5 - PAX Short Course - Macro-Mechanics of Lamina

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 80

© 2003, P.

Joyce
Macromechanical Analysis of a
Lamina
Generalized Hooke’s Law

σ ij = Cijkl ε ij

Cijkl is a 9 x 9 matrix!

© 2003, P. Joyce
Hooke’s Law
¾Assume
¾linear elastic behavior
¾small deformations

σ = Eε
Uniaxial loading

© 2003, P. Joyce
Triaxial Stress State
δyx = εfor
Similarly xσ
dxthe deformationsUniaxial:
in the yσ- xzyand
y = δ xx + δ xy + δ¾Shears
xz are
z-directions
δ σ omitted σ σzy
σ
δ xy yy x σ
δ υdxzyydx
dx= −−ε xυε =dx= −dx
zdy υ dx dy
y x
δ yz δ x = ε x dx = 1dx − υ xy xz
xx
xy
E E
σz ε y E=δ [σ y − υE(¾Consider
σ x + σz )] E shear
E
yx
σ y σzx
σx ε =
σx
−υ −υ
σ y δzyσ σxz δδyxyy = −
yz dyxz dy
separately
ε yυε = = −dy υ dy
x E E
E E E ¾Superimpose
1 x
εz = [σz − υδ(σ xthree +
= −σ υε )] dz
y uniaxialσ=
σ
z − υ σ x dz
y
1σy Ezx δδzxzz =
= −ευεz dz y dz= = −dz υ dz
= δ[σ x −δυzz(σ y + σz )] stresses E
δ zy
εx x
E
E
z
σz δ xzE xx

© 2003, P. Joyce
Triaxial Stress State
Strain-Stress Relations

1
ε x = [σ x − υ(σ y + σz )]
E
1
ε y = [σ y − υ(σ x + σz )]
E
1
εz = [σz − υ(σ x + σ y )]
E

© 2003, P. Joyce
Triaxial Stress State
Stress-Strain Relations

E
σx =
(1 + υ)(1 − 2υ)
[(1 − υ)ε x + υ(ε y + εz )]

E
σy =
(1 + υ)(1 − 2υ)
[(1 − υ)ε y + υ(ε x + εz )]

E
σz =
(1 + υ)(1 − 2υ)
[(1 − υ)εz + υ(ε x + ε y )]

© 2003, P. Joyce
Similarly for shear stresses

© 2003, P. Joyce
Shear Stress Strain Relationships
¾Shear stresses are independent of each other
and all axial stresses
¾Shear strains are independent of each other
and all axial strains
¾Each obeys a simple linear elastic model

τ = Gγ E
G=
G – Shear Modulus 2(1 + υ)

© 2003, P. Joyce
Shear Stress-Strain Relationships

E
τxy = Gγ xy = γ xy
2(1 + υ)
E
τxz = Gγ xz = γ xz
2(1 + υ)
E
τ yz = Gγ yz = γ yz
2(1 + υ)

© 2003, P. Joyce
Generalized Hooke’s Law
Compliance matrix

⎡ 1 υ υ ⎤
⎢ E − − 0 0 0⎥
E E
⎢ υ υ ⎥
ε
⎡ x ⎤ ⎢− 1
− 0 0 0 ⎥ ⎡σ x ⎤
⎢ε ⎥ ⎢ E E E ⎥ ⎢σ y ⎥
⎢ y⎥ ⎢ υ υ 1 ⎥ ⎢ ⎥
⎢ ε z ⎥ ⎢− − 0 0 0 ⎢σ ⎥
⎥ z
⎢ ⎥=⎢ E E E
⎥ ⎢τ ⎥
⎢γ yz ⎥ ⎢ 0 0 0
1
0 0 ⎥ ⎢ yz ⎥
⎢γ zx ⎥ ⎢ G ⎥ ⎢τ zx ⎥
⎢ ⎥ ⎢ 1 ⎥ ⎢ ⎥
⎢⎣γ xy ⎥⎦ ⎢ 0 0 0 0
G
0 ⎢τ xy ⎥
⎥⎣ ⎦
⎢ 1⎥
⎢ 0 0 0 0 0 ⎥
⎣ G⎦
Strain-stress relations for an isotropic material/matrix form
© 2003, P. Joyce
Generalized Hooke’s Law
Stiffness matrix

⎡ E (1 − υ ) υE υE ⎤
⎢ 0 0
0⎥
⎡σ x ⎤ ⎢ (1 − 2ν )(1 + v) (1 − 2ν )(1 + v) (1 − 2ν )(1 + v)
⎥⎡ ε x ⎤
⎢σ ⎥ ⎢ υE E (1 − υ ) υE ⎥ ⎢ε ⎥
⎢ y ⎥ ⎢ (1 − 2ν )(1 + v) (1 − 2ν )(1 + v) (1 − 2ν )(1 + v)
0 0 0 ⎢ y⎥

⎢σ z ⎥ ⎢ ⎥⎢ ε z ⎥
⎢ ⎥=⎢ υE υE E (1 − υ )
0 0 0 ⎥ ⎢γ ⎥
⎢τ yz ⎥ ⎢ (1 − 2ν )(1 + v) (1 − 2ν )(1 + v) (1 − 2ν )(1 + v) ⎥⎢ ⎥
yz
⎢τ zx ⎥ ⎢ 0 0 0 G 0 0 ⎥ ⎢γ zx ⎥
⎢ ⎥ ⎢ ⎥ ⎢ ⎥
⎢⎣τ xy ⎥⎦ ⎢ 0 0 0 0 G 0 ⎥ ⎢⎣γ xy ⎥⎦
⎢⎣ 0 0 0 0 0 G ⎥⎦

Stress-strain relations for an isotropic material /matrix form


© 2003, P. Joyce
Stress-Strain Relations in
Composite Materials
Material No. of independent
elastic constants
General anisotropic material 81

Anisotropic material (considering 36


symmetry of stress and strain tensors)
Anisotropic material with energy 21
considerations
General orthotropic material 9

Transversely orthotropic material 5

Isotropic material 2
© 2003, P. Joyce
Stress-Strain Relations in
Composite Materials
¾Orthotropic material ~ has three mutually
perpendicular plans of material symmetry
¾Specially orthotropic material ~ when the
reference system of coordinates is selected
along principal planes of material symmetry
¾Transversely isotropic material ~ one of its
principal planes is a plane of isotropy
(properties are the same in all directions.)

