Vous êtes sur la page 1sur 38

MECH3361/9361 Semester 2, 2016

STRESS
1.1. Definition of stress
1.1.1. External forces on a body
Consider a body of continuous (no voids) and cohesive (no cracks or defects) material
subjected to an arbitrary number of externally applied loads as shown in Fig. 1.1a. It is
supposed that the body is in equilibrium.

F4 Free Body Diagram


n
F3 F
F5 F 
 n F
 t t
Cross section: A A

F2 F2

F1 F1

(a) (b)

Fig. 1.1 External and internal forces in a structural member

If we cut this body, the applied forces can be thought of as being distributed over the cut area
A as in Fig. 1.1b. In general, stress varies throughout the body so values depend on the
location of interest. For each patch A, we define the resultant force over this area as F. F
is also a distributed force; however, when A is extremely small, we can say that F is
effectively uniform.

Definition: Stress is the intensity of the internal force


on a specific plane passing through a point.

Zooming back out to look at the entire sectioned area, we can thus say that area A is subject
to an infinite number of forces, where each one (of magnitude F) acts over a small area of
size A. From here, we can define stress as the magnitude of internal force F divided by the
acting area A.

If we let A approach zero, we obtain the stress at a point. Mathematically, stress at a point
can be expressed as:

1
MECH3361/9361 Semester 2, 2016

𝑛 𝛥𝐹
𝜎 = 𝑇 = lim
𝛥𝐴→0 𝛥𝐴

(1.1)

Stress is one of most important concepts that we introduced in mechanics of solids. Why?
Because the design of structures is largely dependent on stress level for safety reasons.

1.1.2. Normal and shear stress


As we know, force is a vector that has both magnitude and direction, but the definition of
stress so far only considers the magnitude of the resultant. Looking at patch A as an
example, we can see that force F is not necessarily perpendicular to the sectional area. If we
only take the magnitude of the force into account, the stress may not reflect the true
mechanical status at this point. Therefore, we need refined concepts that consider both the
magnitude and the direction of the internal force.

We can resolve F in the normal (Fn) and tangential (Ft) directions of the acting area as
per Fig. 1.1b. The intensity of the force (or force per unit area) acting normal to section A is
called normal stress, nn, and it is expressed as:

𝛥𝐹𝑛
𝜎𝑛𝑛 = lim
𝛥𝐴→0 𝛥𝐴
(1.2)

The intensity or force per unit area acting tangentially to A is called shear stress, nt, and it
is expressed as:

𝛥𝐹𝑡
𝜎𝑛𝑡 = lim
𝛥𝐴→0 𝛥𝐴
(1.3)

The SI unit for stress is the Pascal (Pa), which is equivalent to N/m2. In engineering practice,
KPa = 103 Pa, MPa = 106 Pa, and GPa = 109 Pa are commonly used.

1.2. Stress notation


The notation described above is not sufficiently robust for use in general, because:

1. the direction of surface A can change; and


2. there are an infinite number of tangential directions on any given surface.

The normal stress nn can vary with the direction change of n, and shear stress nt can be in
any tangential direction of the surface. We need to be more specific.

2
MECH3361/9361 Semester 2, 2016

1.2.1. Cartesian coordinate system


Recall that we often solve engineering problems under a reference coordinate system,
typically a Cartesian (xyz) coordinate system as shown in Fig. 1.2 below.

For analytical purposes, it would be convenient to discuss stress at a point of interest P on an


infinitesimal plane through P with its external normal n in one of the directions along a
reference coordinate. As an example, we can consider an infinitesimal sectional plane
through P where the normal vector n is coincident with the z direction. The traction can be
resolved along the three Cartesian axes as follows:

𝑛 𝛥𝐹 𝑑𝐹 𝑑𝐹𝑥 + 𝑑𝐹𝑦 + 𝑑𝐹𝑧 𝑑𝐹𝑥 𝑑𝐹𝑦 𝑑𝐹𝑧


𝑇 = lim ( ) =( ) = ( ) =( + + )
𝛥𝐴→0 𝛥𝐴 𝑧 𝑑𝐴 𝑧 𝑑𝐴 𝑧
𝑑𝐴 𝑑𝐴 𝑑𝐴 𝑧
= (𝜎𝑥 i + 𝜎𝑦 j + 𝜎𝑧 k)𝑧

Resolution direction (i.e. co-ord. direction)


𝑛
𝑇 = 𝜎𝑧𝑥 i + 𝜎𝑧𝑦 j + 𝜎𝑧𝑧 k

(1.4)
z-sectional plane

The first suffix of a stress component indicates the direction of the sectional plane (z, here).
The second denotes the direction along which the stress component is aligned.

1.2.2. Positive/negative planes


If the normal to the plane points in the same direction as the positive coordinate, the plane is
positive (Fig. 1.2b). Otherwise, it is negative (Fig. 1.2c).

Similarly, there are infinitesimal planes through P in the other coordinate directions. There
are six planes in total: positive/negative x; positive/negative y; positive/negative z.

F4
z
F3
zz
F5 z
n
+ z
-
Positive
Z-plane
P z-plane
P P
zy Negative
zx y y
z-plane
y
-n
o o o
x x x
F1
Normal in the same Normal in the opposite
F2 direction of coordinate (z) direction to coordinate (z)

Fig. 1.2 Stress in Cartesian coordinate

3
MECH3361/9361 Semester 2, 2016

1.2.3. The infinitesimal cube


For an intuitive visualisation, we often use an infinitesimal cube formed by the six
infinitesimal planes mentioned above, as shown in Fig. 1.3. On each plane, we have one
normal stress component and two shear stress components.

z zz
zy
zx yz

xz yy
xy
yx
xx
y
o
x

Fig. 1.3 Stress in infinitesimal cube

From above, we know that the stress state at point P consists of nine stress components. We
can arrange these stress components into a matrix, with the following form:

Direction of stress component


(co-ordinate direction)
x y z
Plane normal to x 𝜎𝑥𝑥 𝜎𝑥𝑦 𝜎𝑥𝑧
Plane normal to y 𝜎𝑦𝑥 𝜎𝑦𝑦 𝜎𝑦𝑧
Plane normal to z 𝜎𝑧𝑥 𝜎𝑧𝑦 𝜎𝑧𝑧

In engineering, we call this the stress tensor:

𝜎𝑥𝑥 𝜎𝑥𝑦 𝜎𝑥𝑧


𝜎 = [𝜎𝑦𝑥 𝜎𝑦𝑦 𝜎𝑦𝑧 ].
𝜎𝑧𝑥 𝜎𝑧𝑦 𝜎𝑧𝑧

Example 1.1
The stress states at two different points in a machine component are measured as:

16 18 0 19 20 0
𝐴 = [18 17 −15] 𝐵 = [20 −25 0]
0 −15 19 0 0 20

Draw these stress states on infinitesimal cubes.

Solution

To clarify the stress components, you can first compare the corresponding notation:

4
MECH3361/9361 Semester 2, 2016

16 18 0 𝜎𝑥𝑥 𝜎𝑥𝑦 𝜎𝑥𝑧


𝐴 = [18 17 −15] = [𝜎𝑦𝑥 𝜎𝑦𝑦 𝜎𝑦𝑧 ]
0 −15 19 𝜎𝑧𝑥 𝜎𝑧𝑦 𝜎𝑧𝑧

For point A, 𝜎𝑥𝑥 = 16, 𝜎𝑥𝑦 = 18, 𝜎𝑥𝑧 = 0 arbitrary units, which are the three stress
components on the x-plane in x, y, and z directions. Thus, on the x-plane, draw an arrow of
magnitude 16 in the positive x-direction, 18 in the positive y-direction, and nothing in z-
direction, as shown in the front face of the left infinitesimal element. Similarly, y and z plane
stresses can be drawn as shown below.

19 z-section 20 z-section
z z
15

18 18 17 20 25
15 20
x-section 16 x-section 19
y-section y-section
y y
o o
x x

1.3. Positive/negative stresses


It is necessary to define positive and negative sense of stresses for convenience.

1.3.1. Positive direction of normal stress


Consider normal stress nn on an infinitesimal plane An, whose external normal is n, We
define nn as positive if its direction is in the normal direction of section, as in Fig. 1.4 (left).
This stress “pulls” on the area and stretches it, so it is known as a tensile stress.

Conversely, if 𝜎𝑛𝑛 acts in the opposite direction (“pushing” into the surface), it is considered
a negative stress (also known as compressive stress) (Fig. 1.4, right).

Tensile = Positive Compressive= Negative


z z
+
y y
o o
n n
x x
xx positive xx negative

Fig. 1.4 Sign of normal stress (left – the same direction as the normal direction of section
plane; right – opposite to the normal direction of section plane)

5
MECH3361/9361 Semester 2, 2016

1.3.2. Positive direction of shear stress


For a shear stress nt in the infinitesimal plane An, where t is a tangential direction of An,
the directions depend on the polarity of the plane. In a positive plane (e.g. Fig. 1.5a, where
the external normal n of An has the same direction as a coordinate axis x), the positive nt
should have the same direction as coordinate axis t (see bold arrow in Fig. 1.5a). In a negative
plane, the positive nt should have the opposite direction to coordinate axis t (Fig. 1.5d).

z z
(a)
+ (b)

y y
o xy positive o xy negative
x n x n
(Positive-section) (Positive-section)
z z
(c) n (d) + n
negative positive
xy
xy

y y
o o
x (Negative-section)
x (Negative-section)

Fig. 1.5 Sign conventions for shear stress

1.4. Symmetry of the stress matrix


Are all nine stress components independent, or can we use a smaller number to describe a
stress state?

Let’s check the infinitesimal element shown before. If we cut the infinitesimal element
through the middle, i.e. a z-section as shown by the dashed line below, we get the sectional
model on the right hand side.

yy
z zz
zy yx
zx yz
C B
D C xy
xz yy O
xx y
xy
A yx xy
B x
xx D A
y
y yx

x yy
o
x

Fig. 1.6 Equilibrium of element

For this 2D element, the sum of moments can be written as:

6
MECH3361/9361 Semester 2, 2016

1 1 1 1
𝜎𝑥𝑦 (𝛥𝑦)(𝛥𝑧) [( 𝛥𝑥)] + 𝜎𝑥𝑦 (𝛥𝑦)(𝛥𝑧) [( 𝛥𝑥)] − 𝜎𝑦𝑥 (𝛥𝑥)(𝛥𝑧) [( 𝛥𝑦)] − 𝜎𝑦𝑥 (𝛥𝑥)(𝛥𝑧) [( 𝛥𝑦)] = 0
2 2 2 2

∴ 𝜎𝑥𝑦 − 𝜎𝑦𝑥 = 0

Thus,

𝜎𝑥𝑦 = 𝜎𝑦𝑥

Similarly, by checking equilibrium conditions in the yz- and xz planes, we find that:

𝜎𝑥𝑧 = 𝜎𝑧𝑥
𝜎𝑦𝑧 = 𝜎𝑧𝑦

Thus we have shown that the stress tensor is symmetric, i.e.

