Thanks to visit codestin.com
Credit goes to www.scribd.com

0% found this document useful (0 votes)
5 views55 pages

Binder 1

The document provides an overview of viscous stress, focusing on its definitions, views, and implications in fluid dynamics, particularly in granular materials. It discusses the macroscopic and microscopic perspectives of viscosity, including the relationship between stress and velocity gradients, and introduces relevant equations such as Newton's Law of viscosity. The author invites readers to engage further through email or a discussion forum for more detailed inquiries.

Uploaded by

hidayat tullah
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
5 views55 pages

Binder 1

The document provides an overview of viscous stress, focusing on its definitions, views, and implications in fluid dynamics, particularly in granular materials. It discusses the macroscopic and microscopic perspectives of viscosity, including the relationship between stress and velocity gradients, and introduces relevant equations such as Newton's Law of viscosity. The author invites readers to engage further through email or a discussion forum for more detailed inquiries.

Uploaded by

hidayat tullah
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 55

All I wanna know ’bout Viscous Stress

(but I never dare to ask !)


Have a nice Night!

You will find the basic facts about viscous phenomenon - no details -. If you wanna know more just email me or feel
free to ask in the Discussion Forum. I purposely erased all the bibliographical references and detailed equations to
keep the text simple and easy to read.

If you need an official reference for the content of this website, please, use:
Dartevelle, S., Numerical and granulometric approaches to geophysical granular flows, Ph.D. thesis,
Michigan Technological University, Department of Geological and Mining Engineering, Houghton, Michigan,
July 2003.

You will find herewith:


I. Introduction: Have you said viscosity?
II. Stress view of viscosity
III. Granular kinetic view of viscosity
IV. Viscosity and Tensors
V. Bulk viscosity, Shear viscosity and Second coefficient of viscosity
VI. The concept of viscosity within rigid-plastic granular flows

Please, don’t forget to sign our Granular Volcano Guest Book.

Done and updated by Sébastien Dartevelle, The WebMaster, Saturday November 22 2003.

I. Introduction: Have you said viscosity?

After all, what is viscosity?

Hmmmm, there are many different ways to answer to this question. When I was a High-School
student, my science teacher said that viscosity is a property that owns any fluid and makes it
somehow resistant to the flowing process. When I was Bachelor student, I got another picture of the
viscosity. I was told that the viscosity is the tendencies towards uniformity of mass-velocity
(momentum) throughout the flow. Such tendencies are explained by kinetic theory which attributes
this phenomenon to the motion of molecules or grains from point to point.

I have been long a wee bit confused about those two views or explanations of the phenomenon of
viscosity because I could not reconcile what was for me two different stories. But believe me, it is
the same story but seen from a different point of view. The "resistance to flow" point of view is a
global macroscopic force balance view, while the "mass-velocity uniformity" is a molecular-granular
microscopic view of the viscosity phenomenon.

Herewith, I will only consider the viscosity concept for granular materials, and since granular theory
has evolved from gas-kinetic theory, the following paragraphs also apply to the gas viscosity
concept.

[Back to Top]

II. Stress view of viscosity

A moving fluid exhibit evidences of internal friction (friction is a good word, it gives us somehow the
feeling of resistance), called viscosity. To see this, let’s consider a moving fluid where the velocity is
minimal at the bottom and tend to be higher towards the top of the fluid:
[Figure 1a: the Y axis is perpendicular to your screen, hence the XY plane is parallel to your table and the ZX plane is nothing but
your screen]
[Figure 1b: Variation of the velocity in the X-direction along the Z axis]
[Be aware that those two figures do not exactly say the same thing, see their respective axis. However, those two figures show that
the X-velocities (velocities parallel to the X-direction) are not the same along the vertical Z-direction. Those unequal velocities in one
specific direction is a "velocity gradient", namely the gradient of X-velocity along the Z-direction]

Now, let’s imagine we can subdivided the moving fluid into many very thin layers parallel to the XY
plane (your table in the Fig. 1a). If the height of any individual layers is small enough, we can see
the motion of the individual layer as that of a rigid body. This means that each layer can be
characterized with a single velocity along the Z-direction. Obviously, in this particular example, each
layer moves with a different velocity. Now, there are many, many physical processes that can cause
those velocity differences (or velocity gradient) between fluid layers but, one thing is sure, because
of this velocity gradient, viscosity will act, and act as long as those velocity differences exist. Be
aware!! Viscosity is not responsible of this velocity gradient, it rather tends to counteract it.

Since some layers move a little bit faster, and others a little bit slower, we might expect that each
layer will experience a forward drag from the layer above (and will exert a backward drag on this
upper layer) and, at the same time each, layer experience a backward drag from the layer below
(and will exert a forward drag on the lower layer). This is represented on the following figure which
focuses on the j-layer. In Green, we have what the j-layer does or exerts to its neighboring layers
and, in Black what the j-layer experiences or suffers from its neighbors:
[Figure 2: What the j-layer does and what it suffers]
[in Black what it suffers from its neighbors and in Green what it exerts to its neighbors]

Now, what we have in the Figure 2 is a quite complicate situation, so, let’s adopt the following
convention:

- If the frictional forces acting on a plane of a given layer are parallel to it, they are referred as
shearing viscous force (see Fig. 2),

- If the frictional force acting on a plane are normal to it, they are referred as normal viscous force
(not shown on Fig. 2),

- Now, since those forces act on a surface, it is much more useful to work with a new quantity, call
stress, which is a Force per unit of Surface Area (N/m2). Hence, we have shearing stress and
normal stress. If you try to imagine Fig. 2 in 3D instead, you will see that we can deal with many
different planes (XY, XZ, YZ) and directions (X, Y, Z). We need a further convention:

- To properly define a stress, we need to specify its magnitude (which is the magnitude of the force),
its direction (which is the direction of the force), and the surface on which the stress acts (which will
be given by the normal vector of that surface). For instance, Tzx is the stress in the X-direction
acting on the plane XY (whose normal vector is parallel to the Z-direction). Hence, in Tzx, the first
subscript (z) gives the direction of the normal vector of the XY surface, and the second subscript (x)
gives the direction of the force. Tzz, is the stress in the Z-direction acting on the XY plane (whose
normal is in the Z-direction). Tzz is an example of a normal stress. All this can be easily understood
from Figure 3:
[Figure3: Principal directions of stress acting on the XY plane. The first subscript (Z) indicates the normal direction of the XY plane,
the second subscript indicates the main direction of the stress acting on the plane. Note that ous sign convention is postive in
compression.]

Whenever the two subscripts are the same (Txx, Tyy, Tzz), we deal with normal stress (the force is
perpendicular to the plane, like, for instance, the pressure). Whenever the two subscripts are
different (Txy, Txz, Tzy, Tzx, so forth), we deal with shearing stress (the force is tangential to the
plane). Be aware that the order of the two subscripts is fundamentally important, e.g., xy is the not
the same as yx !!!

Now, let’s go back to Fig. 2. It illustrates the forces acting along the top and bottom planes of a
given layer (j-layer). Recall that the flow velocity increases upward. Hence, the j-layer exerts a
backward shearing stress, , on the faster moving fluid layer above (it tends to decelerate the
layer j+1). But, according to Newton’s Third Law, the fluid above (j+1) also exerts a forward stress
( ) equal in magnitude to . Similarly, the j-layer exerts a forward stress ( ) on
the layer below, while the layer below (j-1) exerts an equal in magnitude but opposed backward
stress on the j-layer ( ). What we have is nasty, since the j-layer is sheared by the forces in
Black on Fig. 2. At the top, the force tends to make the j-layer moving faster, while at its bottom, the
force tends to decelerate it. The shearing forces generated by the velocity difference try to reduce
those velocity differences between adjoining layers of fluid. It has been found empirically that the
decelerating viscous stress is proportional to the velocity gradient (assuming we have a Newtonian
fluid):

Eq. 1

Eq.1 is the Newton’s Law of viscosity. In this equation, the proportionality factor ( ) of this empirical
law is called the coefficient of viscosity of the fluid. The negative sign in Eq.1 (that is not often seen
in the literature) expresses the fact that the viscous forces generated by the velocity gradient is
trying to reduce the physical effect which generated it. In other words, a force must be applied to the
lower wall to keep it at rest (or viscosity effects will take over and make it move faster) and a force
must be applied to the upper wall to maintain its uniform motion (or viscosity effect will take over
and make it move slower). Hence, we feel this notion of resistance to the natural flowing. If we don’t
do anything, viscosity effects will take over until the velocity gradient dwindles.
What is clear is that the fluid viscosity tends to reduce the velocity difference between adjacent fluid
layers. In terms of vector, the viscous shearing are opposite to the velocity direction (as they are
positive in the opposite direction of the velocity vectors, see Fig.2).

Be aware that viscosity effects do not necessarily make the fluid stop after some time, but rather, it
only reduces the velocity gradient within the fluid. Actually, for understanding the full effect of
viscosity, we must also understand what happen at the boundaries (e.g., on the ground). For
properly modeling viscous flows, we must set up with great care the boundary conditions (i.e., free-
slip wall, no-slip wall, or partial-slip wall). This is beyond the scope of this web page, I will present
boundary models later and elsewhere, when I will have time for this …

In a nutshell, viscous stress try to stop relative motion between different small part of the fluid … I
can even say whenever there is a strain rate (i.e., deformation that changes with time and is caused
by the velocity gradient) within the fluid, a viscous stress act to reduce that rate-of-strain. By the
way, Eq. 1 says that in a Newtonian fluid, the viscous stress is simply a linear function of the rate-of-
strain (the higher the rate-of-strain, the higher the viscous stress). If the fluid is non-Newtonian (like
granular flow) the relationship between strain and viscous stress is not simply linear, but more
complicate …

Now, I am nor sure how you feel here, but I can tell you that I still feel not quite at ease … Why? you
may ask. Well, the problem is that we know what does viscosity but we still do not know why it is
like that, we did not have explain the deep reason of the viscosity phenomenon, how it works. But
this can only be done with a molecular-granular kinetic model, which lead us to the next paragraph.

