Chapter 6

Nonlinear Elasticity

Abstract

This chapter presents constitutive equations appropriate for a broad class of engineering materials known as nonlinear elastic solids. Common examples include rubber, elastomers (rubber-like polymers), and soft biological tissues. Nonlinear elastic solids are characterized by their ability to undergo large recoverable deformations and their highly nonlinear stress-strain response. Hence, they exhibit geometric nonlinearity (i.e., strain-displacement nonlinearity) due to finite elastic deformations, and material nonlinearity (i.e., stress-strain nonlinearity) due to nonlinear constitutive response. We examine nonlinear elastic materials in the context of the mechanical (isothermal) theory as well as the thermomechanical theory. In the latter case, we explicitly illustrate how the constitutive equations must satisfy the second law of thermodynamics, invariance, conservation of angular momentum, and material symmetry (isotropy).

Keywords

constitutive equations

finite elasticity

nonlinear elasticity

hyperelasticity

strain energy

rubber

compressible

Blatz-Ko

This chapter presents constitutive equations appropriate for a broad class of engineering materials known as nonlinear elastic solids. Common examples include rubber, elastomers (rubberlike polymers), and soft biological tissues. Nonlinear elastic solids are characterized by their ability to undergo large deformations before yielding, and their highly nonlinear stress-strain response.1,2 Hence, they exhibit geometric nonlinearity (or strain-displacement nonlinearity) owing to finite elastic deformations, and material nonlinearity owing to nonlinear constitutive response. Sections 6.1 and 6.2 discuss nonlinear elastic materials in the context of the mechanical (isothermal) theory and the thermomechanical theory, respectively. In the latter case, we explicitly illustrate how the constitutive equations must satisfy the second law of thermodynamics, invariance, conservation of angular momentum, and material symmetry (isotropy).

6.1 Mechanical theory

Recall from Section 5.1 that in the Eulerian formulation of the mechanical theory, a constitutive assumption is made on the Cauchy stress T as a function of the motion χ, and possibly its rates, gradients, and history, i.e.,

T=˘T(motion).

si42_e

For an elastic material, which has a perfect memory of some reference configuration, the Cauchy stress T at position x and time t depends on the motion only through the strain at x and t with respect to the reference configuration. The deformation gradient F = Grad χ (refer to Section 3.3.1) is a measure of this relative strain, so we may write

T=˘T(F).

si43_e  (6.1)

We emphasize that the stress at position x and time t in an elastic material depends only on the strain at x and t, and neither the history nor the rate of the strain (refer to Appendix B).

Equation (6.1) implies that the stress T at position x and time t is an explicit function of the deformation gradient F = Grad χ with respect to some reference configuration X, evaluated at the same x and t. Therefore, implicit in (6.1) is a dependence of T on a reference configuration X. Also implicit in (6.1) is a dependence of T on x and t, and a possible dependence of T on the present density ρ, since ρ is an algebraic function of the reference density ρR and the deformation gradient F:

ρ=ρRdetF.

si44_e

At this stage, it is useful to appeal to the mechanical energy balance

ρ˙ε=T·D,

si45_e  (6.2)

which is the first law of thermodynamics (5.26) specialized to a mechanical setting. Essentially, (6.2) indicates that in the mechanical theory, the rate of change of the internal energy ε is due solely to the stress power T · D; refer to Section 4.7. Arguing that the internal energy ε (like the Cauchy stress T) is a function of the strain F alone, i.e.,

ε=˘ε(F),

si46_e

and using the results of Problem 6.2 (which appear in the upcoming section), we have

˙ε=d˘εdF·˙F=d˘εdFFT·L.

si47_e

Hence, the mechanical energy balance (6.2) becomes

(Tρd˘εdFFT)·L=0,

si48_e  (6.3)

noting that T · L = T · D (refer to Problem 6.1 in the upcoming section). Since the coefficient of L in (6.3) is independent of L, and L is arbitrary, it follows that

T=ρd˘εdFFT.

si49_e  (6.4)

That is, the derivative of the internal energy ε with respect to the strain F gives the stress T. Hence, the internal energy is a strain energy. The class of materials for which the stress-strain response is derived from a strain energy potential, or stored elastic energy, is called hyperelastic.