© 2003, P. Joyce
Contracted Notation
¾Thanks to symmetry of the stress and strain
tensors, the compliance matrix reduces to a
6 x 6 matrix, introduce a contracted notation.
σ11 = σ1,σ 22 = σ 2 ,σ 33 = σ 3 ,
σ 23 = σ 4 ,σ 31 = σ 5 ,σ12 = σ 6
ε11 = ε1, ε 22 = ε 2 , ε33 = ε1,
2ε 23 = ε 4 ,2ε31 = ε5 ,2ε12 = ε 6
C1111 = C11, C1122 = C12, C1133 = C13, C1123 = 2C14, C1131 = 2C15, C1112 = 2C16
C2211 = C21, C2222 = C22, C2233 = C23, C2223 = 2C24, C2231 = 2C25, C2212 = 2C26
© 2003, P. Joyce
Stress-Strain Relations
for Thin UD Lamina
Assumed to be under a state of plane stress
⎡σ1 ⎤ ⎡Q11 Q12 0 ⎤ ⎡ε1 ⎤
⎢σ ⎥ = ⎢Q Q 0 ⎥ ⎢ε ⎥
⎢ 2 ⎥ ⎢ 12 22 ⎥⎢ 2 ⎥
⎢⎣τ 6 ⎥⎦ ⎢⎣ 0 0 Q66 ⎥⎦ ⎢⎣γ 6 ⎥⎦

Fully characterized by 4 independent constants,


Qij ~ reduced stiffnesses

© 2003, P. Joyce
Reduced Stiffness Matrix?
¾If the stiffness matrix is the inverse of the
compliance matrix, what is the reduced
stiffness matrix?
C i 3C j 3
Qij = Cij − (i, j = 1, 2, 6)
C33

© 2003, P. Joyce
Relations Between Mathematical
and Engineering Constants

E1
Q11 =
1 − υ12υ 21
E2
Q22 =
1 − υ12υ 21
υ 21 E1 υ12 E2
Q12 = =
1 − υ12υ 21 1 − υ12υ 21
Q66 = G12

© 2003, P. Joyce
You Said Four Independent
Constants?
¾ From symmetry of the compliance matrix
υ ij υ ji
=
Ei Ej

¾ The above can also be deduced from Betti’s


reciprocal law according to which transverse
deformation due to a stress applied in the
longitudinal direction is equal to the longitudinal
deformation due to an equal stress applied in the
transverse direction.
© 2003, P. Joyce
Stress-Strain Relations
¾Also expressed in terms of compliances

⎡ ε1 ⎤ ⎡ S11 S12 0 ⎤ ⎡σ 1 ⎤
⎢ε ⎥ = ⎢ S S 22 0 ⎥⎥ ⎢⎢σ 2 ⎥⎥
⎢ 2 ⎥ ⎢ 12
⎢⎣γ 6 ⎥⎦ ⎢⎣ 0 0 S 66 ⎥⎦ ⎢⎣ τ 6 ⎥⎦

© 2003, P. Joyce
Relations between Mathematical
and Engineering Constants
1
S11 =
E1
1
S 22 =
E2
υ12 υ 21
S12 = − =−
E1 E2
1
S 66 =
G12

That’s Better!

© 2003, P. Joyce
Sample Calculation
¾ For a graphite/epoxy UD
laminate, find the following:
¾ 1.) Compliance matrix
¾ 2.) Minor Poisson’s ratio τ12
¾ 3.) Reduced Stiffness Matrix
¾ 4.) Strains in the 1-2 coordinate
system if the applied stresses are σ1

σ1 = 2 MPa, σ2 = -3MPa, τ12 = 4 MPa

© 2003, P. Joyce σ2
Sample Data
Glass/ Boron Graphite
Property Symbol Units
epoxy /epoxy /epoxy
Fiber
volume Vf - 0.45 0.50 0.70
fraction
Long.
elastic E1 GPa 38.6 204 181
modulus
Trans.
elastic E2 GPa 8.27 18.50 10.30
modulus
Major
Poisson’s ν12 - 0.26 0.23 0.28
ratio
Shear
G12 GPa 4.14 5.59 7.17
Modulus © 2003, P. Joyce
Sample Calculation
The compliance matrix elements are
calculated as follows:
1 1 −11
S11 = = 9
= 0. 5525(10 )
E1 181(10 )
ν 12 0.28
S12 = = 9
= −0.1547(10 −11 )
E1 181(10 )
1 1 −10
S 22 = = 9
= 0 .9709(10 )
E1 10.3(10 )
1 1 −9
S 66 = = 9
= 0. 1395(10 )
G12 7.17(10 )

(all terms have units of Pa-1)


© 2003, P. Joyce
Sample Calculation
From Betti’s reciprocal law:

ν 21 ν 12
=
E2 E1
(0.28)
ν 21 = (10. 3)(10 9
) = 0.01593
181(109 )

© 2003, P. Joyce
Sample Calculation
The stiffness matrix elements are calculated
as follows:
E1 181(109 )
Q11 = = = 181.8(109 )
1 −ν 12ν 21 1 − (0.28)(0.01593)
ν12 E2 ν 21E1 (0.28)(10.3)(109 )
Q12 = = = = 2.897(109 )
1 −ν 12ν 21 1 −ν 12ν 21 1 − (0.28)(0.01593)
E2 10.3(109 )
Q22 = = = 10.35(109 )
1 −ν 12ν 21 1 − (0.28)(0.01593)
Q66 = G12 = 7.17(109 )

(all terms have units of Pa)


© 2003, P. Joyce
Sample Calculation
The stiffness matrix can also be calculated
by inverting the compliance matrix of Part 1:
−1
⎡ 0.5525 (10 −11 ) − 0.1547 (10 −11 ) 0 ⎤
[Q ] = ⎢⎢− 0.1547 (10 −11 ) 0.9709 (10 −10 ) 0


⎢ 0 0 −9 ⎥
0.1395 (10 ) ⎦

⎡181 .8(10 9 ) 2.897 (10 9 ) 0 ⎤
[Q ] = ⎢⎢ 2.897 (10 9 ) 10 .35(10 9 ) 0

⎥ Pa
⎢ 0 0 7 . 17 (10 9 ⎥
)⎦

(all terms have units of Pa)


© 2003, P. Joyce
Sample Calculation
The strains in the 1-2 coordinate system are
calculated as follows:
⎡ ε 1 ⎤ ⎡ S11 S12 ) 0 ⎤ ⎡σ 1 ⎤
⎢ ε ⎥ = ⎢S S 22 0 ⎥⎥ ⎢⎢σ 2 ⎥⎥
⎢ 2 ⎥ ⎢ 12
⎢⎣γ 12 ⎥⎦ ⎢⎣ 0 0 S 66 ⎥⎦ ⎢⎣τ 12 ⎥⎦
⎡ ε 1 ⎤ ⎡ 0.5525 (10 ) ⎤ ⎡ 2(10 6 ) ⎤
−11
− 0.1547 (10 −11 ) 0
⎢ ε ⎥ = ⎢ − 0.1547 (10 −11 ) 0.9709 (10 −10 ) 0
⎥⎢
− 3(10 6 ⎥
)⎥
⎢ 2⎥ ⎢ ⎥⎢
⎢⎣γ 12 ⎥⎦ ⎢⎣ 0 0 0.1395 (10 − 9 ) ⎥⎦ ⎢⎣ 4(10 6 ) ⎥⎦
⎡ ε 1 ⎤ ⎡ 15 .69 ⎤
⎢ ε ⎥ = ⎢ − 294 .4 ⎥ (10 -6 )
⎢ 2⎥ ⎢ ⎥
⎢⎣γ 12 ⎥⎦ ⎢⎣ 557 .9 ⎥⎦

(10-6 is microstrain)
© 2003, P. Joyce
Stress-Strain Relations
for Thin Angle Lamina
Generally, a laminate does not consist only of UD laminae because
of their stiffness and strength properties in the transverse direction.

Hence, in most laminates, some laminae are placed at an angle.


2 y 1

Global and material axes of an angle lamina.


© 2003, P. Joyce
Stress-Strain Relations
for Thin Angle Lamina
¾ The axes in the x-y coordinate system are called the global
axes of the off-axes.
¾ The axes in the 1-2 coordinate system are called the
material axes or the local axes, where direction 1 is parallel
to the fibers (also called the longitudinal direction) and
direction 2 is is perpendicular to the fibers (also called the
transverse direction.)
¾ The angle between the two axes is denoted by the angle θ.

© 2003, P. Joyce
Stress-Strain Relations
for Thin Angle Lamina
¾ The stress-strain relationship in the 1-2 coordinate system has already
been established.
¾ From Mechanics of Materials, the stresses in the global and material
axes are related to each other through the angle of the lamina, θ.

⎡σ x ⎤ ⎡σ 1 ⎤
⎢ ⎥ −1 ⎢ ⎥
σ
⎢ y⎥ = [T (θ ) ] σ
⎢ 2⎥
⎢τ xy ⎥ ⎢⎣τ 12 ⎥⎦
⎣ ⎦
Where [T(θ)] is called the transformation matrix and is defined as

⎡ c2 s2 2 sc ⎤ ⎡c 2 s2 − 2 sc ⎤
[T (θ ) ] = ⎢⎢ s 2 c2
⎥ ⎢
− 2 sc ⎥ thus [T (θ ) ] = ⎢ s 2
−1
c2

2 sc ⎥ = [T(-θ ) ]
⎢ − sc sc c 2 − s 2 ⎥⎦ ⎢ sc − sc c 2 − s 2 ⎥⎦
⎣ ⎣
© 2003, P. Joyce
Stress-Strain Relations
for Thin Angle Lamina
¾Using the stress-strain equation in the
material axes together with the
transformation equation we obtain:

⎡σ x ⎤ ⎡ ε1 ⎤
⎢ ⎥ ⎢ε ⎥
σ
⎢ y⎥ = [T ]−1
[Q ]⎢ 2⎥
⎢τ xy ⎥ ⎢⎣γ 12 ⎥⎦
⎣ ⎦

© 2003, P. Joyce
Stress-Strain Relations
for Thin Angle Lamina
¾ Similarly, the strains in the global and material coordinate
axes are related through the transformation matrix
⎡ ε1 ⎤ ⎡ εx ⎤
⎢ ε ⎥ = [T ]⎢ ε ⎥
⎢ 2 ⎥ ⎢ y ⎥
⎢⎣ 12 γ 12 ⎥⎦ ⎢ 12 γ xy ⎥
⎣ ⎦
which can be rewritten as
⎡ ε1 ⎤ ⎡ εx ⎤
⎢ ε ⎥ = [R ][T ][R ]−1 ⎢ ε ⎥
⎢ 2 ⎥ ⎢ y ⎥
⎢⎣ 12 γ 12 ⎥⎦ ⎢ 12 γ xy ⎥
⎣ ⎦
where [R] is the Reuter matrix and is defined as
⎡1 0 0 ⎤
[R ] = ⎢⎢0 1 0⎥⎥
⎢⎣0 0 2⎥⎦
© 2003, P. Joyce
Stress-Strain Relations
for Thin Angle Lamina
¾ Multiplying out the first five matrices on the RHS of the previous
equation we obtain the transformed reduced stiffness matrix, [Qxy ]
¾ Thus, [σ ]x , y = [Q ]x , y [ε ]x , y
¾ Summarizing,
Qxx = c 4Q11 + s 4Q22 + 2c 2 s 2Q12 + 4c 2 s 2Q66
Q yy = s 4Q11 + c 4Q22 + 2c 2 s 2Q12 + 4c 2 s 2Q66
Qxy = c 2 s 2Q11 + c 2 s 2Q22 + (c 4 + s 4 )Q12 − 4c 2 s 2Q66
Qxs = c 3 sQ11 − cs 3Q22 + (cs 3 − c 3 s )Q12 + 2(cs 3 − c 3 s )Q66
Q ys = cs 3Q11 − c 3 sQ22 + (c 3 s − cs 3 )Q12 + 2(c 3 s − cs 3 )Q66
Qss = c 2 s 2Q11 + c 2 s 2Q22 − 2c 2 s 2Q12 + (c 2 − s 2 ) 2 Q66

¾ The subscript s corresponds to shear stress or strain components in the


x-y system, i.e., τ s = τ xy and γ s = γ xy
© 2003, P. Joyce
Strain-Stress Relations for Thin
Angle Lamina
¾ Similarly, the transformed compliance can be obtained:
S xx = c 4 S11 + s 4 S 22 + 2c 2 s 2 S12 + 4c 2 s 2 S 66
S yy = s 4 S11 + c 4 S 22 + 2c 2 s 2 S12 + 4c 2 s 2 S 66
S xy = c 2 s 2 S11 + c 2 s 2 S 22 + (c 4 + s 4 ) S12 − 4c 2 s 2 S 66
S xs = 2c 3 sS11 − 2cs 3 S 22 + 2(cs 3 − c 3 s ) S12 + (cs 3 − c 3 s ) S 66
S ys = 2cs 3 S11 − 2c 3 sS 22 + 2(c 3 s − cs 3 ) S12 + (c 3 s − cs 3 ) S 66
S ss = 4c 2 s 2 S11 + 4c 2 s 2 S 22 − 8c 2 s 2 S12 + (c 2 − s 2 ) 2 S 66

¾ Thus, [ε ]x , y = [S ]x , y [σ ]x , y

© 2003, P. Joyce
Strain-Stress Relations for Thin
Angle Lamina
¾ How about in terms of Engineering Constants?
¾ If we imagine a series of simple experiments on an element
with sides parallel to the x- and y-axes, we obtain:
⎡ 1 ν yx η sx ⎤
⎢ − ⎥
⎡ε x ⎤ ⎢ x
E Ey Gxy ⎥
⎡σ x ⎤
⎢ε ⎥ = ⎢− ν xy 1 η sy ⎥ ⎢ ⎥
⎢ y⎥ ⎢ E ⎥ ⎢σ y ⎥
Ey Gxy ⎥
⎢⎣γ s ⎥⎦ ⎢ x
⎢⎣ τ s ⎥⎦
⎢ η xs η ys 1 ⎥
⎢ ⎥
⎣ Ex Ey Gxy ⎦

© 2003, P. Joyce
Strain-Stress Relations for Thin
Angle Lamina
¾What is η ? Æ shear coupling coefficient
¾ ηxs , the first subscript denotes normal
loading in the x-direction; the second
subscript denotes shear strain.
η xs = γ s ε x
η ys = γ s ε y
η sx = ε x γ s
η sy = ε y γ s

© 2003, P. Joyce
Strain-Stress Relations for Thin
Angle Lamina
¾ Comparison of equivalent strain-stress relations yields the
following relationships:
1 1
S xx = Ex =
Ex S xx
1 1
S yy = Ey =
Ey S yy
1 1
S ss = G xy =
G xy S ss
ν xy ν yx S yx S xy
S xy = S yx = − =− ν xy = − ; ν yx = −
Ex Ey S xx S yy
η xs η sx S sx S xs
S xs = S sx = = η xs = ; η sx =
Ex G xy S xx S ss
η ys η sy S sy S ys
S ys = S sy = = η ys = ; η yx =
Ey G xy S yy S ss

© 2003, P. Joyce
Sample Calculation
¾ Find the following for a 60° angle lamina of
graphite/epoxy.
¾ Transformed compliance matrix
¾ Transformed reduced stiffness matrix
¾ Global strains
¾ Local strains
¾ If the applied stresses are σx = 2 MPa, σy = -3MPa, τxy = 4 MPa

© 2003, P. Joyce
Sample Calculation
¾From the previous example:

S11 = 0.5525(10 −11 )