[Back to Top]

III. Kinetic view of the Viscosity

Kinetic theory attributes the tendencies towards uniformity of velocity within the fluid to the motion of
gas molecules or grains from one layer to another. This tends to equalize conditions at the two ends
by transporting to a further end amount of momentum (mass-velocity) that is characteristic of the
starting-point.

The motions of grains or gas molecules in the Y and Z-directions is totally random, while in the X-
direction the average velocity of those grains or molecules is equal to the bulk motion of the global
fluid (ux). Gas-molecules or grains randomly cross the interface between two adjoining layers in
both directions with equal frequency:
[Figure 4: In blue, grains or gas molecules moving from the upper level to the lower one with more momentum, and in red molecules
or grains from the lower level moving to the upper one with less momentum]

Of course, in addition to this vertical random motion, each molecule or grain has a velocity in the X-
direction, in other words it has some X-momentum. Therefore, each molecule or grain that crosses
from above carries with it momentum (and energy too) to the lower layer and so does the grains or
molecules moving from below to the upper layer. But the grains or molecules from above carry out a
somewhat larger average X-momentum to the layer below, while the molecules or grains from
below carry out a somewhat smaller average X-momentum to the layer above. As a result of this
microscopic kinetic process, macroscopic momentum is slowly transferred from the upper layer to
the lower one. Viscosity tends to equalize momentums throughout the fluid, hence, equalize
velocities throughout the fluid. We see why the upper layer exerts a forward drag to the layer below.

According to Newton Second Law, the rate of change of the X-momentum of the j-layer is equal to
the force exerted on it. The net transfer of X-momentum per unit of time and unit of surface area is
equal to the shearing viscous stress. That is exactly what says Eq. 1 in a qualitative way. Somehow,
the net variation of X-momentum of the j-layer must be equal to the sum of all the forces acting on it
(e.g., in Green, -Tzx from above and Tzx from below in Fig.2).

[Back to Top]

IV. Viscosity and Tensors

So, now we know that viscous stresses try to reduce the relative motion between near parts of the
fluids. In other words, whenever there is a strain rate (i.e., a velocity gradient within the fluid), a
stress acts to reduce the nasty strain … And the higher the viscosity, the higher will be the stress to
reduce the strain rate. Remember Eq. 1 we wrote earlier? It is a very simple one as it considers only
two-dimensions, but it can be generalized over the entire 3D space. This equation simply states that

the stress ( ) is proportional to the velocity gradient ( ), the constant of proportionality is the
viscosity. Ok, but we have considered two directions (X and Z). In 3D, the Newtonian Law is written
as:
Eq. 2

where now we must consider the shear and normal stresses acting in all the possible directions of
all the surfaces of a given infinitesimal reference volume. So, even if we don’t know anything about
tensor, we may feel that is the stress very much indeed, and D must be somehow related to the
velocity gradient ( ) as it was in Eq. 1. Also, since is a scalar, it is clear that D is also a tensor,
which is named the rate-of-deformation, or usually, rate-of-strain tensor. Its name says all! However,
D is not directly equal to the gradient of the velocity but rather to a somehow more complicate
function. Why it is so? Well, let’s have a deeper look into those tensors ( ).

Let’s imagine a very, very small (infinitesimal) volume of fluid which shows no spin. As I explained
earlier, there are two different sort of stresses or forces acting upon a plane, normal and tangential
(shear). The stress tensor can be written as a 3x3 matrix in which the diagonal elements are the
normal stresses and off-diagonal element are the tangential stresses:

Eq. 3

where ex, ey, ez are unit vectors in the X, Y and Z-directions respectively. In the 3x3 matrix, the first
line indicates all the forces in the 3-directions of space (X, Y, Z, second indices) acting on a surface
whose normal is parallel to the X-direction (first indices), the second lines is all the 3-directional
forces acting on a surface whose normal is parallel to the Y-direction, and the third lines is for all the
forces acting on a surface whose normal is parallel to the Z-direction. The first column indicates the
different values of forces systematically parallel to the X-direction (but acting on different planes),
the second column is all the values of the forces parallel to the Y-direction (but on three different
planes), and the third column is all the forces parallel to the Z-direction. The diagonal elements are
the normal stress (they have the same two indices), and may be associated to a "pressure", while
the off-diagonal elements represents the shear forces. Now since this volume is spinless, we must
have:

Eq. 4

The off-diagonal elements which have the same pair of indices are equal, you may transpose
columns and rows in this matrix, it won’t change anything. This tensor is said to be symmetric. The
physical reasons is because if the shear stress having the same indices were not equal, the fluid
volume would spin.

Now, let’s see the velocity gradient tensor as it can be also written as a tensor (but a little bit
different than the stress tensor):
Eq. 5

In the 3x3 matrix, the first line indicates the variation along the X-direction of the three velocities (ux,
uy, and uz), the second line is the variation along the Y-direction of the same three velocities, and
the third line indicates their variation along the Z-direction. Therefore, a column considers the
variation of a given velocity (ux, uy or uz) in the three directions of space.

This velocity gradient tensor is not exactly similar to the stress tensor in that the velocity gradient
tensor is not symmetric, if you transpose rows and columns you will have something totally different.
The physical interpretation of this is that the velocity gradient measures not only the deformation of
the fluid body (by compression and/or shearing), but also the relative orientation of the fluid body
(rotation). When I’ll have some more time, I will demonstrate this. Therefore, we cannot write a
direct mathematical relationship between stress tensor ( ) and velocity gradient ( ) since the
stress tensor doesn’t care at all of the rotation of the body, which has nothing to do with pure
deformation. However, the velocity gradient tensor like any similar tensor can be broken into two
parts, one symmetric (D) and the other one, antisymmetric (S). Hence, we can simply write the
velocity gradient tensor as the sum of those two parts:

Eq. 6

The symmetric tensor is the rate-of-strain (or rate-of-deformation) tensor (that’s what we are looking
for) and is equal to (for now, I’m not gonna write the unit tensor matrices, which are therefore
implied):

Eq. 7

where the "T" describes the operation of Transpose (i.e., switching rows and columns). It can be
seen that D is indeed a symmetrical tensor as (look inside the 3x3 matrix). You may be
confused by the minus sign in Eq.7. Well, I should demonstrate this but for now it simply indicates
that the rate-of-deformation is positive whenever deformation is compressive. The antisymmetric
tensor is the vorticity or spin tensor and is equal to:
Eq. 8

I will not comment this spin tensor any further as it has nothing to do with the scope of this page
(viscous phenomenon). Again, note the minus sign in Eq.8, which indicates that the spin tensor is
positive clockwise in the inward direction.

So, I guess you can see where I’m heading for, the stress tensor is not a direct function of the
velocity gradient but rather a function of the symmetric part of the velocity gradient tensor, i.e., D.
Indeed, D is not only symmetric as but represents all the rate-of-deformation suffered by the fluid
body. Youpie! We have all we need to write the Newton’s law of viscosity in terms of tensors (in 3D),
the stress is proportional to the rate-of strain:

Eq. 9

Now, notice the minus sign. Recall Eq. 1, where we have already seen this minus sign. In Eq. 1, it
expresses two main ideas. The first one is that momentum flux "moves" from region of high
momentum to region of lower momentum in order to equalize momentum throughout the whole flow.
The second idea is that the stress is positive in the opposite direction of increasing velocity to
precisely counteract this velocity increase. Now, from Eq. 9, we can add a new idea to the meaning
of this minus sign. It can be demonstrated (I’ll do that whenever I’ll have time but you may have a
good clue if you read till the bottom of this page) that whenever we are in compression, D is positive
and so is the compressive stress tensor. This means that if I compress the fluid body, the
compressive stress will be positive, and the rate-of-compression will be positive as well.

The idea of this minus sign in that particular place in Eq. 1 and Eq. 9 is extremely powerful, smart
and physically sane. However, it is sometimes sad too see many people confusing the exact
meaning of this minus sign and/or too many times forgetting to add this sign. For example, it can be
sometimes seen in some (bad) books that the stress tensor is defined positive in compression but
the rate-of-strain is defined negative in compression or that the stress tensor is positive in
compression and in the next chapter positive in extension, all this is horribly confusing, complicate
and pure non-sense. So, we should take care of the meaning of the directions of the stress tensor
AND at the same time the rate-of-strain tensor. Of course, all this is a matter of convention, but this
one is really the best. Note that Eq. 9 is not yet quite accurate (see next paragraph).

The very last thing you may ask is how from this complicate tensorial equation (Eq.9) we can come
back to the simple Newton’s equation seen in Eq. 1. Well that’s not too difficult to do. Eq. 1 says that
the stress in the plane XY is proportional to the X-velocity gradient in the Z-direction. So, it means
we are only interested in the velocity in the X-direction and its variation along the Z-direction. So,
let’s assume in this case that the X-velocity is only a function of the Z-direction, and that all the other
velocities are equal to zero:

ux = f(z); uy = 0 and uz = 0

Hence,
It follows from Eq. 7 that:

Eq. 10

Finally, we have formally the original Eq.1:

where the partial derivative is unnecessary since ux only varies in the Z-direction, and therefore
could be replaced by a total derivative as in Eq. 1.

[Back to Top]

V. Bulk, Shear, and Second coefficient of viscosity

Ok. We have accomplished tremendous things, we know what is viscosity and why. We also know
the relationship between strain and stress (I should say rate-of-strain and stress). Basically, I
explained that from a microscopic point of view, viscosity is due to the random motion of gas
molecules (or grains in a granular flow) between neighboring fluid layers. This random motion also
carries some momentum, and therefore is responsible for homogenizing the momentum throughout
the whole flow. This momentum flux explains the stress within the fluid. The problem is that this nice
picture is not quite accurate yet, as we have two different viscosities, and both don’t do exactly the
same thing to the fluid. Those two different viscosities are the bulk viscosity (or dilatational viscosity)
and shear viscosity (or dynamic viscosity). In consequence, Eq. 9 is not yet very accurate.

