3.5 Hyperelasticity  - time independent behavior of
rubbers and foams subjected to large strains
 
Hyperelastic constitutive laws
are used to model materials that respond elastically when subjected to very
large strains. They account both for nonlinear material behavior and large
shape changes.  The main applications of
the theory are (i) to model the rubbery behavior of a polymeric material, and
(ii) to model polymeric foams that can be subjected to large reversible shape
changes (e.g. a sponge).  
 
In
general, the response of a typical polymer is strongly dependent on
temperature, strain history and loading rate. 
The behavior will be described in more detail in the next section, where
we present the theory of viscoelasticity. 
For now, we note that polymers have various regimes of mechanical
behavior, referred to as ‘glassy,’ ‘viscoelastic’ and ‘rubbery.’   The various regimes can be identified for a
particular polymer by applying a sinusoidal variation of shear stress to the
solid and measuring the resulting shear strain amplitude.  A typical result is illustrated in the
figure, which
shows the apparent shear modulus (ratio of stress amplitude to strain
amplitude) as a function of temperature.
 
At a critical temperature known
as the glass transition temperature, a polymeric material undergoes a
dramatic change in mechanical response. 
Below this temperature, it behaves like a glass, with a stiff response. Near
the glass transition temperature, the stress depends strongly on the strain
rate.  At the glass transition
temperature, there is a dramatic drop in modulus.  Above this temperature, there is a regime
where the polymer shows ‘rubbery’ behavior - the response is elastic; the stress does not
depend strongly on strain rate or strain history, and the modulus increases
with temperature.  All polymers show
these general trends, but the extent of each regime, and the detailed behavior
within each regime, depend on the solid’s molecular structure.  Heavily cross-linked polymers (elastomers)
are the most likely to show ideal rubbery behavior.   Hyperelastic constitutive laws are intended
to approximate this behavior.
 
 
Features of the behavior of a solid
rubber: 
 
1. The material is close to ideally
elastic. i.e. (i) when deformed at constant temperature or adiabatically,
stress is a function only of current strain and independent of history or rate
of loading, (ii) the behavior is reversible: no net work is done on the solid
when subjected to a closed cycle of strain under adiabatic or isothermal
conditions.
 
2. The material strongly resists volume
changes.  The bulk modulus (the ratio of
volume change to hydrostatic component of stress) is comparable to that of
metals or covalently bonded solids;
 
3. The material is very compliant in
shear  shear modulus is of the order of  times that of most metals;
 
4. The material is isotropic  its stress-strain response is independent of
material orientation. 
 
5. The shear modulus is temperature
dependent: the material becomes stiffer as it is heated, in sharp contrast to
metals;
 
6. When stretched, the material gives
off heat.
 
 
Polymeric foams (e.g. a sponge) share some of these
properties:
 
1. 
They are close to reversible, and
show little rate or history dependence.
 
2. In contrast to rubbers, most foams
are highly compressible  bulk and shear moduli are comparable.
 
3. Foams have a complicated true
stress-true strain response, generally resembling sketch in the figure. The
finite strain response of the foam in compression is quite different to that in
tension, because of buckling in the cell walls.
 
4. Foams can be anisotropic, depending
on their cell structure.   Foams with a
random cell structure are isotropic.  
 
The literature
on stress-strain relations for finite elasticity can be hard to follow, partly
because nearly every paper uses a different notation, and partly because there
are many different ways to write down the same stress-strain law.   You should find that most of the published
literature is consistent with the framework given below  but it may take some work to show the
equivalence.
 
 
All hyperelastic models are
constructed as follows:
 
1. Define the stress-strain relation for
the solid by specifying its strain energy density W (which is related to the Helmholtz free energy of the solid by , where  is the mass per unit reference volume)   as a function of the deformation gradient
tensor (or some strain measure derived from the deformation gradient): W=W(F).    This
ensures that the material is perfectly elastic, and also means that we only
need to work with a scalar function.  The
general form of the strain energy density is guided by experiment; and the
formula for strain energy density always contains material properties that can
be adjusted to describe a particular material.
 