Problem 6.1

In direct notation, show that T · L = T · D.

Solution

T·L=T·(D+W)(decomposition(3.57))=T·D+T·W(property(2.6)3)=T·D(result(2.44)).

si1_e

In the second step, we were able to employ the distributive property (2.6)3 of the inner product since it was demonstrated (in Problem 2.34) that the set of all second-order tensors is an inner product space. In the third step, we have exploited that T · W = 0, since T is symmetric (by conservation of angular momentum; see (5.25)) and W is skew (by construction; see (3.58)2).

Problem 6.2

Prove in direct notation that ψF·˙F=ψFFT·Lsi2_e.

Solution

ψF·˙F=ψF·(LF)(result(3.60)1)=tr[ψF(LF)T](definition(2.41))=tr[ψF(FTLT)](result(2.14)2)=tr[(ψFFT)LT](associativity of tensor multiplication)=ψFFT·L(definition(2.41)).

si3_e

Recall from Section 5.2 that the constitutive equation (6.4) for the Cauchy stress T must satisfy conservation of angular momentum, invariance requirements, and material symmetry conditions. We postpone the exploration of these requirements until Section 6.2.

6.2 Thermomechanical theory

Recall from Section 5.3 that in formulation (5.36)(5.38) of the thermomechanical theory, constitutive assumptions are made on the Cauchy stress T, heat flux vector q, Helmholtz free energy ψ, and entropy η as functions of the motion χ and temperature Θ, and possibly their rates, gradients, and histories, i.e.,

T=ˉT(motion and temperature),q=ˉq(motion and temperature),ψ=ˉψ(motion and temperature),η=ˉη(motion and temperature).

si50_e

For a thermoelastic material, which has a perfect memory of its reference configuration and temperature, the list of arguments in (6.1) is expanded from χ (implicit) and F = Grad χ (explicit) to include the analogous thermal quantities Θ and gR = GradΘ. Therefore, for a thermoelastic material, we write

T=ˉT(F,Θ,gR),q=ˉq(F,Θ,gR),ψ=ˉψ(F,Θ,gR),η=ˉη(F,Θ,gR).

si51_e  (6.5)

This notation means that T, q, ψ, and η at position x and time t are explicit functions of F, Θ, and gR evaluated at x and t (and, thus, implicit functions of x and t). There is an implicit dependence on the reference configuration X through F = Grad χ and gR = Grad Θ. Recall from (4.65) the relationship gR = FTg between the referential and spatial temperature gradients, so gR is a function of g and F. We may therefore replace (6.5) with

T=˘T(F,Θ,g),q=˘q(F,Θ,g),ψ=˘ψ(F,Θ,g),η=˘η(F,Θ,g).

si52_e  (6.6)

Note that ˉTsi53_e in (6.5) and ˘Tsi54_e in (6.6) denote two different response functions for T. Also note that the constitutive functions for T, q, ψ, and η in (6.6) do not depend on the history or the rate of the strain or temperature (refer to Appendix B).

Now that a list of arguments has been specified for the constitutive functions that characterize a thermoelastic material, the number of independent constitutive functions, and the list of arguments itself, can be reduced. As described in Section 5.4 and illustrated in the following sections, this reduction is accomplished via the second law of thermodynamics, invariance requirements, conservation of angular momentum, and material symmetry considerations.

