You are on page 1of 30

Introduction to barodesy


Dimitrios Kolymbas

April 26, 2016

Barodesy is a frame for constitutive modeling of soils based on their asymp-


totic properties. This frame allows to derive the constitutive relation by rea-
soning on general properties of granular materials. The so obtained constitu-
tive relation is a single tensorial equation of outstanding simplicity that ex-
presses the evolution of stress in dependence of the deformation. Being more
than a mere equation, barodesy is rather a constitutive theory to provide
a frame for understanding the intricate mechanical behavior of condensed
granular matter. As compared with previous publications on barodesy, the
present article contains considerable improvements in terms of conceptual
rigor, simplicity and performance.

Key words
Constitutive models, hypolasticity, barodesy, critical states, asymptotic behavior.

1 Introduction
Barodesy is a frame of constitutive models for soils that deviates from elastoplasticity. In
this article the structure of the theory and a simple calibration procedure are outlined,
and some results of simulations of element tests are also presented. This article refers
to sand, barodesy however holds also for clay. Clay, being also a particulate material
consisting of minute particles has a behavior very similar to sand. Both exhibit critical
states. In addition, an asymptotic behavior similar to the one described by Goldscheider
for sand (see Sect. 3) was observed by Topolnicki in experiments with Kaolin clay in a
Hambly type biaxial apparatus (Topolnicki, 1987). However, there are some differences
that mainly arise from the fact that the stiffness of sand in monotonic compression is
much higher than that of virgin consolidated clay. In other words, sand has the tendency
to jam at monotonic compression.


Division of Geotechnical and Tunnel Engineering, University of Innsbruck, dim-
itrios.kolymbas@uibk.ac.at

1
Since the first communication in 2009,1 several versions of barodesy have been pub-
lished (Kolymbas, 2011, 2012, 2014). The here published version contains a series of
substantial novelties:

1. The function R(D), which is of central importance, has been changed from the
original form R(D) := (trD0 ) 1 + c1 exp (c2 D0 ) to the considerably simpler
form of equation (4), which gives better simulations. Moreover, a rationale has
been added in Section 4 to explain why this exponential function was chosen.

2. The stiffness factor h(σ) was initially chosen as a power function of the type h = σ c
or h = (σ/p0 )c . This was inspired by the known relations of Ohde and Janbu, but
also by results from resonant column tests by Hardin and Richard. However, the
power function has the disadvantage that the stiffness vanishes at σ = 0, i.e. at
the surface of the ground, and this causes numerical problems. An additional dis-
advantage refers to the related non-integer powers of dimensions of the material
constants or the introduction of the reference stress p0 , which is virtually an ad-
ditional material parameter. All these shortcomings are removed by the equation
(23), which has the additional advantage of becoming ∞ for e → emin .

3. The critical void ratio ec and its dependence on σ (i.e. the ”critical state line”) have
been derived from barodesy in Section 8 and explicitly stated in equation (26).

As for the name ’barodesy’, true, too many neologisms create confusion. But some-
times, new names are needed to denote ideas that substantially differ from the main-
stream. To denote constitutive models, usually the word ’plasticity’ is used with prefixes
such as hypo-, para-, hyper- etc. For granular materials, however, the words ’elastic-
ity’ and ’plasticity’ (to the extend the latter is associated with notions such as yield
surface, elastic regime etc., originally created for metals) can (and perhaps should) be
avoided. Of course, soil behaviour can—in principle—also be described in the frame-
work of the theory of elastoplasticity if appropriate extensions are used. However, yield
surfaces and the other concepts of plasticity theory may prejudice our perception and
sometimes obscure soil mechanics, which suffers from a long lasting fragmentation in
constitutive modelling (Kolymbas, 2000). The name ’barodesy’ has been coined by the
author motivated by the fact that granular materials gain their stiffness (δ έσις = bond,
hence stiffening, hardening) from externally applied pressure (β άρoς). Thus, the names
’barodesy’ and ’barodetic’ are proposed for granular materials to distinguish them from
what traditionally is denoted as ’elastic’ or ’plastic’.

1.1 Early quests for alternatives to plasticity theory


For soils, the generally observed inelastic behavior within the yield surface contradicts
a basic ingredient of the theory of plasticity.2 This theory, based on the notions of yield
surface, flow rule, consistency, decomposition of strain into elastic and plastic parts,
1
At the Symposium on Mechanics of Natural Solids, Horto/Greece, 2009 (Kolymbas, 2011).
2
I do not refer here to elastic response at very small strain.

2
etc. was for a long time the only mathematical tool to describe irreversible deformation
of solids. Thus, also soil mechanics has been developed along the principles of plasticity
theory. The first consistent model for soil, the Cam-Clay model, is a particular plasticity
theory adapted to clay. However, several researchers have called to depart from plasticity.
In 1973 Palmer and Pearce published a paper titled ”Plasticity theory without yield
surfaces” (Palmer & Pearce, 1973). Some sentences of this paper deserve being quoted
here:

. . . it might be useful to idealise clay as a material in which the yield surface


has shrunk to a point, so that all deformations are plastic . . . The devia-
toric stress has two components. The magnitude of the first component is a
function of the octahedral shear strain, and its direction coincides with the
principal strain vector (referring strain to an isotropically-consolidated initial
state). The magnitude of the second component is constant, and its direction
coincides with the current strain rate . . .
Reversal of the strain path would reverse the second component but not the
first . . . ”

Palmer and Pearce refer to clay, but their arguments are equally valid for sand. The
very last sentence strongly resembles to a basic concept of hypoplasticity and barodesy,
to which pressumably the authors would have concluded, had they used rate equations
instead of finite ones.

1.2 Barodesy and hypoplasticity


As in hypoplasticity (Bauer, 1996, 1997; Gudehus, 1996; Herle & Kolymbas, 2004; Kolym-
bas, 1977, 1985; Mašı́n, 2005; Niemunis & Herle, 1997; Niemunis, 2003; von Wolffersdorff,
1996; Wu, 1992, Bauer & Kolymbas,1996), the barodetic constitutive equation has the
general form
Ṫ = h(T, D, e) (1)
However, barodesy has a different design than hypoplasticity. Besides intuition, trial
and error is a basic tool in developing constitutive models. In the original versions
of hypoplasticity published by the author, trial and error has been guided by general
principles of objectivity and representation theorems for tensor-valued functions. In bar-
odesy, the amount of trial and error has been further reduced in favour of reasoning on
asymptotic behaviour of granulates. Asymptotic states are interesting not only from
conceptual reasons but also from the experimental viewpoint: If we consider long mono-
tonic deformations, initial disturbances, related e.g. with sample preparation, fade out
and do no more influence the measurements.

