You are on page 1of 11

J Elast (2020) 142:383–393

https://doi.org/10.1007/s10659-020-09798-1

Pure Shearing and Pure Distortional Deformations Are


Not Equivalent

M.B. Rubin1

Received: 24 May 2020 / Accepted: 26 September 2020 / Published online: 7 October 2020
© Springer Nature B.V. 2020

Abstract This paper attempts to clarify the notions of a state of pure shear stress and pure
shearing deformations. Specifically, it is shown that pure shearing deformations and pure
distortional deformations are not equivalent. Attention is limited to isotropic, compressible,
hyperelastic materials. Differences between the distortional deformations of pure shearing,
pure shear caused by tension and compression, and plane strain extension and contraction
defined as pure shear by Rivlin and Saunders (Philos. Trans. R. Soc. Lond. Ser. A, Math.
Phys. Sci. 243(865):251–288, 1951) are discussed. It is shown that these deformations are
physically different and should not be expected to test the same features of a proposed form
of the strain energy function. It is also shown that two deformations of pure shearing and
two deformations of pure shear caused by tension and compression are nearly universal
distortional deformations valid for all strain energy functions.

Keywords Finite deformation · Isotropic hyperelasticity · Pure distortional deformation ·


Pure shearing deformation · Nearly universal distortional deformations

Mathematics Subject Classification 74A05 · 74A10 · 74B20

1 Introduction
Finite deformation pure shear and related shear deformations in isotropic, compressible,
hyperelastic materials have been considered at least since the experiments conducted by
[14] on rubber. Analytical work on a state of pure shear stress can be found in [12, 15]. [3]
considered a fixed deformation and determined pairs of material line elements which are
unsheared as well a pairs of line elements which suffer the maximum shear, and a class of
universal relations was developed in [1]. A more recent discussion of universal results can
be found in [18]. Restrictions on constitutive equations related to the Poynting effect were
considered in [11], the response to pure shear was revisited in [6], and distortional isochoric

B M.B. Rubin
mbrubin@tx.technion.ac.il
1 Faculty of Mechanical Engineering, Technion-Israel Institute of Technology, 32000 Haifa, Israel
384 M.B. Rubin

plane strain deformations, called (pure) shear (stretch), were analyzed in [19]. These later
deformations should not be confused with the three-dimensional deformations associated
with a state of pure shear stress discussed in this paper.
By way of background, it is recalled that [7] introduced the notion of a unimodular tensor
to characterize pure distortional deformation. Specifically, the deformation gradient F from
a zero-stress initial configuration is used to define the dilatation J , its unimodular part F
and the unimodular part B of the left Cauchy-Green deformation tensor by

J = detF , F = J −1/3 F , B = F FT , detB = 1 ,


(1)
α1 = B · I ≥ 3 , α2 = B2 · I ≥ 3 .

In these expressions, J is a pure measure of dilatation, B is a pure measure of distortional


deformation, α1 and α2 are two independent nontrivial invariants of B , I is the second order
unit tensor and A · B = tr(ABT ) is the inner product of two second order tensors A and B.
Also, recall that the symmetric Cauchy stress T tensor can be expressed in the form

1
T = −pI + T , p = − (T · I) , (2)
3
where T is its deviatoric part and p is the pressure.
The discussions in [2, 4, 8, 9, 13, 20] can be used to prove that an orthonormal triad
of vectors ni exits for which the components of a general stress tensor T and its general
deviatoric part T can be represented in the special forms
⎡ ⎤ ⎡ ⎤
−p T12 T13 0 T12 T13
⎢ ⎥ ⎢ ⎥
T · ni ⊗ nj = ⎢
⎣ T12 −p T23 ⎥
⎦, T · ni ⊗ nj = ⎢
⎣ T12 0 T23 ⎥
⎦, (3)
T12 T23 −p T12 T23 0