6.2.1 Restrictions imposed by the second law of thermodynamics

Substitution of the constitutive assumptions (6.6) into the Clausius-Duhem inequality (5.35) gives

ρ˘ψF·˙F+˘T·Dρ(˘ψΘ+˘η)˙Θρ˘ψg·˙g1Θ˘q·g0,

si55_e  (6.7)

where we have used the chain rule (refer to Section 2.5.2):

˙ψ=˘ψF·˙F+˘ψΘ˙Θ+˘ψg·˙g.

si56_e

It can be shown (refer to Problems 6.1 and 6.2) that

˘T·D=˘T·L,˘ψF·˙F=˘ψFFT·L,

si57_e

so inequality (6.7) becomes

(˘Tρ˘ψFFT)·Lρ(˘ψΘ+˘η)˙Θρ˘ψg·˙g1Θ˘q·g0.

si58_e  (6.8)

Inequality (6.8) must hold for all processes. In particular, it must hold for the family of processes with g = 0, ˙g=0si59_e, and L = 0, but ˙Θsi60_e arbitrary, at particular position x and time t. (An example is the family of processes (D.1) in Appendix D with A = 0, a = 0, and g0 = 0, but a any real number.) For members of this family of processes, (6.8) simplifies to

ρ(˘ψΘ+˘η)˙Θ0.

si61_e

The density ρ is positive, which implies

(˘ψΘ+˘η)˙Θ0.

si62_e  (6.9)

The coefficient of ˙Θsi63_ein (6.9) is independent of ˙Θsi64_e. For (6.9) to hold for all members of this family, the coefficient of ˙Θsi65_e must vanish, i.e.,

η=˘ψΘ.

si66_e  (6.10)

Thus, the partial derivative of the response function for the Helmholtz free energy ψ with respect to temperature Θ gives the value of the entropy η, so ˘ηsi67_e is not an independent response function. Result (6.10) holds for all processes, not just the special family considered above. Hence, inequality (6.8) reduces to

(˘Tρ˘ψFFT)·Lρ˘ψg·˙g1Θ˘q·g0.

si68_e  (6.11)

Inequality (6.11) must hold for all processes, in particular the family of processes with g = 0 and L = 0, but ˙gsi69_e arbitrary, at x and t (e.g., the family (D.1) in Appendix D with A = 0 and g0 = 0, but a any vector). For members of this family, (6.11) simplifies to

ρ˘ψg·˙g0.

si70_e

The density is positive; hence,

˘ψg·˙g0.

si71_e  (6.12)

The coefficient of ˙gsi72_e in (6.12) is independent of ˙g si73_e. Hence, the rate ˙gsi74_e may be chosen to violate the inequality (6.12) unless its coefficient is the zero vector. Therefore,

˘ψg=0.

si75_e

Thus, if the response function for ψ depended on the temperature gradient g, the second law would be violated for some processes. Since the second law must be obeyed for all processes, we conclude that a thermoelastic response function for ψ must be independent of g. Inequality (6.11) thus reduces to

(˘Tρ˘ψFFT)·L1Θ˘q·g0.

si76_e  (6.13)

which must hold for all processes.

Consider those processes with g = 0 but L arbitrary at x and t (e.g., the family (D.1) in Appendix D with g0 = 0, F0 fixed and invertible, and A any tensor). For these processes, (6.13) simplifies to

(˘Tρ˘ψFFT)·L0.

si77_e  (6.14)

The coefficient of the rate L in (6.14) is independent of rates. Hence, we obtain

T=ρ˘ψFFT.

si78_e

Thus, the partial derivative of the response function for the Helmholtz free energy ψ with respect to strain F gives the value of the stress T, i.e., the Helmholtz free energy is a strain energy. Recall that the class of materials for which the stress-strain response is derived from a strain energy potential is called hyperelastic. The inequality (6.13) thus reduces to

1Θ˘q·g0

si79_e

or

˘q·g0,

si80_e  (6.15)

since the absolute temperature Θ is strictly positive. Inequality (6.15) cannot be simplified any further since ˘qsi81_e depends on g. Thus, the Clausius-Duhem inequality has been reduced for thermoelastic materials to an intuitive statement of the second law: heat flows against the temperature gradient, from hot to cold.