2 Symbols and notation


As usual in mechanics, the symbol σ denotes the stress, and ε denotes the strain. Tensors
are usually denoted either in the index notation or in the symbolic notation. Compared to

3
the index notation, the symbolic notation facilitates insight into the prevailing relation-
ships. Therefore, in this paper the symbolic notation, as introduced in the ”Non-Linear
Field Theories of Mechanics” (Truesdell & Noll, 1965), is mainly followed. Thus we use:
T : Cauchy-stress. Its components are σij . Its principal components are denoted as
σ1 , σ2 , σ3 .

D : stretching tensor, i.e. the symmetric part of the velocity gradient ∇v. It can be set
approximately equal to the strain rate, Dij ≈ ε̇ij .

trace of D is the scalar quantity trD = D11 + D22 + D22 , which denotes the rate of
volume change: trD = ε̇v = ε̇11 + ε̇22 + ε̇33 .

e : void ratio, i.e. the ratio Vp /Vs , where Vp and Vs are the volumes of pores and solids
(grains), respectively.

value of a tensor is its Euclidean norm: |A| := trA2 .

normalisation of a tensor A is denoted by the exponent 0, i.e. A0 := A/|A|. Clearly,


A0 is a unit tensor with |A0 | = 1. A tensor can de represented by its value and
the corresponding unit tensor, e.g. T = σT0 , D = ε̇D0 .

σ : |T|, the value of stress

ε̇ : |D|, the value of stretching



 : trD0 . This quantity is a measure of dilatancy.3 Its minimum value, − 3, is obtained
for isotropic compression, i.e. for zero deviatoric deformation.  increases with
increasing part of deviatoric deformation.

Ṫ : time rate of stress. Objective stress rates are addressed in Section 12.6

c1 , . . . , c5 : material constants. It facilitates reading to clearly denote the material


constants as such and to distinguish them from quantities that depend on other
variables. The present version of barodesy needs 5 material constants and, in
addition, two particular void ratios: the critical void ratio at zero stress, ec0 , and
the minimum void ratio emin .

3 Empirical basis of barodesy


Of basic importance for the following is the notion of a proportional path. Proportional
stress and strain paths are characterized by constant ratios of the principal values σ1 :
σ2 : σ3 and ε1 : ε2 : ε3 , respectively.
3
In (Vermeer, 2000) the following measure of dilatancy is used: sin ψ := −(ε̇1 + 2ε̇2 )/(ε̇1 − 2ε̇2 ) which
implies !
ε̇1 + 2ε̇2 4 sin ψ
 = p = p (2)
2 2 2
ε̇1 + 2ε̇2 6 sin ψ − 4 sin ψ + 6
.

4
There are two basic experimental findings for sand:
1. Starting from the stress-free state, proportional strain paths lead to proportional
stress paths.
2. Starting from a non-vanishing stress state and applying a proportional strain path
leads asymptotically to the proportional stress path that would be obtained start-
ing from the stress-free state.
These two rules are inferred by Goldscheider from his test results obtained with rec-
tilinear extensions of sand (Goldscheider, 1976) (cf. also (Mašı́n, 2012)). These tests
have been carried out in a so-called true triaxial apparatus, which allows to apply rec-
tilinear extensions (i.e. motions without rotation of the principal axes of deformation)
independently in all three directions of space.

4 Proportional paths A
Let us first consider proportional strain paths starting form the stress-free state. Such
paths can be volume-decreasing (we will call them ’consolidations’), characterized by
trD< 0, or volume preserving (’isochoric’ or ’undrained’), characterized by trD= 0, or
volume increasing, characterized by trD> 0. Clearly, the latter are not feasible with
cohesionless sand.
Let us denote with R a tensor that has the direction of a proportional stress path. The
question arises, how does R depend on the direction of the corresponding proportional
strain path? The latter is characterized by the direction of stretching D, i.e. by the
normalized stretching D0 . How can we determine the relation R(D0 )? This question
can be easily answered if we observe that all consolidations are mapped into a specific
part of the principal stress space, which is formed by the stress components σ1 , σ2 and
σ3 . This part is the octant, where all principal stresses are compressive, i.e. negative.
Hence, the product σ1 σ2 σ3 must also be negative. Now, for a proportional stress path
we have σi = µRi (D), i = 1, 2, 3, where µ simply denotes the proportionality of σi and
Ri .4 With R(Di ) := Ri , the following condition must hold:

R(D1 )R(D2 )R(D3 ) < 0 for trD = D1 + D2 + D3 < 0 (3)

This implies that the product R(D1 )R(D2 )R(D3 ) must be a function of the sum D1 +
D2 + D3 , a requirement which is fulfilled by the exponential mapping

R(D) = − exp(aD0 ) (4)

It should be added that Goldscheider’s rules have, as every rule, some exceptions in
detail. Thus, it is known that stress paths obtained with oedometric deformation are
not strictly proportional, i.e. K0 := σ2 /σ1 is not strictly constant. Such ’details’ can be
captured by slight modifications of barodesy (Fellin, 2013), but they are omitted here
for simplicity.
4
Herein, Ri (D) are the principal values of R(D) .

5
deviatoric plane

Figure 1: Cross section of the R-cone with a deviatoric plane.

