You are on page 1of 29

9 Mechanical Reliability

9.1 INTRODUCTION
Usually, the reliability of electronic parts is predicted on the assumption that their
failure times are exponentially distributed (i.e., their failure rates are constant). The
past experience also generally supports this assumption. Nonetheless, this assump-
tion was derived from the bathtub hazard rate concept which states that during the
useful life of many engineering items, the failure rate remains constant. However,
in the case of mechanical items, the assumption of constant failure rate is not
generally true. In fact, in many instances their increasing failure rate patterns can
be represented by an exponential function.
The history of mechanical reliability may be traced back to World War II with
the development of V1 and V2 rockets by the Germans. In 1951, Weibull published
a statistical function to represent material strength and life length [1]. Today it is
known as the Weibull distribution and it has played an important role in the devel-
opment of mechanical reliability and reliability in general. Also, in the 1950s
Freudenthal reported many advances to structural reliability [2-4].
In 1963 and 1964, the National Aeronautics and Space Administration (NASA)
lost SYNCOM I and Mariner III due to busting of a high pressure tank and mechan-
ical failure, respectively. As a result of these and other mechanical failures, research-
ers in the field called for the improvement in reliability and longevity of mechanical
and electromechanical parts. In years to come, NASA spent millions of dollars to
test, replace, and redesign various items, including mechanical valves, filters, actu-
ators, pressure gauges, and pressure switches [5].
Some of the major projects initiated by NASA in 1965 were entitled “Designing
Specified Reliability Levels into Mechanical Components with Time-Dependent
Stress and Strength Distributions”, “Reliability Demonstration Using Overstress
Testing”, and “Reliability of Structures and Components Subjected to Random
Dynamic Loading” [6]. After the mid-1960s, many other people [7-9] contributed
to mechanical reliability and a comprehensive list of publications on the subject is
given in Reference 10.
This chapter presents various different aspects of mechanical reliability.

9.2 REASONS FOR THE DISCIPLINE OF MECHANICAL


RELIABILITY AND MECHANICAL FAILURE MODES
There have been many reasons for the development of the mechanical reliability
field. Five such reasons are presented in Figure 9.1 [11]. They are: influence from
electronic reliability, stringent requirements and severe application environments,
lack of design experience, cost and time constraints, and optimization of resources.

© 1999 by CRC Press LLC


FIGURE 9.1 Basic reasons for the development of the mechanical reliability field.

Influence from electronic reliability for the stimulation of similar developments


in mechanical engineering is an important factor because of vast advances made in
predicting and analyzing reliability of electronic equipment. Stringent requirements
and severe application environments are other important reasons because of many
expensive systems to be used in military and space applications. Cost and time
constraints make it difficult for designers to learn from past mistakes. Lack of design
experience occurs because of rapid changes in technology; consequently, mechanical
designers do not have the time to master the design, in particular design of complex
equipment for use in aerospace or military environment. Optimization of resources
is another good reason as the workable design is no longer considered satisfactory.
Thus, a given design has to be optimized under such constraints as reliability,
performance, weight, size, and cost.
There are many different types of failure modes associated with mechanical
items. Good design practices help to reduce or eliminate the occurrence of these
failure modes. Nonetheless, these failure modes include [12-14]:

• Bending failure. This may also be called a combined failure because it


occurs when one outer surface is in tension and the other outer surface is
in compression. The tensile rupture of the outer material is a representative
of the bending failure.
• Metallurgical failure. This may also be called a material failure and is
the result of extreme oxidation or operation in corrosive environment. The
environmental conditions such as heat, nuclear radiation, erosion, and
corrosive media accelerate the occurrence of metallurgical failures.

