev surface to points close to the origin. There is some evidence that at very low effective stress levels the strength of the clay is determined by tencracking, when zero normal stress occurs on any plane. If the test does indeed start at virtually zero mean normal stress and preconsolidation pressure then the only prediction that can be made is that all stresses will remain virtually zero throughout this This result would, however, be very between the to the exact , and any minute variation from constant volume could result in any of a variety of stress paths. Problems of unsaturation are also liable to introduce significant uncertainties at very low stress levels.
>

The model makes use of isotropic harden ing, with the parameter pc' which establishes the size of the above loci being linked to the plastic volumetric strain. The relationship is most conveniently given in incremental form: vP
U

PC K*) I Pc

. . . (A3)

Plastic potentials may be defined for the model. For the Roscoe surface the flow rule is so that the plastic potential is
P' (P'-Pj)

. . . (A4)

For the Hvorslev surface a non-associated flow rule with zero dilation is used, so that the plastic potential is given by:
q qc

(A5)

APPENDIX A BRIEF SPECIFICATION OF THE MODEL The assumptions embodied in the model are given in detail in sections 2 and 3 of the paper. This appendix contains only a mathematical specification. A.1 FUNDAMENTALS The model is elastoplastic in concept. Within the yield locus the material behaves elastically with a bulk modulus proportional to pressure and a constant shear modulus. The yield locus of two One section is elliptical in T q~P' ( oct~P') space and involves an associ ated flow rule and work hardening linked The to the plastic volumetric i other section is a straight line in q-p (Toct-p') space and involves perfectly plastic behaviour with a non-associated flow rule. The yield surfaces in the model are Roscoe surface
P' (P'-Pc')
+

where qc is a constant dependent on the current stress state. The model does not make use of yield (although at the critical state which forms the transition from the Roscoe to Hvorslev surfaces there is a corner in the yield locus). Coaxiality of strain rate and stress is assumed during plastic behaviour. If the Hvorslev surface is not included in the model, it is then able to predict softening on the "dry" side of the critical state. This process is thought to be unstable, with localisation into narrow shear bands occurring. The use of the Hvorslev surface in effect precludes this softening by modelling the formation of narrow shear bands by a perfectly plastic, non-dilatant behaviour. The model fulfils all the requirements of a physically meaningful and mathematically consistent model in continuum mechanics. As a consequence of the above the model should generate well posed problems subject to appropriate boundary conditions. Ill conditioning could of course occur if certain combinations of conditions are imposed, for example, a rapid rate of loading resulting in "undrained" conditions linked to entirely deformation-controlled boundary conditions may result in ill conditioning if a bulk modulus of the pore fluid is not admitted. A.2 INCREMENTAL RELATIONS Elastic behaviour

3 2 M

0,

(Al)

Hvorslev surface M + + q 1+a

(A2)

by:

116

p'

K*

-e

(A6) (A7)

p q = 3G e Behaviour on Roscoe surface

The conventional assumption of additive and plastic strains is made. Using the associated flow rule and the conventional formulation: P e.. ID 9g A a 3a. _ 3a. . kl' kl in 3f . . . (A8)

/unloading on the dry side of critical are a particular area where a detailed understanding of the principles of the model are required. The precise numerical formulation used is important as the critical state is approached if numerical difficulties are to be avoided. The details of a subroutine depend too strongly on the constraints of the calling program to give a suitably general routine. For isotropic loading the model simplito the incremental form: P' < Pc' P' P p' > 0
1

v
V

K*P' P' A*P' p' P' Pc '

. .(A15)

which results in V

. .(Al6)

0 p' ( 2 p ' p J ) | f ( 2 P -P ' f p' c P 2 + A 1 03 q (2p'-pJ)M q G _ M "


*
_ |

(A9)

Closed form for the triaxial test may be obtained but is rather lengthy. APPENDIX B UNDRAINED TESTS ON KAOLIN

The work hardening equation (A PC K*) Pc ' . . .(A10)

is combined with the consistency condition obtained by differentiation of the yield locus: P(2p-pc) + M
qq

PcP

. . .(All)

Comparison with the appropriate term from (A10): Alp' (2p'-p-)


2

.(A12)

then yields the value of the hardening parameter A: A (A K*) P'PJ (2p'-p ')

. . .(A13)

Behaviour on Hvorslev surface Similarly the behaviour on the Hvorslev surface may be determined as: V

0 0 0 p' p * P' s + X 1 M e 03 -r 1 q q G _ 1 + a

. . .(Al4)

where, since perfect plasticity is assumed, the value of the scalar multi plier A may be determined, for by Hill's method (Hill 1951). In writing a FORTRAN subroutine to implement the above calculations, great care must be exercised in attention to in the model. Criteria for loading