In the absence of pressure p = 0, (3)2 has been identified in the literature as a state of pure
shear. However, it would be better to identify (3)2 as a pure shear representation of the
deviatoric stress T , since it is valid for an arbitrary stress state.
It is well known that an orthonormal triad of eigenvectors pi exists for which the com-
ponents of a general stress tensor T and its general deviatoric part T can be represented in
the diagonal forms
⎡ ⎤ ⎡ ⎤
σ1 0 0 σ1 0 0
⎢ ⎥ ⎢ ⎥
T · pi ⊗ pj = ⎢
⎣0 σ2 0 ⎥⎦, T · pi ⊗ pj = ⎢
⎣0 σ2 0 ⎥ ⎦, (4)

0 0 σ3 0 0 σ3

where the eigenvalues σi and σi have the properties that

σi = −p + σi , σ1 ≥ σ2 ≥ σ3 , σ1 + σ2 + σ3 = 0 . (5)

This is also a special representation of a general stress state.


Three independent invariants of stress can be defined by

1 3  
p = − T · I, σe = T ·T , detT , (6)
3 2
Pure Shearing and Pure Distortional Deformations Are Not Equivalent 385

where σe is the von Mises stress. Assuming that σe is nonzero, the Lode angle β can be
defined by [5, 10, 16]
27detT π π
sin β = − , − ≤β ≤ , (7)
2σe3 6 6
and the eigenvalues σi of a general deviatoric state of stress can be expressed as

2σe π 2σe 2σe π


σ1 = cos( + β) , σ2 = , σ3 = − cos( − β) . (8)
3 6 3 3 6
In contrast with the representations (3) and (4), which depend on specific choices of the
base vectors, restrictions on the invariants of stress are independent of the choice of the base
vectors. For examples: when the pressure vanishes p = 0, the stress state is pure deviatoric;
when the von Mises stress vanishes σe = 0, the stress state is pure hydrostatic; and when
detT = 0, the Lode angle β and the middle eigenvalue σ2 of the deviatoric stress vanish

detT = 0 ⇒ β = 0, σ2 = 0 . (9)

An invariant definition of a state of pure shear stress, which is used throughout the remainder
of this paper, requires
p = 0, detT = 0 . (10)
It is noted that this state of pure shear stress was considered as a special case in [20].
General isotropic, compressible, hyperelastic response can be modeled by the strain en-
ergy function  per unit mass [e.g., [17]]

 = (J, α1 , α2 ) , (11)

with p and T determined by the hyperelastic equations


∂ ∂  1 ∂ 2 1
p = −ρJ , T = 2ρ (B − α1 I) + 4ρ (B − α2 I) , (12)
∂J ∂α1 3 ∂α2 3
where ρ is the current mass density of the material. In addition, the functional form for  is
restricted so that the unique solution of a zero-stress state requires

T=0 ⇒ J =1 and B = I . (13)

From (12) it follows that a pure hydrostatic stress state is characterized by


∂  1 ∂ 2 1
T = 2ρ (B − α1 I) + 4ρ (B − α2 I) = 0 , (14)
∂α1 3 ∂α2 3
which yields the universal solution of no distortional deformation

T = 0 ⇒ B = I . (15)

This means that an isotropic, compressible, hyperelastic material subjected to a hydrostatic


stress state experiences pure dilatation with no distortional deformation. In contrast, when
the pressure vanishes p = 0, distortional deformations (i.e. B = I) can cause dilatational
deformation with J determined by solving the equation
∂
= 0, (16)
∂J
386 M.B. Rubin

Fig. 1 (PS) Sketch of pure


shearing with the stretches
b ≥ a > 0, shear γ ≥ 0 and the
constant unit normal vector n to
the top material surface

which depends explicitly on the functional form of . Also, it is obvious from the general
form (12) that T commutes with B

B T = T B . (17)