4.1 Some implications of equation 4


Equation (4) maps all volume-reducing (trD < 0) proportional strain paths into a cone
in the stress space with appex at T = 0, which can be called the R-cone. Its boundary
corresponds to paths with trD = 0 and is thus the critical state surface. Consider
the intersection of the R-cone with a plane trT = const, as shown in Figure 1. This
curve expresses the critical limit state in a so-called deviatoric plane of the stress space.
The mathematical representation of this curve can be easily derived from equation (4):
Eliminating D0 from (4) we obtain:
1
D0 =
ln(−R) . (5)
a
The requirement trD0 = 0 results in ln(−R1 R2 R3 ) = 0 or R1 R2 R3 = −1. From the
additional requirement |D0 | = 1 we obtain:
[ln(−R1 )]2 + [ln(−R2 )]2 + [ln(−R3 )]2 = a2 . (6)
For the here considered proportional paths holds: T = µR, 0 < µ < ∞, hence we can
replace in this equation R by T/µ and obtain finally the equation of critical states in
the stress space:5
!2 !2 !2
σ1 σ2 σ3
ln √
3 σ σ σ
+ ln √
3 σ σ σ
+ ln √
3 σ σ σ
= a2 . (7)
1 2 3 1 2 3 1 2 3

Equation (7) is homogeneous of the 0-th degree in T and describes thus a conical surface
in the stress space with apex at T = 0. Its intersection with a plane trT =const is shown
in Figure 1. Its shape practically coincides (Fellin & Ostermann, 2011) with the well-
known curve obtained by the expression of Matsuoka & Nakai:
(σ1 + σ2 + σ3 )(σ1 σ2 + σ1 σ3 + σ2 σ3 )
= const . (8)
σ1 σ2 σ3
5 √
From σ1 σ2 σ3 = −1/µ3 we obtain µ = −1/ 3 σ1 σ2 σ3 and replace in equation (6) Ri by Ti /µ.

6
Equation (4) maps not only isochoric but also volume-reducing proportional strain
paths to the corresponding proportional stress paths. It proves that this mapping gets
more realistic if we let a depend on dilatancy , e.g. as a = c1 exp(c2 ). This modification
does not affect isochoric (i.e. undrained) proportional strain paths with  = 0. Thus, we
replace equation (4) by equation

R(D) = − exp(c1 · exp(c2 ) · D0 ) (9)

Volume increasing proportional strain paths are investigated by Medicus (Medicus,


2014).

5 Proportional paths B
Now we start from a stress state T 6= 0 and apply the stretching D. In order to asymptot-
ically approach the corresponding proportional stress path T = µR(D), the stress rate
Ṫ must be oriented towards a point of the proportional stress path with the direction
R(D), i.e.
T + λṪ = µ1 R(D), (10)
where the positive constants µ, µ1 and λ need not be further specified at the moment.
If we eliminate Ṫ we obtain an evolution equation for the stress:

Ṫ = ν1 R(D) + ν2 T (11)

with scalar quantities ν1 and ν2 . Equation (11) is already the general form of the barode-
tic constitutive relation. To comply with barotropy (i.e. dependence on σ), pyknotropy
(i.e. dependence on e) and rate independence, ν1 and ν2 can be further specified, such
that the barodetic constitutive equation for sand obtains the following specific form:

Ṫ = h · (f R0 + gT0 ) · ε̇ . (12)

The quantities f and g will be specified below. h is responsible for the stiffness and
depends on σ. In the sequel it will be shown how all known concepts of soil mechanics
can be cast in the frame given by equation (12).

6 Limit states and peaks


A limit state is defined by vanishing stiffness:

Ṫ = 0. (13)

Considering stress-strain curves, limit states are manifested either as peak or residual
(critical) limit states, where the curves obtain a horizontal slope. In barodesy, Ṫ = 0
implies:
f R0 + gT0 = 0. (14)

7
This tensorial equation implies:6
R0 = T0 (15)
f +g = 0 (16)
Interestingly, equation (15) contains a flow rule, i.e. a stress-dilatancy relation for peak
states. Critical limit states are obtained with  = 0 and e = ec , whereas peak limit
states are obtained with  > 0 and e < ec . A simple way to fulfill these requirements
for limit states with equation (16) is to set
f + g =  + c3 (ec − e) (17)
with ec being the critical void ratio (see equation 26).

7 Incremental non-linearity
Incremental non-linearity (or ’non-linearity in the small’) means different stiffnesses
at loading and unloading and, consequently, irreversible or hysteretic mechanical be-
haviour. Both, elastoplastic and hypoplastic relations exhibit incremental non-linearity.
The elastoplastic approach consists in introducing (at least) two different stiffnesses, one
for loading and one for unloading. A criterion has to be added to distinguish loading
from unloading. In the frame of hypoplasticity and barodesy, a unique expression for
the stress rate (or stiffness) is used, and the distinction between loading and unloading
is accomplished by the non-linearity of this equation. In barodesy, the difference of stiff-
ness at loading and unloading is modelled by the fact that the second term (i.e. gT0 ) in
equation (12) is not changed if D is switched to −D, whereas the first term (i.e. f R0 )
undergoes a change.

8 Consolidations and Critical State Line


We consider proportional strain paths starting at zero stress, which we call (virgin)
consolidations. Given the low compressibility of sand, experimental results on the stress
strain behavior of consolidations are rather meagre and/or obscured by natural scatter.
However, barodesy leads to interesting results merely by reasoning (this is an example
of how a theory can shed light also into questions that can be hardly answered by
experiment).
Noting that T0 = R0 holds true for consolidations, we obtain from equation (12)
Ṫ = h T0 (f + g) ε̇ (18)
For proportional paths holds also Ṫ = σ̇T0 , hence equation (18) reduces to
σ̇ = h (f + g) ε̇ , (19)
6
The other two possibilities, (i) R0 = −T0 and f −g = 0, as well as (ii) f = 0 and g = 0, can be excluded,
because R always points to the compression octant, and f and g cannot vanish simultaneously, see
equations 29 and 30.

8
Using equation (17) and the relation

trD ė
 ε̇ = ε̇ = trD = , (20)
ε̇ 1+e
which holds for incompressible grains, we obtain for consolidations:
1 ė
σ̇ = h [ + c3 (ec − e)] · · . (21)
 1+e
or
de 1+e 1
= ec −e · (22)
dσ 1 + c3  h
Equation (22) expresses the compressibility at consolidations. h(σ) expresses the stress-
dependence of stiffness (or of compressibility, which is the inverse of stiffness). It is
known that the stiffness of soil increases sub-linearly with σ (Hardin & Richard, 1963;
Janbu, 1963). However, it should not vanish for σ = 0, as is the case with hypoplastic
and previous barodetic formulations,7 because this leads to problems in numerical sim-
ulations of boundary value problems with free surface. Furthermore, it is reasonable to
set a lower bound, emin for void ratio. Clearly, emin is prescribed by the geometry of the
granulate, mainly by the grain size distribution, and (contrary to some hypoplastic for-
mulations e.g. (Gudehus, 2004)) does not depend on stress, if we exclude grain crushing
from our consideration. Note that the value of emin can be considerably smaller than
the conventionally obtained one.
These requirements can be fulfilled e.g. by the relation
c4 + c5 σ
h=− . (23)
e − emin
Equation (23) has the advantage that the void ratio e tends to emin for σ → ∞ and thus
guarantees e ≥ emin . Obviously, the denominator of h has been chosen in such a way
as to vanish for e = emin . Clearly, other expressions, such as (e − emin )β , can possibly
provide better simulations, but they have not been investigated in this paper in view of
the difficulties related with calibration (Section 10).
The fraction in equation (22)
ec − e
(24)

deserves particular attention. It points to the fact that the compressibility of sand
depends also on the dilatancy . Let us distinguish between dense (e < ec ) and loose
(e > ec ) sand. At shearing, dense sand has a tendency to loosening and loose sand has a
tendency to get denser, in other words, a loose sand is more compressible than a dense
one. In accordance to this fact, equation (22) implies that for consolidations carried out
with constant ,