S12 = −0.1547(10 −11 )
S 22 = 0.9709(10 −10 )
S 66 = 0.1395(10 −9 )

(all terms have units of Pa-1)

© 2003, P. Joyce
Sample Calculation
¾ The transformed compliance matrix elements are calculated as
follows:
S xx = c 4 S11 + s 4 S 22 + 2c 2 s 2 S12 + 4c 2 s 2 S 66 = 0.8053(10 −10 )
S yy = s 4 S11 + c 4 S 22 + 2c 2 s 2 S12 + 4c 2 s 2 S 66 = −0.7878(10 −11 )
S xy = c 2 s 2 S11 + c 2 s 2 S 22 + (c 4 + s 4 ) S12 − 4c 2 s 2 S 66 = −0.3234(10 −10 )
S xs = 2c 3 sS11 − 2cs 3 S 22 + 2(cs 3 − c 3 s ) S12 + (cs 3 − c 3 s ) S 66 = 0.3475(10 −10 )
S ys = 2cs 3 S11 − 2c 3 sS 22 + 2(c 3 s − cs 3 ) S12 + (c 3 s − cs 3 ) S 66 = −0.4696(10 −10 )
S ss = 4c 2 s 2 S11 + 4c 2 s 2 S 22 − 8c 2 s 2 S12 + (c 2 − s 2 ) 2 S 66 = 0.1141(10 −9 )
(All terms are in Pa-1)

© 2003, P. Joyce
Sample Calculation
¾ Next, invert the transformed compliance matrix [S] to
obtain the transformed reduced stiffness matrix [Q].
−1
⎡ 0.8053(10 −10 ) − 0.7878(10 −10 ) − 0.3234(10 −10 )⎤
[Q] = [S ]−1 = ⎢⎢− 0.7878(10−10 ) 0.3475(10−10 ) − 0.4696(10−10 )⎥⎥
⎢− 0.3234(10 −10 ) − 0.4696(10 −10 ) 0.1141(10 −9 ) ⎥
⎣ ⎦
⎡0.2365 0.3246 0.2005⎤
[Q] = ⎢⎢0.3246 1.094 0.5419⎥⎥(1011 )
⎢⎣0.2005 0.5419 0.3674⎥⎦

(all terms in Pa)

© 2003, P. Joyce
Sample Calculation
¾The global strains in the x-y plane are given
by [ε ]x, y = [S ]x, y [σ ]x, y
⎡ ε x ⎤ ⎡ 0.8053(10 −10 ) − 0.7878(10 −10 ) − 0.3234(10 −10 )⎤ ⎡ 2 ⎤
⎢ ⎥ ⎢ −10 ⎥ ⎢ ⎥ (106 )
ε
⎢ ⎥ ⎢
y = − 0 . 7878(10 −10
) 0. 3475(10 −10
) − 0 .4696 (10 ) −
⎥⎢ ⎥ 3
⎢γ xy ⎥ ⎢− 0.3234(10 ) − 0.4696(10 ) 0.1141(10 ) ⎥ ⎢⎣ 4 ⎥⎦
−10 −10 − 9
⎣ ⎦ ⎣ ⎦
⎡ ε x ⎤ ⎡ 0.5534(10 − 4 ) ⎤
⎢ ⎥ ⎢ −3 ⎥
ε
⎢ y⎥ ⎢ = − 0 .3078 (10 )⎥
⎢γ xy ⎥ ⎢ 0.5328(10 −3 ) ⎥
⎣ ⎦ ⎣ ⎦

© 2003, P. Joyce
Sample Calculation
¾ The local strains in the lamina can be calculated using the
Transformation equation.

⎡ ε1 ⎤ ⎡ εx ⎤
⎢ ε ⎥ = [T ]⎢ ε ⎥
⎢ 2 ⎥ ⎢ y ⎥
⎢⎣ 12 γ 12 ⎥⎦ ⎢ 12 γ xy ⎥
⎣ ⎦
⎡ ε1 ⎤ ⎡ cos 2 60 sin 2 60 2 cos 60 sin 60 ⎤ ⎡ 0.5534(10 − 4 ) ⎤
⎢ ε ⎥ = ⎢ sin 2 60 cos 2
60 − 2 cos 60 sin 60
⎥⎢
− 0 .3078(10 −3 ⎥
)⎥
⎢ 2 ⎥ ⎢ ⎥⎢
⎢⎣ 12 γ 12 ⎥⎦ ⎢⎣− cos 60 sin 60 cos 60 sin 60 cos 2 60 − sin 2 60⎥⎦ ⎢⎣0.5328(10 −3 ) / 2⎥⎦

⎡ ε 1 ⎤ ⎡ 0.1367(10 ) ⎤
−4

⎢ ε ⎥ = ⎢− 0.2662(10 −3 )⎥
⎢ 2⎥ ⎢ ⎥
⎢⎣γ 12 ⎥⎦ ⎢⎣− 0.5809(10 )⎥⎦
− 3

© 2003, P. Joyce
Transformation of
Engineering Constants
¾ Flow chart for determination of transformed
elastic constants of UD lamina.

[S]1,2 [S]x,y

[E]1,2 θ [E]x,y

[Q]1,2 [Q]x,y

© 2003, P. Joyce
Macromechanical Strength
Parameters
¾From a macromechanical POV, the strength
of a lamina is an anisotropic property.
¾It is desirable, for example, to correlate the
strength along an arbitrary direction to some
basic strength parameters (analogous to
micromechanic definitions before.)

© 2003, P. Joyce
Strength Failure Theories of an
Angle Lamina
¾ Various theories have been developed for studying the failure of an
angle lamina.
¾ Generally based on the normal and shear strengths of a UD lamina.
¾ Need to consider tension and compression
¾ UD lamina has 2 material axes, 1-direction parallel to the fibers and
2-direction which is perpendicular to the fibers.
¾ Hence there are 4 normal strength parameters for UD lamina.
¾ Tensile strength in fiber direction
¾ Transverse tensile strength
¾ Compressive strength in fiber direction
¾ Transverse compressive strength
¾ The fifth strength parameter is the shear strength

© 2003, P. Joyce
Strength Failure Theories of an
Angle Lamina

¾ Unlike the stiffness parameters, these strength parameters


cannot be transformed directly for an angle lamina.
¾ Hence, the failure theories are based on first finding the
stresses in the material axes and then using these five
strength parameters of a UD lamina to find whether the
lamina has failed.

© 2003, P. Joyce
Macromechanical Strength
Parameters
¾ Also predict transverse compressive strength and in-plane
shear strength using micromechanics. . .
¾ Failure mechanisms vary greatly with material properties
and type of loading.
¾ Even when predictions are accurate with regard to failure
initiation at critical points, they are only approximate as far
as global failure of the lamina is concerned.
¾ Furthermore, the possible interaction of failure
mechanisms makes it difficult to obtain reliable strength
predictions under a general type of loading.
¾ A macromechanical or phenomological approach to failure
analysis may be preferable.
© 2003, P. Joyce
Macromechanical Strength
Parameters
¾ This characterization recognizes the fact that most
composite materials have different strengths in tension and
compression.
¾ By convention the sign of the shear stress is immaterial, as
long as the shear strength is referred to the principal
material directions.
¾ Exception, refers to the case when the shear stress is
applied at an angle wrt the principal material directions.
¾ Since most composites have different tensile and
compressive strengths and they are weakest in transverse
tension, it follows that in this case the lamina would be
stronger under positive shear.
© 2003, P. Joyce
Macromechanical Strength
Parameters σ y = τ6

τ6

σx = -τ6
Positive shear stress

Shear stress acting along principal material axes.


© 2003, P. Joyce
Macromechanical Strength
Parameters σ y = τ6

τ6

σx = -τ6
Negative shear stress

Shear stress acting along principal material axes.


© 2003, P. Joyce
Macromechanical Strength
σ = τ6
Parameters 1

τs

σ2 = -τs
Positive shear stress

Shear stress acting at 45° wrt principal material axes.


© 2003, P. Joyce
Macromechanical Strength
Parameters σ = -τs
1

τs

σ2 = τs
Negative shear stress

Shear stress acting at 45° wrt principal material axes.


© 2003, P. Joyce
Macromechanical Failure
Theories
¾ Given a state of stress, the principal stresses and their
directions are obtained by stress transformation
(independent of material properties.)
¾ The principal strains and their directions are obtained by
using the appropriate anisotropic stress-strain relations and
strain transformation.
¾ In general, the principal stress, principal strain, and
material symmetry directions do not coincide.
¾ Since strength varies with orientation, maximum stress
alone is not the critical factor in failure.

© 2003, P. Joyce
Macromechanical Failure
Theories
¾ An anisotropic failure theory is needed.
¾ Failure criteria for homogeneous isotropic materials, such as
¾ Maximum normal stress (Rankine),
¾ Maximum shear stress (Tresca),
¾ Maximum distortional energy (von Mises),
¾ and so forth are well established.
¾ More than 40 anisotropic theories have been proposed – look at the
four most widely used.

© 2003, P. Joyce
Maximum Stress Failure Theory
¾ Related to the Maximum Normal stress theory by Rankine and the
Maximum Shear stress theory by Tresca.
¾ The stresses acting on a lamina are resolved into the normal and shear
stresses in the material axes.
¾ Failure is predicted in a lamina, if any of the normal or shear stresses
in the material axes are equal to or greater than the corresponding
ultimate strengths of a UD lamina.

( )
− σ1C ult
( )
< σ1 < σ1T ult
( )
, − σ 2C ult
( )
< σ 2 < σ 2T ult
, − (τ12 )ult < τ12 < (τ12 )ult

¾ Each component of stress is compared with the corresponding strength


and hence does not have an interaction with the others.

© 2003, P. Joyce
Sample Calculation
¾ Find the off-axis shear strength of a 60° graphite/epoxy
lamina using the Maximum Stress failure criteria.
¾ Assume the following stress state
σ x = 0, σ y = 0, τ xy = τ ,
¾ Find the stresses along the principal material axes, using
the Transformation Equation.
⎡ σ 1 ⎤ ⎡ 0.2500 0.7500 0.8660 ⎤ ⎡0 ⎤
⎢σ ⎥ = ⎢ 0.7500 0.2500 − 0.8660 ⎥⎥ ⎢⎢0 ⎥⎥
⎢ 2⎥ ⎢
⎢⎣τ 12 ⎥⎦ ⎢⎣ − 0.4330 0.4330 − 0.5000 ⎥⎦ ⎢⎣τ ⎥⎦
⎡ σ 1 ⎤ ⎡ 0.866τ ⎤
⎢σ ⎥ = ⎢ − 0.866τ ⎥
⎢ 2⎥ ⎢ ⎥
⎢⎣τ 12 ⎥⎦ ⎢⎣ − 0.500τ ⎥⎦

© 2003, P. Joyce
Sample Calculation
¾ Applying the Maximum Stress Failure Criteria together with strength
data for graphite/epoxy composites from the Data Sheet,we have
− 1500 < 0.866τ < 1500
− 246 < −0.866τ < 40
− 68 < 0.500τ < 68
or
− 1732 < τ < 1732
− 46.19 < τ < 284.1
− 136.0 < τ < 136.0

© 2003, P. Joyce
Sample Calculation
¾ The off-axis shear strength of a lamina is defined as the minimum of the
positive and negative shear stress which can be applied to an angle lamina
before failure.
¾ Calculations show that τxy = 46.19 MPa is the largest magnitude of shear
stress one can apply to the 60° graphite/epoxy composite.
¾ However, the largest positive shear stress one could apply is 136.0 MPa,
and the largest negative shear stress one could apply is –46.19 MPa.
¾ This shows that the maximum magnitude of allowable shear stress in other
than the material axes direction depends on the sign of the shear stress.
¾ This is because the tensile strength perpendicular to the fiber direction is
much lower than the compressive strength perpendicular to the fiber
direction.

© 2003, P. Joyce
Failure Envelopes
¾ A failure envelope is a 3D plot of the combinations of normal
and shear stresses which can be applied to an angle lamina
before failure.
¾ Drawing 3D graphs is time consuming. . .
¾ One may develop failure envelopes for constant shear stress,
τxy, and then use the 2 normal stresses σx and σy as the 2 axes.
¾ If the applied stress is within the failure envelope, the lamina is
safe; otherwise it has failed.

© 2003, P. Joyce
Failure Envelopes
¾ For a UD lamina at a given ¾ For a 60° lamina at a given
shear stress loading, the failure shear stress loading, the failure
envelope takes the form of a envelope takes the form of a
rectangle as shown. rectangle as shown.
σ2
τxy
σ2 σ1 σ2

( )
− σ 2T ult
( )
− σ 2T ult

σ1 ( )
− σ 1C ult
( )
− σ 1T ult
σ1
( )
− σ 1C ult
( )
− σ 1T ult

( )
− σ 2C ult
( )
− σ 2C ult

© 2003, P. Joyce
Maximum Strain Failure Theory
¾ Based on the Maximum Normal Strain Theory by St. Venant and the Maximum
Shear Stress Theory by Tresca.
¾ The strains applied to a lamina are resolved into the normal and shear stresses in
the material axes.
¾ Failure is predicted in a lamina, if any of the normal or shear strains in the
material axes are equal to or greater than the corresponding ultimate strains of a
UD lamina.
( )
− ε1C ult
< ε1 <( )
ε1
T
ult
, −( )
ε C
2 ult
< ε 2 <( )
ε T
2 ult
, − (γ 12 )ult < γ 12 < (γ 12 )ult
¾ The ultimate strains can be found directly from the ultimate strength parameters
and the elastic moduli, assuming the stress-strain response is linear until failure.
¾ Each component of strain is compared with the corresponding ultimate strain
and hence does not have an interaction with the others.
¾ Yields different results from Maximum Stress Failure Theory, because the local
strains in a lamina include the Poisson’s ratio effect (allows some interaction of
stress components.)
© 2003, P. Joyce
Maximum Strain Failure Theory
¾ Assume a general biaxial state of stress σ2
on an angle lamina.
σ1

¾ Obtain the stress components along the principal material axes by


stress transformation.
⎡σ x ⎤ ⎡σ 1 ⎤
⎢ ⎥ −1 ⎢ ⎥
σ
⎢ y⎥ = [T (θ ) ] σ
⎢ 2⎥
⎢τ xy ⎥ ⎢⎣τ 12 ⎥⎦
⎣ ⎦

© 2003, P. Joyce
Maximum Strain Failure Theory
¾ Then the corresponding strain components can be
calculated by means of the lamina stress-strain relations:
σ1 σ2
ε1 = − ν 21
E1 E2
σ2 σ1
ε2 = − ν 12
E2 E1
τ6
γ6 =
G12

¾ Next calculate the ultimate strains as follows

(ε ) = (σE ) , (ε ) (σ ) , (ε ) (σ ) , (ε ) (σ ) , (γ ) (τ 12 )ult
T C T C
T
1 ult
1 ult C
1 ult = 1 ult T
2 ult = 2 ult C
2 ult = 2 ult
12 ult =
1 E
1 E 2 E
2 G12

© 2003, P. Joyce
Maximum Strain Failure Theory
¾Failure subcriteria restated in terms of the
stresses:
⎧ (σ ) when ε > 0T
σ −ν σ = ⎨ 1 ult 1

⎩− (σ ) when ε < 0
1 12 2 C
1 ult 1

⎧ (σ ) when ε > 0 T
σ −ν σ = ⎨ 2 ult 2

⎩− (σ ) when ε < 0
2 21 1 C
2 ult 2

− (τ 12 )ult < τ 12 < (τ 12 )ult

© 2003, P. Joyce
Maximum Strain Failure Theory
¾ For a 2D state of stress with τ6 = 0, the failure envelope takes
the form of a parallelogram with its center off the origin.
σy

σ2 σx

( )
− σ 2T σ 2 − ν 21σ 1 = (σ 1T )
ult

( )
− σ 1C ult
( )
− σ 1T ult
σ1

σ 2 − ν 21σ 1 = −(σ 1C )
−σ( )C

σ 1 − ν 12σ 2 = (σ 1T )
2 ult

σ 1 − ν 12σ 2 = −(σ C
1 )

© 2003, P. Joyce
Tsai-Hill Failure Theory
¾ Based on the deviatoric or distortional energy failure
theory of von Mises.
¾ Adapted to anisotropic materials by Hill.
¾ Then adapted to a UD lamina by Tsai.
⎡ σ 1 ⎤ ⎡ σ 1σ 2 ⎤ ⎡ σ 2 ⎤ ⎡ τ 12 ⎤
2 2 2

⎢ T ⎥ −⎢ T 2 ⎥+⎢ T ⎥ +⎢ ⎥ <1
( )
σ ( ) σ( ) (τ )
⎣ 1 ult ⎦ ⎢⎣ σ 1 ult ⎥⎦ ⎣ 2 ult ⎦ ⎣ 12 ult ⎦
¾ Given the global stresses in a lamina, one can find the
local stresses in a lamina and apply the above failure
theory to determine whether or not the lamina has
failed.

© 2003, P. Joyce
Tsai-Hill Failure Theory
¾ The failure envelope described by the Tsai-Hill criterion is a closed
surface in the σ1, σ2, τ12 space.
¾ Failure envelopes for constant values of k = τ 12 (τ 12 )ult
σ 12 σ 22 σ 1σ 2
have the form 2
+ 2 − 2 = 1− k 2
F1 F2 F1
¾ Where: ( )
⎧ σ 1T ult when σ 1 > 0
F1 = ⎨ C
( )
⎩ σ 1 ult when σ 1 < 0 (Modified Tsai-Hill Criterion)
( )
⎧ σ 2T ult when σ 2 > 0
F2 = ⎨ C
( )
⎩ σ 2 ult when σ 2 < 0
¾ Graphically represents 4 different elliptical arcs joined at the σ1, σ2 axes.

© 2003, P. Joyce
Tsai-Hill Failure Theory

¾ Considers the interaction between the 3 UD lamina


strength parameters, unlike the Maximum Stress and
Maximum Strain Theories.
¾ Tsai-Hill Failure Theory is a Unified Theory and hence
does not give the mode of failure like the Maximum Stress
and Maximum Strain Theories.

© 2003, P. Joyce
Tsai-Wu Failure Theory
¾ Based on a general failure theory for anisotropic materials
first proposed by Gol’denblat and Kopnov (1965).
¾ Capable of predicting strength under general states of
stress for which no experimental data are available.
¾ Uses the concept of strength tensors.
¾ Has the form of an invariant formed from stress and strain
tensor components
¾ Has the capability to account for the difference between
tensile and compressive strengths

© 2003, P. Joyce
Tsai-Wu Failure Theory
¾ Tsai and Wu (1971) proposed a modified tensor polynomial
theory by assuming the existence of a failure surface in the
stress space of the form —

H1σ 1 + H 2σ 2 + H 6τ 12 + H11σ 1 + H 22σ 2 + H 66τ 12 + 2 H12σ 1σ 2 < 1


2 2 2

¾ The coefficients are obtained by applying elementary loading


conditions to the lamina. Thus —
1 1 1 1
H1 = − H2 = − H6 = 0
(σ )
T
1 ult (σ )
C
1 ult
(σ )
T
2 ult (σ )
C
2 ult
1
1 1 H 66 =
H11 =
(σ ) (σ )
T C
H 22 =
(σ ) (σ )
T
2 ult
C
2 ult
(τ 12 )ult 2
1 ult 1 ult

© 2003, P. Joyce
Tsai-Wu Failure Theory
¾ The remaining coefficient H12 must be obtained by
some type of biaxial testing.
¾ Direct biaxial testing is not easy or practical to
perform.
¾ An easier test producing a biaxial state of stress is
the off-axis tensile test.
¾ For θ = 45° we can measure the off-axis tensile
strength, σ.
¾ Then, H12 = 2 − (H1 + H 2 ) − 1 (H11 + H 22 + H 66 )
σ σ 2
¾ Again produces an elliptical failure envelope.
© 2003, P. Joyce
Comparison with Experimental Results

© 2003, P. Joyce
Comparison with Experimental Results

© 2003, P. Joyce
Comparison with Experimental Results

© 2003, P. Joyce
Comparison with Experimental Results

© 2003, P. Joyce
Comparison with Experimental Results

Observations —
¾ The difference between the Maximum Stress and
Maximum Strain Failure Theories and the experimental
results is quite pronounced.
¾ Tsai-Hill and Tsai-Wu Failure Theories are in good
agreement with experimentally obtained results.
¾ The cusps observed in the Maximum Stress and
Maximum Strain Failure Theories correspond to the
change in failure mode.
¾ The variation of the strength as a function of angle is
smooth in the Tsai-Hill and Tsai-Wu Failure Theories.

© 2003, P. Joyce
Macromechanical Failure Theories
Theory Physical Basis Operational Req’d experimental
Convenience characterization
Maximum Stress Tensile behavior of brittle Inconvenient Few parameters
material By simple testing
No stress interaction
Maximum Strain Tensile behavior of brittle Inconvenient Few parameters by
material simple testing
Some stress interaction
Deviatoric strain Ductile behavior of Can be programmed Biaxial testing is
energy anisotropic materials needed in addition to
uniaxial testing
(Tsai-Hill) “Curve fitting”for Different functions
heterogeneous brittle required for tensile and
composites compressive strengths
Interactive tensor Mathematically General and Numerous parameters
polynomial consistent comprehensive;
operationally simple
Tsai-Wu Reliable “curve fitting” Comprehensive
experimental program
needed.
© 2003, P. Joyce
References
¾ Engineering Mechanics of Composite Materials, Daniel, I.M. and Ishai, O., 1994.
¾ Mechanics of Composite Materials, Kaw, A.K., 1997.
¾ Introduction to Composite Materials, Tsai, S. W. and Hahn, H. T., 1980.
¾ “Application of Advanced Composites in Mechanical Engineering Designs,” Zweben, C.,
Proceedings of the 31st International SAMPE Technical Conference, 1999.

© 2003, P. Joyce

You might also like