We now pause and take stock of our accomplishments. By demanding that the Clausius-Duhem inequality holds for all thermoelastic processes, we have reduced the constitutive assumption (6.6) to

ψ=˘ψ(F,Θ),q=˘q(F,Θ,g)with˘q·g0,

si82_e  (6.16a)

and

T=ρ˘ψFFT,η=˘ψΘ.

si83_e  (6.16b)

Thus, as a consequence of the second law, we have found that:

(1) Only two response functions ( ˘ψsi84_e and ˘qsi85_e ) are needed to characterize a thermoelastic material, rather than the four initially supposed. (Nice! Fewer response functions means fewer unknowns in the characterization problem.)

(2) The response function for ψ is independent of g. (Again, fewer experiments to perform.)

(3) The response function for q is restricted by the second law through the inequality ˘q·g0si86_e, i.e., heat flows against the temperature gradient.

It can be shown (refer to Problem 6.3) that if the heat flux vector ˘qsi87_e in (6.6) is a continuous function of temperature gradient g at g = 0, restriction (3) above has an additional implication: ˘qsi88_e evaluated at g = 0 is zero, i.e., heat does not flow in the absence of a temperature gradient.

Problem 6.3

Show that if the heat flux vector ˘qsi4_e in (6.6) is a continuous function of the temperature gradient g at g = 0, then it follows from ˘q·g0si5_e that ˘qsi6_e evaluated at g = 0 is zero. That is, heat does not flow in the absence of a temperature gradient.

Solution

If ˘qsi7_e is a continuous function of g at g = 0, then it follows that

q=˘q(F,Θ,g)=a(F,Θ)+b(F,Θ,g),

si8_e

with

limα0b(F,Θ,αg)=0,

si9_e

so

˘q(F,Θ,g)|g=0=limα0˘q(F,Θ,αg)=a(F,Θ).

si10_e

Then ˘q·g0si11_e implies

a(F,Θ)·g+b(F,Θ,g)·g0.

si12_e

Replacing g by αg, taking α > 0, and dividing by α leads to

a(F,Θ)·g+b(F,Θ,αg)·g0.

si13_e

Taking the limit as α → 0 gives

a(F,Θ)·g0,

si14_e

which must hold for all processes. Since a(F, Θ) is independent of g, this implies that a(F, Θ) = 0, or

˘q(F,Θ,g)|g=0=0.

si15_e

6.2.2 Restrictions imposed by invariance under superposed rigid body motions and conservation of angular momentum

Our appeal to the second law has proven fruitful. We now appeal to invariance requirements under superposed rigid body motions (refer to Sections 5.2.1 and 5.4.1) and obtain further restrictions. (Note that any restrictions on the response functions are very helpful, as they reduce the number of experiments needed to characterize a particular material.) In particular, we must have

ψ+=ψ,q+=Qq

si89_e

when

F+=QF,Θ+=Θ,g+=Qg

si90_e

for all proper orthogonal tensors Q(t). It can be shown (refer to Problems 6.4 and 6.5) that these invariance requirements demand

ψ=˘ψ(F,Θ)=ψ(C,Θ)

si91_e  (6.17)

and

q=˘q(F,Θ,g)=Fq(C,Θ,gR).

si92_e  (6.18)

Note that ˘ψsi93_e and ψsi94_e denote two different response functions for ψ. Equation (6.17) implies that the response function for ψ can depend on F only through the combination C = FTF, where C is Green's deformation tensor; the dependence on Θ is unrestricted. Equation (6.18) implies that the dependence of the response function for q on g must be through gR = FTg (where gR is the referential temperature gradient), there must be a linear dependence of q on F, and any further dependence on F must be through the combination C = FTF.

Problem 6.4

Prove that the invariance requirement ψ+ = ψ is satisfied if and only if

ψ=˘ψ(F,Θ)=ψ(C,Θ).

si16_e

Solution

The phrase “if and only if” requires that we demonstrate both necessity and sufficiency in our proof. To demonstrate necessity, we must show that the invariance requirement ψ+ = ψ implies that ψ=ψ(C,Θ)si17_e. To demonstrate sufficiency, we must show that ψ=ψ(C,Θ)si18_e satisfies the invariance requirement ψ+ = ψ.