A similar expression was used in [12] to develop the universal relationship (22) for pure
shearing in the next section and was shown in [1] to generate a class of universal relations.
For a state of pure shear stress (10) with constant direction (i.e. a constant direction of
T ), the eigenvectors of T remain constant and the two nonzero eigenvalues of T are equal
in magnitude and opposite in sign. Consequently, fixed rectangular Cartesian base vectors
ei can be specified so that the eigenvectors p1 and p2 of T associated with these nonzero
eigenvalues lie in the e1 -e2 plane. Then, the Cauchy stress T can be written in the form

T = T = τ (e1 ⊗ e2 + e2 ⊗ e1 ) = τ (p1 ⊗ p1 − p2 ⊗ p2 ) , (18)

where τ is the shear stress and a ⊗ b denotes the tensor product of two vectors a and b.
Also, the fixed orthonormal eigenvectors pi of T are defined by

1 1
p1 = √ (e1 + e2 ) , p2 = √ (−e1 + e2 ) , p3 = e3 . (19)
2 2

Since  can depend on all three invariants J , α1 and α2 , and the pressure vanishes for a
state of pure shear stress (10), the dilatation J determined by the solution of (16) need not be
unity. Consequently, the deformation associated with a state of pure shear stress need not be
a pure distortional deformation, which is isochoric. This seems to be a cause for confusion
in the literature. For example, [19] introduce the notion of (pure) shear (stretch) based on
plane strain, isochoric, pure distortional deformation right and left stretch tensors.
In this paper, two nearly universal distortional deformations called pure shearing (PS),
sketched in Fig. 1 and discussed in Sect. 3, and two nearly universal distortional deforma-
tions called (PS-TC), sketched in Fig. 2 and discussed in Sect. 4, caused by pure shear
stress due to tension and compression are shown to be consistent with the same state of pure
shear stress (18). The deformation (PS) is called a pure shearing deformation since it has the
property that the angle between any two distinct material line elements in the shearing plane
Pure Shearing and Pure Distortional Deformations Are Not Equivalent 387

Fig. 2 (PS-TC) Sketch of pure


shear caused by tension and
compression with the extension
stretch (a ≥ 1) being orthogonal
to the contraction stretch (b ≤ 1)
and with n and s being,
respectively, the unit normal
vector and unit tangent vector to
the top surface

Fig. 3 (PS-EC): Plane strain


extension (a ≥ 1) and contraction
(b ≤ 1) defined as pure shear in
[14]

changes with the amount of shearing throughout the shearing process. In contrast, the de-
formation associated with (PS-TC) has the property that the three material line elements in
the principal directions of deformation remain orthogonal and aligned with these principal
directions throughout the deformation process. Also, for (PS), the traction vector t applied
to the material surface with constant unit normal n = e2 remains a shear stress throughout
the deformation process (see Fig. 1)

t · n = 0. (20)

[14] conducted experiments on rubber to study the constitutive response for large defor-
mations. One of the experiments was denoted as pure shear. In this experiment the specimen
is subjected to plane strain deformation with a kinematic boundary condition that eliminates
deformation in the e3 direction and a kinetic boundary condition that enforces the T22 com-
ponent of Cauchy stress relative to ei to vanish. Neglecting complications near the clamped
boundary, the deformation caused by application of the T11 > 0 stress component can be
approximated as being homogeneous (see Fig. 3). However, as shown in (39), the associ-
ated stress field does not satisfy the condition (10)1 for being a state of pure shear stress.
Consequently, this deformation is referred to here as Plane Strain Extension and Contraction
(PS-EC) instead of pure shear.
An outline of this paper is as follows. Sections 2, 3 and 4 present the deformations and
results for (PS), (PS-TC) and (PS-EC) loadings, respectively, and conclusions are recorded
in Sect. 5. Furthermore, attention is limited to homogeneous deformations with tensors de-
pending on time only.
388 M.B. Rubin

2 Pure Shearing (PS)


Using the work in [15] it can be shown that the deformation for pure shearing (18) relative
to ei , can be characterized by (see Fig. 1)