𝜎𝑖𝑗 = 𝜎𝑗𝑖 , (where 𝑖, 𝑗 = 𝑥, 𝑦, 𝑧)

(1.5)

Hence, only six independent stress components are needed to describe the stress state at a
point.

𝜎𝑥𝑥 𝜎𝑥𝑦 𝜎𝑥𝑧 𝜎𝑥𝑥 ⬚ ⬚


𝜎
[ 𝑦𝑥 𝜎𝑦𝑦 𝜎𝑦𝑧 ] ቎𝜎𝑦𝑥 𝜎𝑦𝑦 ⬚቏
𝜎𝑧𝑥 𝜎𝑧𝑦 𝜎𝑧𝑧 𝜎𝑧𝑥 𝜎𝑧𝑦 𝜎𝑧𝑧

Example 1.2
10 0 −40
(i) Draw an infinitesimal cube to show the stress tensor, 𝜎𝐴 = [ 0 −30 0 ]
−40 0 10

Solution

10
+z-plane
40

30

z 40
10 +y-plane

o y +x-plane
x

(ii) Write the stress tensor from the stressed infinitesimal cube (note the signs of the shear
stress are not given in the figure and you need to decide them):

7
MECH3361/9361 Semester 2, 2016

60 60
+x-plane
40 +z-plane 40

60 60
20 20 40 20 20 40
z z
y 60 y 60
- y-plane
o x o x

Solution

Look at the planes in the infinitesimal cube. Obviously, the front y-section is a negative plane
(its normal direction is opposite to y-positive). The other two faces shown are positive planes.

In +x-plane: 𝜎𝑥𝑥 = −60 (comp.) , 𝜎𝑥𝑦 = −20 (opp. y), 𝜎𝑥𝑧 = −40 (opp. z)

In –y-plane: 𝜎𝑦𝑥 = −20 (same as x), 𝜎𝑦𝑦 = 60 (tension), 𝜎𝑦𝑧 = 0

In +z-plane: 𝜎𝑧𝑥 = −40 (opp. x), 𝜎𝑧𝑦 = 0, 𝜎𝑧𝑧 = 60 (tension)

So the stress tensor can be written as:

−60 −20 −40


𝜎𝐵 = [−20 60 0 ]
−40 0 60

No units are specified in this example, but units should be included outside of the matrix.

Example 1.3
(i) Complete the infinitesimal cube to show the stress tensor [σA].

10 20 −40
𝜎𝐴 = [ 20 −30 0 ]
−40 0 10

z O x

Solution

Step 1: Note that the co-ordinate system is different from that in Example 1.2.

8
MECH3361/9361 Semester 2, 2016

Determine the convention of section planes (all + in the visible planes).

Step 2:

+y-plane 30
20
20
40
10
40
y 10 +x-plane

z o x +z-plane

Step 3:

In +x-plane: 𝜎𝑥𝑥 = 10, 𝜎𝑥𝑦 = 20, 𝜎𝑧𝑥 = −40

In +y-plane: 𝜎𝑦𝑥 = 20, 𝜎𝑦𝑦 = −30, 𝜎𝑦𝑧 = 0

In +z-plane: 𝜎𝑧𝑥 = −40, 𝜎𝑧𝑦 = 0, 𝜎𝑧𝑧 = 10

(ii) Write the stress tensor from the stressed infinitesimal cube (note the signs of the shear
stress are not given in the figure and you need to determine them).

60
40

y 60
x o 20 20 40

z 60

Solution

Step 1: Determine the convention of section planes. Look at the visible planes of the
infinitesimal cube. Obviously, the visible x-plane is a negative plane (its normal direction is
opposite to x-positive). Similarly, other two faces shown are also negative sections as shown.

Step 2:

In –x-plane: 𝜎𝑥𝑥 = −60 (comp.), 𝜎𝑥𝑦 = 20 (opp. +y), 𝜎𝑥𝑧 = −40 (same as +z)

In –y-plane: 𝜎𝑦𝑥 = 20 (opp. +x), 𝜎𝑦𝑦 = 60 (tension), 𝜎𝑦𝑧 = 0

In –z-plane: 𝜎𝑧𝑥 = −40 (same +x), 𝜎𝑧𝑦 = 0, 𝜎𝑧𝑧 = 60 (tension)

Step 3: Write the stress tensor.

9
MECH3361/9361 Semester 2, 2016

−60 20 −40
𝜎𝐵 = [ 20 60 0 ]
−40 0 60

60
– x-plane
– z-plane 40

y 60
x o 20 20 40

z 60
– y-plane

1.5. Stress transformation (2D)


1.5.1. Stresses in any direction
Cut a 2D infinitesimal element diagonally, leaving the left and bottom sides and a hypotenuse
inclined at an angle  from the vertical. Two of its surfaces have their normals in the negative
x and y directions; the third has a normal angled  from the x axis, as in Fig. 1.6 (right).

Applying the equilibrium equations about the normal n and tangential t axes, we get:

∑𝐹𝑛 = 0 = 𝐴𝜎𝑛𝑛 − ⏟
(𝜎𝑦𝑦 )(𝐴 sin 𝜃) sin 𝜃
𝐹𝑦

(𝜎𝑥𝑥 )(𝐴 cos 𝜃) cos 𝜃 − (𝜏


−⏟ ⏟ 𝑦𝑥 sin 𝜃)(𝐴) cos 𝜃 − (𝜏
⏟ 𝑥𝑦 cos 𝜃)(𝐴) sin 𝜃
𝐹𝑥 𝑉𝑥 𝑉𝑦

yy n
t
xy tn
nn
Acos

xx xx xx A

xy Asin
y y
yx
yx

x yy x yy

Fig. 1.6 Stress in a different direction

Since 𝜎𝑥𝑦 = 𝜎𝑦𝑥 , the above equation can be simplified to:

𝜎𝑛𝑛 = 𝜎𝑥𝑥 cos2 𝜃 + 𝜎𝑦𝑦 sin2 𝜃 + 2𝜎𝑥𝑦 cos 𝜃 sin 𝜃

(1.6a)

Using the following trigonometric relations:

10
MECH3361/9361 Semester 2, 2016

1
cos2 𝜃 = (1 + cos 2𝜃)
2
1
sin2 𝜃 = (1 − cos 2𝜃)
2
sin 2𝜃 = 2 cos 𝜃 sin 𝜃

we can obtain the following:

(𝜎𝑥𝑥 + 𝜎𝑦𝑦 ) (𝜎𝑥𝑥 − 𝜎𝑦𝑦 )


𝜎𝑛𝑛 = + cos 2𝜃 + 𝜎𝑥𝑦 sin 2𝜃
2 2
(1.6b)

In a similar way, by applying equilibrium in the tangential (t) axis and using the
trigonometric functions, we can get:

∑𝐹𝑡 = 0 = 𝐴𝜎𝑛𝑡 − ⏟
(𝜎𝑦𝑦 )(𝐴 sin 𝜃) cos 𝜃
𝐹𝑦

(𝜎𝑥𝑥 )(𝐴 cos 𝜃) sin 𝜃 − (𝜏


−⏟ ⏟ 𝑦𝑥 sin 𝜃)(𝐴) sin 𝜃 − (𝜏
⏟ 𝑥𝑦 cos 𝜃)(𝐴) cos 𝜃
𝐹𝑥 𝑉𝑥 𝑉𝑦

1 1
𝜎𝑛𝑡 = [2𝜎𝑦𝑦 sin 𝜃 cos 𝜃] − [2𝜎𝑥𝑥 sin 𝜃 cos 𝜃] − 𝜎𝑦𝑥 sin2 𝜃 + 𝜎𝑥𝑦 cos 2 𝜃
2 2
(1.7a)

Using trigonometry, this simplifies to:

1 1
cos 2 𝜃 − sin2 𝜃 = (1 + cos 2𝜃) − (1 − cos 2𝜃) = cos 2𝜃
2 2

(𝜎𝑦𝑦 − 𝜎𝑥𝑥 )
𝜎𝑛𝑡 = sin 2𝜃 + 𝜎𝑥𝑦 cos 2𝜃
2
(1.7b)

With these two equations above, we can determine the stress in any rotated plane.

1.5.2. Stresses with coordinate rotation


Consider a 2D stress state undergoing coordinate rotation, from xoy to x’oy’ (Fig. 1.7).

11
MECH3361/9361 Semester 2, 2016

y
y’ y’y’ yy
x’y’
x’x’ x’
xy

xx
+90° x
o

Fig. 1.7 Stresses with coordinate rotation

𝜎𝑥′𝑥′ can be determined from the previous section where the coordinate x’ coincides with
normal n. Thus the stress in an inclined plane of  can be calculated by:

𝜎𝑥′𝑥′ = 𝜎𝑥𝑥 cos2 𝜃 + 𝜎𝑦𝑦 sin2 𝜃 + 2𝜎𝑥𝑦 cos 𝜃 sin 𝜃

(1.8)

𝜎𝑦′𝑦′ can be determined by viewing the inclined plane with angle of (+90):

𝜎𝑦 ′ 𝑦 ′ = 𝜎𝑥𝑥 cos2 (𝜃 + 90°) + 𝜎𝑦𝑦 sin2 (𝜃 + 90°) + 2𝜎𝑥𝑦 cos(𝜃 + 90°) sin(𝜃 + 90°)
= 𝜎𝑥𝑥 sin2 𝜃 + 𝜎𝑦𝑦 cos2 𝜃 + 2𝜎𝑥𝑦 (− sin 𝜃) cos 𝜃
= 𝜎𝑥𝑥 sin2 𝜃 + 𝜎𝑦𝑦 cos2 𝜃 − 2𝜎𝑥𝑦 sin 𝜃 cos 𝜃

Thus,

𝜎𝑦′𝑦′ = 𝜎𝑥𝑥 sin2 𝜃 + 𝜎𝑦𝑦 cos 2 𝜃 − 2𝜎𝑥𝑦 sin 𝜃 cos 𝜃

(1.9)

Similarly,

(𝜎𝑦𝑦 − 𝜎𝑥𝑥 )
𝜎𝑥′𝑦′ = sin 2𝜃 + 𝜎𝑥𝑦 cos 2𝜃
2
(1.10)

If we add (1.8) to (1.9),

𝜎𝑥 ′ 𝑥 ′ + 𝜎𝑦 ′ 𝑦 ′ = 𝜎𝑥𝑥 (sin2 𝜃 + cos2 𝜃) + 𝜎𝑦𝑦 (sin2 𝜃 + cos 2 𝜃) + 2𝜎𝑥𝑦 sin 𝜃 cos 𝜃


− 2𝜎𝑥𝑦 sin 𝜃 cos 𝜃
= 𝜎𝑥𝑥 + 𝜎𝑦𝑦

which means that the summation of two normal stress components is independent of the
orientation of the coordinate system. We will show this again in 3D (that 𝜎𝑥𝑥 + 𝜎𝑦𝑦 + 𝜎𝑧𝑧 =
constant).