• the compressibility of dense sand decreases with increasing 


7 √
which adopted the widespread relation of stiffness being proportional to σ

9
• the compressibility of loose sand increases with increasing 
Note that increasing  means increasing deviatoric part of the deformation.
Even more interesting is the special case of equation (22) for e = ec and  = 0. The
aforementioned fraction then obtains the non-defined value 0/0, and the compression
relation e(σ) refers to an ’isochoric (or undrained) compression’, which is clearly a con-
tradictional notion. However, the related curve e(σ) is nothing but the so-called Critical
State Line (CSL). Let us first consider the aforementioned fraction in detail. We set
e = ec and observe that the fraction has the value 0 for all consolidations  < 0. There-
fore, it appears reasonable to set its value equal to 0 also for  = 0. Introducing this into
equation (22) we obtain the following differential equation for the CSL:
dσ c4 + c5 σ
=− (25)
dec (1 + ec )(ec − emin )
which can be easily integrated to yield the following equation of the CSL:
emin + B
ec = (26)
1−B
with the abbreviation
− 1+emin
ec0 − emin c4 + c5 σ

c5
B := . (27)
ec0 + 1 c4
In other words, the critical state line ec (σ) is contained in barodesy and can be derived
from it.
The implications of barodesy on consolidations are elucidated in Fig. 2, where the
following compression lines e(σ) are plotted:
dashed (red): hydrostatic compression
full line (blue/black): deviatoric compression, i.e. compression with a large deviatoric
component,  
−2 0 0
D =  0 0.7 0  (28)
 
0 0 0.7
dash-dotted (black): critical state line, ec (σ)
The red and blue curves start at a
dense state: e = 0.75 < ec0 (= 0.90)
loose state: e = 0.95 > ec0 (= 0.90).
An interesting implication of equation (22) is obtained when the denominator vanishes.
This is obtained for a particular void ratio e = ê, that depends on . For this void ratio,
the slope de/dσ (which is negative) vanishes, and for e > ê this slope becomes positive.
This means that if we start a virgin compression at a too high (and, hence, non-feasible)
void ratio, then the simulated compressibility is zero or even negative. This implies that
unfeasibly high void ratios e > ê (with ê being the maximum void ratio) can be detected
by barodesy.

10
0.95 0.95
hydrostatic compression hydrostatic compression
deviatoric compression deviatoric compression
0.9 CSL 0.9 CSL

0.85 0.85
e

0.8 0.8

0.75 0.75

0.7 0.7

0.65 0.65
0 2000 4000 6000 8000 10 2 10 3 10 4
σ σ

Figure 2: Compression curves calculated with barodesy (see Section 8), calibrated as
shown in Section 10 (natural and semilogarithmic plots). Note that these
curves are not affine.

11
9 Barodetic constitutive equation
The final step in specifying the barodetic constitutive equation consists in partitioning
equation (17) into f and g. A possibility is:

f =  + c3 ec (29)
g = −c3 e (30)

such that the barodetic constitutive equation for sand finally reads:
c4 + c5 σ h i
T̊ = − ( + c3 ec )R0 − c3 eT0 ε̇ , (31)
e − emin
where R is given by equation (9) and ec is given by equation (26) .

10 Calibration
In general, calibration, i.e. the determination of the material constants of a constitu-
tive relation on the basis of experimental results, is difficult as neither the experimental
results nor the constitutive relation mirror perfectly the reality. In terms of approxi-
mation theory, the space composed of values (T, D, e) is not a Haar-space (Wendland,
2005), hence calibration of the material constants c1 , c2 , . . . , c5 based on some points of
this space will not necessarily give the same values as if it were based at some other
points. Scattering data impede not only calibration but also the subsequent validation
of a constitutive model on the basis of experimental results, and this illuminates what
we should expect from a constitutive relation, extreme precision rather or clarity and
sound prediction of tendencies?
In the shown equations of barodesy appear two quantities, ec and emin , that are not
directly measurable (so-called internal variables). Note that the critical void ratio ec
is not a constant but depends on σ. The experimental determination of the relation
ec (σ) is difficult, cf. the statement by D. Muir Wood (Muir Wood, 1990): The paths of
tests on loose and dense samples head towards a somewhat diffuse, but clearly pressure
dependent, zone of critical void ratios.
In view of this difficulty, a tentative calibration for Hostun sand (Vermeer, 2000),
is presented here, whereas an optimized and robust calibration procedure for barodesy,
possibly based on new type experiments, is postponed to a forthcoming paper. It consists
of the following steps:
1. The critical friction angle is obtained from the undrained (D2 /D1 = −0.5) triaxial
test ’host-l-triaxc-cu-600 (cn dhfl 61)’ (with initial lateral (effective) stress σ2 =
−600 kPa) as ϕc = 33.8◦ . Hence Kc := σ2 /σ1 = 0.285. Using equation (9) we
obtain: r
2
c1 = ln Kc = −1.025 (32)
3
c2 is determined by trial and error to c2 ≈ 0.8 on the basis of the curvature of the
numerically obtained stress-strain curves of triaxial tests.

12
2. To calibrate the constants c4 and c5 we resort to the oedometric compression test
’host-l-oed’.