F = a(e1 ⊗ e1 ) + b(e2 ⊗ e2 ) + c(e3 ⊗ e3 ) + aγ (e1 ⊗ e2 ) ,


1
B = ae2 (1 + γ 2 )(e1 ⊗ e1 ) + be2 (e2 e2 ⊗ e2 ) + (e3 ⊗ e3 )
ae2 be2
(21)
+ ae be γ (e1 ⊗ e2 + e2 ⊗ e1 ) ,
J 1/3
a = J 1/3 ae > 0 , b = J 1/3 be > 0 , c= > 0, γ ≥ 0,
ae be
where a, b and c are total stretches, γ controls shearing and ae and be are distortional
stretches. This deformation is the same as that considered in [6, 12]. Now, using (18) and
(21), the condition (17) yields the universal relation [12]

b2 b2
2
= e2 = 1 + γ 2 . (22)
a ae

It follows from this expression that the shearing angle φ in Fig. 1 is given by
aγ γ
tan φ = = ≤ 1. (23)
b 1+γ2

Moreover, it was observed in [6, 11, 12] that φ remains less than 45◦ .
Next, using (1), (21) and (22), it can be shown that

1
a 6 (1 + γ 2 )2 − 1
B − α1 I = e 4 (e1 ⊗ e1 + e2 ⊗ e2 − 2e3 ⊗ e3 )
3 3ae (1 + γ 2 )

+ ae2 γ 1 + γ 2 (e1 ⊗ e2 + e2 ⊗ e1 ) , (24a)
1
a 12 (1 + 2γ 2 )(1 + γ 2 )3 − 1
B2 − α2 I = e (e1 ⊗ e1 + e2 ⊗ e2 − 2e3 ⊗ e3 )
3 3ae8 (1 + γ 2 )2
+ 2ae4 γ (1 + γ 2 )3/2 (e1 ⊗ e2 + e2 ⊗ e1 ) . (24b)

Thus, with the help of (2) and (12), the state of pure shear stress (18) requires
∂
=0 (25a)
∂J
∂
a 6 (1 + γ 2 )2 − 1 ∂
ae12 (1 + 2γ 2 )(1 + γ 2 )3 − 1
e
+2 = 0, (25b)
∂α1 3ae4 (1 + γ 2 ) ∂α2 3ae8 (1 + γ 2 )2

and the shear stress τ is given by



∂ ∂ 2
τ = 2ρ +4 ae (1 + γ 2 ) ae2 γ 1 + γ 2 . (26)
∂α1 ∂α2
Although this distortional deformation satisfies the universal relation (22), it cannot sat-
isfy the condition (25b) exactly for an arbitrary strain energy . To see this, the coefficient
Pure Shearing and Pure Distortional Deformations Are Not Equivalent 389

Fig. 4 (PS): Comparison of the


stretches A1 and A2 for pure
shearing

of ∂/∂α1 in (25b) vanishes when ae = A1 and the coefficient of ∂/∂α2 in (25b) vanishes
when ae = A2 , with
1
A1 = , A2 = AA1 ,
(1 + γ 2 )1/3
(27)
1 + γ 2 1/12
0.9438 < A(∞) ≤ A(γ ) = ≤ A(0) = 1 .
1 + 2γ 2
Fig. 4 shows that these functions are not identical but are nearly the same for a large range of
γ , which indicates that either of the functions A1 or A2 for ae will produce a nearly universal
distortional deformation solution. Moreover, the total stretches a, b and c in (21) depend on
the dilatation J determined by the solution of (25a), which depends on the functional form
of the strain energy .
The deformation (PS) is called a pure shearing deformation since it has the property that
the angle between any two distinct material line elements in the shearing plane changes with
the amount of shearing throughout the shearing process. To prove this statement, consider
two distinct material line elements in the shearing plane characterized by the unit vectors