12
MECH3361/9361 Semester 2, 2016

Example 1.4
1 −1 0
Rotate the following stress tensor about z-axis for =90 : 𝜎 = [−1
o
2 0] MPa.
0 0 3

Solution

Since the rotation is about z-axis and 𝜎𝑥𝑧 = 𝜎𝑦𝑧 = 0, we do not need to change z-directional
stress:

𝜎𝑥 ′ 𝑥 ′ = 𝜎𝑥𝑥 cos2 𝜃 + 𝜎𝑦𝑦 sin2 𝜃 + 2𝜎𝑥𝑦 cos 𝜃 sin 𝜃


= (1) cos2 90° + (2) sin2 90° + 2(−1) cos 90° sin 90°
= (1)(0) + (2)(1) + 2(−1)(0)(1)
= 2 MPa

𝜎𝑦 ′ 𝑦 ′ = 𝜎𝑥𝑥 sin2 𝜃 + 𝜎𝑦𝑦 cos2 𝜃 − 2𝜎𝑥𝑦 cos 𝜃 sin 𝜃


= (1) sin2 90° + (2) cos2 90° − 2(−1) cos 90° sin 90°
= (1)(1) + (2)(0) + 2(1)(0)(1)
= 1 MPa

(𝜎𝑦𝑦 − 𝜎𝑥𝑥 )
𝜎𝑥 ′ 𝑦 ′ = sin 2𝜃 + 𝜎𝑥𝑦 cos 2𝜃
2
(2 − 1)
= sin(2 × 90°) + (−1) cos(2 × 90°)
2
1
= (0) + (−1)(−1)
2
= 1 MPa

𝜎𝑧′𝑧′ = 𝜎𝑧𝑧 = 3 MPa, since the rotation is purely about the z-axis.

2 1 0
Thus, 𝜎 ′ = [1 1 0] MPa
0 0 3

1.6. Principal stresses (2D)


To find where the maximum normal stresses occur, we solve for stationary points from Eq.
1.6b.

𝜕𝜎𝑛𝑛 𝜕 (𝜎𝑥𝑥 + 𝜎𝑦𝑦 ) (𝜎𝑥𝑥 − 𝜎𝑦𝑦 )


= [ + cos 2𝜃 + 𝜎𝑥𝑦 sin 2𝜃] = 0
𝜕𝜃 𝜕𝜃 2 2
−(𝜎𝑥𝑥 − 𝜎𝑦𝑦 ) sin 2𝜃 + 2𝜎𝑥𝑦 cos 2𝜃 = 0

13
MECH3361/9361 Semester 2, 2016

2𝜎𝑥𝑦
tan 2𝜃𝑝 =
𝜎𝑥𝑥 − 𝜎𝑦𝑦

(1.11)

We call the maximum and minimum 𝜎𝑛𝑛 the principal stresses. From Eq. 1.10, 2𝜎𝑥′𝑦′ =
−(𝜎𝑥𝑥 − 𝜎𝑦𝑦 ) sin 2𝜃 + 2𝜎𝑥𝑦 cos 2𝜃, we can obtain 2𝜎𝑥′𝑦′ = (𝜕𝜎𝑛𝑛 /𝜕𝜃) = 0, meaning that
when 𝜎𝑛𝑛 reaches its extrema (principal stresses) on the plane, 𝜎𝑛𝑡 = 0. In other words, a
zero shear plane is a principal plane.

( xx yy ) 2+ 2
xy
2

xy
2 p+180° 2 p
( xx yy )
2

Fig. 1.8 2D principal stress directions

For a 2D case, there are two roots to Eq. 1.11: p1 and p2. Mathematically, 2p1 and 2p2 are
180 apart, and thus p1 and p2 are 90 apart (or orthogonal), i.e.

2𝜎𝑥𝑦
tan 2𝜃𝑝 = tan(2𝜃𝑝 + 180°) =
𝜎𝑥𝑥 − 𝜎𝑦𝑦

Solving for p1,

𝜎𝑥𝑦
sin 2𝜃𝑝1 =
2
(𝜎 − 𝜎𝑦𝑦 )
√[ 𝑥𝑥 2
] + 𝜎𝑥𝑦
2
(𝜎𝑥𝑥 − 𝜎𝑦𝑦 )
cos 2𝜃𝑝1 = 2
2
(𝜎 − 𝜎𝑦𝑦 )
√[ 𝑥𝑥 2
] + 𝜎𝑥𝑦
2

(1.12)

For p2 (=p1+90),

14
MECH3361/9361 Semester 2, 2016

−𝜎𝑥𝑦
sin 2𝜃𝑝2 =
2
(𝜎 − 𝜎𝑦𝑦 )
√[ 𝑥𝑥 2
] + 𝜎𝑥𝑦
2
(𝜎𝑥𝑥 − 𝜎𝑦𝑦 )

cos 2𝜃𝑝2 = 2
2
(𝜎 − 𝜎𝑦𝑦 )
√[ 𝑥𝑥 2
] + 𝜎𝑥𝑦
2

(1.13)

Substituting the above two trigonometric relations into Eq. 1.6b, we can find the associated
principal stresses:

(𝜎𝑥𝑥 + 𝜎𝑦𝑦 ) (𝜎𝑥𝑥 − 𝜎𝑦𝑦 )


𝜎𝑛𝑛 = + cos 2𝜃 + 𝜎𝑥𝑦 sin 2𝜃
2 2

(𝜎𝑥𝑥 − 𝜎𝑦𝑦 )
2
(𝜎𝑥𝑥 + 𝜎𝑦𝑦 ) (𝜎𝑥𝑥 − 𝜎𝑦𝑦 ) 2 𝜎𝑥𝑦
𝜎𝑛𝑛 = 𝜎𝑝 = ± ±
2 2 2 2
(𝜎 − 𝜎𝑦𝑦 ) (𝜎𝑥𝑥 − 𝜎𝑦𝑦 )
√[ 𝑥𝑥 2
] + 𝜎𝑥𝑦 √[ 2
] + 𝜎𝑥𝑦
2 2
2
(𝜎𝑥𝑥 − 𝜎𝑦𝑦 ) 2
[ ] + 𝜎𝑥𝑦
(𝜎𝑥𝑥 + 𝜎𝑦𝑦 ) 2
= ±
2 2
(𝜎 − 𝜎𝑦𝑦 )
√[ 𝑥𝑥 2
] + 𝜎𝑥𝑦
2

2
(𝜎𝑥𝑥 + 𝜎𝑦𝑦 ) (𝜎𝑥𝑥 − 𝜎𝑦𝑦 )
𝜎𝑝 = ± √[ 2 =𝜎
] + 𝜎𝑥𝑦 1,2
2 2

(1.14)

1.7. Maximum shear stresses (2D)


To find the maximum shear stress, mathematically, we can solve for

𝜕𝜎𝑛𝑡 𝜕 (𝜎𝑦𝑦 − 𝜎𝑥𝑥 )


= [ sin 2𝜃 + 𝜎𝑥𝑦 cos 2𝜃] = 0
𝜕𝜃 𝜕𝜃 2
−(𝜎𝑥𝑥 − 𝜎𝑦𝑦 ) cos 2𝜃 − 2𝜎𝑥𝑦 sin 2𝜃 = 0

𝜎𝑥𝑥 − 𝜎𝑦𝑦
tan 2𝜃𝑠 = −
2𝜎𝑥𝑦

(1.15)

15
MECH3361/9361 Semester 2, 2016

For a 2D case, there are two roots to Eq. 1.15 as for the max normal stress (principal stress)
directions.

𝜎𝑥𝑥 − 𝜎𝑦𝑦
tan 2𝜃𝑠 = tan(2𝜃𝑠 + 180°) = −
2𝜎𝑥𝑦

In contrast with the principal stress directions, each root (2s) is 90 to 2p. Thus the
maximum shear and s and p are 45 apart. The planes for max shear stress can be
determined by rotating 45 from the principal plane.

( xx yy ) 2+ 2
xy
2

xy
2 s

( xx yy ) 2 s+180°
2

Fig 1.9 2D maximum shear stress directions

Referring to Fig. 1.9, we have:

(𝜎𝑥𝑥 − 𝜎𝑦𝑦 )

sin 2𝜃𝑠1 = 2
2
√(𝜎𝑥𝑥 − 𝜎𝑦𝑦 ) + 𝜎 2
2 𝑥𝑦

𝜎𝑥𝑦
cos 2𝜃𝑠1 =
2
(𝜎𝑥𝑥 − 𝜎𝑦𝑦 )
√[ 2
] + 𝜎𝑥𝑦
2

Thus the maximum shear stress is calculated as follows:

(𝜎𝑦𝑦 − 𝜎𝑥𝑥 )
(𝜎𝑛𝑡 )max = sin 2𝜃𝑠 + 𝜎𝑥𝑦 cos 2𝜃𝑠
2
(𝜎𝑥𝑥 − 𝜎𝑦𝑦 )
(𝜎𝑦𝑦 − 𝜎𝑥𝑥 ) − 𝜎𝑥𝑦
= 2 + 𝜎𝑥𝑦
2 2 2
(𝜎𝑥𝑥 − 𝜎𝑦𝑦 ) (𝜎𝑥𝑥 − 𝜎𝑦𝑦 )
√[ 2
] + 𝜎𝑥𝑦 √[ 2
] + 𝜎𝑥𝑦
2 2

16
MECH3361/9361 Semester 2, 2016

2
(𝜎𝑥𝑥 − 𝜎𝑦𝑦 )
(𝜎𝑛𝑡 )max = √[ 2
] + 𝜎𝑥𝑦
2

(1.16)

From the definition of the principal stresses, we have:

(𝜎𝑛𝑡 )max 1
{ = ± (𝜎1 − 𝜎2 )
(𝜎𝑛𝑡 )min 2
(1.17)

Example 1.5
Determine the principal and maximum shear stresses for the following stress tensor and
illustrate magnitude and direction on the infinitesimal elements.