Necessity: First, note that

ψ=˘ψ(F,Θ),ψ+=˘ψ(F+,Θ+).

si19_e

The invariance requirement

ψ+=ψwhenF+=QF,Θ+=Θ

si20_e

for all proper orthogonal Q implies that

˘ψ(QF,Θ)=˘ψ(F,Θ).

si21_e

Since this must hold for all proper orthogonal Q, it must hold for the particular case Q = RT, so

˘ψ(F,Θ)=˘ψ(RTF,Θ).

si22_e

(Recall from Section 3.3.3 that R is the proper orthogonal rotation tensor in the polar decomposition of F.) Since

RTF=RTRU=IU=U=C1/2.

si23_e

it follows that

ψ=˘ψ(F,Θ)=˘ψ(RTF,Θ)=˘ψ(U,Θ)=˘ψ(C1/2,Θ)=ψ(C,Θ).

si24_e

Sufficiency: It follows from ψ=ψ(C,Θ)si25_e that

ψ+=ψ(C+,Θ+)=ψ(C,Θ)=ψ,

si26_e

where we have used the results C+ = C and Θ+ = Θ from Sections 5.2.1.2 and 5.4.1.1.

Problem 6.5

Prove that the invariance requirement q+ = Qq is satisfied if and only if

q=˘q(F,Θ,g)=Fq(C,Θ,FTg).

si27_e

Solution

As was the case in Problem 6.4, the phrase “if and only if” requires that we demonstrate both necessity and sufficiency in our proof. To demonstrate necessity, we must show that the invariance requirement q+ = Qq implies that q=Fq(C,Θ,FTg)si28_e. To demonstrate sufficiency, we must show that q=Fq(C,Θ,FTg)si29_e satisfies the invariance requirement q+ = Qq.

Necessity: First, note that

q=˘q(F,Θ,g),q+=˘q(F+,Θ+,g+).

si30_e

The invariance requirement

q+=QqwhenF+=QF,Θ+=Θ,g+=Qg

si31_e

implies that

˘q(QF,Θ,Qg)=Q˘q(F,Θ,g),

si32_e

which must hold for all proper orthogonal Q, and, in particular, Q = RT. Since

RTF=RTRU=IU=U,

si33_e

it follows that

˘q(F,Θ,g)=R˘q(U,Θ,RTg).

si34_e

We define a new function ˉqsi35_e by

˘q(U,Θ,RTg)=Uˉq(U,Θ,URTg),

si36_e

which gives

˘q(F,Θ,g)=RUˉq(U,Θ,URTg).

si37_e

We have

F=RU,FT(RU)T=UTRT=URT,U=C1/2,

si38_e

which imply that

˘q(F,Θ,g)=Fˉq(C1/2,Θ,FTg)=Fq(C,Θ,FTg).

si39_e

Sufficiency: It follows from q=Fq(C,Θ,FTg)si40_e that

q+=F+q(C+,Θ+,(F+)Tg+)=QFq(C,Θ,FTQTQg)=QFq(C,Θ,FTg)=Qq,

si41_e

where we have used the results F+ = QF, C+ = C, g+ = Qg, and Θ+ = Θ from Sections 5.2.1.2 and 5.4.1.1.

Now that we have established ψ=ψ(C,Θ)si95_e, it can be shown, through a change of independent variable from F to C, that (6.16b)1 becomes

T=ρ˘ψ(F,Θ)FFT=ρF(ψ(C,Θ)C+ψ(C,Θ)CT)FT.

si96_e  (6.19)

If ψ(C,Θ)si97_e is a symmetric function of C, then (6.19) becomes

T=2ρFψ(C,Θ)CFT.

si98_e  (6.20)

It can be shown that (6.19) and (6.20) ensure that the Cauchy stress T is symmetric, so conservation of angular momentum (5.25) is satisfied.