© 1999 by CRC Press LLC


• Tensile-yield-strength failure. This failure occurs under pure tension and,
more specifically, when the applied stress exceeds the yield strength of
the material. Usually, its consequence is permanent set or permanent
deformation in the structure and seldom catastrophic. The recommended
value of safety factors for this kind of failure could be as low as 1 (i.e.,
design to yield) or a high of 4 or 5 for pressure vessels and bridges.
• Instability failure. This type of failure occurs in structural members such
as columns and beams, particularly the ones manufactured using thin
material where the loading is usually in compression. However, instability
failure may also result because of torsion or by combined loading, i.e.,
including bending and compression. Usually this type of failure is crip-
pling or leads to a total failure of the structure. The past experience
indicates that a safety factor value of 2 is usually sufficient but avoid using
less than 1.5.
• Ultimate tensile-strength failure. This leads to a complete failure of the
structure at a cross-sectional point and it occurs (i.e., ultimate tensile-
strength failure) when the applied stress is greater than the ultimate tensile
strength. The recommended values of safety factors for this kind of failure
could be as low as unity (i.e., design to fail) or a high of 5 or 6 for pressure
vessels and bridges.
• Material flaw failure. This type of failure usually occurs due to factors
such as poor quality assurance, fatigue cracks, weld defects, and small
cracks and flaws. These factors reduce the allowable material strength
and, as a consequence, lead to premature failure at the flawed spots.
• Compressive failure. This failure is similar to the tensile failures dis-
cussed earlier but with one exception, i.e., under compressive loads. The
compressive failure causes permanent deformation/cracking/rupturing.
• Fatigue failure. This type of failure occurs because of repeated loading
or unloading (or partial unloading) of an item. In order to prevent the
occurrence of this kind of failure, make sure the proper materials are
selected, e.g., steel outlasts aluminum under cyclic loading.
• Bearing failure. This type of failure is similar in nature to the compressive
failure and it usually occurs due to a round cylindrical surface bearing on
either a flat or a concave surface like roller bearings in a race. If repeated
loading occurs, it is important to consider the possibility of the occurrence
of fatigue failures.
• Stress concentration failure. This type of failure occurs in circumstances
of uneven stress “flow” through a mechanical design. Usually, the con-
centration of stress occurs at sudden transitions from thick gauges to thin
gauges, at sudden variations in loading along a structure, at various attach-
ment conditions, or at right-angle joints.
• Shear loading failure. The occurrence of ultimate and yield failure takes
place when the shear stress becomes greater than the strength of the
material when applying high torsion or shear loads. Normally, these fail-
ures occur along a 45° axis in relation to the principal axis.

© 1999 by CRC Press LLC


FIGURE 9.2 Seven possible causes of failure.

• Creep/rupture failure. Normally, long-term loads make elastic materials


stretch even when they are less than the normal yield strength of the
material. Nonetheless, material stretches (creeps) if the load is maintained
continuously and usually it ultimately terminates in a rupture. It is to be
noted that at elevated temperatures, creep accelerates.

9.3 GENERAL AND GEAR FAILURE CAUSES


There are many causes of failure. Seven of them are shown in Figure 9.2 [15].
A study conducted over a period of 35 years reported 931 gear failures and their
major classifications were as follows [15]:

• Breakage: 61.2%
• Surface fatigue: 20.3%
• Wear: 13.2%
• Plastic flow: 5.3%

Furthermore, the failure causes were grouped into five distinct categories: (1) service
related (74.7%), (2) heat-treatment related (16.2%), (3) design related (6.9%),
(4) manufacturing related (1.4%), and (5) material related (0.8%). In turn, each of
these five categories were further broken into elements as follows:

Service-Related (74.7%)

• Continual overloading: 25%


• Improper assembly: 21.2%
• Impact loading: 13.9%
• Incorrect lubrication: 11%
• Foreign material: 1.4%
• Abusive handling: 1.2%

© 1999 by CRC Press LLC


• Bearing failure: 0.7%
• Operator errors: 0.3%

Heat-Treatment Related (16.2%)

• Incorrect hardening: 5.9%


• Inadequate case depth: 4.8%
• Inadequate core hardness: 2%
• Excessive case depth: 1.8%
• Improper tempering: 1%
• Excessive core hardness: 0.5%
• Distortion: 0.2%

Design-Related (6.9%)

• Wrong design: 2.8%


• Specification of suitable heat treatment: 2.5%
• Incorrect material selection: 1.6%

Manufacturing-Related (1.4%)

• Grinding burns: 0.7%


• Tool marks or notches: 0.7%

Material-Related (0.8%)

• Steel defects: 0.5%


• Mixed steel or incorrect composition: 0.2%
• Forging defects: 0.1%

9.4 SAFETY FACTOR AND SAFETY MARGIN


Safety factor and safety margin are used to ensure the reliability of mechanical items
and, basically, are arbitrary multipliers. These indices can provide satisfactory design
if they are established with care using considerable past experiences. Nonetheless,
in today’s environment as new designs involve new applications and new materials,
more consistent approaches are required. Furthermore, the mechanical design totally
based upon safety factors could be misleading and may be costly due to over-design
or catastrophic because of under-design. Both safety factor and safety margin are
described below.