S1 = cos 1 e1 + sin 1 e2 , S2 = cos 2 e1 + sin 2 e2 ,


(28)
cos = S1 · S2 = ±1 , = 2 − 1 ,

where the angles 1 , 2 and define these two material line elements and the angle be-
tween them in the initial configuration. Using standard formulas, it follows that the angle θ
between these same two material line elements in the deformed configuration is determined
by the equations
C · S1 ⊗ S2
cos θ = √ √ ,
C · S1 ⊗ S1 C · S2 ⊗ S2
(29)
cos( 1 − 2 ) + γ sin( 1 + 2 ) + 2γ 2 sin( 1 ) sin( 2 )
cos θ = ,
1 + γ sin(2 1 ) + 2γ 2 sin2 ( 1 ) 1 + γ sin(2 2 ) + 2γ 2 sin2 ( 2 )

where C = FT F is the unimodular part of the right Cauchy-Green deformation tensor. In
particular, it can be seen that for any constant values of 1 and 2 associated with distinct
material line elements, the angle θ is a function of the shear γ , which proves the statement.
Also, for this deformation, the traction vector t remains a pure shear stress τ parallel to
the material surface with constant unit normal n = e2 throughout the deformation process
(see Fig. 1)
t = Te2 = τ e1 , t · n = 0. (30)
390 M.B. Rubin

3 Pure Shear Stress Caused by Tension and Compression (PS-TC)

The deformation tensors for pure shear (18) relative to pi , caused by tension and compres-
sion (PS-TC) can be expressed in the forms

F = a(p1 ⊗ p1 ) + b(p2 ⊗ p2 ) + c(p3 ⊗ p3 ) ,


 
 ae + be 1
F = (e1 ⊗ e1 + e2 ⊗ e2 ) + (e3 ⊗ e3 )
2 ae be
 
a e − be
+ (e1 ⊗ e2 + e2 ⊗ e1 ) ,
2
 4 2 
1 ae be + ae2 be4 − 2
B − α1 I = (e1 ⊗ e1 + e2 ⊗ e2 − 2e3 ⊗ e3 )
3 6ae2 be2
 2  (31)
ae − be2
+ (e1 ⊗ e2 + e2 ⊗ e1 ) ,
2
 8 4 
1 ae be + ae4 be8 − 2
B2 − α2 I = (e1 ⊗ e1 + e2 ⊗ e2 − 2e3 ⊗ e3 )
3 6ae4 be4
 4 
a − be4
+ e (e1 ⊗ e2 + e2 ⊗ e1 ) ,
2
J 1/3
a = J 1/3 ae > 0 , b = J 1/3 be > 0 , c= > 0,
ae be
where the total stretches a, b and c and distortional stretches ae and be are different from
those for (PS). From this form of F it is clear that material line elements parallel to pi remain
orthogonal throughout the deformation process. Furthermore, it follows that the state of pure
shear stress (18) requires (25a) and

∂ ae4 be2 + ae2 be4 − 2 ∂ ae8 be4 + ae4 be8 − 2


2 2
+2 = 0, (32)
∂α1 6ae be ∂α2 6ae4 be4

and the shear stress is given by



∂ ∂ 2 a 2 − b 2
τ = 2ρ +2 (ae + be2 ) e e
. (33)
∂α1 ∂α2 2
Similar to the case of pure shearing, this deformation for (PS-TC) cannot satisfy the
condition (32) exactly for an arbitrary strain energy . To see this, the coefficient of ∂/∂α1
in (32) vanishes when be = B1 and the coefficient of ∂/∂α2 in (32) vanishes when be = B2 ,
with
1
1  1/2
B1 = √ (−ae3 + ae6 + 8) , B2 = BB1 ,
ae 2
1/4 . (34)
− 2ae6 + 2 ae12 + 8
0.8408 < B(∞) ≤ B(ae ) = 1/2 ≤ B(1) = 1 .
− ae3 + ae6 + 8
Figure 5 shows that these functions are not identical but are nearly the same for a large
range of ae , which indicates that either of the functions B1 or B2 for be will produce a
Pure Shearing and Pure Distortional Deformations Are Not Equivalent 391