To summarize the results of this section thus far: A thermoelastic material has a perfect memory of its reference state. On physical grounds, it is assumed that such a material is characterized by at most four response functions, each of which could conceivably depend on the deformation gradient F (with respect to some reference configuration), temperature Θ, and temperature gradient g, all evaluated at present time t (i.e., the response at present time t depends on the present values of F, Θ, and g). The four response functions and their lists of arguments are

T=˘T(F,Θ,g),q=˘q(F,Θ,g),ψ=˘ψ(F,Θ,g),η=˘η(F,Θ,g).

si99_e

Note that the dependence on deformation does not involve its history, but rather just the present value of the deformation gradient from some reference configuration.

Because these response functions must be consistent with conservation of angular momentum, the second law of thermodynamics for all thermomechanical processes, and invariance requirements under any superposed rigid body motion, we find that only two of the response functions are independent, i.e.,

ψ=ψ(C,Θ),q=Fq(C,Θ,gR),

si100_e  (6.21a)

from which

T=2ρFψ(C,Θ)CFT,η=ψ(C,Θ)Θ

si101_e  (6.21b)

can be deduced, with the restrictions

q(C,Θ,0)=0,q·g0.

si102_e  (6.21c)

Note that the Helmholtz free energy, Cauchy stress, and entropy are independent of temperature gradient. Also note that the Clausius-Duhem inequality reduces to the intuitive second law statements (6.21c)1 and (6.21c)2, i.e., heat does not flow in the absence of a temperature gradient, and in the presence of a temperature gradient, heat flows opposite the temperature gradient, from hot to cold.

Exercises

1. Using indicial notation, verify through a change of independent variable from F to C that

T=ρ˘ψ(F,Θ)FFT=ρF(ψ(C,Θ)C+ψ(C,Θ)CT)FT=2ρFψ(C,Θ)CFT,

si103_e

where the last equality holds if ψ(C,Θ)si104_e is a symmetric function of C, i.e.,

ψ(C,Θ)C=ψ(C,Θ)CT.

si105_e

2. Verify that (6.19) and (6.20) ensure that the Cauchy stress T is symmetric, so conservation of angular momentum is satisfied.

6.2.3 Restrictions imposed by material symmetry: isotropy

In this section, we apply material symmetry considerations (refer to Section 5.2.2) to further reduce and simplify the form of the response functions. In particular, we consider thermoelastic materials that are isotropic. Loosely speaking, isotropic materials have properties that are identical in all directions.

Recall from Section 6.2.1 that by demanding the Clausius-Duhem inequality hold for all thermoelastic processes, we found that only two response functions are necessary to characterize a thermoelastic material, i.e.,

ψ=˘ψ(F,Θ),q=˘q(F,Θ,g),

si106_e

where the dependence of the deformation gradient F on a particular reference configuration κ is understood. Here, κ is chosen to be the stress-free configuration. By analogy to (5.22), we have

˘ψ(F,Θ)=˘ψ(FH1,Θ)for allHG,

si107_e  (6.22)

where Gsi108_e is the symmetry group of the material relative to the stress-free reference configuration κ, and H is a symmetry transformation with respect to κ. Since a thermoelastic material is a solid, we have GOsi109_e, where Osi110_e is the full orthogonal group. Then (6.22) becomes

˘ψ(F,Θ)=˘ψ(FHT,Θ)for allHG.

si111_e  (6.23)

Recall from Section 6.2.2 that in order to satisfy invariance under superposed rigid body motions and conservation of angular momentum, we must specify ψ to be a function of F in the particular combination C = FTF, i.e.,

ψ=ψ(C,Θ),

si112_e

where C is Green's deformation tensor. The condition on ψsi113_e corresponding to (6.23) is

ψ(C,Θ)=ψ(HCHT,Θ)for allHG.

si114_e  (6.24)

For isotropic thermoelastic solids, the symmetry group Gsi115_e relative to the stress-free reference configuration κ is the full orthogonal group Osi116_e. Therefore, (6.24) becomes