Table 1: Experimentally obtained relation between e and σ for Hostun sand from test
’host-l-oed’

e 0.8703 0.8609 0.8516 0.8478 0.8398 0.8291


σ (kPa) 0 10.8 39.5 59.9 119.9 246.9

In view of the aforementioned difficulty with respect to ec , we neglect the fraction


(24) in equation (22), i.e. we set it equal to 0. This is admittedly imprecise, but it
is hoped that it will help to roughly fix c4 and c5 . The advantage is that it renders
equation (22) integrable leading to the same e(σ)-relation as the ec (σ)-relation of
equation (26). Fitting to the obtained results shown in the table 1 (using the
MATLAB program fsolve) yields:

c4 ≈ 465 kPa (33)


c5 ≈ 28 (34)

3. The aforementioned difficulty with respect to the CSL makes the calibration of the
remaining constants c2 and c3 impossible without further assumptions. Therefore,
we directly proceed to their determination by simple trial and error, obtaining thus

c2 ≈ 1 (35)
c3 ≈ −2.3 . (36)

10.1 On the physical meaning of material constants


The often encountered question ”what is the physical meaning of a material constant?”
is not well-posed. Material constants do not need to have a particular physical meaning,
they simply serve to adjust an equation to the outcomes of experiments obtained with
a particular material. For some simple cases,8 a material constant is the outcome of a
single experiment, and this gives rise to the expectation that every material constant
should be the outcome of a single experiment, having thus a ’physical meaning’.

11 Cyclic loading, limit cycles and shake-down


Proportional paths are not the only attractors in the constitutive relation presented
so far. Being ordinary differential equations, constitutive relations ”of the rate type”

8
E.g., in barodesy, where c1 is uniquely related to the critical friction angle ϕc by means of equation
32.

13
(Truesdell & Noll, 1965) may exhibit also limit cycles or cyclic orbits as further attrac-
tors. In fact, equation 12 exhibits periodic orbits (limit cycles) at cyclic loading. In
terms of mechanics, this effect is called ’shake-down’ and implies that stress cycles lead
asymptotically to cyclic changes of void ratio. In case of e.g. oedometric deformation
(but not for conventional triaxial tests), this implies also cyclic strain, i.e. strain due to
cyclic stress is bounded, i.e. it does not increase to infinity. Generally, shake-down is one
of the possible responses of sand to cyclic loading, the other one being ’incremental col-
lapse’ (i.e. unlimited growth of strain with increasing number of cycles). It is yet unclear
when exactly shake-down and when incremental collapse are to be expected. However,
equation 12 exhibits shake-down (and periodic orbits) e.g. at cyclic oedometric loading:
If the axial stress component σ1 is periodically changed between a lower and an upper
limit, then the corresponding radial stress component σ2 , which should be bounded, will
also become cyclic, i.e. a limit cycle will eventually be obtained in the stress space.9
Cyclic stress, asymptotically obtained with strain cycles of infinitesimally small am-
plitude, is related with a void ratio ě, which can be called the cyclic void ratio. Little is
known from experiments on the dependence of ě on actual stress T and on the direction
D0 of strain cycles.
Considering strain cycles with infinitesimal amplitude with the constitutive relation
12 and denoting with ”+” and ”−” loading and unloading, respectively, at a limit cycle
must hold: Ṫ+ = −Ṫ− . Hence, the condition for cyclic response reads

(f + R0+ + f − R0− ) + 2gT0 = 0 . (37)

This equation constitutes a relation between the direction of the strain amplitude, D,
the cyclic void ratio ě and the stress σT0 , around which the stress oscillation occurs.
Eliminating T0 from equation 37 yields:
−1 + 0+
T0 = (f R + f − R0− ) . (38)
2g

Using this equation and the additional condition |T0 | = 1, makes it possible to determine
for a given D0 the stress direction T0 of the corresponding cyclic state and also the
pertaining cyclic void ratio ě(σ). Figure (3) shows the numerical simulation of the
evolution of void ratio with small (i) oedometric and (ii) simple-shear stress cycles, as
obtained with barodesy. Interestingly, cycles applied to dense samples lead to loosening.
It should be added that barodesy exhibits ratcheting at cycles of small amplitude e.g. in
conventional triaxial tests.
The implications of this section could be far-reaching, but their detailed study is left
for a future paper.

9
Integrity of grains (or permanence of the grain size distribution) has not been assumed for the deriva-
tion of the constitutive relation so far, and a constitutive relation that does not contain the strength
of grains excludes grain crushing. In reality, of course, grain crushing is inevitable, especially at
higher stresses.

14
0.95

0.9

0.85
e

0.8

0.75

0.7
0 1000 2000 3000 4000 5000 6000 7000 8000
number of time steps

Figure 3: Evolution of void ratio with small stress cycles. Oedometric stress cycles (blue;
they converge to the upper limit), simple-shear stress cycles (red; they converge
to the lower limit)

12 Simulation of element tests


The simulation of element tests can be achieved with numerical integration of the pre-
sented constitutive relation. The initial void ratio einit must be known at each test.
Then, the stress and strain paths can be traced, e.g. with Euler forward integration,
which is perhaps not very efficient, but good enough for this purpose, provided that
the time steps are sufficiently small. Kinematically determined tests (such as oedo-
metric and undrained triaxial tests) are simple to simulate, because the full D-tensor is
given (its value, |D|, is immaterial, as long as rate-independent constitutive relations are
used). For mixed boundary conditions, a stretching component (e.g. the lateral stretch-
ing D2 ≡ D3 for conventional triaxial test) has to be determined at each step so as to
fulfil the pertinent stress condition (e.g. σ2 ≡ σ3 =const).
This is a ’proof-of-principle’ paper. Its intention is to demonstrate a new constitutive
frame rather than to show a perfect coincidence of simulations and measurements. Hence,
a systematic comparison of the predictions of the here presented theory with real tests
and other theories, accompanied with appropriate calibration, is here omitted. The same
holds for a sensitivity analysis to check the role of the included material parameters.
Only a few comparisons of simulations with experimental results on Hostun sand have
been added (see Figures 8, 10 and 11) to provide a sense of the predictive capacities
and also of some deficiencies, that refer mainly to the stiffness at oedometric tests. The
deficiencies can be attributed partly to the still incomplete calibration and partly to the
model itself. Both shortages, so the author believes, can be removed by further research.

15
600

500

400
σ1 (kN/m2)

300

200

100

0
0 100 200 300
2 400 500 600
σ2 (kN/m )

Figure 4: Stress paths corresponding to PεP’s starting from a non-vanishing stress state.
Note that stress paths corresponding to consolidations asymptotically ap-
proach PσP’s, whereas stress paths corresponding to dilations (i.e. ė > 0)
tend to T = 0.