Any soil test, whether performed in the laboratory or the field, can be considered as a miniature boundary value problem. A triaxial test is no exception. An initially cylindrical sample is subjected to a uniform overall axial deformation through the rigid end platens, and to a uniform lateral total pressure by means of the cell fluid acting on the encapsulating rubber membrane. Only in certain circumstances will the boundary conditions result in the soil deforming as a single homogeneous element and it will not necessarily be correct to assume that conditions of stress and strain within the sample are uniform. Test procedures can be chosen in order to try and encourage the development of uniform conditions : for example, lubricated end platens and short samples may be desirable; a slow rate of testing will give the pore pressures within the sample an opportunity to reach equilibrium in drained or undrained tests. It is also important to appreciate that the so-called undrained test, conven tionally interpreted as a constant volume is really a constant overall mass the term isomassic might be appro priate to it. Assuming no leakage occurs, once the drainage valves are closed, no mass of soil or pore fluid can move into or out of the sample, but this does not prevent the sample from changing in volume. If the pore fluid is not incompressible, for example, because the soil is not satu, then a soil sample can change in volume as it is sheared, even though the deformations and stresses within the sample remain uniform. On the other hand, even if

117

the overall sample does not change in volume, it is quite possible for local changes in volume to occur within the sample, as a consequence of end restraint or rate of resting, as shown by the analyses of Carter (1982). A true constant volume test path, the so-called isochoric, is best achieved as a drained test in a device such as a true triaxial apparatus in which a combination of deformations can be imposed on a sample, such that no change in the volume occupied by the soil skeleton can occur, always assuming that the true triaxial sample can itself be treated as a single element of soil. When a new testing procedure is to be used to study the undrained behaviour of soil, whether it uses a new pore pressure probe or merely a new rate of testing, it is important to check that the results produced by this new procedure are not inconsistent with existing knowledge of soil behaviour, or at least not without some good explanation. There are only four parameters to measure in the undrained triaxial test: the total radial stress (cell pressure), the total axial stress (cell pressure plus deviator stress), the pore pressure and the axial strain. The background of critical state soil mechanics makes it clear that the behaviour of soils is best understood in terms of effective stresses, and that significant patterns are more likely to emerge when soil response is studied in terms of effective stresses and volumetric packings, which are limited parameters, rather than in terms of stress:strain curves, shear strain being an unlimited parameter. The best way to examine the results of an undrained test is therefore to plot the effective stress path in a convenient stress space, such as deviator stress q:
9

mean effective stress p'. Plotting q, p' or pore pressure u against strain is unlikely to show up problems, particularly if the test has been taken to large strains; the significant changes in effective stress occur at small deformations. The effective stress paths for an undrained compression test and an undrained extension test performed by Kuntsche on kaolin at Karlsruhe are shown in Fig.Bl. For comparison, two pairs of effective stress paths for undrained compression and extension tests performed by Nadarajah (1973) on kaolin in Cambridge are shown in Fig.B2. Some points of dissimilarity may be noted: (i) Nadarajah's compression and extension paths are broadly of the same shape; Kuntsche's are very different. For an initially isotropic material the effect of a small increase or decrease in deviator stress should be essentially identical, there can be no theoretical explanation for the difference; (ii) Nadarajah finds a lower ratio of q to p' at failure in extension than in compression and lower values of q at failure in extension than in compression; Kuntsche finds a higher strength in extension than in compression and the same value of q/p*. The total stress paths for Kuntsche's tests have been shown in Fig.Bl. It is clear that, for the first part of the compression test and for virtually the whole of the extension test, zero pore pressure was recorded. Kuntsche measures pore pressures by means of a rigid porous probe extending from the base to the centre of the sample. Evidently the average pore pressure recorded by this probe is zero. Perhaps it provides a drainage path for positive pore pressures near the centre of the sample to compensate for negative pore

Table Bl Material Index Cambridge kaolin (Nadarajah (1973) LL 72.1% 40.4% PL 2.63 Gs mix at 160% 2 ID to 195 kN/m over 8* days 2 2 from 103 kN/m in 69 kN/m /I2h steps then 24 h rest (typically 3$-4i days) B values 0.97-0.99 load controlled: 2.25 kgf/h reducing to 0.45 or 0.22 kgf/h (typical initial strain rate 0.125%/h Karlsruhe kaolin (Kuntsche 1982)

Sample

50% LL 16% PL 2.72 Gs mix at 100% 2 to 80 kN/m over 7 days 2 to 400 kN/m in 1 step B values 0.97-0.99 strain controlled: 12%/h