ψ(C,Θ)=ψ(HCHT,Θ)for allHO.

si117_e  (6.25)

The requirement (6.25) on the functional form of ψsi118_e is equivalent to the condition that ψsi119_e depend on C only through the principal scalar invariants of C, i.e.,

ψ=ˉψ(I1,I2,I3,Θ),

si120_e  (6.26a)

where

I1=trC,I2=12[(trC)2tr(C2)],I3=detC.

si121_e  (6.26b)

Recalling that the Cauchy stress T is obtained from the Helmholtz free energy ψ by

T=2ρFψ(C,Θ)CFT,

si122_e

we have, through a change of independent variable,

T=2ρF(ˉψI1dI1dC+ˉψI2dI2dC+ˉψI3dI3dC)FT.

si123_e  (6.27)

Recall from Section 2.5.1 the results

dI1dC=I,dI2dC=I1IC,dI3dC=I3C1.

si124_e  (6.28)

Substitution of (6.28) into (6.27) gives

T=α0I+α1B+α2B2,

si125_e  (6.29a)

where B = FFT is the Finger deformation tensor and

α0=2ρI3ˉψI3,α1=2ρ(ˉψI1+I1ˉψI2),α2=2ρˉψI2.

si126_e  (6.29b)

Note that since B and C have the same eigenvalues, they also have the same principal invariants (refer to Sections 2.3 and 3.3.5). Use of the Cayley-Hamilton theorem (2.75) gives the alternative form

T=β0I+β1B+β1B1,

si127_e  (6.30a)

where

β0=2ρ(I2ˉψI2+I3ˉψI3),β1=2ρˉψI1,β1=2ρI3ˉψI2.

si128_e  (6.30b)

In nonlinear elasticity it is common to speak of the strain energy W per unit reference volume, i.e., the strain energy density, defined by

W=ρRψ,

si129_e

rather than the Helmholtz free energy. For isotropic thermoelastic materials, we have

W=ˉW(I1,I2,I3,Θ).

si130_e

It follows that (6.30b) becomes

β0=2I123(I2ˉWI2+I3ˉWI3),β1=2I1/23ˉWI1β1=2I1/23ˉWI2,

si131_e  (6.31)

where we have used

ρρR=1J=1detF=I1/23,

si132_e

which follows from conservation of mass (4.56a). Recall that ρ is the density in the present configuration, ρR is the density in the reference configuration, and J is the determinant of the deformation gradient F.

Exercises

1. Verify that substitution of (6.28) into (6.27) gives (6.29a) and (6.29b).

2. Verify that (6.30a) and (6.30b) follow from (6.29a) and (6.29b) and use of the Cayley-Hamilton theorem.

3. Verify (6.31).

6.3 Strain energy models

We found in Section 6.2 that the Cauchy stress for an isotropic nonlinear elastic material can be expressed as

T=β0I+β1B+β1B1,

si133_e

where

β0=2I1/23(I2ˉWI2+I3ˉWI3),β1=2I1/23ˉWI1β1=2I1/23ˉWI2,

si134_e

B = FFT is the Finger deformation tensor, W=ˉW(I1,I2,I3,Θ)si135_e is the strain energy density, and I1, I2, and I3 are the principal invariants of B. We now specialize this constitutive model to the mechanical (isothermal) theory by eliminating the temperature dependence of W, so W=ˉW(I1,I2,I3)si136_e. It remains only to specify the dependence of the strain energy ˉWsi137_e on the invariants I1, I2, and I3.