2. The undeformed material is often
assumed to be isotropic  i.e the behavior of the material is independent of the
initial orientation of the material with respect to the loading.  If the strain energy density is a function of
the Left Cauchy-Green deformation tensor  the constitutive equation is automatically
isotropic.  To see this, note that if we
subject the solid to a rigid rotation R before applying the deformation F
we find that   so B
is unchanged by changing the orientation of the specimen.  But if B is used as the deformation
measure, then the strain energy must be a function of the invariants of B to ensure that the constitutive equation is frame
indifferent (see Sect 2.7 and 3.1).  This
is because under a change of reference frame represented by an orthogonal
tensor Q the components of B change to , but the strain energy must be
independent of Q to satisfy frame indifference.  The invariants of  and  are equal.
 
3. For an anisotropic constitutive equation we can make the strain energy
density a function of the Right Cauchy-Green tensor .  
This is automatically frame indifferent because  under a change of reference frame.   But under a rotation R before
deformation a deformation F the Cauchy Green tensor is . 
Unlike B, it is not invariant to the orientation of the specimen.
 
4. Formulas for stress in terms of
strain are calculated by differentiating the strain energy density as outlined
below.
 
 

 
3.5.1 Deformation Measures used in finite elasticity
 
Suppose that a
solid is subjected to a displacement field , as shown in the figure. Define
 
· The deformation gradient and its Jacobian
      
 
· The Left and Right Cauchy-Green
deformation tensors
 
· Invariants of B (these are the
conventional definitions)
 
· An alternative set
of invariants of B (more
convenient for models of nearly incompressible materials  note that  remain constant under a pure volume change)
 
· Principal stretches
and principal stretch directions, specified as follows
1. Let  denote the three eigenvalues of B.  The principal stretches are 
2. Let  denote three, mutually perpendicular unit eigenvectors of B. These
define the principal stretch directions. 
(Note: since B is symmetric its eigenvectors are automatically
mutually perpendicular as long as no two eigenvalues are the same.  If two, or all three eigenvalues are the
same, the eignevectors are not uniquely defined  in this case any convenient mutually
perpendicular set of eigenvectors can be used).
3. Recall that B can be expressed in terms of its
eigenvectors and eigenvalues as
 
 
 
3.5.2 Stress Measures used in finite elasticity
 
Usually stress-strain laws are given as equations relating Cauchy stress (`true’
stress)  to left Cauchy-Green deformation tensor.  For some computations it may be more
convenient to use other stress measures. 
They are defined below, for convenience. 
 
· The Cauchy (“true”) stress represents
the force per unit deformed area in the solid and is defined by
 
· Kirchhoff stress  
 
· Nominal (First Piola-Kirchhoff) stress  
 
· Material (Second Piola-Kirchhoff) stress  
 
 
 
 
3.5.3 Calculating
stress-strain relations from the strain energy density
 
The
constitutive law for an isotropic hyperelastic material is defined by an
equation relating the strain energy density of the material to the deformation
gradient, or, for an isotropic solid, to the three invariants of the strain
tensor
The
stress-strain law must then be deduced by differentiating the strain energy
density.   This can involve some tedious
algebra.  Formulas are listed below for
the stress-strain relations for each choice of strain invariant.  The results are derived below
 
 
· Strain energy density in terms of 
 
· Strain energy density in terms of 
 
· Strain energy density in terms of 
 
· Strain energy density in terms of 
 
· Strain energy density in terms of 
 
 
 
Derivations:   We start by deriving the general formula for
stress in terms of : 
 
1. Note that, by
definition, if the solid is subjected to some history of strain, the rate of change of the strain
energy density W (F)  must equal the rate of mechanical work
done on the material per unit reference volume.
 
2. Recall that the rate of work done per unit undeformed volume by body forces and
surface tractions is expressed in terms of the nominal stress  as .
 