Testing procedure

118

pressures developing near the base. It is worthwhile to compare the procedures used by Nadarajah and Kuntsche for their tests: these are summarised in Table Bl. The kaolins being tested are rather different. The plasticities are about the same but the Cambridge kaolin has a much higher liquid limit. Nadarajah's isotropic consolidation procedure was much slower than Kuntsche's and one may wonder whether the Karlsruhe kaolin is completely consolidated at the start of the tests. A condition of initial under-consolidation could well give rise to a tendency for continuing pore pressure development later in the tests, thus obscuring the attainment of critical states. The final isotropic consolidation, from 2 2 80 kN/m to 400 kN/m in one step, could conceivably leave the sample initially nonuniform as a result of the Mandel-Cryer effect. The rate of testing used by Kuntsche is initially typically about 100 times faster than that used by Nadarajah (who was conducting the tests with careful load control)

and this high rate of testing may have led to inaccurate observation of pore pressures. For an isotropic sample deforming uniformly, an excessive rate of testing should, however, lead at worst to an effective stress path with no initial change in effective mean stress (6p'=0, 6u=6q/3) as for the elastic material but not a path with an increase in p'. Effective stress paths predicted by CamClay models for undrained tests have the same shape, independent of consolidation pressure. Using volume change due to lack of saturation as an explanation of the results that are shown in Fig.Bl, an estimate of the lack of saturation can be made by fitting curved undrained effective stress paths through as much as possible of the paths obtained experimentally; these curves are also shown in Fig.Bl. The compression and extension paths expected for kaolin con2 solidated to 400 kN/m are shown. Much of the compression path is fitted by using the undrained effective stress path for a 2 sample consolidated to 440 kN/m .

t i i

Karlsruhe kaolin

Fig. Bl.

119

compression

Cambridge kaolin

o
o

o o
i

o o
o o

kNAn

500

o
o

Fig. B2

finV

ZnV/Vo

\*
0.1225

in p ' / p '

X*
V

1.665 at p'

2 400 kN/m

In p
for change of p vol strain 6v 6V Sr
1

4 0 0 4 4 0 kN/m 1.17% 0.0194 0.97 ?

400 3.90%

2 550 kN/m

0.0637 0.90 ?

Fig. B3

120

The normal compression of the clay can be described by the expression (Fig.B3): J i n V/V0 = X* In p'/p0' with A* = 0.1225 and V = 1.665 at p' = 2 400 kN/m . A change of p' from 400 to 2 440 kN/m implies a volumetric strain of 6v = 1.17% (cf. 1.2% used in prediction exercise), a change in specific volume 6V = -0.0194 and a saturation Sr = 0.97. The last part only of the extension test is fitted by using the undrained effective stress path for a sample consolidated to 2 550 kN/m . A change of p' from 400 to 2 550 kN/m implies a volumetric strain of 6v = 3.90% (cf. 3% used in prediction exercise) , a change in specific volume 6V = -0.0637 and a saturation Sr = 0.90. REFERENCES Atkinson, J.H. & P.L.Bransby 1978. The mechanics of soils: an introduction to critical state soil mechanics. London: McGraw Hill. Burland, J.B. 1967. Deformation of soft clay. University of Cambridge: PhD thesis. Butterfield, R. 1979.A natural compression law for soils (an advance on e-log p'). G^otechnique 29:469-480. Carter, J.P. 1982. Predictions of the nonhomogeneous behaviour of clay in the triaxial test. G^otechnique 32:55-58. Hill, R. 1951. The mathematical theory of plasticity. Oxford University Press. Houlsby, G.T. 1981. A study of plasticity theories and their applicability to soils. University of Cambridge: PhD thesis. Hvorslev, M.J. 1969. Physical properties of remoulded cohesive soils. Vicksburg: US Army Engineer Waterways Experiment Station, Translation 69-5. Nadarajah, V. 1973. Stress-strain properties of lightly overconsolidated clays. University of Cambridge: PhD thesis. Parry, R.H.G. 1960. Triaxial compression and extension tests on remoulded saturated clay. G^otechnique 10:166-180. Roscoe, K.H. & J.B.Burland 1968. On the generalised behaviour of 'wet' clay. In J.Heyman&F.A.Leckie (eds.), Engineering plasticity, p.536-609. Cambridge University Press. Schofield, A.N. & C.P.Wroth 1968. Critical state soil mechanics. London: McGraw Hill. Wroth, C.P. & G.T.Houlsby 1980. A critical state model for predicting the behaviour of clays. In Workshop on Limit Equili- * brium, Plasticity and Generalised Stress
m

-Strain in Geotechnical Engineering, p.592-627. Wroth, C.P., M.F.Randolph, G.T.Houlsby & M.Fahey 1979. A study of the engineering properties of soils with particular reference to the shear modulus. University of Cambridge: Technical Report (Soils TR75).

121

You might also like