Several examples of invariant-based strain energy models for compressible rubberlike materials are the Blatz-Ko model [10],

W=μ2{f[(I13)+1γ(Iγ31)]+(1f)[(I2I33)+1γ(Iγ31)]},

si138_e  (6.32)

the compressible neo-Hookean model [11, p. 247],

W=μ2[(I13)+1γ(Iγ31)],

si139_e  (6.33)

the compressible Mooney-Rivlin model [11, p. 247],

W=c1(I13)+c2(I23)+c3(I(1/2)31)2(c1+2c2)lnI3,

si140_e  (6.34)

and the Levinson-Burgess (polynomial) model [12],

W=μ2[f(I13)+(1f)(I2I33)+2(12f)(I(1/2)31)+(2f+4v112v)(I(1/2)31)2].

si141_e  (6.35)

In (6.32)(6.35), γ = v/(1 − 2v); μ and v are the shear modulus and Poisson's ratio evaluated at small strains; and f, c1, c2, and c3 are parameters that can be adjusted to fit experimental data for a particular rubbery material. Note that (6.33) can be obtained as a special case of (6.32) by setting f = 1.

In contrast to the strain energy models (6.32)(6.35) that are based on the principal invariants I1, I2, I3 of B, the compressible Ogden model [13] is based on the principal stretches λ1, λ2, λ3:

W=nμn[λαn1+λαn2+λαn33αnlnJ]+λβ2(βlnJ+Jβ1),

si142_e  (6.36)

where J is the determinant of the deformation gradient F, λ is the second Lamé constant evaluated at small strains, and μn, αn, β, and n are adjustable parameters. Recall from Sections 2.3 and 3.3.5 that the principal stretches λ1, λ2, λ3 are related to the principal invariants I1, I2, I3 of B through

I1=λ21+λ22+λ23,I2=λ21λ22+λ21λ23+λ22λ23,I3=λ21λ22λ23.

si143_e  (6.37)

Other models, such as those developed by Anand [14] and Bischoff et al. [15], are based on statistical mechanics, and thus account for the underlying deformation physics of the polymer chains.

Many other invariant-based, stretch-based, and statistical-mechanics-based strain energy models for compressible rubbery materials—beyond the representative few presented here—can be found, for instance, in books by Holzapfel [11], Treloar [16], and Ogden [17], and review articles by Ogden [18] and Boyce and Arruda [19].

References

[10] Blatz PJ, Ko WL. Application of finite elastic theory to the deformation of rubbery materials. Trans. Soc. Rheol. 1962;6(1):223–251.

[11] Holzapfel GA. Nonlinear Solid Mechanics: A Continuum Approach for Engineering. Chichester: John Wiley & Sons; 2000.

[12] Levinson M, Burgess IW. A comparison of some simple constitutive relations for slightly compressible rubber-like materials. Int. J. Mech. Sci. 1971;13(6):563–572.

[13] Ogden RW. Large deformation isotropic elasticity: On the correlation of theory and experiment for compressible rubberlike solids. Proc. R. Soc. Lond. A. 1972;328(1575):567–583.

[14] Anand L. A constitutive model for compressible elastomeric solids. Comput. Mech. 1996;18(5):339–355.

[15] Bischoff JE, Arruda EM, Grosh K. A new constitutive model for the compressibility of elastomers at finite deformations. Rub. Chem. Tech. 2001;74(4):541–559.

[16] Treloar LRG. The Physics of Rubber Elasticity. Oxford: Clarendon Press; 1975.

[17] Ogden RW. Non-Linear Elastic Deformations. Chichester: Ellis Horwood; 1984.

[18] Ogden RW. Elastic deformations of rubberlike solids. In: Hopkins HG, Sewell MJ, eds. Mechanics of Solids, The Rodney Hill 60th Anniversary Volume. Oxford: Pergamon Press; 1982:499–537.

[19] Boyce MC, Arruda EM. Constitutive models of rubber elasticity: a review. Rub. Chem. Tech. 2000;73(3):504–523.


star “To view the full reference list for the book, click here

1 An elastic material has a perfect memory of its reference configuration. So, if an elastic material deforms away from its reference configuration, it returns once the force responsible for the deformation is removed. Further, the response of an elastic material does not depend on the history or the rate of the deformation.

2 Large deformations are also referred to as finite deformations, much like small deformations are referred to as infinitesimal deformations.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset