You are on page 1of 52

SOLUTIONS MANUAL

Engineering Materials I
An Introduction to Properties, Applications and Design, Fourth Edition

Solutions to Examples

rA
2. 1. (a) For commodity A P (t ) = CA exp t,
100
rB
and for commodity B Q(t ) = CB exp t,
100
where CA and CB are the current rates of consumption (t = t0) and P(t) and Q(t) are the
values at t = t. Equating and solving for t gives

100 C 
t= ln  A  .
rB − rA  CB 

(b) The doubling time, tD, is calculated by setting C(t = t) = 2C0, giving

100 70
tD = ln 2 ≈ .
r r

Substitution of the values given for r in the table into this equation gives the doubling
times as 35, 23 and 18 years respectively.

(c) Using the equation of Answer (a) we find that aluminium overtakes steel in 201
years; polymers overtake steel in 55 years.

2.2. Principal conservation measures (see Section 2.7):

Substitution

Examples: aluminium for copper as a conductor; reinforced concrete for wood, stone
or cast-iron in construction; plastics for glass or metals as containers. For many
applications, substitutes are easily found at small penalty of cost. But in certain
specific uses, most elements are not easily replaced. Examples: tungsten in cutting
tools and lighting (a fluorescent tube contains more tungsten, as a starter filament,
than an incandescent bulb!); lead in lead-acid batteries; platinum as a catalyst in
chemical processing; etc. A long development time (up to 25 years) may be needed to
find a replacement.

Recycling

The fraction of material recycled is obviously important. Products may be re-designed


to make recycling easier, and new recycling processes developed, but development
time is again important.

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
More Economic Design

Design to use proportionally smaller amounts of scarce materials, for example, by


building large plant (economy of scale); using high-strength materials; use of surface
coatings to prevent metal loss by corrosion (e.g. in motor cars).

2.3. (a) If the current rate of consumption in tonnes per year is C then exponential
growth means that

dC r
= C,
dt 100

where r is the fractional rate of growth in % per year. Integrating gives

 r (t − t0 ) 
C = C0 exp  ,
 100 

where C0 was the consumption rate at time t = t0.


(b) Set
Q
= ∫ C dt ,
t1/2

2 0

where
 rt 
C = C0 exp  .
100 
Then
t1/2
Q  100  rt  
= C0 exp   ,
2  r 100   0

which gives the desired result.

2.4. See Chapter 2 for discussion with examples.

3.1. Refer to the results at the end of Chapter 3 for the elastic buckling of struts (pp 52 and
53), and second moments of area (pp 49 and 50). Appropriate situation is probably
Case 2 (left hand side drawing).

 EI  πr 4
Fcr = 9.87 2 , I = ,
l  4
9.87πE  r  2 9.87π × 2 × 10 4 N mm–2  8.5 mm 
2 2

Fcr =   r =   8.5 2 mm2


4  l 4  750 mm 
= 1439 N = 148 kgf.

This gives a factor of safety of about 148/90 = 1.65, so he should be OK.

3.2. Refer to the results at the end of Chapter 3 for the mode 1 natural vibration
frequencies of beams (pp 50 and 51), and second moments of area (pp 49 and 50).

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
The appropriate situation is Case 2.

EI bd 3
f = 0.560 ,I= , M = ρlbd .
Ml 3 12
 Ed 2 1/ 2  d  E 1/ 2
f = 0.560 4 = 0.1617 2   ,
 12 ρl   l  ρ 
 f 2 l 2 
2

E =    ρ.
 0.1617   d 

Because the frequency of natural vibration involves a force acting on a mass to give it
an acceleration, it is crucial when real numbers are put into the governing equation
that the basic SI units are used as follows

 440  2  0.085 2 
2

E =   × 7.85 × 10 = 204 × 10 N m = 204 GN m .


3 9 –2 –2
 0.1617   0.00386 

3.3. This is a consequence of the equations of static equilibrium.

3.4. Principal planes have no components of shear stress acting on them. Principal
directions are normal to principal planes. Principal stresses are normal stresses acting
on principal planes. The shear stress components all vanish.

 σ 1 0 0  σ1 0 0 –p 0 0  σ1 0 0 
3.5. (a)  0 0 0 , (b) 0 σ 0 , (c)  0 –p 0  , (d) 0 σ 0.
   2
    2

 0 0 0 0 0 0  0 0 – p  0 0 σ3

 0 τ 0
3.6.  τ 0 0 . No.
 
 0 0 0

 σ1 0 0
3.7.  0 0.5σ 0 . There are no shear stress components normal to these axes.
 1

0 0 0

3.8. Because the two shear strain terms on any given axis plane are defined so there is no
rotation.

3.9. Principal strains are axial strains. The shear strain components all vanish.

 ε1 0 0   ε 0 0
3.10. (a)  0 –υε1 0  . (b)  0 ε 0 .
   
0 0 –υε1   0 0 ε

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
3.11. Compare the right hand diagram on p 44 with the centre diagram in Fig. 3.5.

e13 = γ and e31 = 0.


1 γ
ε13 = ε 31 = (e13 + e31) = .
2 2

∆V
3.12. From the bottom diagram in Fig. 3.5, the dilatation is defined as ∆ = .
V

Consider a cube of material of unit side. For each of the three principal (axial) strains,
the increase in the volume of the cube is equal to the strain. For example, a strain ε1
makes the cube longer in the 1 direction by ε1. The increase in the volume of the cube
is ε1 x 1 x 1 = ε1. The dilatation produced by ε1 is therefore (ε1 x 1 x 1)/(1 x 1 x 1) =
ε1. Therefore,

∆= ε1 + ε 2 + ε 3 .

3.13. Take a cube of the material having its faces normal to the principal directions, and
σ1 σ1
apply a stress σ1 along the 1 direction. Since
= E = , ε1 . Now apply a stress σ2
ε1 E
σ
along the 2 direction. This will produce a strain along the 2 direction of ε 2 = 2 . In
E
σ2
=
turn, this strain will produce a strain along the 1 direction of ε1 –=
υε 2 –υ . The
E
σ1 σ2
net strain along the 1 direction now becomes ε1 = . Next apply a stress σ3
–υ
E E
along the 3 direction, and repeat as before to find the total net value of ε1, as given by
the first equation. Repeat this procedure to find the equations for ε2 and ε3.

Sum the three principal strains using the three equations. It is then straightforward to
show that

(1 – 2υ )
=∆ (σ 1 + σ 2 + σ 3 ).
E

The dilatation is zero when υ = 1/2.

(1 – 2υ )
3.14. =
For uniaxial tension, ∆ = σ 1 (1 – 2υ )ε1.
E

The volume changes are therefore 0.4ε1, ε1, and 0.

p
3.15. From Eqn (3.8), K = – . For hydrostatic pressure loading, the equation

(1 – 2υ ) (1 – 2υ )
=∆ (σ 1 + σ 2 + σ 3 ) becomes ∆ =– (3 p ). Then
E E

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
E E
K=
–p× =.
–(1 – 2υ )3 p 3(1 – 2υ )

This shows that K = E when υ = 1/3.

3.16. The solid rubber sole is very resistant to being compressed, because it is restrained
against lateral Poisson’s ratio expansion by being glued to the relatively stiff sole.
However, the moulded surface has a much lower resistance to being compressed,
because the lateral Poisson’s ratio expansion of each separate rubber cube can occur
without constraint (provided the gaps between adjacent cubes do not close-us
completely). So your colleague is correct.

3.17. The axial force applied to the cork to push it into the bottle results in a zero lateral
Poisson’s ratio expansion, so it does not become any harder to push the cork into the
neck of the bottle. However, the axial force applied to the rubber bung results in a
large lateral Poisson’s ratio expansion, which can make it almost impossible to force
the bung into the neck of the bottle.

4. 1. Refer to Fig. 4.11.


dU
 Force  F between atoms = .
dr

At the equilibrium distance, ro, the energy U is a minimum (i.e. F is zero, and U is the
“dissociation energy” Uo).

dU mA nB
= − = 0,
dr r m +1 r n +1
m
or B = ron − m A.
n
A 1 m
U o = − m + n ⋅ ron − m A
ro ro n
A  m
=− 1 −  .
rom n

Now, for ro = 0.3 nm, Uo = –4 eV.


5
A = 4 ⋅ (0.3) 2 = 0.45eV nm 2
4
= 7.2 × 10 −20 J nm 2 .

−5 −25 d 2U
B = (0.3) × 0.45 = 0.59 × 10 eV nm = 9.4 × 10
1
5
8 10 10
J nm . Max force is at =0.
dr 2
i.e. at value of r given by

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
1

m(m + 1) A n(n + 1) B  B n (n + 1)  n−m


− + = 0 which is r =  
r m+ 2 r n+ 2  A m (m + 1) 
1 1

 n + 1  n−m  11 8
=  ro =   × 0.3 = 0.352 nm.
 m + 1 3
dU mA  ron − m 
and Force = = 1 − 
dr r m +1  r n − m 
2 × 0.45  (0.3)8 
= 3 
1 − 8
= 14.9 eV nm −1
(0.352)  (0.352) 
14.9 × 1.602 × 10 −19
= −9
J m −1
10
= 2.39 × 10 −9 N.

4.2. The term –A/rm is an attractive potential which depends on the type of bonding. The
B/rn term is a repulsive potential due to charge-cloud overlap and diminished
screening of the nuclei (see Section 4.2).

4.3. The values of A are shown below. The mean is 88. The calculated values of the
moduli are

Material Calculated from 88 kTM/Ω Measured


Ice 1.0 × 1010 N m–2 7.7 × 109 N m–2
Diamond 9.0 × 1011 N m–2 1.0 × 1012 N m–2

The calculated values, for these extremes of elastic behaviour, are close to the
measured values. The important point is that the moduli are roughly propertional to
absolute melting temperatures.
A values: Ni, 98; Cu, 78; Ag, 76; Al, 89; Pb, 51; Fe, 96; V, 61; Cr, 116; Nb, 48; Mo,
138; Ta, 72; W, 127.

dU q2 nB
4.4. From eqn (4.3), = – n +1 .
dr 4πε 0 r 2
r
 dU  q2 nB
Setting   = 0 gives = n +1 .
 dr  r = r 4πε 0 r0 r0
2
0

q 2 r0n –1
Therefore, B = as required.
4π nε 0

 d 2U 
From eqn (4.8), S0 =  2  .
 dr  r = r
0

From eqn (4.3) and the result derived above for B,


q2 q 2 r0n –1 1
U (=
r) Ui – + × .
4πε 0 r 4π nε 0 r n

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
dU q2 q 2 r0n –1
= – .
dr 4πε 0 r 2 4πε 0 r n +1
d 2U 2q 2 q 2 (n + 1)r0n –1
= – + .
dr 2 4πε 0 r 3 4πε 0 r n + 2
2q 2 q 2 (n + 1)  q 2 
=
Therefore, S0 – + =  ( –2 + n + 1) .
4πε 0 r03 4πε 0 r03  4πε 0 r03 
(n –1)q 2 α q2
=
Finally, S0 = .
4πε 0 r03 4πε 0 r03

4.5. See Section 4.2, last paragraph.

4.6. See Fig. 4.10.

4.7. See Section 4.5, paragraphs 1 to 3.

5. 1. (a) Let the spheres have a diameter of 1. Then (referring to Fig. 5.3) the unit cell has
an edge length 2 , and a volume 2 2 . It contains 4 atoms, with a total volume
4π/6. Hence the density, ρ , is given by

ρ= = 0.740.
6.2 2

(b) Glassy nickel is less dense than crystalline nickel by the factor 0.636/0.740. The
density is therefore 8.90(0.636/0.740) = 7.65 Mg m −3 .

5.2. (a)

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
(b)

(c)

(d)

5.3. (a) If the atom diameter is d, then the lattice constant for the f.c.c. structure is

2d1
a1 = = 0.3524 nm.
2

(b) The weight of one atom is

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
Atomic wt. 58.71
= = 9.752 × 10 −26 kg.
Avogadro’s number 6.022 × 10 26

There are 4 atoms per unit cell,

4(9.752 × 10 −26 )
so density = −9 3
= 8.91Mg m −3 .
(0.3524 × 10 )

(c) If the atom diameter is d2, then the lattice constant for the b.c.c. structure is

2d 2
a2 = = 0.2866 nm .
3

(d) One atom weighs

55.85
= 9.274 × 10 −26 kg.
6.022 × 10 26

There are 2 atoms per unit cell.

2(9.274 × 10 −26 )
So density = −9 3
= 7.88 Mg m −3 .
(0.2866 × 10 )

5.4. (a) Copper

1 1
Have 4 atoms per unit cell (8 × from cube corners = 1 + 6 × from cube faces = 3)
8 2

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
Atoms touch along cube-face diagonal:

2
This gives atom radius r = 4 a.
4
4 × π r 3 × 100
3 16π 2 2a 3 × 100
Required percentage = = 3⋅
a3 3a 16 × 4
π 2
× 100 = 74%.=
6
Answer is same for magnesium — both are close-packed srructures.

(b) Copper
4×m
Density = where m is atomic mass.
a3
Atomic weight for Cu = 63.54.
63.54 kg
= m = 10.55 × 10 −26 kg.
6.023 × l0 26

4 × 10.55 × 10 −26
8.96 Mg m −3 = kg.
a3
a 3 = 4.71 × 10 −29 m3 .
a = 3.61 × 10 −10 m, or 0.361 nm.

Magnesium

1 1
Have 6 atoms per unit cell (12 × from corners = 2 + 2 × from end faces = 1 + 3
6 2
inside)

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
 3 
Volume of unit cell = 3 a 2 c .
 2 
2 × 6m 4m
Density = = .
2
3 3a c 3a 2 c
24.31 kg
m= = 4.04 × 10 −26 kg.
6.023 × l0 25

4 × 4.04 × 10−26 kg
1.74 Mg m −3 = .
3a 2 c
a 2 c = 5.36 × 10 −29 m3 .

In face-centred structure, plane spacing

3a 2a
= , where r = .
3 4
3 4r
Spacing = .
3 2

c 3 4r 3 4 a
In close-packed-hexagonal structure, plane spacing.= = = .
2 3 2 3 22

c
= 1.633,
a
c = 1.633a.

Using value for a2c of 5.36 × 10–29 m3 we obtain:

a = 3.20 × 10 −10 m or 0.320 nm;


c = 5.23 × 10 −10 m or 0.523 nm.

5.5. See Section 5.10.

5.6. Let V be the volume fraction of the polyethylene which is crystalline. Then,

(a) 1.014 V + 0.84(1 – V) = 0.92, giving V = 0.46, or 46%.


(b) 1.014 V + 0.84(1 – V) = 0.97, giving V = 0.75, or 75%.

6. 1. The two sets of values for the moduli are calculated from the formulae

Ecomposite = Vf Ef + (1 − Vf ) Em (upper values);


1
Ecomposite = (lower values);
Vf
Ef + 1E−Vmf

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
where Vf is the volume fraction of glass, Em the modulus of epoxy, and Ef that of
glass.

Values are given in the table and plotted in the figure. The data lie near the lower
level. This is because the approximation from which the lower values are derived (that
the stress is equal in glass and epoxy) is nearer reality than the approximation from
which the upper values are derived (that the strains are equal in the two components).

Note that the sets of values are widely separated near Vf = 0.5. Fibreglass tested
parallel to the fibres, or wood tested parallel to the grain, lie near the maximum
composite modulus. Both materials, tested at right angles to the fibres or grain lie near
the lower modulus. They are, therefore, very anisotropic: the ratio of the two moduli
can be as much as a factor of 4.5 for fibreglass (see the figure, at Vf = 0.5); it can be
more for woods.

Volume Ecomposite Ecomposite


fraction of Ecomposite (upper values) (lower values)
glass, Vf (GN m–2) (GN m–2) (GN m–2)
0 5.0 5.0 5.0
0.05 5.5 8.8 5.2
0.10 6.4 12.5 5.5
0.15 7.8 16.3 5.8
0.20 9.5 20.0 6.2
0.25 11.5 23.8 6.5
0.30 14.0 27.5 7.0

6.2. Ec = EfVf + (1 – Vf) Em.


ρc = ρfVf + (1 – Vf) ρm.

(a) ρc = 0.5 × 1.90 Mg m–3 + 0.5 × 1.15 Mg m–3 = 1.53 Mg m–3.


(b) ρc = 0.5 × 2.55 Mg m–3 + 0.5 × 1.15 Mg m–3 = 1.85 Mg m–3.
(c) ρc = 0.02 × 7.90 Mg m–3 + 0.98 × 2.40 Mg m–3 = 2.51 Mg m–3.

(a) Ec = 0.5 × 390 GN m–2 + 0.5 × 3 GN m–2 = 197 GN m–2.


(b) Ec = 0.5 × 72 GN m–2 + 0.5 × 3 GN m–2 = 37.5 GN m–2.
(c) Ec = 0.02 × 200 GN m–2 + 0.98 × 45 GN m–2 = 48.1 GN m–2.

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
6.3.
 V (1 − V1 ) 
−1

E⊥ =  1 +  ; see Section 6.4


 E1 E2 
E|| = V1 E1 + (1 − V1 )E2 ; see Section 6.4 .
 V (1 − V1 ) 
= {V1 E1 + (1 − V1 )E2 } 1 +
E\\

E⊥  E1 E2 
E E 
= 1 − 2V1 + 2V1 + (1 − V1 )V1  1 + 2 .
2

 E2 E1 
d  E\\   E1 E 2 
  = (1 − 2V1 )  + − 2.
dV1  E⊥   E 2 E1 
For E1 ≠ E2 ,
d  E\\ 
  = 0 when V1 = 0.5 (the only turning po int ).
dV  E⊥ 

6.4. to 6.8 Refer to Chapters 4, 5, 6 and the appropriate References.

7.1. By inspection from the results given on pp 48 and 49, the values of C are respectively:
Case 1 = 1/3; Case 2 = 1/8; Case 3 = 1/48; Case 4 = 5/384; Case 5 = 1/192; Case 6 =
1/384.

7.2. Refer to the results at the end of Chapter 3 for the elastic bending of beams (pp 48 and

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
49), and second moments of area (pp 49 and 50).

(a)
 Fl 3  F 1 EI 1 E a 4
=δ C =
, = ,
 EI  δ C l
3
C l 3 12
=M a=
2
l ρ , a 4 M 2 /l 2 ρ 2 ,
F 1 E 1 M2
= ,
δ C l 3 12 l 2 ρ 2
1/2 1/2
 ρ2   ρ2 
1/2
F
=M   (12 C )1/2 l 5/2   ∝  .
δ   E   E 

(b)
1 E bd 3
F
= ,
δ C l 3 12
= =
M bdl ρ, d 3 M 3 / b3l 3ρ 3 ,
F 1 E b M3
= ,
δ C l 3 12 b3l 3 ρ 3
1/3 1/3
 ρ3   ρ3 
1/3
 F
=M   1/3 2 2/3
(12 C ) l b  E  ∝  .
δ  E

(c)
1 E πr4
F
= ,
δ C l3 4
=M π=r 2 l ρ, r 4 M 2 / π 2 l 2 ρ 2 ,
F 1 E π M2
= ,
δ C l 3 4 π 2l 2 ρ 2
1/2 1/2
 ρ2   ρ2 
1/2
 F
=M   (4π C ) l
1/2 5/2
 E  ∝  .
δ  E

(d)
1 E 3 F
π r t,
=
δ C l3
=M 2=π rtl ρ, t M / 2π rl ρ,
F 1 E 3 M
= πr ,
δ C l3 2π rl ρ
 F l4  ρ   ρ 
=M   2C 2   ∝   .
δ r  E  E

7.3. By inspection from the results given on pp 52 and 53, the values of C are respectively:
Case 1 = 2.47; Case 2 = 9.87; Case 3 = 39.5.

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
7.4. Refer to the results at the end of Chapter 3 for the elastic buckling of struts (pp 52 and
53), and second moments of area (pp 49 and 50).

(a)
E E a4
=Fcr C= I , Fcr C 2 ,
l2 l 12
=M a= 2
l ρ, a 4 M 2 / l 2 ρ 2 ,
E 1 M2
Fcr = C ,
l 2 12 l 2 ρ 2
1/2 1/2
 ρ2   ρ2 
1/2
 12 
=M ( Fcr )   1/2 2
l   ∝  .
C  E  E

(b)
E bd 3
Fcr = C
,
l 2 12
= =
M bdl ρ, d 3 M 3 / b3l 3ρ 3 ,
E b M3
Fcr = C 2 ,
l 12 b3l 3 ρ 3
1/3 1/3
 ρ3   ρ3 
1/3
 12 
=M ( Fcr )   1/3 5/3 2/3
l b  E  ∝  .
C  E

(c)
E πr4
Fcr = C,
l2 4
=M π=
r 2 l ρ, r 4 M 2 / π 2l 2 ρ 2 ,
E π M2
Fcr = C 2 ,
l 4 π 2l 2 ρ 2
1/2 1/2
 4π   ρ2   ρ2 
1/2

=M ( Fcr )   1/2 2
l   ∝  .
C  E  E

(d)
E 3
π r t,
Fcr = C
l2
=M 2= π rtl ρ, t M / 2π rl ρ,
E M
Fcr = C 2 π r 3 ,
l 2π rl ρ
 2 l  ρ  ρ
3
=M ( Fcr )   2   ∝   .
 C r  E  E

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
7.5.
2
t
pb = 0.3E   .
 r
M b = 4π r 2t ρ.
2
t pb
  = .
r 0.3E
1/2
t  pb 
=  .
r  0.3E 
1/2
 p 
t= b  r.
 0.3E 
 ρ 
1/2
 p 
= πr ρ  b 
M b 4= 3
22.9r 3 pb1/2  1/2  .
 0.3E  E 
 E1/2 
Merit index is  .
 ρ 

7.6. The correct equation for the whirling speed is given in Case 4 on p 51, and is

 EI 1/ 2  E 1/ 2
f = 1.571 3  = 1.571 3 πr t  . 3
 Ml   Ml 
M
=M 2= π rtl ρ, t ,
2π rl ρ
1/2
 E M 
f = 1.571  3 π r 3 ,
 Ml 2π rl ρ 
1/2
1.571  r   E 
f =  2   ,
2 l   ρ
2 2  ρ
1/2

r= f l   ,
1.571  E 
2 2  ρ
1/2

= π rtl ρ 2π tl ρ f
M 2= l   ,
1.571  E 
1/2
2 2  ρ3 
M= π tl 3 f   .
1.571  E

When all the other parameters in this equation are fixed, the mass is proportional to
the wall thickness t. The earlier equation

1/2
1.571  r   E 
f =  2  
2 l   ρ

shows that r will also be fixed. The cross sectional area of the tube (2πrt ) will thus be

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
proportional to t. The minimum value of M corresponds to the value of t which gives
a cross sectional area just sufficient to carry the torque which the prop shaft is
required to transmit.

8.1. Nominal stress = Hardness/3 (1 + εn) where εn is the nominal strain as given in the
question plus 0.08.

Data for nominal stress–nominal strain curve

Nominal stress (MN m–2) 129 171 197 210 216 217 214 209 188
Nominal strain 0.09 0.18 0.28 0.38 0.48 0.58 0.68 0.78 1.08

8.2. (a) From graph, tensile strength is 217 MN m–2 (the stress maximum of the curve).
(b) The strain at the stress maximum is 0.6 approximately.
l
(c) From eqn. (8.4), Aolo = Al, and l = AAo .

(l − lo )/ lo = 0.6; A/ Ao = 1/1.6.
A
1− = 0.38; ( Ao − A)/ Ao = 0.38.
Ao

Thus percentage reduction in area = 38% .


(d) 109 MJ from graph.

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
8.3. During a tensile test, unstable necks develop when the maximum in the nominal stress–
nominal strain curve is reached. Neck growth then leads rapidly to failure. Physically,
necks become unstable when the material in the elongating neck work hardens
insufficiently to make up for the decrease in load-bearing capacity at the neck. In
rolling, the material is deforming mainly in compression, and the load-bearing area is
always on the increase. Tensile instabilities cannot therefore form, and failure occurs at
the much larger strains required to cause failure by cracking.

8.4.

(a) Tensile strength = 101.5 kN/160 mm2


= 634 MN m–2.
−2
Working stress = 160 MN m .

(b) Load corresponding to 0.1%


Proof strain = 35 kN.
∴ 0.1% Proof stress = 219 MN m–2.
−2
Working stress = 131 MN m .

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
8.5. The indentation hardness is defined by

F
H= ,
π a2

where a is the radius of the circle of indentation. Simple geometry (see figure) gives

( r − h) 2 + a 2 = r 2 ,

or

2rh − h 2 = a 2 ,

or

a2
h= if h << r
2r

F
Thus H = .
2π rh

The indenter penetrates a distance h which is proportional to the load.

8.6. (a) Conservation of energy gives mgh = Uel, so m = Uel /gh. Thus

(1000N × 15m + 0.5 × 1500N × 15m)


mmax = = 89.2 kg.
9.81 × 30

(b) Taking the maximum extension, in one cycle of loading/unloading, 500N × 15


m = 7500 J is dissipated by hysteresis in the rope, compared to a Uel of 26250 J.

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
Thus 29% of the energy input is lost. This results in the jumper rebounding to a
position which is well below the bridge deck (this is obviously essential for a
safe jump).

9.1. See Section 9.2.

9.2. The fractional volume change is

∆V
= (8.9323 − 8.9321) / 8.9323 = 2.24 × 10 −5.
V

(a) Define the dislocation density as ρ (m/m3 or m–2). Then

∆V 1 2
= b ρ,
V 4

from which
ρ = 1.4 × 1015 m −2 .

(b) The energy is:

1 2 ∆V 3
U= Gb ρ = 2 ⋅ E = 2.1 MJ m −3 .
2 V 8

9.3. to 9.7. Refer to Chapter 9.

10.1. (a) See Section 10.5.


(b) See Section 10.3.
(c) See Section 10.4.

10.2. (a) Balance line tension against force on dislocation (Section 10.4 and Fig. 10.2 (c)):
τ ybL = 2T .
2T Gb
So τ y = ≈ .
bL L

10.3. (a) For the alloy:


L = 5 ×10−8 m,
b = 2.86 ×10−10 m,
G = 26 ×109 Nm −2 .
Gb
So τ y = = 1.5 ×108 N m −2 .
L
But σ y = 3τ y for polycrystals. Hence
σ y ≈ 450 MN m −2 .

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
(b) New L = 15 × 10–8 m.

Repeating the calculation, σ y = 149 MN m −2 .


Drop in yield strength ∆σ y ≈ 300 MN m −2 .

10.4. d = 4 × 10–6 m, d –1/2 = 103 / 2 = 500 m–1/2. (120 – 20) MN m–2 = β500 m–1/2. β = 100
MN m–2/500 m–1/2 = 0.2 MN m–3/2.

10.5. See Section 10.7, paragraph 1 and Fig. 10.5. The ratio is 1.5 – see Section 10.7,
paragraph 2.

10.6. See Sections 10.3, 10.4, 10.5 and 10.7 (Hall-Petch effect).

11.1. Suppose first that the shaft yields. The stress in the shaft is F/π ·102. If this is equal to
the tensile yield stress σy, we have

F = 100πσ y = 200π k , since σ y = 2k ,

where k is the shear-yield strength.

Now consider shearing-off of the head, as shown. At yield

F = 2πrtk
= 180πk

Thus the head will shear off.

11.2. (a) Lubricated Anvils

Work balance gives, from upper-bound theorem,

wL u
Fu ≤ 8 × 2×k × 2 = 2 wLku.
4 2
F ≤ 2 wLk .

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
(b) Welded Interfaces

wL u
Fu ≤ 2 wLku + 4 × ×k ×
2 2
= 2 wLku + wLku = 3wLku.
F ≤ 3wLk .

w
General formula gives, for = 2,
d
F  1
≤ 2k 1 +  = 3k .
wL  2
F ≤ 3wLk , verification demonstrated.

 w
11.3. 4000 MN m–2 = 350 MN m–2. 1 +  .
 4d 
w
= 10.43.
4d
w
= 41.7.
d

w w
Take = 42 to produce an integral value of in a safe direction.
d 2d
w 20μm
d= = μm. = 0.48
42 42

(This gives a volume fraction of cobalt of 7.2% — a typical value for a rock-drilling
grade of WC-Co cement.)

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
11.4. (a) See Section 8.4.
(b) See Section 11.4.


= σ.

11.5. σ = 350ε0.4 MN m–2.



At onset of necking =σ .

350ε 0.4 ,
350 × 0.4ε 0.6 =
or ε = 0.4.
σ=
350 × (0.4)0.4 MN m −2 =
242.6 MN m −2 .

σ
Nominal stress, σ n = , where ε n is the nominal strain.
(1 + ε n )
ε = ln(1 + ε n ).

242.6
Tensile strength, σ TS = MN m −2
antiln (0.4)
242.6
= MN m −2
1.492
−2
= 163 MN m .


ε
Work = ∫ σ dε per unit volume
0
0.4
= ∫ 350ε 0.4 dε MN m −2
0
0.4
 ε 1.4 
= 350   MN m −2
 1.4  0
(0.4)1.4
= 350 MN m −2
1.4
= 69.3MJ per 1m3 .

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved

11.6. (a) Given σ = Aεn. From eqn (11.4), =σ at onset of necking. Thus


= Anε n −1 = Aε n , so ε = n and σ = Ann. From eqn (8.13),

σ An n
σ TS = = .
(1 + ε n ) (1 + ε n )

From eqn (8. 15), ε = ln (1 + εn) = n.


An n
Thus (1 + εn) = en. Finally, σ TS = n .
e
800(0.2)0.2
(b) Inserting A = 800 MN m and n = 0.2 gives σ TS
–2
= 0.2
= 475 MN m −2 .
e

σ = σ TS (1 + ε n ) = σ TSe n = 580 MN m −2 .

11.7. We have that σy at 8% plastic strain is H/3 or 200 MN m–2. Thus

200
A= = 334 MN m −2 .
(ln1.08)0.2

Using the result of Example 11.6(a),

An n
σ TS = n
= 198 MN m −2 .
e

12.1.
t
pf = 2σ f   .
 r
M f = 4π r 2t ρ.
t pf
  = .
r 2σ f
pf r
t= .
2σ f
4π r 3 pf ρ  ρ 
Mf = = 2π r 3 pf  σ  .
2σ f f

σ 
Merit index is  f  .
 ρ

12.2.

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
 ρ 
M b = 22.9r 3 pb1/2  1/2  .
E 
Mb
tb = .
4π r 2 ρ
 ρ
M f = 2π r 3 pf   .
 σf 
Mf
tf = .
4π r 2 ρ

Set r = 1 m and pb = pf = 200 MPa. Values for ρ, E and σf are to be taken from the
table of data.

Material Mb (tonne) tb (mm) Mf (tonne) tf (mm) Mechanism


Al2O3 2.02 41 0.98 20 Buckling
Glass 3.18 97 1.63 50 Buckling
Alloy steel 5.51 56 4.90 50 Buckling
Ti alloy 4.39 74 4.92 83 Yielding
Al alloy 3.30 97 6.79 200 Yielding

Optimum material is Al2O3 with a mass of 2.02 tonne. The wall thickness is 41 mm
and the limiting failure mechanism is external-pressure buckling.

12.3. When the bolts yield, the connection can be approximated as a mechanism which
hinges at X. The cross-sectional area of one bolt is π (1.25/2)2 in2 = 1.23 in2. The
yield load of one bolt is 1.23 in2 × 11 tsi = 13.5 tons. The moment at yield is given by

M ≈ (2 × 13.5 × 25) + (2 × 13.5 × 19) + (2 × 13.5 × 9) + (2 × 13.5 × 3)


= 1512 ton in = 126 ton ft.

The hinge must react the total load from the bolts, which is 108 tons. This means that
in practice the hinge will extend over a finite area of contact. X will lie in from the
outer edge of the flange by about 1 in to 2 in but the effect on the bending moment
will be small.

12.4. The yield load of each link plate in tension is given approximately by the minimum
cross-sectional area multiplied by the yield strength. The total breaking load of the
two links in parallel is double this figure and is given by

=T 2(1500 N mm −2 ×1 mm × 4.5 mm)


= 1.35 × 104 N.

From eqn (11.2), the shear yield strength k = σy/2. k for the pin is therefore 1500/2 =
750 MN m–2.

The yield load of the pin in double shear is obtained by multiplying k by twice the
cross-sectional area of the pin to give

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
  3.5 
2

T = 2 750N mm × π   mm 2  = 1.44 × 104 N.
−2

  2  

This load is 7% greater than the load needed to yield the links and the strength of the
chain is therefore given by the lower figure of 1.34 × 104 N.

To estimate the tension produced in the chain during use we take moments about the
centre of the chain wheel to give

190 mm
90 kgf × 170 mm ≈ T × ,
2
T ≈ 161 kgf ≈ 1.58 × 103 N.
The factor of safety is then given by

1.35 × 104 N
= 8.5.
1.58 × 103 N

Comments

(a) The factor of safety is calculated assuming static loading conditions. The
maximum loadings experienced in service might be twice as much due to
dynamic effects.

(b) The chain must also be designed against fatigue and this is probably why the
factor of safety is apparently so large.

12.5. The cross-sectional area of the pin is

A = π (22 − 1.22 ) mm 2 =
8.0 mm 2 .

The force needed to shear this area is

fs = kA = 750N mm −2 × 8.0 mm 2 = 6000 N.

Finally, the failure torque is

Γ = 2( fs × 10 mm) = 2 × 6000 N × 10 mm =
1.2 × 105 N mm
= 120 N m ≈ 12 kgf m.

pr 6.303 × 71.76 σ y 1244


12.6. σ
From Example 3.7, = = = 944 MN m –2 . =
S = = 1.32 .
t 0.479 σ 944

12.7.

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
pr
σ= ,
t
M
= π rtl ρ , t
M 2= ,
2π rl ρ
2π rl ρ
σ = pr ,
M
ρ  ρ   ρ 
M 2π pr 2l  =
=  S 2π pr l  =
2
 1.32 × 2π pr l   .
2

σ  σy  σy 
ρ /σ=
y = 6.3 ×10 –3
7.8 /1244

12.8. Because the applied stress is not unidirectional — it is equi-biaxial (see Eqn 12.10,
and Fig. 12.5). When the applied stress is unidirectional — i.e. when the only stress
acting is the hoop stress, and there is no axial stress. This stress state is satisfied when
the cylinder has no end caps — the opposite of the conditions shown in Example 3.7.
A good example is a long pressurised water main, with the axial load removed by
fixing the pipe to reinforced concrete thrust blocks. In such an application, CFRP is
clearly the winner, resulting in a pressure vessel approximately three times lighter
than one made from alloy steel or aluminium alloy. However, cost is still a major
issue, and may well rule out CFRP for other than special applications (e.g. aerospace).

12.9. Refer to the results at the end of Chapter 3 for the elastic bending of beams (p 47),
and second moments of area (pp 49 and 50). Refer to the results at the end of Chapter
11 for plastic moments (p 169).

(a)
M yc d 12 6 M y
σy
= = My = ,
I 2 bd 3 bd 2
σ ybd 2
My = .
6
σ ybd 2
Mp = .
4
Mp 6
= = 1.5.
My 4

(b)

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
M yc 4 4M y
σy
= = M yr = ,
I π r4 π r3
πσ y r 3
My = .
4
4σ y r 3
Mp = .
3
Mp 4 4
= × = 1.70.
My 3 π

(c)
M yc 1 My
σy
= = M yr = ,
I π r t π r 2t
3

M y = πσ y r 2t.
M p = 4σ y r 2t.
Mp 4
= = 1.27.
My π

12.10. See the section “Plastic buckling” on p 170. Once the metal yields, the slope of the
stress/strain curve decreases markedly. Because the buckling load is determined by
the slope of the stress/strain curve (Young’s modulus below yield, and tangent
modulus above) it decreases markedly as well.

13.1. The maximum tensile stress at the surface of a beam loaded in three-point bending
(see eqn (12.3)) is

3Fl
σ=
2bt 2

at the mid-span of the beam. Fracture occurs when

Yσ π a = Kc ,

i.e. when

Y 3Fl
π a = Kc .
2bt 2

Hence, the maximum load which can be sustained by the adhesive joint is

2bt 2 K c
F= .
Y 3l π a

For the joint shown

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
2 × (0.1)3 × 0.5
=F = 4.65 kN.
0.64 × 3 × 2 π × 0.001

13.2. Calculate the stress for failure by (a) general yield and (b) fast fracture.

(a) σ = 500 MN m–2 for general yield.


Kc 40
(b) =σ = = 285 MN m –2 ,
Y π a 1.12 π × 0.005

for fast fracture, assuming that a crack on the limit of detection is present. The plate
will fail by fast fracture before it fails by general yield.

pr 7000 1
13.3. σ= = 0.06 × × = 70 MN m −2 . K c = Y σ π a , Y = 1 .
t 2 3

2 2
1K  1  100 
a=  c = 
π  a  π  70 
= 0.65 m.

13.4.
Kc = Yσ π a ,
Y = 1,
Kc = σ π a ,
2
1  30 MPa m 
2
1  Kc 
a=   =   = 0.080 m = 80 mm.
π  σ  π  60 MPa 
2a = 160 mm.

13.5. (3 × 1.0 + 4) = 7.0 MN m–2. K = Yσ πa = 1.2 × 7.0 π × 0.010 = 1.49 MN m–3/2 .

This is 15% more than the value of Kc obtained from tests. Experimental scatter in the
test data, and errors in the stress analysis are more than enough to account for this
small discrepancy.

13.6. The total stress on the crack is 285 + 100 = 385 MN m–2. Fast fracture occurs when
the crack length ac is given by Kc = Yσ πa c . For a semicircular surface crack in a
thick plate (a = 6 mm << 62 mm), Y = 1.12 x 0.64 = 0.72 (see Case 10 on p 203).
1  Kc  1 2
2
40
Therefore, a c =   =   = 0.0066 m = 6.6 mm.
π  Yσ  π  0.72 × 385 

This is of the same order as the crack depths found in practice, verifying the fracture
mechanics analysis of the failure.

14.1. (a) See Sections 14.2 and 14.3. (b) See Section 14.4.

14.2. Classic features of tensile failure by microvoid coalescence. See Fig. 14.2.

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
14.3. See Fig. 14.3 and Section 14.3, paragraph 2. Atomically flat cleavage planes can be
seen. Many fracture facets have “river markings”, produced by fracture on multiple
parallel cleavage planes.

14.4. The mode of failure is shown in the diagram and the photograph.

A horizontal outward force applied to The crack path in a failed end post
the top rail tends to split the end post.

The collapse of the railing was caused by the propagation of pre-existing cracks in the
end posts, which are safety-critical components. The collapse was caused by a lack of
awareness by both architect and maintenance personnel of the poor along-grain
fracture toughness of wood, the potential sources of crack-opening forces in the
design, the propensity of externally exposed timber to develop cracks and the
significance of cracks for the integrity of the structure.

You could recommend fixing each end of the top rail directly to the brick wall, using
a steel bracket bolted to both the top rail and the brick wall. This designs-out the end
posts and their problems entirely.
14.5. The low fracture toughness of wood along the grain allows wood to be split very
easily along the grain. This permits easy splitting of logs into kindling and wood for
fires, production of wood shingles for roofing material, finishing/sizing by planing,
shaping by routing, turning and chiselling, and even pencil sharpening.

15.1.
From eqn (15.1)

Kc 3MP a m1/2
σ TS = = = 276 MPa.
1.12 π a 1.12 π × 30 ×10−6 m

15.2. From eqn (15.6)

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
6M r F l 1
σr = 2
= 6× × × 2
bd 2 2 bd
330 N 50 mm 1
= 6× × × 3
2 2 5 mm3
N
= 198 = 198 MPa.
mm 2

V π (11/ 2) 2 mm 2 × 50mm
15.3 = = 9.7 .
Vo π (5 / 2) 2 mm 2 × 25mm

For the test specimens, eqn (15.4) gives

 V  σ  m 
0.5 = exp − o    .
V σ
 o  0  

For the components, eqn (15.4) gives

 V  σ ′  m 
0.99 = exp −    .
 Vo  σ 0  

Thus

ln(0.5)  Vo  σ    Vo  σ 0  
m m

= −    −    .
ln(0.99)  Vo  σ 0    V  σ ′  
 
1 σ 
m

69.0 =   .
9.7  σ ′ 
1
(σ ′) m = (σ ) m .
669
σ ′ = 0.272 ×120 MPa =
32.6 MPa.

15.4. See Answer to this Example.

15.5. Specimen measuring 100 mm × 10 mm × 10 mm will have median σTS of 300


Mpa/1.73.

Specimen volume V = 104 mm3. Eqn (15.4) then gives

 10 4 mm3 10 
0.5 = exp − σ .
 Vo σ o
10 TS

Taking natural logs gives

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
10 4 mm3
−0.69 = − σ TS10 .
Vo σ o10

Component volume = 1.25 × 103 mm3. Pf = 10–6 gives Ps = 1 – 10–6. Thus

 1.25 ×103 mm3 10 


1 − 10−6 = exp − σ .
 V o σ o
10

Taking logs, and assuming that ln (1 – x) = – x for small x, gives

1.25 ×103 mm3 10


−10−6 = − σ .
Voσ o10

Thus

0.69 104 mm3 Vo σ o10 1


= σ 10
×
−6
Vo σ o 1.25 ×10 mm σ 10
10 TS 3 3
10
104  σ TS 
10

=   .
1.25 ×103  σ 
104 10−6
σ 10 = × σ TS10 .
1.25 ×103 0.69
300 MPa
σ = 3.21×10−1σ TS = 3.21×10−1 ×
1.73
= 55.7 MPa

15.6. (a) Weight of material below section at x is

π
ρ g × (α x) 2 x.
3

Cross-sectional area = π (ax)=


2
=
. Stress force/area 1
3 ρ gx .

(b) Integrate over the volume, using disc of thickness dx with volume dV = π(αx)2 dx.
Eqn (15.5) then gives

  σ m dV   L  ρ gx m π (α x) 2 dx 
Ps ( L) = exp  ∫ −    = exp  ∫ −   .
σ σ
V  0  0   0  0 
V 3 V 0 

Integrating gives

  ρ g m πα 2 Lm +3 
Ps ( L) = exp −   .
  3σ 0  (m + 3)V0 

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
The probability of survival falls with increasing α because, although the stresses are the
same, the amount of material which is stressed increases with α, and hence the chances
of meeting a critical flaw increase.

16.1. From Section 13.3, K = Y σ π a .

(a) From Fig. 16.3, a / W = 5 /10 = 0.5, Y = 3. K = 3 ×100 π × 0 005 = 37.6 MN m −3/2

(b) From Section 13.3, Y = 1.


(c)
K = 1 × 100 π × 0.020 = 25.1 MN m −3/2 .

16.2. See Section 16.3 (“Failure analysis” and “Conclusions”). PMMA is a poor choice of
material because it has a very low fracture toughness. Under a tensile hoop stress, the
connector is liable to suffer catastrophic fast fracture from a small defect.

16.3. Reinforce the foam with polymer fibres. These will bridge any incipient cracks, and
prevent crack propagation. Layers of fibre mesh can be incorporated into the foam as it
is sprayed on.

16.4. From Section 16.3 (p 235), the maximum (hoop) stress occurs at the bore of the
cylinder, and is given by

A2 + B 2 2792 + 127 2
σ=
p 2 =
p ==
1.52 p 1.52 × 83 MN m –2 =
126 MN m –2 .
A – B2 2792 –127 2

Fracture will commence at the bore of the cylinder when the hoop stress reaches the
tensile strength of the material, i.e. the tensile strength = 126 MN m–2.

Grey cast iron is too brittle a material to use for making a pressure vessel. The fracture
toughness is typically 10 MN m–3/2 (see Table 13.1 on p 195). The critical defect size is
given by the equation Kc ≈ σ πa (see p 235), i.e.
2 2
1  Kc  1  10 
= = = 0.002 =
π  σ  π  126 
 
a m 2.0 mm . This is far too small a critical defect

size — a low-tech casting process cannot achieve this degree of control over casting
defects.

17.1. Answers obtained by reading directly off the graph in the diagram.

17.2. Each time the iron was moved backwards and forwards the flex would have expe
rienced a cycle of bending where it emerged from the polymer sheath. The sheath is
intended to be fairly flexible to avoid concentrating the bending in one place. Possibly
the sheath was not sufficiently flexible and the flex suffered a significant bending stress
at the location of failure. The number of cycles of bending is well into the range for
high-cycle fatigue and fatigue is the likely cause. The scenario is that the individual
strands in the live conductor broke one by one until the current became too much for the

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
remaining strands to carry. At this stage the last strands would have acted as a fuse and
melted, causing the fire. If 23 strands are rated to carry 13 A, then a single strand should
carry about 0.57 A safely. The iron draws 4.8 A, which is 8.4 times the safe capacity of
one strand. It is therefore not surprising that, when only a few wires were left intact, the
flex was no longer able to take the current without overheating. Failures of this sort
have also occurred with appliances such as vacuum cleaners. However, these tend to
have a smaller current rating and failure does not always result in a fire.

17.3. Fast fracture occurs when the (final) crack length a2 is given by K c = Y σ max π a2 ,
where σmax is the maximum tensile stress in the loading cycle. For a long surface crack
in a thick plate, Y = 1.12 (see Case 3 on p 199). Therefore,
54 =1.12 ×180 π a2 , a2 =
0.0228 m .

The initial crack length a1 is 2 mm, which we need to write as 0.002 m in the crack
growth equation in order to get the units right.

We integrate the crack growth equation over the range of crack lengths from 0.002 m to
0.0228 m to find the number of cycles. This is done as follows

da
= A(∆K )3 = A(Y ∆σ π a )3 = A(Y ∆σ π )3 a 3/2 ,
dN
a2
da
∫ a=
a1
3/2
A(Y ∆σ π )3 ∫ dN ,

 1 
a2
 1 1 
= 2  – =  A(Y ∆σ π ) N ,
3
–2 
 a  a1  a1 a2 
 1 1 
2 – = 31.48 = 6 × 10 –12 (1.12 × 180 π )3 N ,
 0.002 0.0228 
N=
1.15 ×105 , rounded up to 1.2 ×105.

17.4. (a) Circular = 0730 hrs position. Rectangular = 0600 hrs position.
(b) Circular = 1300 hrs position. Rectangular = 1200 hrs position.

The maximus stress is greater in the rectangular tool than in the circular tool.

17.5. ∆σ(Nf)a = C.

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
280(105 ) a = 200(107 ) a .
a
280  107 
= 1.4 =  5  = (102 ) a .
200  10 
log 1.4 = a 2 = 0.146.
 1 
a = 0.073  or . .
 13.7 
C = 280(105 )0.073 MN m −2 .
C = 649 MN m −2 .
13.7
C1/ a
−2  649 
At 150 MN m , N f = =  = 5.2 ×108 cycles.
∆σ 1/ a
 150 

17.6. The 1 µm marker at the bottom LHS of the photograph measures 3.0 mm in length.
Take a length of 20 mm, and count the number of striation spacings over this length on
various parts of the photograph. The number of spacings is typically about 14. One
spacing therefore corresponds to about 20 mm divided by 14, i.e. about 1.43 mm. This
corresponds to a real length of about 1.43/3.0 µm, i.e. about 0.48 µm, which should be
rounded up to 0.5 µm.

17.7. The maximum thermal shock that could be experienced by the boiler plate next to the
feed water inlet pipe is 180 – 10 = 170°C. This produces a thermal strain in the plate
surface of α∆T = 12 × 10–6 × 170 = 2.04 × 10–3. The strain required to give yield is
σ y 250 MN m –2
ε=y = = 1.25 ×10 –3 . This is only 61% of the maximum thermal strain.
E 200 GN m –2
Since the thermal shock is cyclic (it occurs every time water is put into the boiler) it is
likely that the cracks formed and grew by low-cycle fatigue, with ≤ 104 cycles required
for failure (see Fig. 17.4). Over a three-year period, we would only have to put water
into the boiler ten times a day to achieve this number of cycles. In practice the number
of cycles could have been much more. The baffle would have made the cold water
sweep sideways over the inner surface of the boiler plate, ensuring efficient cooling.
The problem would have been much less serious if a baffle had not been fitted.

18.1. (a) A good surface finish will increase the fatigue life by increasing the time required
for fatigue-crack initiation.

(b) A rivet hole will cause a local stress concentration which will increase ∆σ and
reduce the fatigue life.
(c) A mean tensile stress will decrease Nf (see eqns (17.3) and (17.4)).
(d) Corrosion may reduce the fatigue life, by creating pits in the surface from which
fatigue cracks can initiate more easily.

18.2. The maximum pressure in the cylinder occurs at the point of admission. The maximum
force acting on the piston is therefore given by

0.7 N mm −2 × π (45 mm) 2 =


4453 N.

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
The stress in the connecting rod next to the joint is

4453 N / (28 mm ×11 mm) = 14.5 N mm −2 .

Since the locomotive is double-acting the stress range is twice this value, or 29 N mm–2.

The number of revolutions that the driving wheel is likely to make in 20 years is (20 ×
6000 × 1000m) / (π × 0.235m) = 1.6 × 108.

Data for the fatigue strengths of welded joints are given in Fig. 18.4. The type of weld
specified is a Class C. The design curve shows that this should be safe up to a stress
range of 33 N mm–2, which is slightly more than the figure calculated above. Of course,
our calculations have ignored dynamic effects due to the reciprocating masses of the
piston and connecting rod and these should be investigated as well before taking the
design modification any further.

18.3. Data for the fatigue strengths of welds are given in Fig. 18.4. The weld is a surface
detail on the stressed plate and the weld classification is Class F2. We extrapolate the
curve following the dashed line for Class F2 until we hit the stress range of 8 N mm–2.

The mean-line fatigue curve gives the data for a 50% chance of cracking. For the stress
range of 8 N mm–2 the cycles to failure are 3 × 109. The time to failure is (3 × 109) / (20
× 60 × 60 × 12 × 6 × 52) = 11 years.

The design curve gives the data for a 2.3% chance of cracking. The number of cycles to
failure is 109 and the time to failure is therefore 4 years.

18.4.
(a) Because the shoulder combined with the zero fillet radius produced a stress
concentration (see the bottom diagram in Fig. 18.1).
(b) At the top of the 4 mm diameter section – the position of maximum tensile
bending stress.
(c) Eliminate the shoulder, by making the arm from rod having a uniform diameter of
2 mm, and fixing it into the roller with an adhesive like Loctite (this has the
counter-intuitive consequence that removing material actually increases the
fatigue strength).
(d) Titanium alloy (see Table 18.1). This has a maximum fatigue strength of 600 MN
m–2 compared to a maximum fatigue strength for aluminium alloy of 170 MN m–2,
yet is only about 50% heavier (see Table 5.1).

4 Fl 4 × 9.81 × 1 × 57
σ= = = 90 N mm–2 = 90 MN m–2 .
πc 3
π × 23

The stress amplitude is therefore 45 MN m–2 and the mean stress is 45 MN m–2
(tensile).

Refer to Eqn (17.4),

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
∆σσm ∆σ(σ'f – σm ) ∆σ  σm 
= = 1 – ,
2 2σ'f 2  σ'f 
∆σσm ∆σ  45  ∆σ
= 1 –  = 0.82 ,
2 2  250  2
∆σ 1 ∆σσm 1
= = × 45 = 55 MN m–2 .
2 0.82 2 0.82

18.5.
∆F ∆F
∆σplate = , ∆σthroat =.
Lt 2 gL
∆F ∆F 2a
For ∆σplate = ∆σthroat , = , and t = 2 g = = 2 a,
Lt 2 gL 2
t
= 2.
a

Refer to the weld fatigue curves of Fig. 18.4 (50% survival). At 107 cycles, ∆σ for class
W = 32 MN m–2, and ∆σ for class F2 = 49 MN m–2. For the same fatigue lives in the
“T” joint, we therefore need

∆σplate 49 ∆F 2 gL 2 g 2a
= = × = = ,
∆σthroat 32 Lt ∆F t t
t 32
= 2× = 0.92.
a 49

Note that the same result would be obtained for any given number of cycles, because
the fatigue curves are plotted on log-log scales, and the lines for the two weld classes
are parallel to one another, i.e. the ratios of the fatigue strengths are independent of the
number of cycles.

18.6. Because the hole produced a stress concentration. At the side of the hole (see the top
diagram in Fig. 18.1). In uniaxial loading, a hole produces a stress concentration factor
of 3 at the sides of the hole (see the top diagram in Fig. 18.1). Therefore the margin of
safety would have been approximately 3.

18.7. See Section 18.7 for how to design bolted connections so the bolts see only a small
fraction of the applied loading cycle. The bolts should be made from high tensile steel,
and should be “waisted”. This increases their “give” under tension, and also eliminates
unnecessary threads. For maximum fatigue resistance, the threads should be rolled, and
the fillet between the shank and the head should be roller peened (see Section 18.6,
paragraphs 1 and 2). Contact between the bolt head and the flange should only take
place on the flat part of the head, and not on the under-head fillet. This may require the
hole in the flange to be chamfered back to clear the fillet.

In spite of the above measures, the load shielding of the bolts is limited by their short
length, which is defined by the combined thickness of the flanges. The design
modification involves making the bolts much longer, so they have more “give”, and see
less of the applied loading cycle. This can be done by inserting a length of thick-walled

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
tube between the underside of the nut and the outer face of the flange.

18.8. The fracture occurred at the zero fillet radius at the shoulder, i.e. at the bottom of the
run-out groove at the lower end of the thread. Low-cycle fatigue, since <<104 cycles
(see Fig. 17.4). From the given equation,

π × 540 ×12.53
=
M elastic = 8.28 × 105 Nmm.
4

The applied bending moment is 9.81 x 1800 x 60 = 10.6 x 105 Nmm. This is more than
sufficient to make the steel yield at the crack initiation site.

Do away with the lifting eye altogether, and extend the top of the pulley block to
provide a horizontal hole to take the pin of a shackle. Because the rotational degree of
freedom provided by the lifting eye assembly has now been lost, it will be necessary to
insert an in-line swivel-link (a standard item) between the shackle and the crane boom.

19.1. (a) Because the end of the crack was subiected to a large tensile stress every time the
cyclist pushed the pedal down. There were clearly enough cycles of stress of
sufficient amplitude to make fatigue cracks initiate and grow to final failure.

(b) There are two fatigue cracks, one on either side of the hole. They are smooth and
dark in appearance, and are located in the lower half of each fracture surface. The
dark appearance is caused by a compact layer of iron oxide produced by slow
long-term corrosion, indicating that the cracks had been present for a long time. In
addition, the smooth appearance of the crack surfaces is consistent with high-cycle
fatigue.

(c) The final fracture surfaces are grainy and bright in appearance, and are located in
the upper half of each fracture surface. They consist of fresh, un-corroded metal,
indicating that the final fracture was not exposed to a corrosive environment after
failure. In addition, the grainy appearance of the crack surfaces is consistent with
a single overload failure.

(d) The RHS crack probably formed first, because it is larger than the LHS crack. It
probably initiated at the surface of the hole, because the local stress would have
been larger than the average stress over the whole cross-section. The LHS crack
appears to have initiated at the 0630 hrs position, since the crack appears to
radiate from this position.

(e) Moderate, since the fatigue cracks had spread across 50% of the cross section on
average before they reached the critical size for fast fracture.

19.2. (a) In the reduced section of the pivot pin, at the lower end of the reduced section.

(b) The horizontal force from the top of the door pulls the bottom of the pivot pin to
the right. As a result, the pin is made to rotate around the point at which it touches
the bottom of the frame housing. This rotation pushes the top of the pin to the left,
against the left hand wall of the hole in the frame housing. The reaction to this
force at the top of the pin generates a bending stress in the reduced section of the

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
pin. The maximum value of this bending stress occurs at the lower end of the
reduced section. In addition, the sharp change of section at this location inrroduces
a large SCFeff (see Sections 18.3 and 18.4).

(c) The bending stress at the failure location cycled from tension to compression
every time the door swung from one extreme position to the other (e.g. from fully
open inwards to fully open outwards).

(d) When the fracture took place at the lower end of the reduced section, the length of
pivot pin below the fracture fell down into the hole in the door housing, and came
clear of the bottom of the frame housing.

19.3. See Fig. 17.6, Fig. 17.8 and Section 17.4, paragraph 1.

19.4. The cracks were located in the welds between the trunnion shaft and the gussets on the
driving side. This is because the gussets are narrower on the driving side than the idling
side (100 mm compared to 180 mm), so the stress in the welds is greater on the driving
side than the idling side. The exact geometry of the cracking is given in Example 18.5,
where it is shown as either a class W or a class F2. In practice, the cracking is most
likely to be a class W, since “T” joints rarely have a leg length sufficiently large for
class F2 cracking to occur.

19.5. The effective fatigue stress amplitude at each fillet is given by the actual stress
amplitude multiplied by the effective stress concentration factor for fatigue loading (see
Section 18.4). Since failure did not occur in the 90 mm section after 10 million cycles
(and will not occur even if the number of cycles goes above 10 million), we can
determine the fillet radius required to avoid failure at the 81 mm section by equating the
effective fatigue stress amplitudes for the two fillets as follows

∆σ 90 ∆σ
× SCFeff,90 = 81 × SCFeff,81 .
2 2
∆σ 90
Then, SCF= × SCFeff,90 .
∆σ 81
eff,81

∆σ 90  81 
3

From the equation given for the maximum bending stress, =   = 0.729 .
∆σ 81  90 
Eqn (18.1) gives SCFeff = S (SCF – 1) + 1.
The look-up table can be used to find values of S, and SCF corresponding to the two
steps in diameter, for a range of fillet radii. For the 2.6 mm fillet at the 180/90 step
(diameter ratio = 2), S = 0.81, and SCF = 2.64 (linear interpolation). This gives SCFeff,90
= 2.33. Then, SCFeff,81 = 0.729 × 2.33 = 1.70.

We now need to use the look-up table to reverse engineer the fillet radius at the 90/81
step (diameter ratio = 1.11) which gives a SCFeff of 1.70. As a first trial, a fillet radius
of 4.5 mm gives S = 0.85 and SCF = 1.83. SCFeff is then found to be 1.71, so the
required radius is indeed 4.5 mm.

If the actual fillet radius has to be limited to 2.5 mm, the situation is not as bad as it
might appear, because the 2.6 mm fillet had obvious lathe tool markings. As long as the
surface at the 2.5 mm fillet does not show lathe tool markings, it will have a better

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
fatigue strength, other things being equal, than the 2.6 mm fillet. The relative fatigue
strength at the 2.5 mm fillet can be further improved by polishing the machined surface
(see Section 18.2, paragraph 1) and/or subjecting it to roller peening in order to
introduce a compressive residual stress (see Section 18.6, paragraph 2).

20.1.
Temperature (°C) T (K) 1/T (K–1) ε (s −1 ) ln ε
–7
618 891 0.00112 1.0 ×10 –16.12
640 913 0.00110 1.7 ×10−7 –15.59
660 933 0.00107 4.3 ×10−7 –14.66
−7
683 956 0.00105 7.7 ×10 –14.08
707 980 0.00102 2.0 ×10−6 –13.12
–10
510 783 0.00128 8.3 ×10 From graph –20.90

The hoop stress in the tube is

pr 6 MN m −2 × 2 0mm
σ= =
t 2 mm
= 60 MN m −2 .
at 510° C, ε =
8.3 ×10−10 s −1 at σ =
200 MN m −2
5
 60  −1
Under 60 MN m , 2
ε = 8.3 × 10  −10
 s . .
 200 
= 2.0 × 10−12 s −1.
Strain in 9 years = 2.0 ×10−12 × 60 × 60 × 24 × 365 × 9
= 5.7 ×10−4 or 0.0057.

Design safe

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
20.2. At 25 MN m–2 and 620°C, ε = 3.1 × 10−12 s −1 .
5
 30 
–2
At 30 MN m and 620 C, o
ε =   × 3.1×10−12 s −1
 25 
= 7 ⋅ 71 × 10−12 s −1.
ε = Aσ 5e − Q / RT .
Q1 1 
ln ε1 − ln ε2 = −  −  .
R  T1 T2 
Q1 1 
ln ε2 =  −  + ln ε1 ,
R  T1 T2 
160 × 103  1 1 
=  − 
8.313  893 923 
+ ln(7 ⋅ 71×10−12 s −1 )
= 0.700 − 25.58 = −24.88.

ε2 = 15.5 ×10−12 s −1 at 30 MN m −2


and 650°C.
70
ε = 3.1×10−12 s −1 × + 15.5
100
30
× 10−12 s −1 ×
100
−12 −1
= 6.82 ×10 s .

20.3. From Table 20.1, the softening temperature of soda glass is in the range 700–900 K.
The operating temperature of window glass is rarely more than 293 K, so T/Ts is at most
0.42. Ceramics only begin to creep when T/TM > 0.4. (see Section 20.1), and then only
under a large stress (far greater than the self weight stress). Thus the flow marks cannot
possibly be due to creep. In fact, the flow marks come from the rather crude high-
temperature processes used to manufacture panes of glass in the past.

20.4. See Section 20.3.


20.5. See Section 20.4 and Fig. 20.10.
20.6. A major fire would increase the temperature of the steelwork to the point at which it
would creep under the applied loads, and the subsequent deformation could trigger the
collapse of the building. This is why the World Trade Center towers collapsed on 9/11.

20.7. The linear elastic stress analysis ignores the fact that the material will deform by plastic
deformation and creep, which will keep the actual stress well below the tensile strength.
For this reason, the correct failure criterion is the one based on strain, not stress.

M
21. 1. Measure the rate by the mass injected per second, . Then, if this rate follows an
t
Arhennius Law, we have

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
M  Q 
= A exp − .
t  RT 
Or, combining the constants M and A:
1  Q 
= B exp − .
t  RT 
Then, converting temperatures to kelvin:
1  Q  −1
= B exp − s ,
30  R 450 

1  Q  −1
= B exp − s .
81.5  R 430 
Solving for Q and B, using R = 8.31J mo 1−1 K −1 , we have
Q = 8.04 ×104 J mo1−1 ,
B = 7.25 ×107 s −1.
We may now calculate the time required to inject the mass M of polymer at
227 o C(500K) :
1  8.04 ×104  −1
= 7.25 ×10 exp −
7
s ,
t  8.313 × 500 
giving t = 3.5 seconds.

21.2. See Sections 21.2 and 21.4 (“Fast diffusion paths: grain boundary and dislocation core
diffusion”).

21.3. See Section 21.4. More specifically,


(a) Carbon forms an interstitial solid solution in iron at room temperature containing up
to 0.007% by weight carbon. Even at this maximum concen tration there is a large
proportion of alternative interstitial sites remaining unfilled by carbon. The
probability of a carbon atom being next to an alternative interstitial site is therefore
high, and hence the probability of the carbon atom moving into a different position
is also high, i.e. it will diffuse relatively rapidly. Chromium forms a random
substitutional solid solution in iron at room temperature (that it should not form an
interstitial solution is evident from its large atomic radius, comparable to that of
iron). The probability of a vacancy appearing next to a chromium atom is therefore
small, and the diffusion of chromium is correspondingly slow.
(b) Diffusion in grain boundaries is generally more rapid than in the grains themselves
because of the geometrically more open structure of grain boundaries. A small grain
size produces a larger contribution from grain boundary diffusion than does a large
grain size, and thus increases the overall diffusion coefficient of the polycrystal.

21.4. See Section 21.4 (“A useful approximation”).


x = Dt , so t = x 2 / D .
 −159 ×103 J mo1−1 
D = Do e − Q / RT = 9.5mm 2s −1 exp  −1 −1 
 8.313J K mo1 1023 K 
9.5mm 2s −1
= = 7.20 ×10−8 mm 2s −1.
1.32 ×10 8

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
x2 (10 −2 mm) 2
t= =
D 7.2 ×10−8 mm 2s −1

10−4 mm 2 108
= × s
7.2 mm 2

10 4
= s = 0.1389 ×10 4 s
7.2

= 1389 s, or 23 min.

22.1. See Fig. 22.2.


22.2. See Fig. 22.3.
22.3. See eqn (22.1).
22.4. See Fig. 22.4 and eqn (22.2).

22.5. The rate of bulk vacancy diffusion decreases rapidly with decreasing temperature
because the activation energy Q is large for bulk diffusion (see Table 21.1). At a
sufficiently low temperature, short-circuit diffusion (in this case along dislocation cores,
see Fig. 21.10) takes over from bulk vacancy diffusion. Because the activation energy Q
is small for short-circuit diffusion, the rate of short-circuit diffusion decreases only
slowly with decreasing temperature, so creep continues to occur at temperatures well
below the transition from bulk to short-circuit diffusion.

22.6. For reasons analogous to those in Example 22.5. However, in the present case, the
short-circuit diffusion takes place along grain boundaries (see Fig. 21.9).

22.7. See Section 22.2 (“Designing metals and ceramics to resist power-law creep” and
“Designing metals and ceramics to resist diffusional flow”).

22.8. See Section 22.3.

23.1. See Section 23.1.

23.2. See Section 23.3, paragraph 4, and Figs 23.4 and 23.5.

23.3. See Section 23.5.

23.4. See Section 23.4.

23.5. See Section 23.4, next to last paragraph.

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
24. 1.

dc
Fick’s first law J = − D .
dx
δ (∆m) = AJ δ t where A = constant.
 dc  c −c 
δ (∆m) = − A  D  δ t = AD  1 2  δ t.
 dx   b 
bδ (∆m) = AD(c1 − c2 )δ t.
b = B∆m where B = constant.
B∆mδ (∆m) = AD(c1 − c2 )δ t.
B
(∆m) 2 = AD(c1 − c2 )t ,
2
i.e., (∆m) 2 = CDt , where C = constant.
At constant temperature (∆m) 2 = kP t.

24.2. Ohm’s Law V = IR .


δ (∆m) = PI δ t where P = constant.
PV δ t
δ (∆m) = PI δ t = .
R
Rδ (∆m) = PV δ t.
R = Qb where Q = constant
= S ∆m where S = constant.
S ∆mδ (∆m) = PV δ t.
S
(∆m) 2 = PVt , i.e., (∆m) 2 = kP t.
2

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
24.3.
 138 ×103 J mo1−1  2 −4 −1
AT 500o C, kP = 37 exp − −1 −1  kg m s
 8.313J K mo1 773K 
= 1.74 ×10−8 kg 2 m −4s −1
(∆m) 2 = kpt
= 1.74 ×10−8 kg 2 m −4s −1 × 3600 × 24 × 365 s
= 0.50 kg 2 m −4 .
∆m = 0.74 kg m −2

i.e., each square metre of metal surface absorbs 0.74 kg of oxygen from the
0.74 kg m −2
atmosphere in the form of FeO. Number of oxygen atoms absorbed = ,
16 / N A
where NA is Avogadro’s number.
0.74 kg m −2
∴ Number of iron atoms removed from metal as FeO = .
16 / N A
0.74 kg m −2
Weight of iron removed from metal = × 55.9 / N A = 2.59kg m–2.
16 / N A
2.59 kg m −2 m3
Thickness of meral lost =
7870 kg
= 0.33 mm.
at 600o C, kP = 2.04 ×10−7 kg 2 m −4s −1.
(∆m) 2 = 2.04 ×10−7 × 3600 × 24 × 365kg 2 m −4 ,
= 6.43kg 2 m −4 .
∆m = 2.54 kg m −2 ,
giving loss = 1.13 mm.

24.4. As shown in Table 24.1, gold is the only metal which requires energy to make it react
with oxygen (80 kJ mol–1 of O2 in fact). It therefore remains as un-reacted metal.

24.5. If there is any contact resistance across a pair of silver contacts, the surfaces of the
contacts will be heated up by the current passing through the contact resistance. If the
temperature goes above 230°C, any oxide will decompose to leave pure metal-to-metal
contact. Gold contacts would not form an oxide film at any temperature, but silver is
used because it is much cheaper than gold.

24.6. See Section 24.3, paragraphs 1 to 3.

24.7. See Section 24.3, paragraph 4, and Fig. 24.2.

25.1. See Section 25.3, next-to-last paragraph, and Fig. 25.2.

25.2. See Section 25.2, paragraphs 1 and 2. The protective oxide film is Cr2 O3 , produced by

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
the chromium content of the stainless steel.

25.3. See Sections 25.2 and 25.3. Examples are Cr in stainless steel ( Cr2 O3 film), Cr in nickel
alloys ( Cr2 O3 film), Al in aluminium bronzes ( A12 O3 film).

25.4. See Table 24.2. The refractory metals oxidize very rapidly in air at high temperature.
Lamp filaments are surrounded by a glass bulb, which is either evacuated or filled with
an inert gas to remove any oxygen which would attack the filament.

25.5. The oxide layer prevents the molten solder or brazing alloy from wetting the surfaces to
be joined. When the copper connection tabs are soldered to pre-tinned copper wire, the
pre-tinning solder melts and this protects the copper surfaces from oxidation.

25.6. The chromium and nickel form a protective oxide layer on the wire, which resists
oxidation of the alloy at the high running temperature. Mild steel would oxidize far too
rapidly at the running temperature, and would burn out in a matter of days (see Table
24.2).

25.7. Many are oxides already, e.g. MgO,SiO 2 , A12 O3 (see Table 24.1). Others, e.g. SiC and
Si3N4 form protective oxide layers of SiO 2 when exposed to oxygen at high
temperature.

da
26.1. ∝ a at constant σ (4 MN m–2).
dt
da
∝ σ 2 at constant a(0.25 mm).
dt
da
∴ = Aπσ 2 a = AK 2 ; n = 2.
dt
 da  1 0.3mm year −1
πA=  2 =
 dt  σ a 16(MN m ) 0.25mm
−2 2

A = 0.0239 mm year −1 (MPa mm ) −2 .

26.2 Since water and air were in contact with the surface of the pipe, the cathodic oxygen-
reduction reaction would have taken place easily. The temperature of the pipe would
have varied from approximately 20o C (summer time, heating off) to 70o C (winter
time, heating on). The corrosion rate at 70o C will be approximately twice that at 20 o C.
Thus putting the heating on will double the rate of external corrosion. The pipe did not
rust from the inside because there is little or no oxygen inside the heating water circuit.

26.3. Pitting attack. See Section 26.7 and Fig. 26.12.

26.4. Stress corrosion cracking can lead to complete fracture even though the surface of the
component appears free from corrosion. See Section 26.7 and Fig. 26.12. Even if the
stress corrosion cracks do not travel right across the component, they can still propagate
to failure by fatigue or fast fracture.

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
26.5. Typical applications are as follows.
(a) Plastic water pipes in plumbing and drainage systems.
(b) Plastic gas pipes for underground use.
(c) Rubber hoses in automobile cooling systems, flexible brake hoses, windscreen
washer hoses.
(d) Plastic containers for storing water, acids, alkalis, etc.
(e) Plastics for exterior architectural use, e.g. window frames, roofing sheets, rainwater
gutters and downpipes.
(f) Plastics for marine use, e.g. boat hulls (matrix of GFRP composite), mooring
buoys, small fittings.
(g) Plastics and rubbers in domestic appliances, e.g. water pumps, washing machine
drums, flexible hoses, seals.

26.6. Stress corrosion cracking. See Section 26.7 and Fig. 26.12. Austenitic stainless steels
are prone to SCC in hot chloride solutions. The solution in the present case was very hot
and contained chloride ions from the zinc chloride corrosion inhibitor. There was also a
large tensile stress to drive the initiation and growth of SCC cracks.

26.7.
(a) See Section 26.6 and Table 26.1. The standard electrode potentials for these three
metals are all greater than the SEP of the oxygen reduction reaction.
(b) Because of the protective effect of some oxide films – See Section 26.4, last
paragraph.
(c) See Section 26.5, next to last paragraph, and Fig. 26.10. The surface oxide film is
only stable between these two pH values.
(d) See Section 26.5, paragraph 3, and Fig. 26.7. The surface oxide film is only stable
between these two pH values.

26.8.
(a) See Section 26.4, first paragraph – the three bullet points taken in that order define
the terms (i), (ii) and (iii).
(b) See Section 26.5, paragraph 3.
(c) See Section 26.5, last two paragraphs, and Fig. 26.11.
(d) Because it eliminates the most common cathodic (electron accepting) reaction – the
oxygen reduction reaction.

27. 1. See Section 26.2 and Fig. 26.2. Because there was no oxygen in the system, the
oxygen-reduction reaction could not take place. Therefore the anodic reaction could not
take place and the steel was protected from corrosion.

27.2. See Section 27.1


27.3. See Section 27.1 and Fig. 27.4.
27.4. Corrosion of zinc:
Zn → Zn ++ + 2e.
Number of electrons to give a current of 6 ×10−3 A m–2 for 5 years
6 ×10−3 × 5 × 3.15 ×107
= −19
= 5.89 ×1024 elecrrons
1.6 ×10

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
5.89
×1024 atoms Zn.
=
2
5.89 ×1024 × 65.4
∴ Mass zinc = = 0.320 kg m −2
2 × 6.02 ×10 26

0.320
Thickness = m = 4.5 ×10−5 m
7130
= 0.045 mm.
27.5. Following 27.4, if steel were lost uniformly over a square metre at 2 ×10−3 A m −2 ,
thickness lost by reaction:
Fe → Fe ++ + 2e
would be 0.0116 mm on each side of the plate.
If this loss is concenrrated over 0.5% of the surface, the loss there will be
100 99.5
0.0116 × × mm = 2.31 mm.
0.5 100
The sheet will thus rust through.

27.6. See Section 27.4, and Fig. 27.12.

27.7. The following table gives the mass of iron per unit area of corrosion product for each
flake.

Sample number Mass of iron (mg mm–2)


1 1.195
2 0.632
3 0.629
4 0.761
5 0.540
6 0.983
7 0.670
8 0.864

The average value is 0.784 mg mm–2.

The mill scale that covers an area of 1 mm2 has a volume of 1 mm2 x 0.10 mm = 0.1
mm3. The mass of Fe2O3 in this volume is 5.26 g cm–3 x 0.1 mm3 = 0.526 mg. Given
that the atomic weight of iron is 55.85, the mass of iron in the volume of scale is

 2 × 55.85 
  × 0.526 mg=0.368 mg.
 2 × 55.85 + 3 × 16 

The mass of iron taken up by the corrosion process is therefore 0.784 – 0.368 =
0.416 mg. The mass of iron is then divided by the density of iron (7.78 g cm–3) to give
the volume of iron removed by corrosion = 0.053 mm3. The surface area over which
the volume of iron is lost is 1 mm2, so the thickness of the steel lost by corrosion is
0.053 mm, which is rounded down to 0.05 mm.

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
27.8. At temperatures above 70°C, austenitic stainless steel is susceptible to stress corrosion
cracking (SCC) in the presence of tensile stress and chloride ions. Failures have been
reported even with low levels of stress and chloride ion concentration. In the present
case, the tensile stress was provided by the residual stresses from the welding process.
In addition, large concentrations of chloride ion were found. The chloride ions
originated from the sea water and were carried under the cladding by leaking rain water.
There is little doubt that chloride-induced SCC must have been the cause of the failure.
See Section 26.7, last paragraph, and Fig. 26.12. See also Example 26.6.

The nitric acid, because it is a powerful oxidising agent, actually protects the stainless
steel by artificially thickening the protective oxide film. See Section 27.2, paragraph 4.

28.1. See Section 28.2.

28.2. If P is the radial pressure that the olive exerts on the outside of the pipe then we can
write
Pr
σY =
t
provided we neglect the srrengthening effect of the sections of pipe that lie outside the
olive. If we assume that the end of the pipe far away from the fitting has an end cap (or
a bend that functions as an end cap) then the force trying to push the pipe out of the
fitting is pwπ r 2 . This force is balanced by the frictional force between the olive and the
pipe so we can write
pwπ r 2 = µ P 2π rl.
Combining the two equations to eliminate P gives
 t  l 
pw = 2 µσ y    .
 r  r 
Using the data given we get
 0.65   7.5 
pw = 2 × 0.15 ×120 MPa    = 3.1 MPa.
 7.5   7.5 
The hydrostatic head of water in a seventy-floor building is about 2 bar, so the joint
could actually cope with pumping water to the top of a seventy-floor skyscraper and
still have a factor of safety of 1.5. However, the pressure in water systems frequently
exceeds the static head, often substantially, because of “water hammer”. This is the
dynamic overpressure that arises when taps are suddenly shut off.

28.3. (a) Typical examples are as follows. Car tyres/road surfaces, brake pads/brake discs,
clutches, shoe soles/walking surfaces, climbing shoes/rock faces, knots in ropes,
V-belt drives, interference fits, compression joints (see Example 28.2).
(b) Typical examples are as follows. Bearings and sliding surfaces in machinery,
sledges and skis on snow, actuating mechanisms (e.g. car window mechanisms),
door latches, ceramic discs in water taps, clock and watch mechanisms.

28.4. (a) Typical examples are as follows. Metal finishing by linishing, grinding and
polishing. Wood finishing by sanding. Removal of surface scale by grit blasting.
Bedding-in of brake pads, clutch linings and plain bearings.

(b) Typical examples are as follows. Bearings and sliding surfaces in machinery, car

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
tyres, brake pads and discs, clutch linings, shoe soles, tools for metalworking and
woodworking, grinding wheels and abrasive belts/discs.

28.5. Let δ be the depth of metal lost from the surface of each contacting/loadbearing part
(pin or hole). Then it can easily be shown (draw a sketch of one pin, plus the two
links it passes through) that the elongation or “stretch” per 12.5 mm repeat unit of the
chain is equal to 4δ.

The total stretch is 4δ x 110 = 440δ = 25 mm.

Therefore, δ = 25 mm/440 = 0.057, or 0.06 mm rounded up.

28.6. The seal is only perfectly gas-tight or liquid-tight when there is atom-to-atom contact
between the surfaces over the whole of the nominal area of contact. In this case, eqn
(28.3) is modified to Py = Aσy , where A is the nominal area of contact, and Py – the
normal force – is also the yield load.

The yield load of the bolt is π × 7.5 2 mm2 × 280 N mm–2 = 49,480 N .
The yield load of the washer is π (112 – 7.5 2 )mm2 × 50 N mm–2 = 10,150 N.

The required ratio is therefore 10,150/49,480 = 0.21.

When the gasket is compressed, it work hardens. The next time it is used, it will not
yield when the bolt is torqued up. It is heated to dull red heat to anneal it - to remove
the work hardening, and restore the yield strength to its original value.

28.7. Also see Example 24.5. In contrast to the situation for silver, with copper, brass – and
in fact all metals other than silver and gold – the oxide film is stable when the
contacts overheat. In fact, if the contacts overheat, the oxide films will grow, and the
contact resistance will increase. This can lead to a run-away situation, ultimately
causing connection burn-out. The answer is to do the screws up tightly, to increase
the actual area of contact a between the surfaces, and make sure that the contacts can
transmit the current without heating up to a temperature at which the oxide films can
grow significantly.

28.8. Adhesive wear between the bore of the wheel and the surface of the axle. Because the
interference of fit was reduced by the additional pressing-on and pressing-off
operations.

28.9. Use a shrink-fit assembly process. The machined interference is 0.25 mm at a


workshop temperature of, say, 20°C. Heat the compressor wheel up to, say, 140°C (a
temperature rise of 120°C). Then the expansion of the bore of the wheel is
110 mm × 22.5 × 10 –6 × 120 = 0.297 mm . This is greater than the machined
interference by approximately 0.05 mm (2 thou), so the bore of the wheel will clear
the shaft. Position the shaft vertically, with the compressor wheel seat uppermost. It
should then be possible to drop the compressor wheel over the shaft (it will stop when
it comes up against the step at the end of the seat). It is important that the wheel does
not lose any heat to the shaft during this process, otherwise it will contract and “grab”
the shaft during assembly - not a good idea! Hence the need to position the wheel

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
above the shaft so it can be allowed to fall down under gravity to the correct final
position.

28.10. The cycle tyres slip over the galvanised steel bars when they are wet because the
contact areas between the rubber and the zinc are partially separated by a thin film of
water. The contact areas are effectively lubricated by boundary lubrication (see
Section 28.4, paragraph 3). This should have been foreseen. Ironically, if the steel
had not been galvanised, but had been allowed to rust, the roughness of the rust layer
might have helped to prevent sliding (but note that heavy traffic might have worn
away the rust layer, and polished the surface of the steel).

The solution which was adopted is shown in the photograph.

The steel bars were all cut away from the frame, and new bars having a thread
machined along their length were welded into the frame (the threaded bars were also
galvanised to stop them rusting). The crests of the threads indent the rubber of the
tyres (elastically), and stop them sliding sideways. More information on why this
works is given in Chapter 29, Section 29.4.

29.1. See Sections 28.4 and 29.2, paragraph 1.

29.2. From the slope of the roof, the coefficient of static friction is:
µs = tan 24o = 0.45.
If the slope of the roof is greater than 24° the static frictional force is exceeded and
the snow will slide off. On a 2° slope, with a ski already moving, it is the sliding
friction which counts:
µk = tan 2o = 0.035.
29.3. The work done = 2(100gµk ) = 68J .
This melts a volume of water
69
V= = 2.1×10−7 m3 .
330 ×10 6

If spread uniformly over the undersurface of two skis this would give a film of
thickness
2.1×10−7
m = 0.5µ m.
2 × 2 × 0.1
29.4. See Section 29.4.

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved
29.5. Secondary roads are often covered in ice, because it is uneconomical to treat them
with rock salt. As shown in Example 29.3, as soon as the tyre starts to slide on the ice,
it will melt the surface of the ice, reducing the friction coefficient to zero. The hard
studs bite into the ice, so the (non-zero) friction coefficient is determined by the
resistance of the ice to ploughing (see Fig. 28.9).

29.6. See Section 29.2, paragraphs 1 and 2.

29.7. See Section 29.2, paragraphs 4 to 9.

29.8. When only one interface has sticking friction,


wL u wLku
Fu ≤ 2 wLku + 2 × ×k× = 2 wLku + = 2.5 wLku ,
2 2 2
F ≤ 2.5wLk .

This result was derived for the case of w/d = 2 (see Example 11.2). Inserting this
 w
value of w/d into the general formula, F ≤ 2 wLk1+  , gives F ≤ 2.5 wLk as
 8d 
required.

For d = 0.5 mm and w = 40 mm, w/8d = 10. Then


F ≤ 2wLk(1+10) = 22wLk .

The yield load of the unconstrained sample is obtained by setting w/8d = 0, giving
F ≤ 2wLk . The ratio of the constrained to unconstrained yield loads is then
22wLk/2wLk = 11.

29.9. The main features of the damage are that large thin “pancakes” of white metal have
come away from the bearing. Extensive cracks have formed parallel to the bearing
surface, at a relatively small distance below the surface. Other cracks have initiated at
the bearing surface and propagated into the white metal at right angles to the surface.
These cracks have allowed the detachment of the white metal pancakes.

The likely mechanism of cracking is fatigue (see Section 29.2, last paragraph). This
could have been caused by misalignment of the propeller shaft (the vessel had been
involved in a grounding incident, which had bent the blades of the propeller).

29.10. When the rails are damp, the contact areas between the wheels and the rail are
partially separated by a thin film of water (which probably also contains oily residues
dropped from locomotives which have passed over the track before). The contact
areas are effectively lubricated by boundary lubrication (see Section 28.4, paragraph
3).

Sand particles are small and hard, and tend to “bite” through the surface film. Any
slipping between wheels and rail will be resisted by the “ploughing” action of the
sand over the metal surfaces (see Section 28.4, paragraphs 8 to 10, and Fig. 28.9).
Sand continues to be used on modern diesel locomotives to gain traction on slippery
rails when pulling heavy trains.

Copyright ©2012, Michael F. Ashby and David R.H. Jones. Published by Elsevier Ltd. All rights reserved

You might also like