nearly universal distortional deformation solution. Again, the total stretches a, b and c in
(31) depend on the dilatation J determined by the solution of (25a), which depends on the
functional form of the strain energy .
Also, the unit tangent vector s and the unit normal vector n to the deformed top material
surface shown in Fig. 2 are given by
(a + b)e1 + (a − b)e2 −(a − b)e1 + (a + b)e2
s= , n= , (35)
2(a 2 + b2 ) 2(a 2 + b2 )
so the traction vector on this material surface takes the form
2ab b2 − a 2
t = Tn = τ s + τ n. (36)
a 2 + b2 a 2 + b2
This shows that this material surface is not loaded by a pure shear since the normal compo-
nent of the traction vector applied to this surface does not vanish.

4 Plane Strain Extension and Contraction (PS-EC)


For the plane strain extension and contraction deformation (PS-EC) associated with the pure
shear experiment in [14], the deformation is shown in Fig. 3 and is characterized by

F = a(e1 ⊗ e1 ) + b(e2 ⊗ e2 ) + (e3 ⊗ e3 ) ,


1
F = ae (e1 ⊗ e1 ) + be (e2 ⊗ e2 ) + (e3 ⊗ e3 ) , (37)
ae be
a = J 1/3 ae > 0 , b = J 1/3 be > 0 , J = ab = (ae be )3 > 0 ,

where the total stretches a and b, and distortional stretches ae and be are different from those
for (PS) and (PS-TC). Also, the condition that the component T22 of stress vanishes requires

∂ ∂
−ae4 be2 + 2ae2 be4 − 1 ∂
−ae8 be4 + 2ae4 be8 − 1
J +2 + 4 = 0, (38)
∂J ∂α1 3ae2 be2 ∂α2 3ae4 be4
which depends explicitly on the form of the strain energy function. Furthermore, when (38)
is satisfied, it follows from (12) that
∂
p = −ρJ = 0 . (39)
∂J
This means that the response is not in a state of pure shear stress (10).
For rubber materials, which are nearly incompressible, the response to any loading causes
nearly pure distortional deformation. Consequently, the response to (PS-EC) produces data
that supplements data for uniaxial stress loading, which helps determine the functional de-
pendence of the strain energy  on the distortional deformation invariants α1 and α2 .

5 Conclusions
The representation (3)2 has been identified as a pure shear representation of the deviatoric
stress T instead of a state of pure shear since it is valid for an arbitrary stress state. Also,
the conditions (10) are an invariant definition of a state of pure shear.
392 M.B. Rubin

Fig. 5 (PS-TC): Comparison of


the stretches B1 and B2 for pure
shear caused by tension and
compression

It has been shown that the pure shearing deformation (PS), defined in (21) and shown in
Fig. 1, associated with the state of pure shear stress (18) relative to ei , is a nearly universal
distortional deformation when ae equals either of the functions A1 or A2 in (27). Similarly,
the (PS-TC) deformation, defined in (31) and shown in Fig. 2, associated with the state of
pure shear stress (18) relative to pi , caused by tension and compression, produces a nearly
universal distortional deformation when be equals either of the functions B1 or B2 in (34).
However, for both of these deformations the dilatation J needs to be determined by the con-
dition (25a), which depends on the strain energy function . Consequently, pure shearing
and pure distortional deformations are not equivalent.
The deformation (PS) is called a pure shearing deformation since it has the property that
the angle between any two distinct material line elements in the shearing plane changes
with the amount of shearing throughout the shearing process. In contrast, for the (PS-TC)
deformation, the material line elements oriented in the fixed orthogonal directions pi remain
orthogonal throughout the deformation process. Similarly, for the (PS-EC) deformation, de-
fined in (37) and shown in Fig. 3, the material line elements in the directions ei remain
orthogonal. Also, for the (PS) deformation, the traction vector t acting on the material sur-
face with constant unit normal n = e2 remains a pure shear (30) throughout the deformation
process. In contrast, for (PS-TC), the deformation (31) causes the traction vector acting on
the material surface with unit normal n in (35) to develop a normal component in addition
to the shearing component (36).
In conclusion, the deformations (PS) in (21), (PS-TC) in (31) and (PS-EC) in (37) are
physically different and should not be expected to test the same features of a proposed strain
energy function for isotropic, compressible, hyperelastic materials.