3. Therefore, for any deformation
gradient Fij, 
This must hold for all possible ,so that
 
4. Finally, the formula for Cauchy
stress follows from the equation relating  to 
 
For an
isotropic material, it is necessary to find derivatives of the invariants with
respect to the components of F in order to compute the stress-strain
function for a given strain energy density. 
It is straightforward, but somewhat tedious to show that:
                                      
Then,
and
 
When using a
strain energy density of the form , 
we will have to compute the derivatives of the invariants  with respect to the components of F in
order to find
                                                         
We find that
 
Thus,
 
Next, we derive the stress-strain
relation in terms of a strain energy density  that is expressed as a function of the principal
strains.  Note first that
                                         
so that the chain rule gives
Using this and the expression that relates the stress
components to the derivatives of U,
we find that the principal stresses  are related to the corresponding principal
stretches  (square-roots of the eigenvalues of B)
through
The spectral decomposition for B in terms of
its eigenvalues  and eigenvectors 
now allows the stress tensor to be written as
Finally, if  is used for an anisotropic material then
 
 
 
3.5.4 A
note on perfectly incompressible materials
 
The preceding formulas assume that the material has some
(perhaps small) compressibility  that is to say, if you load it with
hydrostatic pressure, its volume will change by a measurable amount.   Most rubbers strongly resist volume changes,
and in hand calculations it is sometimes convenient to approximate them as
perfectly incompressible.   The material
model for incompressible materials is specified as follows:
 
· The deformation must satisfy J=1 to preserve volume.
 
· The strain energy density is
therefore only a function of two
invariants; furthermore, both sets of invariants defined above are identical.  We can use a strain energy density of the
form .
 
· Because you can apply any pressure to
an incompressible solid without changing its shape, the stress cannot be uniquely
determined from the strains.  
Consequently, the stress-strain law only specifies the deviatoric stress . 
In problems involving quasi-static loading, the hydrostatic stress  can usually be calculated, by solving the
equilibrium equations (together with appropriate boundary conditions).   Incompressible materials should not be used
in a dynamic analysis, because the speed of elastic pressure waves is infinite.
 
· The formula for stress in terms of  has the form
The hydrostatic stress p is an
unknown variable, which must be calculated by solving the boundary value
problem.
 
 
 
3.5.5 Specific
forms of the strain energy density 
 
· Generalized Neo-Hookean solid  (Adapted from Treloar 1948)
where  and  are material properties (for small
deformations,  and  are the shear modulus and bulk modulus of the
solid, respectively). Elementary statistical mechanics treatments predict that , where N is the number of polymer
chains per unit volume, k is the Boltzmann constant, and T is temperature.  This is a rubber elasticity model, for
rubbers with very limited compressibility, and should be used with . 
The stress-strain relation follows as
The fully incompressible limit can be obtained by setting  in the stress-strain law.
 
· Generalized Mooney-Rivlin solid (Adapted from Mooney 1940)
where  and  are material properties.  For small deformations, the shear modulus and
bulk modulus of the solid are  and . 
This is a rubber elasticity model, and should be used with . The stress-strain relation follows
as
 
· Generalized polynomial
rubber elasticity potential
where  and  are material properties.  For small strains the shear modulus and bulk
modulus follow as . This model is implemented in many finite
element codes.  Both the neo-Hookean
solid and the Mooney-Rivlin solid are special cases of the law (with N=1 and appropriate choices of  ). 
Values of  are rarely used, because it is difficult to
fit such a large number of material properties to experimental data.  
 
· Generalized Gent model
                                                  
where  and  are material properties.   The stress-strain law is
Since  as , the Gent material has a finite
stretchability.   It reduces to the
Neo-Hookean material in the limit .
 
 
· Ogden model (adapted
from Ogden, 1972)
where  , and  are material properties.  For small strains the shear modulus and bulk
modulus follow as . This is a rubber elasticity model,
and is intended to be used with . 
The stress can be computed using the formulas in 3.4.3, but are too
lengthy to write out in full here.
 
· Arruda-Boyce 8 chain model (Adapted from Arruda and Boyce,
1992)
where  are material properties.  For small deformations  are the shear and bulk modulus, respectively. This
is a rubber elasticity model, so .    The potential was derived by calculating the
entropy of a simple network of long-chain molecules, and the series is the
result of a Taylor expansion of an inverse Langevin function.  The reference provided lists more terms if
you need them.  The stress-strain law is
 
· Ogden-Storakers hyperelastic foam (Storakers, 1986)
where  are material properties.   For small strains the shear modulus and bulk
modulus follow as . 
 This is a foam model, and can
model highly compressible materials.  The
shear and compression responses are coupled.
 
· Blatz-Ko foam rubber (Blatz and Ko, 1962)
                                                            
where  is a material parameter corresponding to the
shear modulus at infinitesimal strains. The corresponding Poisson’s ratio for
such a material is 0.25.  The general stress-strain law is
 
 
 
3.5.6 Calibrating nonlinear
elasticity models
 
To use any of these constitutive relations, you will need to
determine values for the material constants. 
In some cases this is quite simple (the incompressible neo-Hookean
material only has 1 constant!); for models like the generalized polynomial or
Ogden’s it is considerably more involved. 
 
Conceptually, however, the procedure is straightforward.  You can perform various types of test on a
sample of the material, including simple tension, pure shear, equibiaxial
tension, or volumetric compression. It is straightforward to calculate the
predicted stress-strain behavior for the specimen for each constitutive
law.  The parameters can then be chosen
to give the best fit to experimental behavior. 
 
Here are some
guidelines on how best to do this:
 
1. When modeling the behavior of rubber
under ambient pressure, you can usually assume that the material is nearly incompressible,
and don’t need to characterize response to volumetric compression in detail.  For the rubber elasticity models listed above,
you can take  MPa. To fit the remaining parameters, you can
assume the material is perfectly incompressible.
 
2. If rubber is subjected to large
hydrostatic stress (>100 MPa) its volumetric and shear responses are
strongly coupled. Compression increases the shear modulus, and high enough
pressure can even induce a glass transition (see, e.g. D.L. Quested, K.D. Pae,
J.L. Sheinbein and B.A. Newman, (1981)). 
To account for this, you would have to use one of the foam models: in
the rubber models the volumetric and shear responses are decoupled. You would
also have to determine the material constants by testing the material under
combined hydrostatic and shear loading.  
 
3. 
For the simpler material models,
(e.g. the neo-Hookean solid, the Mooney-Rivlin material, or the Arruda-Boyce
model, which contain only two material parameters in addition to the bulk
modulus) you can estimate material parameters by fitting to the results of a
uniaxial tension test.  There are various
ways to actually do the fit  you could match the small-strain shear modulus
to experiment, and then select the remaining parameter to fit the stress-strain
curve at a larger stretch.  Least-squared
fits are also often used.  However, models
calibrated in this way do not always predict material behavior under multiaxial
loading accurately.
 
4. A more accurate description of
material response to multiaxial loading can be obtained by fitting the material
parameters to multiaxial tests.  To help
in this exercise, the nominal stress (i.e. force/unit undeformed area)
v- extension predicted by several
constitutive laws are listed in the table below (assuming perfectly
incompressible behavior, as suggested in item 1.). Specimen dimensions are
illustrated in the figure.
 

 
 
 
 
 
3.5.7
Representative values of material properties for rubbers 
 
The properties of rubber are strongly
sensitive to its molecular structure, and for accurate predictions you will
need to obtain experimental data for the particular material you plan to
use.    As a rough guide, the experimental
data of Treloar (1944) for the behavior of vulcanized rubber under uniaxial
tension, biaxial tension, and pure shear is shown in the figure. The
solid lines in the figure show the predictions of the Ogden model (which gives
the best fit to the data).
 
Material
parameters fit to this data for several constitutive laws are listed in the
table below.
 
 
 
  | 
   Neo-Hookean 
   | 
  
    MNm-2 
   | 
 
 
  | 
   Mooney-Rivlin 
   | 
  
     MNm-2,   MNm-2 
   | 
 
 
  | 
   Arruda-Boyce 
   | 
  
     MNm-2,
   
   | 
 
 
  | 
   Ogden 
   | 
  
    MNm-2,
   
   MNm-2,
   
   MNm-2,
   
   |