The shear viscosity, or sometimes called the dynamic viscosity, measures the random chaotic
motion of molecules or grains, that is what we have been reviewing here on this page so far.
Usually, in many, many fluid dynamic books, it is the only viscosity we have to worry about. Indeed,
for most of the fluid, the bulk viscosity is equal or approximated to zero (which is called Stoke’s
hypothesis) and/or the fluid is incompressible. Actually, in Eq. 9, it is exactly what I have assumed
(because I wanted simple first). However, in a gas made of huge polyatomic molecules, the bulk
viscosity will not be equal to zero. Indeed, such a molecules may vibrate and also rotate. Since
these molecules may have a dimension large enough relative to their mean free path, their vibration
and/or rotation may somehow affect their translation along their mean free path. This vibrational and
rotational effects are somehow measured by the bulk viscosity. Those supplementary motions
required energy (as the translational motion measured by the shear viscosity) which will be taken at
the expenses of the translational mode. In granular matter, the bulk viscosity is also not equal to
zero, particularly for concentrated granular matter. Of course, in the granular matter case, the bulk
viscosity cannot be explained by the vibration/rotation of the grains. Actually, the granular bulk
viscosity is simply proportional to the granular shear viscosity and is simply related to the random
motion of the grains.

Why do we call those viscosities bulk (or dilatational) and shear (or dynamical)? Good question.

Well, the reason is that the shear viscosity is only associated with the shear stress and shear
deformation, it has nothing to do with compressive stress or compressive deformation, while the
bulk viscosity is only associated with the change of volume of the fluid parcel. The bulk viscosity is
only associated with the normal viscous stress.

Soooooo, whenever we have a shear-rate-of-strain, we have a proportional shear stress where the
proportionality constant is the shear viscosity ( ). Whenever we have a pure volumetric-rate-of-
strain, we have a proportional normal stress, where the proportionality constant is the bulk viscosity
( ). In gas kinetic theory, the shear viscosity is often called the dynamic viscosity as it is simply
associated to the random chaotic motion of the gas molecules. In granular kinetic theory, we use
the name "shear viscosity" rather than dynamic viscosity as the granular bulk viscosity is also
associated to the chaotic random motion of the grains. Hence, for granular matter, the name
"dynamic viscosity" would be pretty loose and unclear as it could refer to the shear viscosity as well
as the bulk viscosity.

So what now? Well it means that we can specifically associate normal deformation with normal
stress and shear deformation with shear stress. Let’s see this …

The rate-of-strain tensor, D, represents all the possible deformation a fluid body can suffer; namely,
volumetric (compressive/extensive) deformations and shear deformations. In terms of tensors, it can
be shown that D can be broken into a spherical part ( ) and deviatoric part ( ). Where only
represents pure volumetric-rate-of-deformation (compression/extension, no shear) and is simply
proportional to the sum of the diagonal element of D (i.e., trace of D, which is ID), while only
represents pure shear-rate-of-deformation (no volume deformation):

Eq. 11

where is the divergence of the velocity field, and is equal and opposite to the sum of the
diagonal element of D (= -Dxx-Dyy-Dzz). If the flow is divergent and represents an extensional
state, while if it represents a compressive state. However, because of our sign convention,
at the very end, is positive in compression and negative in extension. You may see from Eq.11c
that the deviator of D is equal to D minus one-third the trace of D or minus the spherical part of D.
Hence the deviator of D can only represent the shear-rate-of-deformation rate (as we have
subtracted all the volumetric-rate-of-deformation). In other words, the deviator of a tensor is said to
be traceless; i.e., the sum of all the diagonal element of is zero. In terms of the stress tensor we
have:
Eq. 12

It should be noted that the shear viscosity is more often symbolized by the simple Greek letter
(mu) instead of . Eq.12b is the utmost complete mathematical relationship between stress and
rate-of-strain for a Newtonian flow.

Also in the case the divergence of the velocity is equal to zero ( ), then the flow is said to
incompressible, and the bulk viscosity is irrelevant. In that very particular case, you may see that
there is no difference between the deviatoric part of the rate-of-strain tensor and the rate-of-strain
tensor itself ( ). Hence, there is no difference between the deviatoric part of the viscous stress
tensor and the viscous stress tensor itself ( ).

Very often, in many fluid dynamic treatise the equations I have been reviewing are not written that
way as they rather use another viscosity concept, called the second coefficient of viscosity. What is
this second coefficient of viscosity? Let’s see this and rearrange Eq. 12b in gathering all the terms
containing the divergence of the velocity.

Eq. 13

and is the second coefficient of viscosity and represents a combination of all the viscous effects
associated with the volumetric-rate-of-strain (the divergence of the velocity). But the problem now is
that this equation is less clear because D is indeed the total rate-of-strain and also contains the
volumetric-rate-of-strain in addition to the shear-rate-of-strain. That’s why I prefer Eq. 12 as it clearly
splits shear effects from the volumetric ones, which is not the case in Eq.13. The second coefficient
of viscosity concept is rather a mathematical trick than a truly physical one (but sometimes it may be
useful to do so, e.g., for easily computing the work done by viscous dissipation).

If you wonder what is the first coefficient of viscosity … well that is simply another name for the
shear viscosity. Also, in most treatise book of fluid dynamic, all this doesn’t matter as it is assumed
(most of the time) that the bulk viscosity is approximately zero. Hence the second coefficient of
viscosity is proportional to the shear viscosity … In this case, the shear viscosity is the only one you
have to worry about. However, as I said earlier, most of the time you cannot do that for granular
matters.
Maybe an important thing?. The SI unit of viscosity (bulk, shear, second) is Pa s (Pascal second),
that is also kg/m s.

[Back to Top]

VI. The concept of viscosity within rigid-plastic granular flows

From the simple Eq. 1 or from the more comprehensive relations in Eq. 12a, it looks like that there
is a linear (proportional) relationship between viscous stress and rate-of-strain. In other words, if the
components of the rate-of-strain-tensor are increased by some factor, then the components of the
viscous stress tensor will increase accordingly and most importantly proportionally. That is true if we
have a linear rheology, i.e., if the constant of proportionality is indeed held
constant, . This typical for Newtonian and Bingham
rheologies.

If you read the frictional course in this website (which starts here), you will soon find out that within
frictional granular flows (i.e., plastic rheology) viscous stress is first-order rate-of-strain independent.
In other words, if you increase the rate-of-strain, you will not increase the viscous frictional stress.
This means, for a given normal applied stress and a given volumetric grain concentration, it is
impossible to have shear stress within a frictional granular flow higher than its yield shear stress
(i.e., the flow is at yield and flow plastically). Within plasticity, the linear relationship between viscous
stress and rate-of-strain cannot hold anymore. Once at yield, you may have a very high rate-of-
strain or very low, it won’t matter, the stress remains unchanged. Therefore, from Eq. 12a, the shear
viscosity ( ) and the bulk viscosity ( ) must be somehow non-linear functions of the rate-of-
strain, so that when the rate-of-strain increases the viscosity functions decrease to maintain the
viscous stress constant (more details on this page, Constitutive Equations for frictional Granular
Flows). For granular frictional flows, plasticity leads to a non-linear rheology.

There is no problem of having a viscosity not constant anymore and of being a non-linear function of
the rate-of-strain itself. Such rheology is non-linear and completely non-Newtonian indeed. But
some "linear" people are deeply troubled by that and have accused me of "abusing" the concept of
viscosity and/or its word within the rigid-plastic theory. That is because the only thing they know
(and want to hear) is a proportional relationship between viscous stress and rate-of-strain (i.e.,
Newtonian). Within Computational Fluid Dynamic, there is no problem of having non-linear
rheologies and, as a matter of fact, such rheologies are quite common in this Universe. I agree that
within elasto-plasticity the word "modulus" is rather used (bulk modulus for bulk viscosity and shear
modulus for shear viscosity) but it won’t change the basic fact that viscous stress is developed to
reduce the relative motions between near parts of the fluid, even though, at yield, plastic viscous
stress will remain constant no matter what happen to the rate-of-strain.

The conclusion is that there is much more than simple linear rheologies (e.g., Newtonian and
Bingham) and nobody should be afraid of dealing with more complex (but richer and more
interesting) non-linear rheologies as we do throughout this website.

If you have enjoyed this Viscous-Stress webpage, pleaaase, before you leave, sign my Guestbook. It's all I ask!

[Back to Top]

Other Granular Volcano Group Webpages:


- What is a Granular Medium?

A complete review of Plastic-Frictional theories:

- Part 1. Introduction, Mohr-Coulomb, and von Mises Stresses

- Part 2. Plastic Potential Theory

- Part 3. Critical State Theory

- Part 4. Constitutive Equations for frictional granular flow

- All I wanna know about viscous stress!

- Granular Theory: an Overview

- Compute Your Own Atmospheric Profile

Numerical Results:

- Go to the Numerical Results Page (Introduction and all Results)

- Go to the Plinian Cloud simulations Page

Welcome Page

You may also enjoy those pages:

A Review of Plastic-Frictional Theory


Part. 4
Constitutive Equations for Frictional Granular Flows
Have a nice Night!

You will find the basic facts about Plastic-Frictional Theories (Part. 4) - no details -. If you wanna know more just
email me or feel free to ask in the Discussion Forum. I purposely erased all the bibliographical references and
detailed equations to keep the text simple and easy to read.

If you need an official reference for the content of this website, please, use:
Dartevelle, S., Numerical and granulometric approaches to geophysical granular flows, Ph.D. thesis,
Michigan Technological University, Department of Geological and Mining Engineering, Houghton, Michigan,
July 2003.

We have seen on the preceding sections:


I. Introduction
II. Stress space, Slip Planes, Mohr-Coulomb and von Mises stresses
II.1. Mohr-Coulomb case: a 2D representation of stress (particular case)
II.2. von Mises case: a 3D representation of stress (general case)
III. Plastic Potential Theory
IV. Dilatation, consolidation, yield locus, and critical state
IV.1. Dilatancy
IV.2. Consolidation
IV.3. Yield locus and critical state

On this page, you will find the last part of this Frictional-Plastic course:
V. Constitutive Equations for Frictional Granular Flows
V.1. Beyond the extended von Mises Yield locus
V.2. Within the extended von Mises Yield locus

Please, don’t forget to sign our Granular Volcano Guest Book.