50 20 0
𝜎 = [20 10 0 ] MPa
0 0 30

Solution

Step 1: Identify the principal stress in the z-direction (30 MPa), as there are no shear stresses
on the z-plane (i.e. 𝜎𝑧𝑥 = 𝜎𝑧𝑦 = 0). Hence, to obtain the remaining two principal stresses,
stresses are transformed in the xy-plane only, enabling the use of the 2D Mohr’s circle.

Step 2: Draw Mohr’s circle

C = 30

22=1.72 11=58.28
xx=50
0 yy=10
2
2 R=28.28

xy=20

max=28.28
max

Centre of Mohr’s circle (C):

𝜎𝑥𝑥 + 𝜎𝑦𝑦 50 + 10
C= = = 30 MPa
2 2

17
MECH3361/9361 Semester 2, 2016

Radius of Mohr’s circle (R):

𝜎𝑥𝑥 − 𝜎𝑦𝑦 2 50 − 10 2
R = √( 2 √
) + 𝜎𝑥𝑦 = ( ) + 202 = 28.28 MPa
2 2

Step 3: Determine the orientation of the principle stresses:

2𝜎𝑥𝑦 2 × 20
tan 2𝜃𝑝 = = =1
(𝜎𝑥𝑥 − 𝜎𝑦𝑦 ) (50 − 10)

arctan(1)
𝜃𝑝1 = = 22.5°
2
𝜃𝑝2 = 𝜃𝑝1 + 90° = 112.5°

Step 4: Compute the principal stresses and the maximum shear stress
(𝜎𝑥𝑥 + 𝜎𝑦𝑦 ) 𝜎𝑥𝑥 − 𝜎𝑦𝑦 2
𝜎𝑝 = ± √( 2 =C±R
) + 𝜎𝑥𝑦
2 2
= 30 ± 28.28
58.28 MPa
= {
1.72 MPa

∴ 𝜎1 = 58.28 MPa, 𝜎2 = 30 MPa, 𝜎3 = 1.72 MPa (by convention, 𝜎1 ≥ 𝜎2 ≥ 𝜎3 )

𝜎𝑥𝑥 − 𝜎𝑦𝑦 2
𝜏max = (𝜎𝑛𝑡 )max = R = √( 2
) + 𝜎𝑥𝑦
2
= 28.28 MPa

Step 4: Draw infinitesimal elements indicating the magnitude and orientations.

To associate the principal stress directions with the correct principal stresses, confirm the
results using the stress transformation equations.

𝜎1 = 𝜎𝑥𝑥 cos2 𝜃𝑝1 + 𝜎𝑦𝑦 sin2 𝜃𝑝1 + 2𝜎𝑥𝑦 cos 𝜃𝑝1 sin 𝜃𝑝1
𝜎2 = 𝜎𝑥𝑥 cos2 𝜃𝑝2 + 𝜎𝑦𝑦 sin2 𝜃𝑝2 + 2𝜎𝑥𝑦 cos 𝜃𝑝2 sin 𝜃𝑝2

The angles used to define the principal directions are with respect to the x-axis (with anti-
clockwise rotation). As defined in Eq. 1.15, the max shear direction is 45° from the principal
direction.

18
MECH3361/9361 Semester 2, 2016

Orientation of Principal Stresses Orientation of Maximum Shear Stress

22= 1.72MPa 30MPa

max=28.28MPa

11=58.28MPa

30MPa
= 22.5o = 22.5o

Note that the principal stresses correspond to zero shear; but max shears do not correspond to
zero normal stress.

1.8. Cylindrical systems


1.8.1. Cylindrical coordinate system

𝜎𝑟𝑟 𝜏𝑟𝜃 𝜏𝑟𝑧


[𝜏𝜃𝑟 𝜎𝜃𝜃 𝜏𝜃𝑧 ]
𝜏𝑧𝑟 𝜏𝑧𝜃 𝜎𝑧𝑧

1.8.2. Cylindrical pressure vessels


This analysis will look at thin-walled tubes with an internal pressure and closed ends. Let xx
be the axial stress due to the pressure on the end walls, and  = yy be the hoop stress due
to the pressure acting on the curved surface.

t y Sectioned plane

yy
P xx
xx x

Fig. 1.10 FBD of axial section of vessel

Look at a FBD of the axial section as shown in Fig. 1.10 and check the axial equilibrium.

∑𝐹𝑥 = 0 = −(𝜋𝑟 2 )𝑃 + (2𝜋𝑟𝑡)𝜎𝑥𝑥


𝑃𝜋𝑟 2 = (2𝜋𝑟𝑡)𝜎𝑥𝑥

19
MECH3361/9361 Semester 2, 2016

which gives the equation for axial stress or longitudinal stress.

𝑃𝑟
𝜎𝑥𝑥 =
2𝑡
(1.18)

Look now at a FBD of the circumferential section as shown in Fig 1.11.

t
P

Fig. 1.11 FBD of circumferential section of the vessel

Equating the forces vertically gives:

∑𝐹𝑦 = 0 = −(2𝑟𝐿)𝑃 + (2𝐿𝑡)𝜎𝜃𝜃


2𝑟𝐿𝑃 = (2𝐿𝑡)𝜎𝜃𝜃

which simplifies to give the equation for hoop stress or circumferential stress:

𝑃𝑟
𝜎𝜃𝜃 =
𝑡
(1.19)

Example 1.6
Determine the principal stresses, maximum shear stress, and their orientations for the
𝑃𝑟
pressurised cylindrical vessel in Fig 1.10. Assume = 10 MPa.
2𝑡

Solution

Step 1: Determine the principal stresses.

Since there are no shear stresses, x and y (θ in cylindrical co-ordinates) are the principal
directions. Based on Eq. 1.18–1.19,  is double xx.

∴ 𝜎1 = 𝜎𝑦𝑦 = 2𝜎𝑥𝑥 = 20 MPa and 𝜎2 = 𝜎𝑥𝑥 = 10 MPa (𝜎1 ≥ 𝜎2 )

𝜃𝑝1 = 90° (with respect to the x-axis)

𝜃𝑝2 = 0° (with respect to the x-axis)

20
MECH3361/9361 Semester 2, 2016

Step 2: Determine the maximum shear stress.


1 1
(𝜎𝑛𝑡 )max = (𝜎1 − 𝜎2 ) = (20 − 10) = 5 MPa
2 2

Alternatively, maximum shear stress can be obtained using Mohr’s circle.

2
(𝜎𝑥𝑥 −𝜎𝑦𝑦 )
(𝜎𝑛𝑡 )max = √[ 2 = 5 MPa
] + 𝜎𝑥𝑦
2

Step 3: Determine the maximum shear stress direction.

𝜎𝑥𝑥 − 𝜎𝑦𝑦 10 − 20
tan 2𝜃𝑠 = tan(2𝜃𝑠 + 180°) = − = = −∞
2𝜎𝑥𝑦 2(0)

2𝜃𝑠1 = 270°
𝜃𝑠1 = 135° (with respect to the x-axis)

2𝜃𝑠2 + 180° = 270°


𝜃𝑠2 = 45° (with respect to the x-axis)

1.9. Stress transformation (3D)


Deriving the Mohr’s circle equations for the principal stresses in 2D (Section 1.6) is rather
tedious. For this reason, we require a more flexible stress transformation approach, which can
be easily expanded to 3D cases.

1.9.1. Rationale for using the Eigenvalue method


The approach we will use in this course is the Eigenvalue method.

Let’s begin by summing the forces on a 2D triangular element, shown in Fig. 1.12.

y
A + =90°
Acos
tn nn

xx
n
xy

x
yx
Acos
yy

Fig. 1.12 FBD of triangular element in 2D

∑ 𝐹𝑥 = 0 = (−𝜎𝑥𝑥 )(𝐴 cos 𝜃) + (−𝜎𝑦𝑥 )(𝐴 cos 𝜙) + (−𝜎𝑡𝑛 cos 𝜙)(𝐴) + (𝜎𝑛𝑛 cos 𝜃)(𝐴)
𝜎𝑥𝑥 cos 𝜃 + 𝜎𝑦𝑥 cos 𝜙 = 𝜎𝑛𝑛 cos 𝜃 − 𝜎𝑡𝑛 cos 𝜙

21
MECH3361/9361 Semester 2, 2016

∑ 𝐹𝑦 = 0 = (−𝜎𝑦𝑦 )(𝐴 cos 𝜙) + (−𝜎𝑥𝑦 )(𝐴 cos 𝜃) + (𝜎𝑡𝑛 cos 𝜃)(𝐴) + (𝜎𝑛𝑛 cos 𝜙)(𝐴)
𝜎𝑦𝑦 cos 𝜙 + 𝜎𝑥𝑦 cos 𝜃 = 𝜎𝑛𝑛 cos 𝜙 − 𝜎𝑡𝑛 cos 𝜃

If n is a principal direction, σnn is a principal stress (σp), and σtn will be zero since by
definition there are no shear stresses acting on principal planes.

𝜎𝑥𝑥 cos 𝜃 + 𝜎𝑦𝑥 cos 𝜙 = 𝜎𝑝 cos 𝜃


∴{
𝜎𝑦𝑦 cos 𝜙 + 𝜎𝑥𝑦 cos 𝜃 = 𝜎𝑝 cos 𝜙

In matrix form, this can be written as:

𝜎𝑥𝑥 𝜎𝑦𝑥 𝑙 𝑙
[𝜎 𝜎𝑦𝑦 ] {𝑚} = 𝜎𝑝 {𝑚}
𝑥𝑦

where 𝑙 = cos 𝜃, and 𝑚 = cos 𝜙 (= sin 𝜃 , because θ+ϕ=90°). l and m are the direction
cosines of the principal direction n.