9.4.1 SAFETY FACTOR


There are many different ways of defining a safety factor [9, 10, 16–19]. Two dif-
ferent definitions of the safety factor are presented below.

© 1999 by CRC Press LLC


Definition I

The safety factor is defined by [20]:

MFGST
SF = ≥1 (9.1)
MFGS

where

SF is the safety factor.


MFGST is the mean failure governing strength.
MFGS is the mean failure governing stress.

This safety factor is a good measure when both the stress and strength are
described by a normal distribution. However, it is important to note that when the
variation of stress and/or strength is large, this measure of safety becomes mean-
ingless because of positive failure rate.

Definition II

The safety factor is defined by [12, 13]:

SM
SF = (9.2)
MAWS

where

SM is the strength of the material.


MAWS is the maximum allowable working stress.

The value of this safety factor is always greater than unity. In fact, its value less
than unity indicates that the item under consideration will fail because the applied
stress is greater than the strength.
The standard deviation of the safety factor is expressed by [12, 13]:

[ { ]
σ = (σ th MAWS) + SM (MAWS) σs2 }
2 12
2
(9.3)

where

σ is the standard deviation of the safety factor.


σth is the standard deviation of the strength.
σs is the standard deviation of the stress.

All in all, there are many factors that must be carefully considered in selecting an
appropriate value of the safety factor: uncertainty of load, cost, failure consequence,

© 1999 by CRC Press LLC


uncertainty of material strength, degree of uncertainty involved in relating applied
stress to strength, etc. [21].

9.4.2 SAFETY MARGIN


This is defined as follows [9]:

θ m = SF − 1 (9.4)

where

θm is the safety margin.


SF is the safety factor.

The value of this measure is always greater than zero and its negative value indicates
that the item under consideration will fail. For normally distributed stress and
strength, the safety margin is expressed by [9]:

θ m = (µ th − µ ms ) σ th (9.5)

where

µ th is the average strength.


µ ms is the maximum stress.
σth is the standard deviation of the strength.

The maximum stress, µ ms , is given by

µ ms = µ s + k σs (9.6)

where

µ s is the mean value of the stress.


σs is the standard deviation of the stress.
k is a factor which takes values between 3 and 6.

It is to be noted that just like the safety factor, the safety margin is a random variable.

Example 9.1

Assume that the following data are known for a mechanical item under design:

µ th = 30,000 psi, µ s = 15,000 psi


σth = 1200 psi, σs = 300 psi

For k = 3 and 6, calculate the values of the safety margin.

© 1999 by CRC Press LLC


Substituting the above data into Equation (9.6) yields

For k = 3

µ ms = 15, 000 + 3 (300)

= 15, 900 psi

For k = 6

µ ms = 15, 000 + 6 (300)

= 16, 800 psi

By inserting the above calculated values and the given data into Equation (9.5),
we get

For k = 3

30, 000 − 15, 900


σm =
1200
= 11.75

For k = 6

30, 000 − 16, 800


σm = = 11
1200

Thus, the values of the safety margin for k = 3 and 6 are 11.75 and 11, respectively.

9.5 “DESIGN BY RELIABILITY” METHODOLOGY


AND STRESS-STRENGTH MODELS
In order to design a reliable mechanical item, a systematic series of steps are required.
The “design by reliability” methodology is made up of such steps [8, 22]:

• Define the design problem to be solved.


• Highlight and list all the relevant design variables and parameters.
• Conduct failure modes, effect and criticality analysis (FMECA) as per
MIL-STD-1629 [23].
• Verify selection of critical design parameters.
• Develop relationship between the critical parameters and the failure gov-
erning criteria.

© 1999 by CRC Press LLC


• Determine the failure-governing stress and strength functions and then
the failure-governing stress and strength distributions.
• Calculate the reliability using the failure-governing stress and strength
distributions for each critical failure mode.
• Iterate the design until the set reliability goals are satisfied.
• Optimize design with respect to factors such as reliability, maintainability,
cost, safety, performance, volume, and weight.
• Repeat the design optimization for all associated critical parts.
• Calculate item reliability.
• Iterate the design until the reliability goals of the item are effectively
satisfied.

Before we discuss stress-strength models, let us define stress and strength:

• Stress. This is any factor that tends to produce a failure. Some examples
are mechanical load, electric current, environment, and temperature.
• Strength. This is the ability of an item to accomplish its identified goal
adequately without a failure when subject to the external loading and
environment.

When the probability distributions of both stress and strength of an item are known,
the item’s reliability may be determined through analytical means. Under such
circumstances, the reliability of an item may be defined as the probability that the
failure governing stress will not exceed the failure governing strength. Mathemati-
cally, it is expressed as follows:

R = P (x > y) = P ( y < x) = P (x − y > 0) (9.7)

where

R is the item reliability.


P is the probability.
y is the stress random variable.
x is the strength random variable.

Using Reference 10, we rewrite Equation (9.7) as follows:


 ∞ 
R=

−∞

f ( y)  f (x) d x  d y
 y 
(9.8)

where

f (y) is the probability density function of the stress.


f (x) is the probability density function of the strength.

© 1999 by CRC Press LLC


Equation (9.8) can also be written in different forms and three such forms are as
follows:

 x 
R=

−∞

f (x)  f ( y) d y  d x
 −∞ 
(9.9)


 y

R=

−∞
 −∞ ∫
f (y) 1 − f (x) d x  d y

(9.10)



 
R=
∫ f (x) 1 − ∫ f (y) d y d x
−∞
x
(9.11)

It is to be noted that Equations (9.10) and (9.11) were written using the following
two relationships in Equations (9.8) and (9.9), respectively:

y ∞

∫ −∞ ∫
f (x) d x + f (x) d x = 1
y
(9.12)

x ∞

∫ −∞ ∫
f ( y) d y + f ( y) d y = 1
x
(9.13)

Using the above reliability equations, the following stress-strength models were
developed for defined stress and strength probability density functions of an item:

MODEL I
In this case, stress and strength associated with an item are exponentially distributed.
Thus, we have

f ( y) = λ e − λ y , 0≤y<∞ (9.14)

and

f (x) = α e − α x , 0≤x<∞ (9.15)

where

λ and α are the reciprocals of the mean values of stress and strength,
respectively.

Inserting Equations (9.14) and (9.15) into Equation (9.8) yields

© 1999 by CRC Press LLC



 ∞ 
R=

0 ∫
λ e− λ y  α e−α x  d y
 y 

∫ λe ( ) dy
− α +λ y
= (9.16)
0

λ
=
λ+α
1
For λ = 1 and α = , Equation (9.16) becomes
y x

x
R= (9.17)
x+y

where

x and y are the mean strength and stress, respectively.

Dividing the top and bottom of the right-hand side of Equation (9.17) by x, we get

1
R=
1+ y x
(9.18)
1
=
1+ β

where

β = y/x.

For reliable design x ≥ y, thus β ≤ 1. For given values of β, using Equation (9.18),
we obtained the tabulated values for item reliability as shown in Table 9.1.
Table 9.1 values indicate that as the value of β increases from 0 to 1, the item
reliability decreases from 1 to 0.5. More specifically, as the mean stress increases,
the item reliability decreases, accordingly.

Example 9.2
Assume that stress and strength associated with an item are exponentially distributed
with mean values of 10,000 and 35,000 psi, respectively. Compute the item reliability.
Substituting the given data into Equation (9.17) yields

35, 000
R=
35, 000 + 10, 000
= 0.7778

Thus, the item reliability is 0.7778.

© 1999 by CRC Press LLC


TABLE 9.1
Values for Item Reliability
β Item reliability, R

0 1
0.1 0.9091
0.2 0.8333
0.3 0.7692
0.4 0.7143
0.5 0.6667
0.6 0.6250
0.7 0.5882
0.8 0.5556
0.9 0.5263
1 0.5

MODEL II
In this case, an item’s stress and strength are described by normal and exponential
probability density functions, respectively, as

 1 y − µ2
exp − 
1
f ( y) =  −∞ < y < ∞ (9.19)
σ 2π  2 σ  

f (x) = α e − α x , x≥0 (9.20)

where

µ is the mean stress.


σ is the standard deviation associated with stress.
α is the reciprocal of the mean strength.

Substituting Equations (6.19) and (6.20) into Equation (9.8), we get

∞  1  1 y − µ  2  ∞ 
exp − 
R=
∫ −∞

 σ 2 π  2  σ   
 ∫
 α e − α x d x  dy
y


(9.21)
∞   1 y − µ    2
exp −  

1
= + α y   dy
−∞ σ 2π   2  σ   

© 1999 by CRC Press LLC


since

( y − µ )2 + α y = 2 µ α σ 2 − α 2 σ 4 + ( y − µ + α σ 2 )
2

(9.22)
2 σ2 2 σ2

Equation (6.21) yields

R = exp − 2 µ α − α 2 σ 2 
( )
1
(9.23)
 2 

Example 9.3

Assume that the strength of an item follows an exponential distribution with a mean
of 35,000 psi. The stress acting upon the item is normally distributed with mean
10,000 psi and standard deviation 5,000 psi. Calculate the item’s reliability.
By inserting the given data into Equation (9.23), we get

 1  2 (10, 000) (5, 000)2 


R = exp −  − 
 2  35, 000 (35, 000)2 

= exp − {0.5714 − 0.0204}


1
 2 
= 0.7592

Thus, the item’s reliability is 0.7592.

MODEL III
In this case, the strength of the item is normally distributed and the stress acting
upon it follows the exponential distribution. Thus, we have

1  1  x − µ 2
f (x) = exp −  x
 , −∞ < x < ∞ (9.24)
σx 2 π  2  σ x  

f ( y) = λ e − λ y , y>0 (9.25)

where

λ is the reciprocal of the mean stress.


µ x is the mean strength.
σx is the strength standard deviation.

© 1999 by CRC Press LLC


Using Equations (9.24) and (9.25) in Equation (9.9) yields

∞  1  x − µ 2  x 
∫ ∫
1  −λ y

R= exp −  x
  d x λ e d y
− ∞ σx 2 π  2  σ x    o 
 
(9.26)
∞  1  x − µ 2  


1
= 1− exp −   x
 + λ x d x
− ∞ σx 2 π
  2  σ x   

since

( )
2
1  x − µx 
2
2 µ x λ σ x2 + σ x2 λ − µ x + x − σ x4 λ2
+ λ x = (9.27)
2  σ2  2 σ x2

Equation (9.26) is rewritten to the following form:

R = 1 − exp − 2 µ x λ − σG
( 2 2 
)
1
x λ  (9.28)
 2 

where


 1  σ 2 λ − µ + x2

1
G≡ exp −  x x
  dx (9.29)
σx 2 π  2 σx  
−∞ 

Since G = 1 [24], Equation (9.28) simplifies to

R = 1 − exp − 2 µ x λ − σ x2 λ2 
( )
1
(9.30)
 2 

Example 9.4

An item’s stress and strength are described by exponential and normal distributions,
respectively. The mean of the stress is 10,000 psi and the mean and standard deviation
of the strength are 30,000 and 3,000 psi, respectively. Calculate the item’s reliability.
Inserting the specified data into Equation (9.30), we get

 1  2 (30, 000) (3, 000)2 


R = 1 − exp −  − 
 2  10, 000 (10, 000)2 

© 1999 by CRC Press LLC


= 1 − exp − (6 − 0.09)
1
 2 
= 0.9479

Thus, the item’s reliability is 0.9479.

Other stress-strength models are presented in Reference 10.

9.6 MELLIN TRANSFORM METHOD


This is a useful method for estimating an item’s reliability when its associated stress
and strength distributions cannot be assumed, but there is an adequate amount of
empirical data. This approach can also be used when stress and strength distributions
associated with an item are known. Obviously, the method is based on Mellin
transforms and for Equation (9.8) they are defined as follows:


X = f (x) d x
y (9.31)

= 1 − F1 (y)

and
y


Y = f (y) dy = F2 (y)
o
(9.32)

where

F1 (y) and F2 (y) are the cumulative distribution functions.

Taking the derivative of Equation (9.32) with respect to y, we have

dY
= f ( y) (9.33)
dy

Thus, from Equation (9.33), we get

d Y = f ( y) d y (9.34)

Obviously, it can be easily seen from Equation (9.32) that Y takes values of 0 to 1
(i.e., at y = 0, Y = 0 and at y = ∞, Y = 1). Thus, inserting Equations (9.31) and (9.34)
into Equation (9.8) yields

© 1999 by CRC Press LLC


1

R=
∫ X dY
0
(9.35)

The above equation indicates that the area under the X vs. Y plot represents item
reliability. This area can be estimated by using the Simpson’s rule, expressed below.

k
k−j
∫ f (z) d z ⯝
j
3n
(W0 + 4 W1 + 2 W2 + 4 W3 +) … + 2 Wn−2 + 4 Wn−1 + Wn (9.36)

where

f (z) is a function of z defined over interval (j, k), i.e., j ≤ z ≤ k.


n is the even number of equal subdivided parts of interval (j, k).
k−j = g is the width of the subdivided part.
n

At j = z0, z1, z2, z3, …, zn = k; the corresponding values of W are W0 = f (z0), W1 =


f (z1), W2 = f (z2), …, Wn = f (zn).

Example 9.5

Use the Mellin transform method to estimate the item reliability in Example 9.2.
Comment on the end result in comparison to the one obtained in Example 9.2.
Thus, substituting Equation (9.14) and the given relevant data of Example 9.2
into relationship (9.32) yields

 1 
y − y


1
Y= e  10,000  d y
o 10, 000
(9.37)
 1 
− y
 10 ,000 
= 1− e

Inserting Equation (9.15) and the given relevant data of Example 9.2 into
Equation (9.31), we have

X = 1− 1− e [  1 
− 
 35,000  ]
(9.38)
 1 
− y
 35,000 
=e

Table 9.2 presents values of Y and X using Equations (4.37) and (4.38), respectively,
for assumed values of stress y.

© 1999 by CRC Press LLC


TABLE 9.2
Computed Values of Y and X for
the Assumed Values of Stress y
y (psi) Y X

0 0 1
4,000 0.33 0.89
8,000 0.55 0.80
12,000 0.70 0.71
16,000 0.80 0.63
20,000 0.87 0.57
24,000 0.91 0.50
28,000 0.94 0.45
32,000 0.96 0.40
36,000 0.97 0.36
40,000 0.98 0.32
44,000 0.99 0.29
48,000 0.992 0.25
∞ 1 0

Figure 9.3 shows the plot of X vs. Y for data given in Table 9.2. The area under
the Figure 9.3 curve is estimated using Equation (9.36) as follows:

(1 − 0) W + 4 W + 2 W + 4 W + W
R⯝
3 ( 4)
[ 0 1 2 3 ] 4

1

12
[1 + 4 (0.95) + 2 (0.84) + 4 (]0.68) + 0
⯝ 0.7667

The above result, i.e., R = 0.7667, is quite close to the one obtained using the
analytical approach in Example 9.2, i.e., R = 0.7778.

9.7 FAILURE RATE MODELS


Over the years, many failure rate estimation mathematical models have been devel-
oped to determine failure rates of various mechanical items including brakes,
clutches, pumps, filters, compressors, seals, and bearings [12, 13, 25–27]. This sec-
tion presents failure rate calculation models for such items.

9.7.1 BREAK SYSTEM FAILURE RATE MODEL


The brake system failure rate is expressed by [25]:

λ bs = λ bfm + λ s + λ b + λ se + λ a + λ h (9.39)

© 1999 by CRC Press LLC


FIGURE 9.3 Plot of X vs. Y.

where

λbs is the break system failure rate, expressed in failures/106 h.


λbfm is the failure rate of brake friction materials.
λs is the failure rate of springs.
λb is the failure rate of bearings.
λse is the failure rate of seals.
λa is the failure rate of actuators.
λh is the failure rate of brake housing.

The values of λbs, λbfm, λs, λb, λse, λa, and λh are obtained through various
means [25, 28]. For example, λbfm can be calculated by using the following
equation [25]:

λ bfm = λ bbfm ∏f
i =1
i
(9.40)

where

λbbfm is the base failure rate associated with the brake friction material.
fi is the ith multiplying factor that considers the effects on the base failure
rate of items such as dust contaminants (i = 1), ambient temperature
(i = 2), brake type (i = 3), and counter-surface roughness (i = 4).

© 1999 by CRC Press LLC


In turn, the base failure rate of the brake friction material is given by [29]:

. . ( A C) θ ( L T ) . A
λ bbfm = 3 × 10 −3 . R W . W . n α
2
(9.41)

where

LT is the lining thickness, expressed in inches.


A is the lining area, expressed in inches2.
RW is the friction material’s specific wear rate, expressed in in.3/ft-lbf.
W is the vehicle/aircraft weight, expressed in lbf.
n is the miles traveled per year.
α is the number of brake applications per miles.
AC is the average change in velocity per brake action, expressed in ft/s.
θ is the proportion of weight carried by lining during braking action. For
example, in the case of a 4-wheel vehicle, typically each front brake
carries 3/10 of the breaking load.

9.7.2 CLUTCH SYSTEM FAILURE RATE MODEL


The clutch system failure rate is expressed by [25]:

λ cs = λ cfm + λ s + λ b + λ se + λ a (9.42)

where

λcs is the clutch system failure rate, expressed in failure/106 h.


λcfm is the failure rate of clutch friction material.

Other symbols used in Equation (9.42) are the same as the ones used in Equation (9.39).
The failure rate of clutch friction materials, λcfm, is expressed by

λ cfm = λ bcfm f1 f2 (9.43)

where

λbcfm is the base failure rate of clutch friction material.


f1 is the factor that considers the effects on the base failure rate of multiple
plates.
f2 is the factor that considers the effects on the base failure rate of ambient
temperature.

The clutch friction material base failure rate, λbcfm, is given by [25, 30]:

λ bcfm = k E a 2 A Wm CFMa L T (9.44)

© 1999 by CRC Press LLC


where

k is the number of applications per hour.


LT is the lining thickness, expressed in inches.
AWm is the average wear of the material.
CFMa is the area of the clutch friction material on each disc, expressed in
inches2.
Ea is the average energy dissipated per engagement, expressed in ft–lbf.

9.7.3 PUMP FAILURE RATE MODEL


The pump failure rate is expressed by [26]:

λ p = λ ps + λ pse + λ pb + λ pc + λ pfd (9.45)

where

λp is the pump failure rate, expressed in failures/106 cycles.


λps is the failure rate of the pump shaft.
λpse is the failure rate of all pump seals.
λpb is the failure rate of all pump bearings.
λpc is the failure rate of the pump casing.
λpfd is the failure rate of the pump fluid driver.

The failure rate of the pump shaft, λps, can be calculated by using the following
equation:

λ ps = λ bps ∏f
i =1
i
(9.46)

where

λbps is the base failure rate of the pump shaft.


fi is the ith modifying factor; for i = 1 is associated with material temper-
ature, i = 2 pump displacement, i = 3 casing thrust load, i = 4 shaft
surface finish, i = 5 contamination, and i = 6 material endurance limit.

The pump seal failure rate, λpse, is expressed by

λ pse = λ bpse ∏f
i =1
i
(9.47)

© 1999 by CRC Press LLC


where

λbpse is the pump seal base failure rate.


fi is the ith modifying factor; for i = 1 is for effects of casing thrust load,
i = 2 surface finish, i = 3 seal smoothness, i = 4 fluid viscosity, i = 5
pressure/velocity factor, i = 6 temperature, and i = 7 contaminates.

Similarly, the values of λpb, λpc, and λpfd can be calculated. Reference 26 presents
procedures to calculate these failure rates.

9.7.4 FILTER FAILURE RATE MODEL


The following equation can be used to predict filter failure rate [26]:

λ f = λ bf ∏fi =1
i
(9.48)

where

λf is the filter failure rate, expressed in failures/106 h.


λbf is the filter base failure rate.
fi is the ith modifying factor; for i = 1 is for temperature effects, i = 2
vibration effects, i = 3 water contamination effects, i = 4 cyclic flow
effects, i = 5 cold start effects, and i = 6 differential pressure effects.

9.7.5 COMPRESSOR SYSTEM FAILURE RATE MODEL


The compressor system failure rate is expressed by the following equation [27]:

λ com = λ cs + λ cse + λ cva + λ cdc + λ cc + λ cb (9.49)

where

λcom is the compressor system failure rate, expressed in failures/106 h.


λcs is the failure rate of all compressor shafts.
λcse is the failure rate of all compressor seals.
λcva is the failure rate of valve assy (if any).
λcdc is the failure rate due to design configuration.
λcc is the failure rate of the compressor casing.
λcb is the failure rate of all compressor bearings.

The compressor shaft failure rate is expressed by

λ cs = λ bcs ∏f
i =1
i
(9.50)

© 1999 by CRC Press LLC


where

λbcs is the compressor shaft base failure rate.


fi is the ith modifying factor; for i = 1 is for material temperature, i = 2
material endurance limit, i = 3 displacement, i = 4 shaft surface finish,
and i = 5 contamination.