Done and updated by Sébastien Dartevelle, The WebMaster, Saturday November 22 2003.

V. Constitutive Equations for Frictional Granular Flows

V.1. Beyond the extended von Mises Yield locus

Again, I’d like to reemphasize that we need a flow rule to make the connection between the rate-of-
strain tensor and the stress tensor when the material is just at yield … This can be done using the
Plastic Potential Flow we have just seen this in paragraph III (Part. 2 - Plastic Potential Theory
Page). Also, we now know from paragraph IV (Part. 3 - Critical State Theory Page) that if we adopt
a yield locus that has the properties of Figures 13 in the /S-N plane or the properties of
Figures 14 in the stress plane, we can successfully describe without no physical inaccuracies
and inconsistencies the properties of concentrated frictional granular medium at yield, such as
dilatancy, consolidation and constant density at critical state. The problem is that the Mohr-Coulomb
doesn’t give us a function that looks like those yield loci as seen in Fig. 13 and Fig. 14. The same
would apply for the extended conical von Mises yield function in 3D in the principal
stress space. If we want to use the Plastic Potential theory and predict dilatancy as well as
consolidation, we must modify those laws, otherwise we are screwed (which is not good).
[Figure 13: Yield locus curve for a given bulk density, which is the combination of a failure locus curve and a consolidation locus
curve]
[At the intersection, we define a point called the critical state. The straight line passing through all the critical sates is named the
Termination locus]
[In Fig. 13C, I show two possible situations: dilatation path and consolidation path. Clearly, as the material deformed at yield, it will
change from one yield locus to another till it reaches and finds its critical state]
[See text for further explanation]
[Figure 14: Nearly the same as Fig. 13 but seen from the principal stress perspective. See Part 4. - Critical state Theory page for
detailed explanation]

As you may guess, changing the von Mises and Mohr-Coulomb law is not an easy task but it has
been done in the past. To the best of my knowledge, there is only one function that can be used
and have all the properties we have previously seen (dilatancy, consolidation, critical state in
paragraph IV):

Eq. 1

where IIdT is the second invariant of the deviator of the stress tensor, is the average of the three
principal stresses, is the angle of internal friction and P is a positive function that measures the
compressibility of the granular material and monotonically increases with the bulk density of the
granular material. P will be a normal frictional isotropic/hydrostatic stress (i.e., Pressure). This yield
locus function is shown on Fig.15 for three different bulk densities in the principal stress space. As a
reminder from the previous sections, IIdT and are calculated by:

Eq. 1_2

and,

Eq. 1_3

[Figure 15: Modified von Mises yield function accounting for compressibility effects, failure and consolidation processes]
[Compare with the extended von Mises conical yield function as shown on Fig.9 of paragraph II.2.]
[One apex of this function lies at the origin as for the extended von Mises yield function, but in this case there is a second apex on the
hydrostatic central axis]
[Since the function P increases with the bulk density of the granular material, any yield surface containing smaller yield surfaces
represents a higher density than the ones it contains]
From Eq. 1 and knowing IIdT (Eq. 1_3), we can find the radius of this yield function (for any fixed
value of P, see Fig.15), which is:

Eq. 2

At the two apex of this function, the radius is zero (on the hydrostatic axis for
, and the radius is maximum whenever , which defines a critical state.
A close inspection of Fig. 15 shows that this new yield function has a convexity. Hence the normal
to the surface has a negative projection on the hydrostatic axis for (failure process), a
positive projection for (consolidation process), and an orthonormal projection at
(critical state with no change of the bulk density).

All this can also be shown if we apply the Plastic Potential Theory to this new yield function. Indeed,

Eq. 3

So, we see that the divergence of the velocity will be positive or negative depending on the sign of
. Since for a dilatancy process, we have , the divergence will be positive, which
is an expansion as expected. While, if we have a consolidation process, , the
divergence is negative, which is indeed a compression. And if , at a critical state, then
we have neither expansion, nor compression as expected. So, we are happy very much indeed.

Of course, it is clear that the use of the Plastic Potential theory assumed that the Mohr-
Coulomb/von Mises approach is not right for such a theory. Therefore, we had to modify it … You
may like that or not, it is as it is.

Now, what I want is to use the preceding results for using them in a computer model, that is for
solving the momentum equation. So, I have to rearrange Eq.3 in order to properly and easily use it.
So, what I am gonna do, it’s a little bit algebra but don’t worry, I think it is a lot of fun … you’ll see.
You may see that Eq. 3b may be generalized as:

Eq. 4

And using Eq. 3c, we may rewrite Eq. 4 as:

Eq. 5

where is the deviatoric part of the rate-of-strain tensor (i.e., pure shear rate-of-strain) and is the
deviatoric part of the stress tensor (i.e., pure shear stress). This is an extremely important result as
this flow rule (Eq.5c) gives the relationships between shear rate-of-strain and shear stress. Such a
flow rule, which is a direct result from the Plastic Potential Theory, is often named the Levy-von
Mises flow rule. This Levy-von Mises flow rule necessarily implies that the granular flow is slightly
compressible. If the flow is incompressible, then we would have .

Notice that the Levy-von Mises flow rule implies the co-axiality between the stress directions and
rate-of-strain directions since the shear rate-of-strain is zero on plane where the shear stress is zero
(Please, if you do not know what co-axiality means you must read the previous section of this
course: Part.2 - Plastic Potential Theory). The normality conditions is of course true since this flow
rule is derived from the associated Plastic Potential theory (again see Part.2: Plastic Potential
Theory Page).

Now, we know that the Pitman-Schaeffer-Gray-Stiles yield function states that (see Eq. 1):

Eq. 7

It is worth noting that the Euclidian norm of the deviator of the stress tensor is:

Eq. 8

And we can also rewrite the Levy-von Mises flow rule as (see Eq. 5c):
Eq. 9

Therefore, taking Eq.7, and using Eq.9 along with Eq. 3c, we have:

Eq. 10

Cool, so what? Well, good question, why the hell we have done all this? Typically in a computer
model of granular flow we must solve a momentum equation in which one term will account for the
momentum contribution from all the stresses within the flow. The total stress can be written as:

Eq. 11

where Eq.11a is the total stress tensor, Eq.11b is the spherical part of the stress tensor, Eq.11c is
the deviatoric part of the stress tensor, Eq.11d is the rate-of-strain tensor (note the minus sign which
is explained by the fact that compression is positive following our sign convention), is the
deviatoric part of the rate-of-strain tensor, I is the unit tensor, are the frictional
bulk and shear viscosities. So, we must also find how to calculate the "frictional viscosities". Here
how. We know from Eq. 3c, Eq. 11b and Eq. 10:
Eq. 12

Cool, and now the shear frictional viscosity … a piece of cake! From the Levy-von Mises flow rule
(Eq. 5c), we know:

Eq. 13

Voila! We’re done folks! And the nice thing is that a close inspection of Eq.12 and Eq.13 show that
the frictional stress is independent of the rate-of-strain tensor (D) as required by the frictional theory.
Indeed, if the components of D are multiplied by a factor, the components of the stress tensor
remain unchanged (because at the denominator we have IIdD and ). And that is exactly what we
want since frictional flow must be rate-of-strain independent.

[Back to Top]

V.2. Within the extended von Mises Yield locus

As we have seen in the previous paragraph, we cannot use the Mohr-Coulomb/von Mises yield
function with the plastic potential theory as it leads to physical inconsistencies. Therefore, we had to
modify the von Mises theory in order to make them work within the plastic potential. This is kinda
done often, and it has my favor as I think it makes sense in terms of physics. On the other hand, in
some conditions, we may consider to keep the von Mises yield function. For instance, if we are only
interested into failure processes at very high concentration, the extended von Mises/Mohr-Coulomb
yield function may suffice for our needs.

Let’s see some properties of the extended conical von Mises conical yield function associated with
the Plastic Potential Theory (we have already seen some):
Eq. 14

As a reminder, we see that the divergence is always positive predicting a continued dilatation
(Eq.14c). Now, in using Eq.14c, we can generalize Eq.14b as:

Eq. 15

where IIdD and IIdT are the second invariant of the deviator of the rate-of-strain and the stress
tensors respectively. The last equation is again the famous Levy-von Mises flow rule, which is a
direct consequence of the compressibility of the material in applying the Plastic Potential Theory.
We know at yield we must have a conical function given by:

Eq. 16

And as in the previous paragraph (V.1.), we note that the Euclidian norm of the deviator of the
stress tensor is (for the extended von Mises case):

Eq. 17

Now, thanks to the Levy-von Mises flow rule, we have:


Eq. 18

Now, in the previous paragraph (V.1.), we have seen that it is possible to find an expression for the
bulk viscosity since we knew that , where P would be a frictional isotropic
Pressure, which is a function of the macroscopic density of the granular material. You may see that
now there is an impossibility to find an expression for the bulk viscosity as we deal with an
indeterminate solution for the bulk viscosity (try to use Eq.18 with Eq. 14c and you will see it cannot
be done). Hence, the bulk viscosity cannot be known and we will assume it is equal to zero, hence
.

Now, as we have done in the previous paragraph, we want to use those results for solving the
momentum equations in a computer model for instance. Typically, we must know the total stress at
any time anywhere. The total stress can be written as:

Eq. 19

Since, the bulk viscosity is assumed to be zero, the viscous stress tensor is traceless and is given
by Eq.19c. Now we just have to find an expression for the shear viscosity in using Eq.19c with Eq.
18:

Eq. 20

A close inspection of Eq.19 and Eq.20 shows that the frictional stress is independent of the rate-of-
strain tensor (D) as required. Indeed, if the components of D are multiplied by a factor, the
components of the stress tensor remain unchanged.
Ooooaaaaaah! It’s over, we’re done! I don’t know if you realize that it took me three months of hard
labor for finding all the books, papers, articles and writing down all this (besides drawing all the
figures) … hope you found this frictional course useful. Please, let me know what you think: email
me …

If you have enjoyed this entire Plastic-Frictional webpage, pleaaase, before you leave, sign my
Guestbook. It's all I ask!
A Review of Plastic-Frictional Theory
Part. 1
Introduction, Mohr-Coulomb, and von Mises Stresses
Have a nice Night!