For an unknown principal stress 𝜎𝑝 = 𝜎, the matrix equation can be rewritten as:

𝜎𝑥𝑥 𝜎𝑦𝑥 𝑙 𝑙 𝜎𝑥𝑥 − 𝜎 𝜎𝑦𝑥 𝑙


[𝜎 ]
𝜎𝑦𝑦 𝑚{ } − 𝜎 { } = [ 𝜎𝑥𝑦 𝜎𝑦𝑦 − 𝜎] {𝑚} = 0
𝑥𝑦 𝑚

(1.20)

The equation has non-vanishing solutions if and only if the determinant of coefficient matrix
is zero.

𝜎𝑥𝑥 − 𝜎 𝜎𝑦𝑥
| 𝜎
𝑥𝑦 𝜎𝑦𝑦 − 𝜎| = (𝜎𝑥𝑥 − 𝜎)(𝜎𝑦𝑦 − 𝜎) − 𝜎𝑦𝑥 𝜎𝑥𝑦 = 0

𝜎 2 − (𝜎𝑥𝑥 + 𝜎𝑦𝑦 )𝜎 + 𝜎𝑥𝑥 𝜎𝑦𝑦 − 𝜎𝑥𝑦


2
=0

(1.21)

Solving Eq. 1.21 for σ, we get:

𝜎𝑥𝑥 + 𝜎𝑦𝑦 𝜎𝑥𝑥 − 𝜎𝑦𝑦 2


𝜎1,3 = ± √( 2
) + 𝜏𝑥𝑦
2 2

Therefore, mathematically, principal stresses are the eigenvalues of the stress tensor, and
the corresponding eigenvector is the direction cosine:

𝑙 𝑐𝑜𝑠𝜃
{ }={ }
𝑚 𝑐𝑜𝑠𝜙

(1.22)

22
MECH3361/9361 Semester 2, 2016

This approach can easily be extended to 3D problems, making the calculation of principal
stresses and principal directions standardised and fairly straightforward.

1.9.2. Stresses in any direction


The above method provides an alternative but convenient way to determine principal stresses
and their direction for 3D stress states using “cosines”. For a more general case, the direction
cosines of triangle abc inclined plane normal n are:

Δ𝑥
𝑙 = cos(𝒏, 𝑥) = 𝑐𝑜𝑠𝜃 =
√Δ𝑥 2 + Δ𝑦 2 + Δ𝑧 2
Δy
𝑚 = cos(𝒏, 𝑦) = 𝑐𝑜𝑠𝜙 =
√Δ𝑥 2 + Δ𝑦 2 + Δ𝑧 2
Δ𝑧
𝑛 = cos(𝒏, 𝑧) = 𝑐𝑜𝑠𝜑 =
√Δ𝑥 2 + Δ𝑦 2 + Δ𝑧 2

(1.23)

From here, we obtain the following geometric identity:

𝑙 2 + 𝑚 2 + 𝑛2 = 1
(1.24)

z z
c c nz n
A n
nn

A
Ax ny
Ay p Ay
y Ax y
o b nx o b
z’
x a x a Az
Az

Fig. 1.13 FBD of triangular element in 3D (a-left, b-right)

Consider the areas of each surface in Fig. 1.13a. Let’s start with the xy plane. From the
geometry, we know zop=oz’p=. Therefore, to project abc onto the xy plane, we have
𝛥𝐴𝑧 = 𝐴𝑜𝑎𝑏 = (𝛥𝐴) cos 𝜑 = (𝛥𝐴)𝑛.

Similarly, 𝛥𝐴𝑥 = 𝐴𝑜𝑏𝑐 = (𝛥𝐴)𝑙, 𝛥𝐴𝑦 = 𝐴𝑜𝑎𝑐 = (𝛥𝐴)𝑚, 𝛥𝐴𝑧 = 𝐴𝑜𝑎𝑏 = (𝛥𝐴)𝑛

To apply equilibrium to the tetrahedron, we resolve 𝜎𝑛𝑛 to 𝜎𝑛𝑥 , 𝜎𝑛𝑦 , 𝜎𝑛𝑧 as in Fig. 1.13b.

∑𝐹𝑥 = 𝜎𝑛𝑥 (Δ𝐴) − 𝜎𝑥𝑥 (Δ𝐴𝑥 ) − 𝜎𝑦𝑥 (Δ𝐴𝑦 ) − 𝜎𝑧𝑥 (Δ𝐴𝑧 )

23
MECH3361/9361 Semester 2, 2016

The latter terms are negative because they act opposite to the coordinate positive.

𝜎𝑛𝑥 (Δ𝐴) = 𝜎𝑥𝑥 (Δ𝐴)𝑙 + 𝜎𝑦𝑥 (Δ𝐴)𝑚 + 𝜎𝑧𝑥 (Δ𝐴)𝑛

∴ 𝜎𝑛𝑥 = 𝜎𝑥𝑥 𝑙 + 𝜎𝑦𝑥 𝑚 + 𝜎𝑧𝑥 𝑛

Similarly, 𝜎𝑛𝑦 = 𝜎𝑥𝑦 𝑙 + 𝜎𝑦𝑦 𝑚 + 𝜎𝑧𝑦 𝑛 and 𝜎𝑛𝑧 = 𝜎𝑥𝑧 𝑙 + 𝜎𝑦𝑧 𝑚 + 𝜎𝑧𝑧 𝑛.

This can be written more succinctly in matrix form:

𝜎𝑛𝑥 𝜎𝑥𝑥 𝜎𝑥𝑦 𝜎𝑥𝑧 𝑙


{𝜎𝑛𝑦 } = [𝜎𝑦𝑥 𝜎𝑦𝑦 𝜎𝑦𝑧 ] {𝑚}
𝜎𝑛𝑧 𝜎𝑧𝑥 𝜎𝑧𝑦 𝜎𝑧𝑧 𝑛

(1.25)

1.9.3. Stresses with coordinate rotation


As per the 2D case, we can derive the coordinate rotation. For example, if we consider
rotating axis in (𝒏, 𝒕′, 𝒕′′), with one normal & two perpendicular tangential directions. These
two perpendicular tangential directions have direction cosines as:

𝑡 ′ = (𝑙 ′ , 𝑚′ , 𝑛′ ) and 𝑡 ′′ = (𝑙 ′′ , 𝑚′′ , 𝑛′′ )

Since 𝜎𝑛𝑥 = 𝜎𝑛𝑛 𝑙, 𝜎𝑛𝑦 = 𝜎𝑛𝑛 𝑚 , 𝜎𝑛𝑧 = 𝜎𝑛𝑛 𝑛, (by resolving the normal stress 𝜎𝑛𝑛 to 𝜎𝑛𝑥 ,
σny , 𝜎𝑛𝑧 in the x, y, z directions), we have: 𝜎𝑛𝑥 𝑙 = 𝜎𝑛𝑛 𝑙 2 , 𝜎𝑛𝑦 𝑚 = 𝜎𝑛𝑛 𝑚2 , 𝜎𝑛𝑧 𝑛 = 𝜎𝑛𝑛 𝑛2 .

Adding these three equations gives 𝜎𝑛𝑥 𝑙 + 𝜎𝑛𝑦 𝑚 + 𝜎𝑛𝑧 𝑛 = 𝜎𝑛𝑛 (𝑙 2 + 𝑚2 + 𝑛2 ), and since
𝑙 2 + 𝑚2 + 𝑛2 = 1, we have:

𝜎𝑛𝑥
𝜎𝑛𝑛 = 𝜎𝑛𝑥 𝑙 + 𝜎𝑛𝑦 𝑚 + 𝜎𝑛𝑧 𝑛 = {𝑙 𝑚 𝑛} {𝜎𝑛𝑦 }
𝜎𝑛𝑧

(1.26)

Plugging (1.25) into (1.26) in matrix form, we get:

𝜎𝑛𝑥 𝜎𝑥𝑥 𝜎𝑥𝑦 𝜎𝑥𝑧 𝑙


𝜎𝑛𝑛 = {𝑙 𝑚 𝑛} {𝜎𝑛𝑦 } = {𝑙 𝑚 𝑛} ([𝜎𝑦𝑥 𝜎𝑦𝑦 𝜎𝑦𝑧 ] {𝑚})
𝜎𝑛𝑧 𝜎𝑧𝑥 𝜎𝑧𝑦 𝜎𝑧𝑧 𝑛

Similarly, for other two normal t’ and t’’ components:

𝜎𝑛𝑥 𝜎𝑥𝑥 𝜎𝑥𝑦 𝜎𝑥𝑧 𝑙′


𝜎𝑡 ′ 𝑡 ′ = {𝑙′ 𝑚′ 𝜎
𝑛′} { 𝑛𝑦 } = {𝑙′ 𝑚′ 𝜎
𝑛′} ([ 𝑦𝑥 𝜎𝑦𝑦 𝜎𝑦𝑧 ] {𝑚′})
𝜎𝑛𝑧 𝜎𝑧𝑥 𝜎𝑧𝑦 𝜎𝑧𝑧 𝑛′

24
MECH3361/9361 Semester 2, 2016

𝜎𝑛𝑥 𝜎𝑥𝑥 𝜎𝑥𝑦 𝜎𝑥𝑧 𝑙′′


𝜎𝑡 ′′ 𝑡 ′′ = {𝑙′′ 𝑚′′ 𝑛′′} {𝜎𝑛𝑦 } = {𝑙′′ 𝑚′′ 𝑛′′} ([𝜎𝑦𝑥 𝜎𝑦𝑦 𝜎𝑦𝑧 ] {𝑚′′})
𝜎𝑛𝑧 𝜎𝑧𝑥 𝜎𝑧𝑦 𝜎𝑧𝑧 𝑛′′

And again for all of the shear components. As a result, we can derive:

𝜎𝑛𝑛 𝜎𝑛𝑡 ′ 𝜎𝑛𝑡 ′′ 𝑙 𝑚 𝑛 𝜎𝑥𝑥 𝜎𝑥𝑦 𝜎𝑥𝑧 𝑙 𝑚 𝑛 𝑇


[ 𝜎𝑡 ′ 𝑛 𝜎𝑡 ′ 𝑡 ′ 𝜎𝑡 ′ 𝑡 ′′ ]=[ 𝑙 ′ 𝑚′ 𝑛′ ] [𝜎𝑦𝑥 𝜎𝑦𝑦 𝜎𝑦𝑧 ] [ 𝑙 ′ 𝑚′ 𝑛′ ]
𝜎𝑡 ′′ 𝑛 𝜎𝑡 ′′ 𝑡 ′ 𝜎𝑡 ′′ 𝑡 ′′ 𝑙 ′′ 𝑚′′ 𝑛′′ 𝜎𝑧𝑥 𝜎𝑧𝑦 𝜎𝑧𝑧 𝑙 ′′ 𝑚′′ 𝑛′′

(1.27)

Thus:

𝝈𝑛𝑒𝑤 = 𝑹𝝈𝑜𝑙𝑑 𝑹𝑇
(1.28)

𝑙 𝑚 𝑛
where 𝑅 = [ 𝑙′ 𝑚′ 𝑛′ ] is called the “transformation matrix”.
′′
𝑙 𝑚′′ 𝑛′′

1.10. Principal stresses (3D)


1.10.1. The Eigenvalue method
From 2D, we know that the principal stresses at a point are the eigenvalues of the stress
tensor. We can extend this from 2D to 3D, making it easier to solve for 3D principal stresses.

From Eq. (1.25), the Eigen equation of the 3D stress tensor can be written as:

𝜎𝑥𝑥 𝜎𝑥𝑦 𝜎𝑥𝑧 𝑙𝑝 𝜎𝑛𝑥 𝜎𝑛𝑛 𝑙𝑝 𝑙𝑝


𝜎
[ 𝑦𝑥 𝜎𝑦𝑦 𝜎𝑦𝑧 ] {𝑚𝑝 } = {𝜎𝑛𝑦 } = {𝜎𝑛𝑛 𝑚𝑝} = 𝜎 {𝑚𝑝 }
𝜎𝑧𝑥 𝜎𝑧𝑦 𝜎𝑧𝑧 𝑛𝑝 𝜎𝑛𝑧 𝜎𝑛𝑛 𝑛𝑝 𝑛𝑝

(1.29)

where {𝑙𝑝 𝑚𝑝 𝑛𝑝 }𝑇 is the direction cosine for the principal plane and  is the principal
stress. Re-arranging Eq. (1.30) leads to:

𝜎𝑥𝑥 − 𝜎 𝜎𝑥𝑦 𝜎𝑥𝑧 𝑙𝑝


[ 𝜎𝑦𝑥 𝜎𝑦𝑦 − 𝜎 𝜎𝑦𝑧 ] {𝑚𝑝 } = 0
𝜎𝑧𝑥 𝜎𝑧𝑦 𝜎𝑧𝑧 − 𝜎 𝑛𝑝

(1.30)

To have a non-vanishing solution, the coefficient determinant must be zero, i.e.:

25
MECH3361/9361 Semester 2, 2016

𝜎𝑥𝑥 − 𝜎 𝜎𝑥𝑦 𝜎𝑥𝑧


| 𝜎𝑦𝑥 𝜎𝑦𝑦 − 𝜎 𝜎𝑦𝑧 | = 0
𝜎𝑧𝑥 𝜎𝑧𝑦 𝜎𝑧𝑧 − 𝜎

(1.31)

This gives rise to the characteristic Eigen equation:

𝜎 3 − 𝐼1 𝜎 2 + 𝐼2 𝜎 − 𝐼3 = 0
(1.32)

where the first, second, and third invariants (the coefficients above) are:

𝐼1 = 𝜎𝑥𝑥 + 𝜎𝑦𝑦 + 𝜎𝑧𝑧

2 2 2
𝐼2 = 𝜎𝑥𝑥 𝜎𝑦𝑦 + 𝜎𝑦𝑦 𝜎𝑧𝑧 + 𝜎𝑧𝑧 𝜎𝑥𝑥 − 𝜎𝑥𝑦 − 𝜎𝑦𝑧 − 𝜎𝑧𝑥

2 2 2
𝐼3 = 𝜎𝑥𝑥 𝜎𝑦𝑦 𝜎𝑧𝑧 + 2𝜎𝑥𝑦 𝜎𝑦𝑧 𝜎𝑧𝑥 − 𝜎𝑥𝑥 𝜎𝑦𝑧 − 𝜎𝑦𝑦 𝜎𝑧𝑥 − 𝜎𝑧𝑧 𝜎𝑥𝑦

(1.33)

Since Eq. 1.32 is a third order equation, there are three roots that can be found using standard
analytical solutions. These three roots are the maximum, intermediate, and minimum
principal stresses, 𝜎1 , 𝜎2 , and 𝜎3 respectively, where:

𝜎1 ≥ 𝜎2 ≥ 𝜎3
(1.34)

When principal stresses 𝜎1 , 𝜎2 , and 𝜎3 are known, we can determine the three invariants as:

𝐼1 = 𝜎1 + 𝜎2 + 𝜎3

𝐼2 = 𝜎1 𝜎2 + 𝜎2 𝜎3 + 𝜎3 𝜎1

𝐼3 = 𝜎1 𝜎2 𝜎3

(1.35)

These are three very important quantities in solid mechanics.

1.10.2. Principal direction cosines


To determine the principal direction cosines, we can solve equations 1.30 for a specific
principal stress. For example, for 𝜎1 :

26
MECH3361/9361 Semester 2, 2016

𝜎𝑥𝑥 − 𝜎1 𝜎𝑥𝑦 𝜎𝑥𝑧 𝑙1


[ 𝜎𝑦𝑥 𝜎𝑦𝑦 − 𝜎1 𝜎𝑦𝑧 ] {𝑚1 } = 0
𝜎𝑧𝑥 𝜎𝑧𝑦 𝜎𝑧𝑧 − 𝜎1 𝑛1

But the geometric relationship 𝑙 2 + 𝑚2 + 𝑛2 = 1 must also hold. Therefore, we pick any two
linear equations and combine them with the geometric equation:

(𝜎𝑥𝑥 − 𝜎1 )𝑙1 + 𝜎𝑦𝑥 𝑚1 + 𝜎𝑥𝑧 𝑛1 = 0


(Solve simultaneously) {𝜎𝑦𝑥 𝑙1 + (𝜎𝑦𝑦 − 𝜎1 )𝑚1 + 𝜎𝑦𝑧 𝑛1 = 0
𝑙12 + 𝑚12 + 𝑛12 = 1

(1.36)

The third linear equation (not listed in Eq. 1.36) should be used for checking the calculation.

Example 1.7
At a point in a structure subjected to 3D loading, the stress was measured as:

50 −20 0
𝜎 = [−20 80 60 ] MPa.
0 60 −70

Determine: (1) the stress invariants, (2) principal stresses, (3) direction cosines.

Solution

Step 1: Calculate stress invariants.

𝐼1 = 𝜎𝑥𝑥 + 𝜎𝑦𝑦 + 𝜎𝑧𝑧 = 50 + 80 − 70 = 60

2 2 2
𝐼2 = 𝜎𝑥𝑥 𝜎𝑦𝑦 + 𝜎𝑦𝑦 𝜎𝑧𝑧 + 𝜎𝑧𝑧 𝜎𝑥𝑥 − 𝜎𝑥𝑦 − 𝜎𝑦𝑧 − 𝜎𝑧𝑥
= 50 × 80 + 80 × (−70) + (−70) × 50 − (−20)2 − 602 − 02
= −9100
2 2 2
𝐼3 = 𝜎𝑥𝑥 𝜎𝑦𝑦 𝜎𝑧𝑧 + 2𝜎𝑥𝑦 𝜎𝑦𝑧 𝜎𝑧𝑥 − 𝜎𝑥𝑥 𝜎𝑦𝑧 − 𝜎𝑦𝑦 𝜎𝑧𝑥 − 𝜎𝑧𝑧 𝜎𝑥𝑦
= 50 × 80 × (−70) + 2 × (−20) × 60 × 0 − 50 × 602
−80 × 02 − (−70) × (−20)2
= −432000

Step 2: Write the Eigen equation.

𝜎 3 − 𝐼1 𝜎 2 + 𝐼2 𝜎 − 𝐼3 = 𝜎 3 − 60𝜎 2 − 9100𝜎 + 432000 = 0

Step 3: Solve the Eigen equation for the principal stresses.

Method 1: Use a calculator

27
MECH3361/9361 Semester 2, 2016

Solve the cubic equation directly, then rank the roots in order of magnitude.

𝜎1 = 107.3 MPa, 𝜎2 = 44.1 MPa, 𝜎3 = −91.4 MPa

Method 2: Use MATLAB

You can use the coefficients of the Eigen equation (which are the invariants adjusted
for the signs in the equation):

>> a = [1 -60 -9100 432000]


>> x=roots(a)

x =
-91.3610
107.2882
44.0728
…or determine the Eigenvalue and Eigenvector directly from the stress tensor:

>> A = [50 -20 0; -20 80 60; 0 60 -70]


>>[v,d]=eig(A)

v =
0.0474 0.9483 0.3140
0.3350 0.2810 -0.8993
-0.9410 0.1478 -0.3044

d =
-91.3610 0 0
0 44.0728 0
0 0 107.2882

The d matrix will contain the principal stresses. Do not forget to rank them.

Step 4: Solve for the principal directions.

For each principal stress, substitute the value as per Eq. 1.36.

(50 − 107.3)𝑙1 + (−20)𝑚1 + (0)𝑛1 = 0 𝑙1 0.314


{ (−20)𝑙1 + (80 − 107.3)𝑚1 + 60𝑛1 = 0  {𝑚1 } = {−0.900}
𝑙12 + 𝑚12 + 𝑛12 = 1 𝑛1 −0.303

Similarly for 𝜎2 = 44.1 MPa and 𝜎3 = −91.4 MPa:

𝑙2 0.948 𝑙3 0.048
{𝑚2 } = {0.282}, {𝑚3 } = { 0.337 }
𝑛2 0.146 𝑛3 −0.940

Alternatively, the v matrix from MATLAB will give the corresponding directions
(reading down each column).

28
MECH3361/9361 Semester 2, 2016

Example 1.8
Determine the principal stress and principal planes for the following stress tensor:

−10 15 −25
𝜎 = [ 15 20 −35] MPa.
−25 −35 −30

Solution

Step 1: Calculate the stress invariants.

𝐼1 = 𝜎𝑥𝑥 + 𝜎𝑦𝑦 + 𝜎𝑧𝑧 = −10 + 20 − 30 = −20

𝜎𝑦𝑦 𝜎𝑦𝑧 𝜎𝑥𝑥 𝜎𝑥𝑧 𝜎𝑥𝑥 𝜎𝑥𝑦


𝐼2 = | 𝜎 𝜎𝑧𝑧 | + | 𝜎𝑧𝑥 𝜎𝑧𝑧 | + | 𝜎𝑦𝑥 𝜎𝑦𝑦 |
𝑧𝑦
2 2 2
= 𝜎𝑥𝑥 𝜎𝑦𝑦 + 𝜎𝑦𝑦 𝜎𝑧𝑧 + 𝜎𝑧𝑧 𝜎𝑥𝑥 − 𝜎𝑥𝑦 − 𝜎𝑦𝑧 − 𝜎𝑧𝑥
= (−10) × 20 + 20 × (−30) + (−30) × (−10) − 152 − (−25)2 − (−35)2
= −2575

𝜎𝑥𝑥 𝜎𝑥𝑦 𝜎𝑥𝑧


𝐼3 = | 𝑦𝑥 𝜎𝑦𝑦 𝜎𝑦𝑧 |
𝜎
𝜎𝑧𝑥 𝜎𝑧𝑦 𝜎𝑧𝑧
2 2 2
= 𝜎𝑥𝑥 𝜎𝑦𝑦 𝜎𝑧𝑧 + 2𝜎𝑥𝑦 𝜎𝑦𝑧 𝜎𝑧𝑥 − 𝜎𝑥𝑥 𝜎𝑦𝑧 − 𝜎𝑦𝑦 𝜎𝑧𝑥 − 𝜎𝑧𝑧 𝜎𝑥𝑦
= (−10) × 20 × (−30) + 2 × 15 × (−25) × (−35) − (−10) × (−35)2
−20 × (−25)2 − (−30) × 152
= 38750

Step 2: Solve the Eigen equation.

𝜎 3 + 20𝜎 2 − 2575𝜎 − 38750 = 0

>> a = [1 -60 -9100 432000]


>> x=roots(a)

𝜎1 = 48.9 MPa, 𝜎2 = −14.6 MPa, 𝜎3 = −54.3 MPa

Alternative solution:

Step 1: Solve the Eigenvalue problem

−10 15 −25 𝑙𝑝 𝑙𝑝
[ 15 𝑚 𝑚
20 −35] { 𝑝 } = 𝜎 { 𝑝 }
−25 −35 −30 𝑛𝑝 𝑛𝑝

in MATLAB directly.

>> A=[-10 15 -25; 15 20 -35; -25 -35 -30]


>> e=eig(A)
or

29
MECH3361/9361 Semester 2, 2016

>>[v,d]=eig(A)

v =
0.3761 0.8353 0.4011
0.3317 -0.5255 0.7834
0.8652 -0.1616 -0.4747

d =
-54.2852 0 0
0 -14.6016 0
0 0 48.8867

Example 1.9
A state of stress at a point is defined by xx = zz =10MPa, yy = –20MPa, xy = –4MPa and
yz = zx = 0. (i) Find the principal stresses and their direction cosines with respect to the x-
axis. (ii) Draw the stress state on an infinitesimal isolated cube element.

Solution:

Step 1: Organise the stress tensor.

10 −4 0
𝜎 = [−4 −20 0 ] MPa.
0 0 10

Step 2: Simplify the problem.

Since 𝜎𝑥𝑧 = 𝜎𝑦𝑧 = 0, z is a principal direction (by observation). Thus, the 3D


problem can be reduced to a 2D problem. We simply solve:

10 −4
𝜎=[ ] MPa.
−4 −20

Step 3: Calculate principal stresses.

From Eq. 1.14:

2
(𝜎𝑥𝑥 + 𝜎𝑦𝑦 ) (𝜎𝑥𝑥 − 𝜎𝑦𝑦 )
𝜎𝑝 = ± √[ 2
] + 𝜎𝑥𝑦
2 2
2
(10 + (−20)) (10 − (−20))
= ± √[ ] + (−4)2
2 2
10.52
={
−20.52

Step 4: Rank the principal stresses.

30
MECH3361/9361 Semester 2, 2016

𝜎1 = 10.52 MPa, 𝜎2 = 10 MPa, 𝜎3 = −20.52 MPa

Step 5: Determine the principal direction cosines.

2𝜎𝑥𝑦 2 × (−4) 4
tan 2𝜃𝑝 = = =−
𝜎𝑥𝑥 − 𝜎𝑦𝑦 10 − (−20) 15
−14.93°
2𝜃 = {
−14.93° + 180°
−7.47°
∴𝜃={
82.54°

For 𝜃 = −7.47°,

𝜎−7.47 = 𝜎𝑥𝑥 cos2 𝜃 + 𝜎𝑦𝑦 sin2 𝜃 + 2𝜎𝑥𝑦 𝑐𝑜𝑠𝜃𝑠𝑖𝑛𝜃


= 10 cos2 (−7.47°) − 20 sin2(−7.47°) + 2(−4) cos(−7.47°) sin(−7.47°)
= 10.52 MPa

Thus, 𝜃 = −7.47° is the direction of 𝜎1 relative to the x-axis. We can say that
𝑙1 = cos(−7.47°) = 0.992, the x-direction cosine of the first principal stress. By
elimination, 𝑙3 = cos(82.54°) = 0.130 is the x-direction cosine of the third
principal stress.

Given the 2D simplification, we can calculate the other direction cosines for each
principal stress.

𝑙1 cos(−7.47°) 0.992
For 𝜎1 , {𝑚1 } = {cos(−7.47° − 90°)} = {−0.130}
𝑛1 cos(90°) 0

𝑙2 cos(90°) 0
For 𝜎2 , {𝑚2 } = {cos(90°)} = {0}
𝑛2 cos(0°) 1

𝑙3 cos(82.54°) 0.130
For 𝜎3 , {𝑚3 } = {cos(90° − 82.54°)} = {0.992}
𝑛3 cos(90°) 0

Step 6: Plot the infinitesimal element in 2D x-y plane.

Note that the principal stresses showed should be 𝜎1 = 10.52 and 𝜎3 = −20.52 MPa
(rather than the typical 𝜎1 , 𝜎2 in 2D problems).

31
MECH3361/9361 Semester 2, 2016

3rd principal plane


y 3 1st principal
plane

82.54°
-7.47°
1
y x
-7.47°- 90°
x
-7.47°
Norm of 1st p-plane
Angle for direction cosine

We can check the answers in MATLAB. Note that MATLAB Eigen vector results
may indicate the opposite principal plane.

>> A=[10 -4 0; -4 -20 0; 0 0 10]


>>[v,d]=eig(A)

v= d=
0.1299 -0.0000 -0.9915 -20.5242 0 0
0.9915 0.0000 0.1299 0 10.0000 0
0.0000 1.0000 0.0000 0 0 10.5242

1.11. Maximum shear stresses (3D)


z z
c c nz n
A n
nn

A
Ax ny
Ay p Ay
y Ax y
o b nx o b
z’
x a x a Az
Az

Fig 1.14 FBD of triangular element (3D)

Assume that the coordinate directions are the principal directions (i.e. there are no shear
components).

𝜎1 = 𝜎𝑥𝑥 , 𝜎2 = 𝜎𝑦𝑦 , and 𝜎3 = 𝜎𝑧𝑧


𝜎𝑛𝑥 = 𝑙𝜎1, 𝜎𝑛𝑦 = 𝑚𝜎2 , and 𝜎𝑛𝑧 = 𝑛𝜎3

The total stress on the inclined plane abc becomes:

32
MECH3361/9361 Semester 2, 2016

2
𝜎 2 = (𝜎𝑛𝑥 )2 + (𝜎𝑛𝑦 ) + (𝜎𝑛𝑧 )2 = (𝑙𝜎1 )2 + (𝑚𝜎2 )2 + (𝑛𝜎3 )2

But from Eq. 1.26, i.e. 𝜎𝑛𝑛 = 𝜎𝑛𝑥 𝑙 + 𝜎𝑛𝑦 𝑚 + 𝜎𝑛𝑧 𝑛:

𝜎𝑛𝑛 = 𝜎𝑛𝑥 𝑙 + 𝜎𝑛𝑦 𝑚 + 𝜎𝑛𝑧 𝑛


= (𝑙𝜎1 )𝑙 + (𝑚𝜎2 )𝑚 + (𝑛𝜎3 )𝑛
= 𝜎1 𝑙 2 + 𝜎2 𝑚2 + 𝜎3 𝑛2
2
∴ 𝜎𝑛𝑛 = (𝜎1 𝑙 2 + 𝜎2 𝑚2 + 𝜎3 𝑛2 )2

The above equations have no shear, but since shear 𝜎𝑛𝑡 and normal 𝜎𝑛𝑛 are perpendicular to
each other, we can use Pythagoras’ Theorem to get 𝜎 2 = 𝜎𝑛𝑛 2 2
+ 𝜎𝑛𝑡 . As such, if there was shear
in the inclined plane, the square of the total shear stress on the plane should be:

2
𝜎𝑛𝑡 = 𝜎 2 − 𝜎𝑛𝑛
2
= (𝑙𝜎1 )2 + (𝑚𝜎2 )2 + (𝑛𝜎3 )2 − (𝜎1 𝑙 2 + 𝜎2 𝑚2 + 𝜎3 𝑛2 )2

𝜕𝜎𝑛𝑡 𝜕𝜎𝑛𝑡
= 0 and = 0 (note that only two of them are independent), so one can derive the
𝜕𝑙 𝜕𝑚
extrema shear stresses as follows:

1
± (𝜎1 − 𝜎2 )
2
1
𝜎𝑛𝑡|𝑒𝑥𝑡𝑟𝑒𝑚𝑎 = ± (𝜎2 − 𝜎3 )
2
1
{± 2 (𝜎3 − 𝜎1 )

(1.37)

The largest of these would be the maximum shear stress. This is clearly given by:

1
𝜎𝑛𝑡|𝑚𝑎𝑥 = (𝜎 − 𝜎3 )
2 1
(1.38)

Each extreme shear stress bisects the angle between two principal axes. From vector
subtraction, we can find the corresponding direction cosines. For the maximum shear stress,
these are:

𝑙𝑠 𝑙1 − 𝑙3 1
{𝑚𝑠 } = {𝑚1 − 𝑚3 }
2 2 2
𝑛𝑠 𝑛1 − 𝑛3 √(𝑙1 − 𝑙3 ) + (𝑚1 − 𝑚3 ) + (𝑛1 − 𝑛3 )
(1.39)

33
MECH3361/9361 Semester 2, 2016

1.12. Equations of motion and equilibrium


In the previous sections, we discussed the stress at a point. The question now is how stress
varies from one point to another. Let’s consider a finite element as shown below:

z
zz+ zz G
C
zy+ zy
zx+ zx F yz+ yz
B
xz+ xz
yy+
o
yy
y
yx+ yx
xy+ xy H
xx+ xx

x A E

Fig 1.15 3D finite element under equilibrium

1.12.1. Stress components


This is no longer the stress at a point, but is instead a cube where we must take into account
the stress variation from one side to the other. To do so, we use the following notation:

At x = 0, back plane (OCGH): 𝜎𝑥𝑥 , 𝜎𝑥𝑦 , 𝜎𝑥𝑧

At y = 0, left plane (ABCO): 𝜎𝑦𝑥 , 𝜎𝑦𝑦 , 𝜎𝑦𝑧

At z = 0, bottom plane (AEHO): 𝜎𝑧𝑥 , 𝜎𝑧𝑦 , 𝜎𝑧𝑧

Over a small increment of Δ𝑥, Δ𝑦, Δ𝑧, we have:

At 𝑥 = 𝛥𝑥, front plane (ABFE): 𝜎𝑥𝑥 + Δ𝜎𝑥𝑥 , 𝜎𝑥𝑦 + Δ𝜎𝑥𝑦 , 𝜎𝑥𝑧 + Δ𝜎𝑥𝑧

At 𝑦 = 𝛥𝑦, right plane (EHGF): 𝜎𝑦𝑥 + Δ𝜎𝑦𝑥 , 𝜎𝑦𝑦 + Δ𝜎𝑦𝑦 , 𝜎𝑦𝑧 + Δ𝜎𝑦𝑧

At 𝑧 = 𝛥𝑧, top plane (BCGF): 𝜎𝑧𝑥 + Δ𝜎𝑧𝑥 , 𝜎𝑧𝑦 + Δ𝜎𝑧𝑦 , 𝜎𝑧𝑧 + Δ𝜎𝑧𝑧

1.12.2. Body forces


In addition to the stress components, the cube may also be subjected to a body force
(typically gravity: mg = (V)g, thus the body force is g per unit volume). Assume that the
body force per unit volume is f.

1.12.3. Motion and equilibrium


Consider the sum of forces in the x direction:

∑𝐹𝑥 = 𝑚𝑎𝑥

34
MECH3361/9361 Semester 2, 2016

∴ ∑𝐹𝑥 = [(𝜎𝑥𝑥 + Δ𝜎𝑥𝑥 )Δ𝑦Δ𝑧 − (𝜎𝑥𝑥 )Δ𝑦Δ𝑧] + [(𝜎𝑦𝑥 + Δ𝜎𝑦𝑥 )Δ𝑥Δ𝑧 − (𝜎𝑦𝑥 )Δ𝑥Δ𝑧]
+[(𝜎𝑧𝑥 + Δ𝜎𝑧𝑥 )Δ𝑥Δ𝑦 − (𝜎𝑧𝑥 )Δ𝑥Δ𝑦] + (Δ𝑥Δ𝑦Δ𝑧)𝜌𝑓𝑥
= (𝜌Δ𝑥Δ𝑦Δ𝑧)𝑎𝑥

Dividing by the volume, Δ𝑥Δ𝑦Δ𝑧, we get:

Δ𝜎𝑥𝑥 Δ𝜎𝑥𝑦 Δ𝜎𝑥𝑧


+ + + 𝜌𝑓𝑥 = 𝜌𝑎𝑥
Δ𝑥 Δy Δ𝑧

When Δ → 0:

∂𝜎𝑥𝑥 ∂𝜎𝑥𝑦 ∂𝜎𝑥𝑧


+ + + 𝜌𝑓𝑥 = 𝜌𝑎𝑥
∂𝑥 ∂y ∂𝑧

(1.40a)

Similarly, for the other directions:

∂𝜎𝑦𝑥 ∂𝜎𝑦𝑦 ∂𝜎𝑦𝑧


+ + + 𝜌𝑓𝑦 = 𝜌𝑎𝑦
∂𝑥 ∂y ∂𝑧

(1.40b)

∂𝜎𝑧𝑥 ∂𝜎𝑧𝑦 ∂𝜎𝑧𝑧


+ + + 𝜌𝑓𝑧 = 𝜌𝑎𝑧
∂𝑥 ∂y ∂𝑧

(1.40c)

Under static equilibrium, acceleration is zero. Therefore, the equilibrium equations are:

∂𝜎𝑥𝑥 ∂𝜎𝑥𝑦 ∂𝜎𝑥𝑧


+ + + 𝑏𝑥 = 0
∂𝑥 ∂y ∂𝑧
∂𝜎𝑦𝑥 ∂𝜎𝑦𝑦 ∂𝜎𝑦𝑧
+ + + 𝑏𝑦 = 0
∂𝑥 ∂y ∂𝑧
∂𝜎𝑧𝑥 ∂𝜎𝑧𝑦 ∂𝜎𝑧𝑧
+ + + 𝑏𝑧 = 0
{ ∂𝑥 ∂y ∂𝑧
(1.41)

35
MECH3361/9361 Semester 2, 2016

We can simplify these under certain conditions. For example, if body forces are negligible:

∂𝜎𝑥𝑥 ∂𝜎𝑥𝑦 ∂𝜎𝑥𝑧


+ + =0
∂𝑥 ∂y ∂𝑧
∂𝜎𝑦𝑥 ∂𝜎𝑦𝑦 ∂𝜎𝑦𝑧
+ + =0
∂𝑥 ∂y ∂𝑧
∂𝜎𝑧𝑥 ∂𝜎𝑧𝑦 ∂𝜎𝑧𝑧
+ + =0
{ ∂𝑥 ∂y ∂𝑧

(1.42)

or, for a 2D case:

∂𝜎𝑥𝑥 ∂𝜎𝑥𝑦
+ =0
∂𝑥 ∂y
∂𝜎𝑦𝑥 ∂𝜎𝑦𝑦
+ =0
{ ∂𝑥 ∂y

(1.43)

These can be written in short form as:

𝜎𝑖𝑗,𝑗 + 𝑏𝑖 = 𝜌𝑢̈ 𝑖

(1.44)

or

𝛁 ⋅ 𝝈 + 𝒃 = 𝜌𝒂

(1.45)
𝜎𝑥𝑥 𝜎𝑥𝑦 𝜎𝑥𝑧
∂ ∂ ∂
where divergence 𝛁 = {∂x ∂y ∂z
} and stress tensor 𝝈 = [𝜎𝑦𝑥 𝜎𝑦𝑦 𝜎𝑦𝑧 ].
𝜎𝑧𝑥 𝜎𝑧𝑦 𝜎𝑧𝑧

36
MECH3361/9361 Semester 2, 2016

Example 1.10
A thin rectangular plate as shown is in static equilibrium and body forces are negligible.
Derive the stress function by using the “engineering beam theory” from AMME2301 and
check whether it satisfies the equilibrium equations.

y P

0 x h

b
L

Solution:

Bending moments:

𝑀 = −𝑃(𝐿 − 𝑥)

𝑀𝑦 𝑃(𝐿 − 𝑥)𝑦
𝜎𝑥𝑥 = − = = 𝑞(𝐿 − 𝑥)𝑦
𝐼 𝐼
𝑃 1
where constant 𝑞 = and second moment of area 𝐼 = 12 𝑏ℎ3 .
𝐼

Since there are no other loads, 𝜎𝑦𝑦 = 0 and 𝜎𝑧𝑧 = 0.

Shear stress due to transverse force P:

𝑉(𝑥)𝑄(𝑦) 𝑉𝑄
𝜎𝑥𝑦 = =−
𝐼𝑡(𝑦) 𝐼𝑡

where:

V(x) the shear force carried by the section, found from the shear force diagram;
Q(y) the first moment of area;
I the second moment of area;
t(y) the sectional width at distance y from the neutral axis.

To calculate Q:
𝑦𝑡𝑜𝑝
𝑄(𝑦) = ∫ 𝑦𝑡(𝑦)𝑑𝑦 = 𝑦̅ ′ 𝐴′
𝑦

37
MECH3361/9361 Semester 2, 2016

A’ is the top (or bottom) portion of the member’s cross-sectional area, defined from the
section where t(y) is measured, and 𝑦̅ ′ is the distance to the centroid of A’, measured from the
neutral axis.

Centroid of A’
Parabolic
=0 curve
A’ y’

h max
max

h/2 y

NA
NA

b
Shear Stress distribution

The distribution of the shear stress throughout the cross section due to a shear force V can be
determined by computing the shear stress at an arbitrary height y from the neutral axis.

1 ℎ ℎ 1 ℎ2
𝑄(𝑦) = 𝑦̅ ′ 𝐴′ = (𝑦 + ( − 𝑦)) × (( − 𝑦) 𝑏) = ( − 𝑦 2 ) 𝑏
2 2 2 2 4

With t = b, applying the shear formula, we have:

1 ℎ2
𝑉𝑄 𝑃 × [2 ( 4 − 𝑦 2 ) 𝑏] 6𝑃 ℎ2
𝜎𝑥𝑦 = − =− =− ( − 𝑦2)
𝐼𝑡 𝑏ℎ3 𝑏ℎ3 4
( 12 ) × 𝑏

Now we can check whether the 2D equilibrium equations are satisfied.

For the x component:

∂𝜎𝑥𝑥 ∂𝜎𝑥𝑦 ∂ 𝑃(𝐿 − 𝑥)𝑦 ∂ 6𝑃 ℎ2


+ = [ ] + [− 3 ( − 𝑦 2 )]
∂𝑥 ∂y ∂𝑥 𝐼 ∂y 𝑏ℎ 4

𝑃 12𝑃 𝑃 12𝑃 12𝑃 12𝑃


= 𝑦 [− + 3 ] = 𝑦 [− 3 + 3 ] = 𝑦 [− 3 + 3 ] = 0
𝐼 𝑏ℎ 𝑏ℎ 𝑏ℎ 𝑏ℎ 𝑏ℎ
( 12 )

∂𝜎𝑦𝑥 ∂𝜎𝑦𝑦 ∂ 6𝑃 ℎ2 ∂
+ = [− 3 ( − 𝑦 2 )] + (0) = 0 + 0 = 0
∂𝑥 ∂y ∂𝑥 𝑏ℎ 4 ∂y

Therefore, the stress functions satisfy the equilibrium equations.

38

Vous aimerez peut-être aussi