Acknowledgements The author would like to thank the reviewers for suggesting additional relevant papers
and acknowledge helpful discussions with K Heiduschke.

Publisher’s Note Springer Nature remains neutral with regard to jurisdictional claims in published maps
and institutional affiliations.

References
1. Beatty, M.F.: A class of universal relations in isotropic elasticity theory. J. Elast. 17(2), 113–121 (1987)
2. Bělík, P., Fosdick, R.: The state of pure shear. J. Elast. 52(1), 91–98 (1998)
3. Boulanger, P., Hayes, M.: On finite shear. Arch. Ration. Mech. Anal. 151(2), 125–185 (2000)
4. Boulanger, P., Hayes, M.: On pure shear. J. Elast. 77(1), 83–89 (2004)
5. Cristescu, N.: Dynamic Plasticity. Elsevier, Amsterdam (2012)
6. Destrade, M., Murphy, J.G., Saccomandi, G.: Simple shear is not so simple. Int. J. Non-Linear Mech.
47(2), 210–214 (2012)
Pure Shearing and Pure Distortional Deformations Are Not Equivalent 393

7. Flory, P.J.: Thermodynamic relations for high elastic materials. Trans. Faraday Soc. 57, 829–838 (1961)
8. Gurtin, M.E.: The linear theory of elasticity. In: Linear Theories of Elasticity and Thermoelasticity,
pp. 1–295. Springer, Berlin (1973)
9. Hayes, M., Laffey, T.: Pure shear–a footnote. J. Elast. 92(1), 109–113 (2008)
10. Lode, W.: Versuche über den Einfluß der mittleren Hauptspannung auf das Fließen der Metalle Eisen,
Kupfer und Nickel. Z. Phys. 36, 913–939 (1926)
11. Mihai, L.A., Goriely, A.: Positive or negative Poynting effect? The role of adscititious inequalities in
hyperelastic materials. Proc. R. Soc. A, Math. Phys. Eng. Sci. 467(2136), 3633–3646 (2011)
12. Moon, H., Truesdell, C.: Interpretation of adscititious inequalities through the effects pure shear stress
produces upon an isotropie elastic solid. Arch. Ration. Mech. Anal. 55(1), 1–17 (1974)
13. Norris, A.N.: Pure shear axes and elastic strain energy. Q. J. Mech. Appl. Math. 59(4), 551–561 (2006)
14. Rivlin, R.S., Saunders, D.W.: Large elastic deformations of isotropic materials VII. Experiments on the
deformation of rubber. Philos. Trans. R. Soc. Lond. Ser. A, Math. Phys. Sci. 243(865), 251–288 (1951)
15. Rubin, M.B.: The significance of pure measures of distortion in nonlinear elasticity with reference to the
Poynting problem. J. Elast. 20(1), 53–64 (1988)
16. Rubin, M.B.: Simple, convenient isotropic failure surface. J. Eng. Mech. 117(2), 348–369 (1991)
17. Rubin, M.B.: Plasticity theory formulated in terms of physically based microstructural variables – part I.
Theory. Int. J. Solids Struct. 31, 2615–2634 (1994)
18. Sacommandi, G.: Universal results in finite elasticity. In: Yu, Y.B., Ogden, R.W. (eds.) Nonlinear Elas-
ticity: Theory and Applications, p. 283 (2001). Chap. 3
19. Thiel, C., Voss, J., Martin, R.J., Neff, P.: Shear, pure and simple. Int. J. Non-Linear Mech. 112, 57–72
(2019)
20. Ting, T.: Further study on pure shear. J. Elast. 83(1), 95–104 (2006)

You might also like