You will find the basic facts about Plastic-Frictional Theories (Part. 1) - no details -. If you wanna know more just
email me or feel free to ask in the Discussion Forum. I purposely erased all the bibliographical references and
detailed equations to keep the text simple and easy to read.

If you need an official reference for the content of this website, please, use:
Dartevelle, S., Numerical and granulometric approaches to geophysical granular flows, Ph.D. thesis,
Michigan Technological University, Department of Geological and Mining Engineering, Houghton, Michigan,
July 2003.

You will find on this page:


I. Introduction
II. Stress space, Slip Planes, Mohr-Coulomb and von Mises stresses
II.1. Mohr-Coulomb case: a 2D representation of stress (particular case)
II.2. von Mises case: a 3D representation of stress (general case)

While, on the following pages, you will find the continuation of this frictional
course:
III. Plastic Potential Theory
IV. Dilatation, consolidation, yield locus, and critical state
IV.1. Dilatancy
IV.2. Consolidation
IV.3. Yield locus and critical state
V. Constitutive Equations for Frictional Granular Flows
V.1. Beyond the extended von Mises Yield locus
V.2. Within the extended von Mises Yield locus

Please, don’t forget to sign our Granular Volcano Guest Book.

Done and updated by Sébastien Dartevelle, The WebMaster, Saturday November 22 2003.

I. Introduction

Concentrated granular flows are economically important phenomena, e.g., in pharmaceutical


industry, corn flow in a silo, coal and flour in a bin, granular material flowing under the action of
gravity in a hopper are a few common examples. Unfortunately, those granular flows are poorly
understood. It is common for the industries to deal with enormous difficulties in the withdrawal
process of the granular material from a container (bin, silo, hopper). This cause high financial
losses, and it is not rare to see the complete collapse of silos as seen on Picture 1:
[Picture1: Collapse of a granular (corn) silo due to a poor design. Such a dramatic phenomenon are very common with silos]
[Most of the time, the silo is poorly designed due to a lack of understanding of concentrated granular flow behaviors and properties]
[In particular, not taking into account compressibility effects and using inadequate modeling techniques (like depth-average
technique) are the main responsible causes of such disaster]

Most of the silos and hoppers are designed after computer modeling of granular flows.
Unfortunately, those modeling are based on simple depth-averaged technique, sometimes called in
engineering design, slice analysis. Those modeling techniques are inadequate since they only
depend on one space variable. Flowing granular materials -even highly concentrated and frictional-
display complex instability in time and space, and are fundamentally unsteady and slightly
compressible. Those depth-average techniques -also commonly used in volcanology and
sedimentology- are totally unable to approach such important complexities and common properties
of granular flows.

At very high concentrations and low rate-of-strain, collisions cannot be seen as instantaneous
anymore, since grains suffer longer and permanent contacts in rubbing, rolling on each other.
Therefore, a kinetic stress model based on the Boltzmann equation is irrelevant and a frictional
stress model must be taken into account. This can be done using plasticity and similar theories in
which the material behavior is assumed to be independent of the velocity gradient or the rate-of-
strain. Needless to say, this is atypical for a viscous Newtonian flow where stress specifically
depends on rate-of-strain and, again, this shows that a Newtonian rheology cannot be chosen for
granular flow whatever the concentration. Under a normal stress, a well-compacted granular
material will shear only when the shear stress attains a critical magnitude. This is described by a
Mohr-Coulomb law based on the laws of sliding friction. However, the Mohr-Coulomb law says
nothing about how the granular material deforms and flows, it rather describes the onset of yielding.
The Plastic Potential theory will provide the required constitutive equations for describing the
deformation of a granular material under frictional motion. According to this theory, the concept of
critical state plays a key role. However, a critical state described by a simple Mohr-Coulomb theory
leads to physical inaccuracies (such a infinite dilatancy of the granular material). Actually, the Mohr-
Coulomb law should only be seen as an asymptotical solution after the granular material has
endured long and permanent plastic-frictional deformation. The Plastic Potential theory allied with
the critical state approach can successfully described the phenomenon of dilatancy, consolidation,
the independence between the rate-of-strain-tensor and the stress tensor. It will also become clear
in this manuscript that the Mohr-Coulomb law does not say anything and should not play a central
role for describing flowing and deforming granular materials at yield.

In establishing the constitutive equations for frictional granular media, I review in this page the basic
theoretical concepts behind frictional and plastic theories, and I also review the important properties
of granular flow, such as dilatancy and consolidation phenomena. Then the constitutive equations
are developed. Those equations show a clear independence between the stress tensor and the
rate-of-strain tensor of the flowing material as prescribed by the theory.

It is assumed that the material is slightly compressible, dry, cohesionless, and perfectly rigid-plastic.
Such properties are relevant for modeling the granular flows commonly seen in volcanology-
geophysics. As done throughout this website, and as done in Soil Mechanics, I consider that
compressive stress, compressive strain, and their rates are taken to be positive or, if you prefer,
tensile stress, elongative strain, and their rates are negative (:-).

[Back to Top]

II. Stress space, Slip Planes, Mohr-Coulomb and von Mises stresses

In a 3D with a Cartesian coordinate system, we can write the total stress tensor for a frictional
granular flow as:

Eq. 1

where Txx, Tyy, Tzz represents the normal stress acting on the YZ, XZ and XY planes respectively,
Txy, Txz are the shear stresses acting in the Y-direction and Z-direction respectively on the plane YZ
(whose normal is X), Tyx, Tyz are the shear stress in the X- and Z-directions respectively acting on
the plane XZ (whose normal is Y), and Tzx, Tzy are the shear stress in the X- and Y-direction acting
on the plane XY (whose normal is Z). Since a stress tensor must be symmetric, we have Txy = Tyx,
and Txz = Tzx and Tyz = Tzy (see Fig. 1).

[Figure 1: Positive directions of stress in the different planes]


[The sign convention is that if we assume that the X-, Y-, Z-velocity gradients are positive in the X-, Y- and Z-directions,
then the stress is positive in the opposite directions. In other words, compressive stress is always positive]

Let’s note that the eigenvalues of T are with and their associated
eigenvectors (principal directions) are n1, n2, and n3 respectively. It is always possible to
decompose the stress tensor into two parts, spherical and deviatoric:

Eq. 2

where I is the unit tensor and IT is the first invariant or the "trace" of the stress tensor (i.e., the sum
of the diagonal element). IT is said to be invariant because it is independent of the chosen
coordinate system (e.g., the sum of the diagonal element in the Cartesian or Principal direction
coordinate system is the same) as indicated by Eq.2b. Obviously, the sum of the diagonal elements
of the deviatoric stress tensor (Eq. 2c) is equal to zero, hence the deviatoric is said to be
"traceless".

It is now convenient to write the relationship between the stress tensor and the rate-of-deformation
(or rate-of-strain) tensor. Let’s write this way:

Eq. 3

where are the bulk and shear viscosities, P is an isotropic/hydrostatic Pressure


(normal stress), is the divergence of the velocity, are the gradient and its
transpose of the velocity, is the deviatoric part of the rate-of-strain tensor, and ID is the first
invariant or the trace of the rate-of-strain tensor (i.e., sum of its diagonal elements). Clearly, Eq.3
associates shear stress with shear rate-of-strain (Eq.3b) and normal stress with normal rate-of-
strain (Eq.3a). Many of you may be puzzled by the fact that Eq.3 looks like a rheology very
Newtonian indeed. But as we will see later (paragraph V) the shear and bulk viscosities are not
constant and actually are complicate functions of the velocity gradient, hence Eq.3 is not at all a
Newtonian rheology. Also, if you are surprised by the minus sign in Eq.3c and Eq.3d, please, recall
that my sign convention imposes that compressive stress and compressive strain are positive.

Having said those very general relationships, we may now move on (please, keep them in mind as
we will need them a lot in the followings).

If you already struggle here, you must then visit my web page All I wanna know ’bout viscous stress!
where I explain all this from the very beginning. Also, you may ask any question you want in the
Granular Discussion Forum.

[Back to Top]

II.1. Mohr-Coulomb case: a 2D representation of stress (particular case)

Let’s assume that , and that the principal direction n1 forms an angle measured
counterclockwise with the X-axis (see Fig 2A and Fig. 2B). A close inspection of Fig. 2 shows we
may carry out a 2-dimensional analysis of stress. Since , the plane XZ only suffers normal
stress (which is ), while the plane XY and YZ suffers shear (e.g., Tzx, Txy, so forth) and
normal stress (Txx and Tzz).

[Fig. 2A]
[Fig. 2B]

[Convenient geometry that permits a 2-dimensional analysis of stress]


[In Fig. 2A, one coordinate axis (Y) is parallel to one of the principal coordinates (the eigenvector, n2)]
[Therefore, the principal directions n1 and n3 stand in the plane XZ]
[The great advantage of this geometry is that the stress can simply be analyzed in 2D in the plane XZ as shown in Fig. 2B]

We can now rewrite Eq. 1 as:

Eq. 4

Let’s define the normal stress as the average between the principal stress components (in 2D,
within the plane XZ):

Eq. 5

while the shearing stress will be defined as:

Eq. 6

And it is clear from Eq. 4, 5, and 6, we must have:


Eq. 7

Consider an element of surface whose unit normal vector n is parallel to the XZ plane so that the
projection of n on that plane has components , where is the angle
between the n and the X-direction (see Fig. 3):

[Fig. 3A]
[Fig. 3B]

[Nearly the same situation as in Fig. 2A and 2B]


[In Fig. 3B, we focus on a particular plane (in green) within the infinitesimal volume, where n represents a normal unit vector to that
plane]
[In Fig. 3B, the same situation but projected over the plane XZ in order to 2-dimensionalize the stress analysis]

Then the normal stress (N) and shear stress (S) acting on that surface element in terms of the
normal and shearing principal stresses will be (Fig. 3B):

Eq. 8

Eq.8 defines a circle on the N-S plane centered at the point ( ,0) and of radius equal to (see Fig.
4). The interpretation of the angle can be easily understood from Fig. 3B. The onset of
yielding can be described with a Mohr-Coulomb model which asserts that a material will yield by
shearing on a surface element with normal n if S attains a critical value given by (see Fig. 4):

Eq. 9

where k is a known material property (describing the cohesive state of grains, and is therefore a
cohesive shear), and is the angle of repose (or the angle of internal friction of the material). Most
granular materials are cohesionless, k=0. The angle of friction can be easily understood from Fig. 5.
[Figure 4: Mohr-Coulomb relationships in the N-S plane]

[Figure 5: Angle of internal friction]

This angle of repose is low when grains are smooth, coarse or rounded, and, it is high for sticky,
sharp, or very fine particles. Typically, it is between 15° and 45°. Experiments suggest that this
coefficient of friction drops when motions begins, i.e., the kinetic friction coefficient is less than the
static coefficient. However, no data exist for granular material, and the universal assumption is that
the kinetic and static coefficient of friction are more and less equal.

The Mohr-Coulomb law is a linear law between the Shear (S) and the Normal (N) Stress. This line is
a yielding condition for shearing (see Figure 4 and 6 where Eq. 9 is drawn in the N-S stress plane).
Below this yield line, the material response will be rigid and does not suffer any strain. If the shear
stress is increased for a given normal stress such that the stress state of the material is exactly on
the yield line, then plastic strain or yielding will results. It is impossible to have a state of stress
above this yield Mohr-Coulomb line.

Before continuing any further, it is important to recall that the angles on the physical plane (e.g., Fig.
2 and 3) are doubled on the N-S plane in a Mohr-Coulomb circle (e.g., Fig. 4 and 6).

When the yield stress is reached then particles will slide over each other. Therefore, at each point in
the material, this can formally written as:

Eq. 10

All those relations can be shown and explained on a Mohr’s circle as represented on Fig. 6:

[Figure 6: Mohr-Coulomb circles in the S-N stress plane]

It is clear that the left hand side of Eq. 10 is maximum (only at yield when equality holds) whenever
(recall angle is positive counterclockwise) we have:

Eq. 11

Therefore using Eq. 8, we can calculate the yielding value of the normal stress (N) and the shear
stress (S) in term of :
Eq. 12

And Eq. 10 may be written as:

Eq. 13

which is another way of writing Mohr-Coulomb’s yield condition: Eq.13 shows the admissible states
of stress. Slip may occur if and only if equality holds in Eq.13. The linear relation between and is
of a fundamental importance, this line is called the yield line in the plane. Eq.13 is a yield
condition in the plane. Recall that it is a linear function because of the Mohr-Coulomb relation
given by Eq. 10. The orientations of the surface on which slip occurs are given by Eq. 11 and are
shown on Fig. 7. From Eq. 13, Eq.5, and Eq. 6, at yield, we must have:

Eq. 14

For a given angle of friction, Eq.14 gives a linear relation between the major ( ) and the minor
principal stresses ( ) that follows from Mohr-Coulomb law (Eq. 9) (see Figure 8).

[Figure 7: Orientation of the Slip-planes relative to the major and minor principal stresses.
Phi is the angle of internal friction of the granular material. Those geometrical relations are given by Eq. 11]
[Figure 8: Domain of no-deformation (rigid) and domain of plastic deformation (represented by two lines having slopes given by Eq.
14.
Theoretically, the inside domain can represent the elastic deformation, however, for most of granular materials, elastic strain is
negligible, and this inner domain can be assumed as perfectly rigid.
The central axis on which the principal stresses are equal represents the hydrostatic Pressure.
Compare this figure with Fig. 9 below, which is a generalization in 3D]

So, let’s summarize, at yield (when equality holds) and in 2D, It can be shown that the Mohr-
Coulomb condition (in Eq. 14) can be written as:

Eq. 15

which is four different ways of expressing the exact same Mohr-Coulomb yielding Law.

[Back to Top]

II.2. von Mises case: a 3D representation of stress (general case)

The problem of the Mohr-Coulomb law is that we have to assume that . Therefore, the Mohr-
Coulomb failure criterion is independent of the intermediate principal stress ( ). This a
mathematical model (among others), and such Mohr-Coulomb assumptions may not be always
possible or valid, specially for (some) granular materials. It is now clear that the velocity predictions
based on the Mohr-Coulomb model may greatly differ from what it can be measured. Hence we
must consider other approaches. An alternative approach would be to generalize in 3D without
imposing any conditions on the intermediate principal stress.

Let’s define the normal stress as the average of the three principal stresses:

Eq. 16

and the shear stress as:

Eq. 17

Sometimes those stresses are called octahedral normal stress (Eq.16) and octahedral shear stress
(Eq.17) respectively. In Eq.17 each principal stress difference is equal to the diameter of the
appropriate Mohr’s circle and is therefore equal to twice the greatest shear stress in that set of
planes ( see, for instance, Fig. 6). The octahedral shear stress is
proportional to the root mean square of the three maximum possible shear stress as shown by
Eq.17.

Taking the analogy with the Mohr-Coulomb law, we can postulate that there must be a
proportionality relationship between octahedral shear stress and octahedral normal stress. We
therefore have:

Eq. 18

Which is therefore a generalized yield criterion in 3D (again I have written 3 times the same Law in
Eq.18). Eq.18 gives the state of the stress of the granular medium just at yield. This yield criterion is
named the extended von Mises yield criterion (or conical yield criterion) as opposed to the Mohr-
Coulomb yield criterion as seen in Eq. 15. Whenever the intermediate principal stress is equal to the

average between the minor and major principal stress ( ), we find back the Mohr-
Coulomb yield criterion. It should be noted that the angle of internal friction may not have the same
value depending on whether the material deforms in 2D only (plane strain), or within a volume
(compression or extension in a triaxial test for instance): care with this frictional angle must be taken
when working with the von Mises yield criterion.

In 2D, the Mohr-Coulomb can be represented by two straight line in the principal stress plane as
shown on Fig. 8. In 3D, in the principal stress space, the extended von Mises yield has the
geometrical shape of a cone with its apex at the origin as shown on Fig. 9. Both yields have an axis
of symmetry, which represent the isotropic/hydrostatic Pressure for in the Mohr-Coulomb
case.
[Figure 9: Representation of the von Mises Yield Surface in the principal stress space.
The central axis of symmetry on which all the principal stresses are equal is the hydrostatic/isotropic Pressure.
Theoretically, the inside domain of the cone represents the elastic deformation, however, for most of granular materials, elastic strain
is negligible, and this inner domain can be assumed as perfectly rigid (no-deformation).
Plastic deformation occurs at yield on the surface of this cone. Compare this Figure with the 2D Mohr-Coulomb case as shown on
Fig.8]