12.1 Proportional strain paths


Applying various PεP’s starting at a non-vanishing stress state produces a fan of curved
stress paths, all of which have to stay within the allowed range (fan) in the stress space.
The corresponding graph (Figure 4) is a visualisation of the performance of a constitutive
equation. To check rule 2, in Figure 5 are shown the stress paths corresponding (i) to
hydrostatic compression (as PP) and (ii) to hydrostatic compression following oedometric
compression.

12.2 Response envelopes


The response envelopes, introduced by (Gudehus, 1979), offer a simple visualization of
the material response for axisymmetric cases. Starting from an axisymmetric stress
 
σ1 0 0
T =  0 σ2 0  (39)
 
0 0 σ2

16
4000 250

3500

200
3000
σ1 (kN/m )

σ1 (kN/m )
2

2
2500
150

2000

100
1500

1000
50

500

0 0

σ2 (kN/m2) σ2 (kN/m2)
0 500 1000 1500 2000 2500 3000 3500 4000 0 50 100 150 200 250

Figure 5: Numerical simulation of Gold- Figure 6: Response envelopes.


scheider’s rule 2 obtained with
barodesy

we apply stretchings
 
− sin α 0 √ 0
D= 0 − cos α/ 2 0 √  0◦ < α ≤ 360◦ (40)
 
0 0 − cos α/ 2

and plot the resulting stress rates Ṫ in a polar diagramm. The response envelopes at
three different stress states, all of which have the same trT and the same void ratio e, are
shown in Fig. 6. Note that response envelopes are nothing but a convenient visualisation
of the response according to a particular constitutive model. Their experimental deter-
mination is very difficult and there are not known any theoretical requirements on their
shape (such as convexity etc.). However, it appears reasonable that a good constitutive
relation yields response envelopes of a simple form (e.g. without intersecting themselves).
In linear elasticity, response envelopes are ellipses centered at the reference stress state.
In hypoplasticity, they are ellipses whose center is not occupied by the actual stress
state. Barodesy yields the shown lemon-shaped curves.

12.3 Oedometer test


The simulation results are shown in Fig. 7. The initial stress has been chosen to σ1 =
−100 kN/m2 , σ2 = σ3 = K0 σ1 , σij = 0 else. In comparison with experimental results
the simulations with barodesy are still not satisfactory (Fig. 8). A scaling factor could
help, but it is the general intention of this approach to keep the constitutive relation
simple and not to burden it with factors for individual effects.

12.4 Drained triaxial test


Considering a constitutive relation of the form Ṫ = h(T, D, e) we have to use such a
stretching tensor D that the equation σ̇2 = σ̇3 = 0 is fulfilled. Herein, the component

17
0 500

0.01 400

σ11 (kN/m2)
0.02 300
ε11

0.03 200

0.04 100

0.05 0
0 100 200 300 400 500 0 100 200 300 400 500
2
σ11 (kN/m ) σ22 (kN/m2)

0.048 0.86

0.046 0.84
0.044
0.82
ε11

0.042
e

0.8
0.04

0.038 0.78

0.036 0.76
0 0 200 400 600 800 1000
10
number of load cycles number of time steps

Figure 7: Oedometric test, stress-strain curve, stress path and strain in dependence of
cycle number.

18
σ (kPa)
0 −100 −200 −300
0 0.88
Barodesy
−1 experiment 0.86

−2 0.84
ε (%)

e
−3 0.82

−4 0.8

−5 0.78
0 2
−10 −10
σ (kPa)

Figure 8: Oedometric test, comparison with experimental results (Hostun sand, loose).

D1 can be set to −1 for compression and 1 for extension. Due to rate-independence, it


matters only its sign but not its value. The unknown D2 (= D3 ) has then to be obtained
at each time step by solving the algebraic equation σ̇2 (D2 ) = 0. Note that experimental
results can deviate from the simulation in the post-peak regime due to inhomogeneous
deformation.

12.5 Undrained triaxial test


The simulation results are shown in Fig. 12, 13 and 14. The bend at the so-called ’phase
transformation’ is not modeled by the present version of barodesy, obviously due to the
simplicity of its algebraic structure.

12.6 Simple shear test


The so far considered element tests are rectilinear extensions, i.e. they do not comprise
any rotation. In contrast, simple shear tests comprise rotations. This implies that the
stretching D has non-vanishing components off the principal diagonal. The exponential
mapping in the R-function is still applicable, usual codes such as MATLAB yield the
values of the resulting matrix without any problem. A major difficulty however arises
now with respect to the objective time rate of stress T. In hypoplasticity and also in
previous versions of barodesy, the Zaremba-Jaumann co-rotated stress rate was used.
However, there is a vast literature on its appropriateness and on possible alternatives
(Chein-Shan Liu & Hong-Ki Hong, 1999, 2001; Dafalias, 1983; Meyers Xiao & Bruhns,
2003; Naghdabadi & Sohrabpour, 2003; Romano & Barretta, 2013; Xiao, Bruhns &

19
1200 1
1000 0.95
σ11 -σ22 (kN/m 2)

800 0.9
600 0.85

e
400 0.8
200 0.75
0 0.7
dense dense
-200 loose 0.65 loose
-400 0.6
0 0.05 0.1 0.15 0.2 0 0.05 0.1 0.15 0.2
ǫ 11 ǫ 11

0.05 1
0.8
0.6
0.4
sinφm

0.2
ǫv

0
0
-0.2
-0.4
dense dense
loose -0.6 loose
-0.05 -0.8
0 0.05 0.1 0.15 0.2 0 0.05 0.1 0.15 0.2
ǫ 11 ǫ 11

Figure 9: Drained triaxial tests, solid lines: initially dense, dotted lines: initially loose.
Lateral stress σ2 = σ3 = −300 kPa, initial void ratios e = 0.63 and e = 0.85.

20
5
600
σ1−σ3 (kPa)

εvol (%)
0
400

−5
200 Barodesy Barodesy
experiment experiment
−10
0
0 5 10 15 0 5 10 15
ε1 (%) ε1 (%)

Figure 10: Drained triaxial test, comparison with experimental results (Hostun sand,
loose).

20
1000

800
σ1−σ3 (kPa)

10
εvol (%)

600

400 0

200 Barodesy Barodesy


experiment experiment
−10
0
0 10 20 30 0 10 20 30
ε1 (%) ε1 (%)

Figure 11: Drained triaxial test, comparison with experimental results (Hostun sand,
dense).

21
2500 2000
σ11−σ22 (kN/m )
2

2000
1500

σ11 (kN/m2)
1500
1000
1000

500
500

0 0
0 0.01 0.02 0.03 0 500 1000 1500 2000
ε11 σ (kN/m2)
22

2000 0.7

0.6
1500
0.5
q (kN/m2)

sinφm

0.4
1000
0.3

0.2
500
0.1

0 0
0 500 1000 1500 2000 0 0.01 0.02 0.03
2 ε11
p´ (kN/m )

Figure 12: Undrained triaxial tests with void ratios varying between e = 0.65 and e =
1.0.

22
100 400

50 300
σ11−σ22

σ11
0 200

−50 100

−100 0
−0.05 0 0.05 0.1 0.15 0.2 0 100 200 300 400
ε σ
11 22

100 60

40
50
(σ11−σ22)/2

20
q

0
0

−50
−20

−100 −40
0 50 100 150 200 50 100 150 200 250
p´ (σ11+σ22)/2

Figure 13: Cyclic undrained triaxial test (e = 0.65). The starting point (σ1 = σ2 = σ3 =
−200) kPa is denoted with a circle.

23
40
120

20 100
σ11−σ22

80

σ11
0
60

40
−20
20

−40 0
−0.06 −0.04 −0.02 0 0.02 0 50 100
ε σ
11 22

40 20
(σ11−σ22)/2

20 10
q

0 0

−20 −10

−40 −20

0 20 40 60 80 100 0 20 40 60 80 100 120


p´ (σ11+σ22)/2

Figure 14: Cyclic undrained triaxial test (e = 0.85). The starting point (σ1 = σ2 = σ3 =
−200) kPa is denoted with a circle.

24
50 100

σ12 (kN/m ) 90

σ11 (kN/m2)
80
2

0 70

60

50

−50 40
−0.015 −0.01 −0.005 0 0.005 0.01 0.015 −0.015 −0.01 −0.005 0 0.005 0.01 0.015
ε12 ε
12

Figure 15: Undrained simple shear test (e = 0.83). Left: shear stress σ12 vs. shear
strain ε12 , right: normal stress σ11 vs. shear strain ε12 . The starting point is
denoted with a circle.

Meyers,1997; Zhou & Tamma, 2003). Most of the literature is inconclusive and can’t be
checked with experiments, as it is not possible to conduct element tests with large shear.
For the usual types of simple shear, also for the ones simulated here, the differences
between Ṫ and the several co-rotated stress rates T̊ are altogether small. For cases with
considerable shear, however, a thoroughly defined stress rate has to be applied (Romano
& Barretta, 2011).
The numerically obtained simulation of an undrained simple shear test is shown in
Fig. 15. A drained simple shear is driven by applying D12 6= 0, whereas D22 has to be
determined at each time step in such a way that the condition σ22 =const is fulfilled.
Consequently, a vertical strain ε22 6= 0 is obtained. The results of the numerical simula-
tion are shown in Figures 16 and 17. It is observed that cyclic shear leads to stationary
states (limit cycles), the void ratio of which decreases with increasing normal stress |σ22 |.

13 Conclusions
The present version of barodesy cannot cover all aspects of sand behavior, e.g. the so-
called phase transformation of undrained tests. The memory of past loading is contained
only in the actual stress T and the actual porosity e, and this is not sufficient to cover
all aspects of re-loading, in particular the so-called aspects of ’small strain stiffness’.
Though, it is interesting to note how many aspects of memory can be covered with T
and e only. The cyclic void ratio ě, as asymptotically obtained for small loading cycles
with barodesy, is higher than emin . This is, perhaps, reasonable but implies that cyclic
tests carried out with dense sand tend to dilation, i.e. they lead to upheavals of the

25
400 0.75

300
0.7495
200
σ12 (kN/m2)

100 0.749
0

e
−100 0.7485

−200
0.748
−300

−400 0.7475
−0.03 −0.02 −0.01 0 0.01 0.02 0.03 −0.03 −0.02 −0.01 0 0.01 0.02 0.03
ε12 ε12

Figure 16: Drained simple shear test. σ22 = −500 kN/m2 . The starting point is denoted
with a circle. Initial void ratio e = 0.75.

400 0.95

300
0.9
200
σ12 (kN/m2)

100 0.85
0
e

−100 0.8

−200
0.75
−300

−400 0.7
−0.03 −0.02 −0.01 0 0.01 0.02 0.03 −0.03 −0.02 −0.01 0 0.01 0.02 0.03
ε12 ε12

Figure 17: Drained simple shear test. σ22 = −500 kN/m2 . The starting point is denoted
with a circle. Initial void ratio e = 0.95.

26
sample.
A point of concern refers to the volumetric behavior predicted by barodesy. E.g. in
oedometric compression, barodesy predicts a far too soft behavior. In addition, it is
observed in the laboratory that dilatancy turns to strong contractancy upon loading
reversals of deviatoric deformation. This is reproduced by barodesy for simple shear
tests but not for triaxial tests.
In total, barodesy may be considered as attractive for its simplicity and elegance but
has still potential for improvement, which should however preserve simplicity (that is
not simple to achieve).

Acknowledgment
The author is indebted to DI Getraud Medicus and to Prof. Wolfgang Fellin for valuable
discussion.

References
E. Bauer. Calibration of a comprehensive hypoplastic model for granular materials. Soils
and Foundations, 36(1):13–26, 1996.

Bauer E., The critical state concept in hypopasticity. In: Comp. Meth. and Adv. in
Geomechanics, Yuan (ed.) Balkema 1997, 691-696

Chein-Shan Liu and Hong-Ki Hong, Non-oscillation Criteria for Hypoelastic Models
under Simple Shear Deformation, Journal of Elasticity 57: 201-241, 1999

Chein-Shan Liu and Hong-Ki Hong, Using comparison theorem to compare corotational
stress rates in the model of perfect elastoplasticity, Int. J. Solids and Structures 38
(2001) 2969-2987

Dafalias Y.F., Corotational rates for kinematic hardening at large plastic deformations,
J. Appl, Mech. 50, 561-565, 1983

Fellin, W. and Ostermann A.: The critical state behaviour of barodesy compared with
the Matsuoka-Nakai failure criterion. International Journal Numerical and Analytical
Methods in Geomechanics (2011) DOI: 10.1002/nag.1111.

Fellin, W. (2013): Extension to barodesy to model void ratio and stress dependency of
the K0 value. Acta Geotechnica 8(5): 561-565.

Goldscheider M., Grenzbedingung und Fließregel von Sand, Mech. Res. Comm., Vol. 3,
463-468, 1976

Gudehus, G., A comparison of some constitutive laws for soils under radially symmetric
loading. 3rd Int. Conf. Num. Meth. in Geomech., Aachen, 2-6 April 1979, 1309-1323.

27
G. Gudehus. A comprehensive constitutive equation for granular materials. Soils and
Foundations, 36(1):1–12, 1996.

Gudehus G., A visco-hypoplastic constitutive relation for soft soils. Soils and Founda-
tions, 2004, 44(4):11-26.

Hardin, B. O., and Richard, F. E., Jr. (1963). Elastic Wave Velocity in Granular Soils,
Journal of the Soil Mechanics and Foundations Division, ASCE, Vol. 89, No. SM1,
pp. 33-65.

Herle I. and Kolymbas D., Hypoplasticity for soils with low friction angles, Computers
and Geotechnics 31 (2004) 365-373

Janbu, N., Soil compressibility as determined by oedometer and triaxial tests. In Pro-
ceedings of the European Conference on Soil Mechanics and Foundation Engineering,
1963

Kolymbas D., A rate-dependent constitutive equation for soils. Mech. Res. Comm. 4:367-
372 (1977)

Kolymbas D., A generalised hypoelastic cosntitutive law. In: Proc. XI Int. Conf. Soil
Mechanics and Foundation Engineering, Vol. 5, p. 2626, San Francisco 1985. Balkelma

Kolymbas, D., The misery of constitutive modelling. In: D. Kolymbas (ed.) ”Constitutive
Modelling of Granular Materials”, Springer, 2000, 11-24

Kolymbas, D.: Sand as an archetypical natural solid. In: D. Kolymbas and G. Viggiani
(eds.), Mechanics of Natural Solids, Springer, 2011, pp. 1-26

Kolymbas, D.: Barodesy: A new hypoplastic approach. International Journal for Nu-
merical and Analytical Methods in Geomechanics, 2011, DOI: 10.1002/nag.1051

Kolymbas D., Barodesy: a new constitutive frame for soils, Géotechnique Letters 2,
17-23, 2012

Kolymbas D., Barodesy as a novel hypoplastic constitutive theory based on the asymp-
totic behaviour of sand, Geotechnik, 35 (2012), Heft 3, 187-197

Kolymbas D., Barodesy: The Next Generation of Hypoplastic Constitutive Models for
Soils, In: G. Hofstetter (ed.), Computational Engineering, DOI 10.1007/978-3-319-
05933-4 2, Springer, 2014

Mašı́n D., Hypoplastic Cam-clay model, Géotechnique 62, No. 6, 2005; 549-553

Mašı́n D., Asymptotic behaviour of granular materials. Granular Matter, 14(6):759-774,


2012.

Medicus G., Barodesy and its application for clay. PhD thesis. Faculty for Technical
Sciences, University of Innsbruck, 2014.

28
Meyers A., Xiao H. and Bruhns O., Elastic Stress Ratchetting and Corotational Stress
Rates, Technische Mechanik, Band 23, Heft 2-4, (2003), 92-102

Muir Wood, D., Soil Behaviour and Critical State Soil Mechanics, Cambridge University
Press, 1990

Naghdabadi, R., Sohrabpour, S., Corotational Constitutive Modeling of Isotropic and


Kinematic Hardening Materials, Scientia Iranica Vol. 10, No. 1, 2003, 56-63

A. Niemunis and I. Herle. Hypoplastic model for cohesionless soils with elastic strain
range. Mechanics of Cohesive-Frictional Materials, 2(4):279–299, 1997

Niemunis, A. (2003): Extended hypoplastic models for soils. Heft 34, Schriftenreihe des
Inst. f. Grundbau u. Bodenmechanik der Ruhr-Universitaet Bochum

Palmer A.C. and Pearce J.A., Plasticity theory without yield surfaces. In: Symposium
on Plasticity and Soil Mechanics, (Ed. A.C. Palmer), 188-200, Cambridge 1973.

Romano G. and Barretta R., Covariant hypo-elasticity, European Journal of Mechanics


A/Solids 30 (2011) 1012-1023

Romano G. and Barretta R., Geometric constitutive theory and frame invariance, Inter-
national Journal of Non-Linear Mechanics 51(2013) 75-86

Topolnicki M. Observed stress-strain behaviour of remolded saturated clay and examina-


tion of two constitutive models. Veröffentlichungen des Instituts für Bodenmechanik
und Felsmechanik der Universität Fridericiana in Karlsruhe, 1987

Truesdell, C.A., Noll W., The Non-Linear Field Theories of Mechanics. In: Encyclopedia
of Physics, Vol. IIIc, Springer 1965.

Vermeer P.A. (ed.), Database for Tests on Hostun RF Sand, J. Desrues, B. Zweschper,
P.A. Vermeer, Institutsbericht 13, 2000, Institut für Geotechnik, Universität Stuttgart.

Wendland, H., Scattered Data Approximation, Cambridge University Press, 2005

P.-A. von Wolffersdorff. A hypoplastic relation for granular materials with a predefined
limit state surface. Mechanics of Cohesive-Frictional Materials, 1, 1:251–271, 1996.

Wu W., Hypoplastizität als mathematisches Modell zum mechanischen Verhalten gran-


ularer Stoffe. Veröffentlichungen des Instituts für Bodenmechanik und Felsmechanik
der Universität Fridericiana in Karlsruhe, Heft 129, 1992

Wu, W., Bauer, E., Kolymbas, D.: Hypoplastic constitutive model with critical state for
granular materials. Mechanics of Materials, 23, 1996, 45-69

Xiao H., Bruhns O. T. and Meyers A., Logarithmic strain, logarithmic spin and loga-
rithmic rate, Acta Mechanica 124, 89-105 (1997)

29
Zhou, X., Tamma, K.K., On the applicability and stress update formulations for corota-
tional stress rate hypoelasticity constitutive models, Finite Elements in Analysis and
Design 39 (2003) 783-816

30

You might also like