Uh Oh! Wow, if you have followed me till here then I am amazed because I can tell we have done a
lot. As you may have noticed this page was mainly a "frictional warm up". Basically, we have
learned the state of the stress within a granular material when it is just at yield. But we dont know
how the material will react or move at yield … Basically, we don’t know anything ’bout the state of
strain within the material. And we don’t know either the relationship between rate-of-strain and
stress. Well, that is exactly what I’m gonna do in the second part of this course where I will define
such a relationship thanks to the Plastic Potential theory. Please, follow this link if you wanna know
more:

~~~~~~~~~ Part 2: III. Plastic Potential Theory ~~~~~~~~~

If you have enjoyed this Plastic-Frictional (Part 1) webpage, pleaaase, before you leave, sign my Guestbook. It's all I
ask!

[Back to Top]

Other Granular Volcano Group Webpages:

- What is a Granular Medium?

A complete review of Plastic-Frictional theories:


- Part 1. Introduction, Mohr-Coulomb, and von Mises Stresses

- Part 2. Plastic Potential Theory

- Part 3. Critical State Theory

- Part 4. Constitutive Equations for frictional granular flow

- All I wanna know about viscous stress!

- Granular Theory: an Overview

- Compute Your Own Atmospheric Profile

Numerical Results:

- Go to the Numerical Results Page (Introduction and all Results)

- Go to the Plinian Cloud simulations Page

Welcome Page

Home Page

You may also enjoy those pages:

A Review of Plastic-Frictional Theory


Part. 2
Plastic Potential Theory
Have a nice Night!

You will find the basic facts about Plastic-Frictional Theories (Part. 2) - no details -. If you wanna know more just email
me or feel free to ask in the Discussion Forum. I purposely erased all the bibliographical references and detailed
equations to keep the text simple and easy to read.

If you need an official reference for the content of this website, please, use:
Dartevelle, S., Numerical and granulometric approaches to geophysical granular flows, Ph.D. thesis, Michigan
Technological University, Department of Geological and Mining Engineering, Houghton, Michigan, July 2003.

We have seen on the preceding sections:


I. Introduction
II. Stress space, Slip Planes, Mohr-Coulomb and von Mises stresses
II.1. Mohr-Coulomb case: a 2D representation of stress (particular case)
II.2. von Mises case: a 3D representation of stress (general case)

On this page, you will find:


III. Plastic Potential Theory

And on the following pages, you will find the continuation of this frictional course:
IV. Dilatation, consolidation, yield locus, and critical state
IV.1. Dilatancy
IV.2. Consolidation
IV.3. Yield locus and critical state
V. Constitutive Equations for Frictional Granular Flows
V.1. Beyond the extended von Mises Yield locus
V.2. Within the extended von Mises Yield locus

Please, don’t forget to sign our Granular Volcano Guest Book.

Done and updated by Sébastien Dartevelle, The WebMaster, Saturday November 22 2003.

III. Plastic Potential Theory

The yield condition laws I have reviewed in the preceding section says nothing about the nature of the
motion which is initiated at yield. Alone, the Mohr-Coulomb/von Mises relations are helpless for
describing a granular media in a continued motion. Constitutive relations are needed for describing
the kinematics of motion at yield following the Mohr-Coulomb/von Mises theory. In addition, I must
also account for the famous kinematics feature described for the first time by Osborne Reynolds:
dilatancy. For a given normal stress, the material will sheared only if a critical shear is attained, once
sheared, the granular medium may initially expand. Afterwards, it will settle to a fairly constant
volume. This counter-intuitive behavior is of a key importance in soil mechanics.

The plastic potential theory will provide us a way to predict the velocity distribution within the granular
medium at yield. This theory makes the connection between stress and deformation (or velocity
gradient) in using three concepts: a Yield Function (Y), a Plastic Potential Function (G), and a Flow
Rule.

The yield function (Y) is what we have seen in the previous section (Part. 1). For instance, from the
Mohr-Coulomb and the von Mises yield functions, we have:

Eq. 1

where IIdT is the second invariant of the deviator of the stress tensor T. The other symbols have been
seen in the 1st Part of this course. IIdT can be expressed as:
Eq. 2

It is clear from all I said in the previous section (Part.1 ), at yield, we must have Y=0 for a plastic
material (as indicated by Eq.1). Therefore, at yield, we can see that Eq.1a is the equation of a straight
line (which is, in the mathematical terminology, a level curve for Y=0, see Fig.8) in the principal stress
plane ( ), and Eq.1b is the equation of a cone (which is, in the mathematical terminology, a level
surface for Y=0, see Fig.9) in the principal stress space ( ). Both functions have an central
axis of symmetry which has the property of being the axis of equal principal stresses, i.e., for
Mohr-Coulomb 2D case and for the von Mises 3D case. Hence, this axis of symmetry
represents the hydrostatic/isotropic Pressure. Both functions have their vertex on this hydrostatic axis
at the origin, when .

[Figure 8: Domain of no-deformation (rigid) and domain of plastic deformation for the 2D Mohr-Coulomb case.
The central axis on which the principal stresses are equal represents the hydrostatic Pressure]
[Figure 9: Representation of the 3D von Mises Yield Surface in the principal stress space.
The central axis of symmetry on which all the principal stresses are equal is the hydrostatic/isotropic Pressure.
Plastic deformation occurs at yield on the surface of this cone]

The Plastic Potential Function (G) is defined so that the strain rate in any arbitrary directions (Di j) is
proportional to the derivative of G with respect to the corresponding stress (Ti j):

Eq. 3

where q is a positive scalar sometimes named "plastic multiplier". This scalar is not a property of the
material but rather a property of the flow conditions. Eq.3 can also be written in terms of the principal
directions of D and T (using their respective eigenvalues) as:

Eq. 4

Now, we are almost there, since we now specifically play with the rate of deformation (D), hence the
velocity gradient. So, the remaining thing to do is to find out what is G exactly? This is given by the
flow rule. One of the most widely used flow rule is the associated flow rule, which states that the
Plastic Potential Function (G) is equivalent to the Yield function (Y), G=Y, hence, Eq. 4 can be
rewritten as:

Eq. 5

It is worth saying that there are other flow rules which do not assume G=Y, those are named non-
associated (I won’t comment this any further). It is clear that in Eq. 5, the Right-Hand-Side (RHS) is
nothing but the gradient of the Yield Function with respect to the stresses, which is important to notice
for defining the properties of the plastic potential flow. But before doing so, let’s indeed derived Eq. 5
using the Yield Function defined by Eq. 1b (extended von Mises) and the definition of the second
invariant of the deviator of the stress tensor (IIdT) in Eq. 2c, we have in 3D:

Eq. 6

This equation can be safely generalized as:

Eq. 7

where D, I, and are the rate-of-strain tensor, the unit tensor and the deviator of the stress tensor
respectively. The deviator of the stress tensor is:

Eq. 8

where is the spherical part of the stress tensor which is the average in the 3-direction of space of
the normal stresses (i.e., one third of the sum of the diagonal element of the stress tensor).

From those equations, we can now state the two key properties of the Plastic Potential Theory:

1- the co-axiality or alignment condition: the principal axe of rate of deformation are aligned with those
of stresses. This to be in agreement with the intuitive idea that the material should respond to unequal
stresses by contracting in the direction of greater stress and expand in the direction of lesser stress.
In other words, it is equivalent to say that there is no shear strain on planes on which there is no
shear stress. This is exactly what shows Eq. 6 and Eq. 7 since whenever the shear stress (Ti j) is zero
on a given plane, so is the shear strain (Di j). I will come back to this property latter in this course.

2- the normality condition: which is shown by the original equation of the Plastic Potential Theory (Eq.
5). Indeed, if we consider the particular level surface Y=0 (at yield) in Eq. 5, we know from the
properties of gradient that Di must be a vector perpendicular to the Yield (level) Surface, in the
principal stress space.

Now, I’m pretty so you’re gonna ask: Sooooooooooooo what? Why the hell in the world do we want to
know all that? Well, those last equations actually give us a connection between the velocity gradient
within the granular flow and the state of stress within it. And basically, that’s what we were looking for,
because it can be easily implemented into momentum equations in a computer model as we must
know such relations.

Now, it would be interesting to calculate the divergence of the velocity field, which is nothing but
minus the volumetric rate of deformation of the granular body (i.e., first invariant of D):
Eq. 9

Therefore, with both Mohr-Coulomb and von Mises cases, the granular body, at yield, suffers
dilatation (divergence of the velocity field is positive, i.e., that is extension). On the one hand, that is
good because it is in agreement with Reynolds' Principle of Dilatancy and therefore that is what we
expect from those laws. However, on the other hand, it is bad as Eq. 9 leads to a continued
expansion of the granular medium without any limits, which is physically impossible.

But there is another trouble. Indeed, dilatancy is one phenomenon that can be observed in granular
material at yield but there are other processes that may happen as well depending on the physical
properties of the granular material and on its previous deformation history. If the granular medium is
well-compacted, it will expand when sheared (that is dilatancy), whereas loosely packed, a granular
medium at large normal stress may contract (that is consolidation or contractancy). Moreover, the
expansion or the contraction takes place only until a critical bulk density is reached, afterwards under
continuous stress, deformation will continue without any change in the bulk density. And that is not
shown at all in the Eq. 9. In addition, it can be shown that the rate of energy dissipation is always zero
if we use a simple Mohr-Coulomb/von Mises law, which is an unacceptable results for frictional
processes.

So, we got a problem down here … In conclusion, the Mohr-Coulomb/von Mises theory cannot be
applied straightforwardly for describing flowing frictional granular material without committing
unphysical instabilities and inconsistencies. The simple Mohr-coulomb/von Mises law must be -
somehow- adapted to the plastic potential flow theory for describing flowing granular medium
accurately. But before getting into the constitutive equations and the possible corrections, it is
necessary to review the basic behaviors of granular material at yield. This is what I’m gonna do in this
following section:

A Review of Plastic-Frictional Theory


Part. 3
Critical State Theory
Have a nice Night!

You will find the basic facts about Plastic-Frictional Theories (Part. 3) - no details -. If you wanna know more just
email me or feel free to ask in the Discussion Forum. I purposely erased all the bibliographical references and
detailed equations to keep the text simple and easy to read.

If you need an official reference for the content of this website, please, use:
Dartevelle, S., Numerical and granulometric approaches to geophysical granular flows, Ph.D. thesis,
Michigan Technological University, Department of Geological and Mining Engineering, Houghton, Michigan,
July 2003.

We have seen on the preceding sections:


I. Introduction
II. Stress space, Slip Planes, Mohr-Coulomb and von Mises stresses
II.1. Mohr-Coulomb case: a 2D representation of stress (particular case)
II.2. von Mises case: a 3D representation of stress (general case)
III. Plastic Potential Theory

On this page, you will find:


IV. Dilatation, consolidation, yield locus, and critical state
IV.1. Dilatancy
IV.2. Consolidation
IV.3. Yield locus and critical state

And on the following page, you will find the continuation of this frictional course:
V. Constitutive Equations for Frictional Granular Flows
V.1. Beyond the extended von Mises Yield locus
V.2. Within the extended von Mises Yield locus

Please, don’t forget to sign our Granular Volcano Guest Book.

Done and updated by Sébastien Dartevelle, The WebMaster, Saturday November 22 2003.

IV. Dilatation, consolidation, yield locus, and critical state

Let’s imagine we could perform the following idealized experiment (see Figure 10). Between two
rough (no-slip) plane stands a granular material. The bottom plate is immobile while the upper plate
can move laterally to generate a shear stress (S) within the granular medium, hence the material
deforms (shear strain, ) as the plate moves. The whole material has a uniform bulk density ( ) and
is under a constant load N (normal stress) on the upper plane. Again, I will assume -as I have done
so far- that the granular material is plastic-rigid (i.e., it does not show any elastic deformation under
applied stress) and cohesionless.

[Figure 10: Idealized experiment: granular medium within two rough plates, the bottom one is immobile, while the top one can move
in order to generate a shear stress (S) within the medium. The whole medium is under a load N]

They are two different behaviors of the granular medium that can be observed depending on the
normal loading (N) and/or the initial bulk density ( ): dilatation and consolidation.
[Back to Top]

IV.1. Dilatancy

For high density and/or low normal pressure, the behavior is called failure, dilatation, or dilatancy of
the material and is shown on Figure 11. At first, as the shear stress increases, no noticeable shear
deformation occurs (except maybe a reversible elastic strain that I do not show on Fig. 11A), till we
reach a critical shear stress (Sc) which depends on the applied normal stress and the bulk initial
density (Fig. 11A). Afterwards, instead of having a flat linear line of S vs. , the required applied
shear stress may be asymptotically decreased to some constant value (Sinf) independent of the bulk
density and only a function of the normal stress. Fig. 11B shows the displacement of the granular
material as a function of the position of the plates (only the upper plate moves). Typically the shear
strain occurs in a very thin layer, while the upper part of the flow moves en bloc without no
noticeable shear within it.
[Figure 11: Dilatancy principles and properties]

This behavior is described as a failure of the granular material. The decrease of the value of S
needed to maintain shear after passing the failure point (Sc) is associated with a weakening of the
material during the initiation of the plastic deformation (basically, we need less shear force for
deforming the material). In the thin shear zone, we would measure a decrease of the bulk density
owing to an overall dilatation of the shear layer. This dilatation phenomenon arises from the need of
a densely packed granular material to spread in order to make enough room for allowing grains to
move.

In the plane or in the S-N plane (it doesn’t matter which plane because
, since according to Mohr-Coulomb, we have
), for a given initial bulk density and at failure (or at yield), we
would have a convex curve as shown on Fig. 11C. This kind of curve is called failure locus or yield
locus. So, the higher the normal stress N, the higher the shear stress S for reaching failure. Also,
there is a different failure locus for different initial bulk densities. Typically, for a given normal stress,
the higher the bulk density, the higher S required for reaching yield. It should be noted that once
values of S asymptotically tend to Sinf, the material tends to a constant density (Fig. 11A), and
therefore the dilatation of the material ceases.

You may be surprised to see curved yield loci as we would have expected rather a straight line in
the or in the S-N plane as predicted by the Mohr-Coulomb Law. Yeah, that’s true but as I said
earlier the Mohr-Coulomb is only an ideal state which may not be quite right. In addition, the
curvature of those yield loci is very low, and at the limit those curves can be approached by a
straight line. However, for some materials, the curvature may be significant and a simple straight
Mohr-Coulomb line is indeed inappropriate (then a Warren Spring Failure locus curve should be
used instead). Generally speaking, for a given yield locus, the curvature is only pronounced at low
normal stress.

[Back to Top]

IV.2. Consolidation

When is small and/or N large, the situation is totally different (Figures 12). Again, we need to
reach a critical value of the shear stress (Sc) to initiate deformation (Fig. 12A). Then afterwards,
shear deformation starts but to maintain it, we must asymptotically increase the applied shear stress
to a constant value (Sinf), independent of the bulk density. In Fig. 12B, it can be seen that, in this
case, the shear strain is distributed uniformly throughout the granular layer. In addition, we would
measure a uniform increase of the bulk density in the whole layer as deformation proceeds.
However, once Sinf is approached, the granular material tend to a fairly constant bulk density, hence
consolidation ceases.
[Figure 12: Contractancy principles and properties]

In this case, the material gains strength since to initially maintain the deformation we must increase
the applied shear stress. The increase of bulk density of the flowing material can be seen as the
upper plate moves downwards. This phenomenon is called consolidation, contractancy or negative
dilatancy. In the or S-N plane, for a given initial bulk density and at yield, we would have the
convex curve as shown on Fig. 12C, where the higher the normal pressure, the less shear stress
we need to initiate consolidation. As for the dilatation, those curves in the /S-N plane depend on
the initial bulk density. For a given normal stress, the higher , the higher the shear stress for
reaching yield. Such a kind of curve in the plane is named consolidation locus or
contractancy locus. Once the critical shear stress is reached (Sc), the shear stress asymptotically
tend to a fairly constant value (Sinf). Notice that for very high value of normal stress, , where
the consolidation surface intersect the N-axis (Fig. 12C). At such intersection point, the material
consolidate under normal load alone, without any assistance from shear stress.
[Back to Top]

IV.3. Yield locus and critical state

As we have just seen the yield locus (e.g., failure locus and consolidation locus) is a function of the
bulk density ( ) and the normal stress. For a given density and for a 2D stress representation, the
function yield locus , is convex as shown on Fig. 11C & 12C and Figures 13 in the
/S-N plane. In Figure 13A, where I drew both the failure locus and the consolidation locus, we can
interpret such a curve as a section of a surface in a space. In soil mechanics, the surface
that represents the failure part is called the Hvorlsev Surface, while the consolidation part in this
space is named Roscoe surface. The material is rigid if is less than the threshold and is
perfectly plastic on the curve. As mentioned above, the higher the initial bulk density of the granular
material, the higher its strength (we need to apply more normal and/or shear stress for deforming it)
as seen on Figure 13. In the space, the intersection of the Roscoe and Hvorlsev Surfaces
defines a curve called the Critical State Locus. This curve projected onto the plane defines a
curve called the Termination Locus , which is most of the time a straight line. This straight line
intersect the yield loci at a point named critical state. (Fig. 13A & 13B). Some authors give a
supplementary property of the critical state: on a given yield locus curve (i.e., for a given bulk
density), the critical state is where the derivative of vanished, in other words:

(Fig. 13A). The material will consolidate or expand according to whether ( ) lies
to the right or left side of the line of critical states. Hence, we have on this yield locus an expansion

side (left side, where ) and a consolidation side (right side, where ) (see Fig.
13B). This means that when the granular material reaches the critical state point on a given yield

locus, it will deform without any change of its volume ( ). Since, at yield, the granular
material suffers density change, it implies that the granular material, as it is deforming, will move
from one yield locus curve to another until it finally reached the critical state where it will deform
without any further change of volume.
[Figure 13: Yield locus curve for a given bulk density, which is the combination of a failure locus curve and a consolidation locus
curve]
[At the intersection, we define a point called the critical state. The straight line passing through all the critical sates is named the
Termination locus]
[In Fig. 13C, I show two possible situations: dilatation path and consolidation path. Clearly, as the material deformed at yield, it will
change from one yield locus to another till it reaches and finds its critical state]
[See text for further explanation]

I guess you have notice that the notion of "yield locus" is somehow ambiguous as it refers to two
different things. Yield locus can be used as failure locus and therefore represents the left side of the
whole curve drawn on Fig. 13A (the side that could be represented by a Mohr-Coulomb Law).
However, some authors used yield locus for describing the whole curve, that is failure locus curve
(left-side) and consolidation locus curve (right-side). So, you must be very careful of the context in
which "yield locus" is used within the text (I know it is horribly confusing …).

Let’s go back to Fig. 13C and see what happens to a granular material of initial bulk density, ,
subject to normal stress smaller or greater than the normal stress at its critical state. If the normal
stress is greater than the normal stress defined by the critical state (right-side), then we will have a
consolidation behavior (path X-Y-Z). In this case the granular material is initially at point X, and will
not suffer any strain as long as the shear stress is smaller than the threshold value defined by the
point Y on its consolidation curve . As soon as the shear stress reached that particular
value at the point Y, the granular material will plastically compress and strengthen until it reaches a
termination locus (critical state point) in Z. This material has moved from an initial density to a
higher density . Thereafter, the material will continue to shear at constant bulk density. Because,
the material has increased its density, this behavior is named consolidation. Now, if the normal
stress is smaller than the critical state normal stress (left-side), we will have a dilatation behavior
(path Q-R-S). The material is initially at Q, and will not suffer any strain as long as the shear stress
is smaller than a threshold value of shear stress defined by the point R on its failure curve
. As soon as the material is on that curve, at R, it will plastically shear and dilate and
therefore weaken until it reaches a termination locus at S. This material has decreased its initial
density from to , this behavior is named failure or dilatancy. The dilatation occurs in very
confined local zones which most of the time has already been previously sheared (weak zone).

All this can be also thoroughly analyzed in the stress plane, it is exactly the same story. In this
plane, yield locus may have the shapes as indicated in Figure 14A which also shows the directions
of the normal to the locus in three different point: 1, 2, and C. At point 1, for low values of and
hence for low value of normal stress, the normal vector has a positive projection on the line of slope
45° (which is the isotropic normal Pressure defined as ). Yielding in this particular case is
accompanied by dilatation. At point 2, for high values of (hence high value of normal stress)
the normal vector has a negative projection on the isotropic normal Pressure line, and yielding is
characterized by a compaction. It is easy to see from Fig. 14A, that the point C is the critical state
where the material at yield has no change in volume since the normal is perpendicular to the
isotropic normal stress line. In Fig. 14A, we have the yield locus for only one value of the bulk
density. In Fig. 14B, the same figure is constructed for three different bulk densities and the lines of
critical states is also shown. Different situations are represented (see trajectories Q-R-S for
dilatation and X-Y-Z for consolidation, the same situation as Fig. 13C). If yield is initiated when the
granular material is just at point Q, dilatation must occur, and if the normal stress is maintained
constant, then the material will loose its bulk density (following the path Q-R-S) till it reaches the
critical state at point S, where deformation will proceed without any change in density. On the other
hand, if yield is initiated at a point beyond the critical state (for high values of normal stress at
position X), then consolidation will occur. The material is gaining strength and it sees its bulk density
increasing. For a constant normal stress, it follows the path X-Y-Z till it reaches the critical state in Z.
Again we have the same story as Fig. 13C, but I simply wanted to show you this from another
"stress perspective" (the more the better).
[Figure 14: Nearly the same as Fig. 13 but seen from the principal stress perspective. See text for detailed explanation]

Hope all this makes sense for you, at least it does for me. There is a last thing I’d like to say, hoping
you wont be more confused than ever? What about the von Mises and Mohr-Coulomb Laws in all
this? Well, If you remember what I said in the Plastic Potential Theory page (Part. 2 of this course):
those two laws seem to predict dilatation (failure) of the granular material. Well, for instance, in the
Mohr-Coulomb case it is clear that the failure locus in Fig. 13 (left-side) can be approximated by a
straight-line. Hence, the Mohr-Coulomb law cannot predict neither the consolidation, nor the critical
state. In other words, the Mohr-Coulomb law is only a failure locus. And this is also the case in 3D in
using a conical extended von Mises yield locus, which only predicts dilatancy process since the
projection on the hydrostatic axis of the normal to the yield locus is negative (see Fig. 9 in the Part.2
of this course). Therefore, the Mohr-Coulomb and von Mises yield loci can only predict dilatation
process. If you use those laws, you wont be able to approach the true phenomena seen in granular
maters. Those two laws may be OK for some cases but you should think twice before using those
laws.

So, now, what I am gonna do is to develop constitutive equations … I will also suggest a possible
more general and useful law than the Mohr-Coulomb/von Mises one. All this can be found in the last
Part of this course:
~~~~~~~~~ Part 4: V. Constitutive Equations for frictional granular flow ~~~~~~~~~

You might also like