The compressor seal failure rate can be predicted by using the following equation:

10

λ cse = λ bcse ∏f
i =1
i
(9.51)

where

λbcse is the compressor seal base failure rate.


fi is the ith modifying factor; for i = 1 is for fluid pressure, i = 2 seal face
pressure and gas velocity, i = 3 seal smoothness, i = 4 seal size, i = 5
temperature, i = 6 allowable leakage, i = 7 contaminants, i = 8 flow rate,
i = 9 fluid viscosity, and i = 10 surface finish and other conductance
parameters.

The compressor valve assembly failure rate is expressed by

λ cva = λ so + λ sa + λ pa (9.52)

where

λso is the failure rate of the solenoid (if any).


λsa is the failure rate of sliding action valve assy (if any).
λpa is the failure rate of poppet assy (if any).

Procedures for calculating λcdc, λcc, and λcb are presented in Reference 27.

9.7.6 BEARING FAILURE RATE MODEL


The bearing failure rate is expressed by [12]:

λ br = λ pbs (AL SL) (AE 0.006) (SV OV)0.54 L2c 3


x 2.36
(9.53)

where

λbr is the failure rate of bearing (using actual conditions), expressed in


failure/106 h of operation.
x is a factor; x = 4.0 for ball bearings and x = 3.33 for roller bearings.

© 1999 by CRC Press LLC


Lc is the actual contamination level, µg/m 3/60 µg/m 3.
AE is the alignment error, radians.
SV is the specification lubricant viscosity, lb.min/in2.
OV is the operating lubricant viscosity, lb.min/in2.
AL is the actual load, psi.
SL is the specification load, psi.
λpbs is the pseudo failure rate of the bearing and is expressed as follows:

λ pbs = 10 6 M 0.9 (5.45) RL (9.54)

where

M is the number of bearings in system.


RL is the rated life in revolutions and is given by:

RL = 16700 rpm (BLR ERL )


α
(9.55)

where

rpm is the shaft rotating velocity, expressed in rev/min.


α is a constant; α = 10/3 for roller bearings and α = 3 for ball bearings.
BLR is the basic load rating, lb.
ERL is the equivalent radial, load, lb.

9.8 FAILURE DATA SOURCES


FOR MECHANICAL PARTS
Just like in the case of electronic parts, the availability of failure data for mechanical
parts is very important. Usually, it is more difficult to obtain useful failure data for
mechanical parts than for the electronic parts. Reference 10 lists over 50 sources for
obtaining failure data for various mechanical items. Some of the important sources
for obtaining mechanical failure related information are as follows:

• RADC Non-Electronic Reliability Notebook [31].


• Government Industry Data Exchange Program (GIDEP) [32].
• AVCO Corporation Data [33].
• Book: Mechanical Reliability [10].
• Martin Marietta Corporation Data [34].

Table 9.3 presents failure rates for selected mechanical items. Chapter 4 discusses
the subject of failure data in more depth.

© 1999 by CRC Press LLC


TABLE 9.3
Failure Rates for Selected Mechanical Parts [10, 12]
Failure rate
No. Mechanical part Use environment (failures/106 h)

1 Hose, pneumatic Not applicable (N/A) 29.3


2 Pressure regulator Ground, fixed 2.4
3 Fuel pump Ground, fixed 176.4
4 Compressor (general) Ground, mobile 33.6
5 Clutch (friction) Ground, mobile 38.2
6 Filter, gas (air) Ground, mobile 3.2
7 Heavy-duty ball bearing N/A 14.4
8 Seal, O-ring N/A 0.2
9 Brake assembly N/A 16.8
10 Guide pin N/A 13

9.9 OPTIMUM MODELS FOR MECHANICAL


EQUIPMENT REPLACEMENT OR MAINTENANCE
During the mechanical equipment use, its maintenance and replacement are two
important considerations. Over the years, professionals in the field have developed
various types of mathematical models to determine optimum time intervals for
equipment replacement or maintenance. This section presents two such models [10].

9.9.1 EQUIPMENT REPLACEMENT MODEL


This mathematical model is concerned with determining the optimum time interval
between replacements by minimizing the mean annual cost associated with the
equipment with respect to the time between replacements or the equipment life
expressed in years. The model assumes that the equipment mean annual cost is made
up of three components: (1) average operating cost, (2) average maintenance cost,
and (3) average investment cost. Thus, the mean annual cost of the equipment is
expressed by

CI (x − 1)
MC = C0 + C m + + (IOC + IMC) (9.56)
x 2

where

MC is the mean annual total cost of equipment.


C0 is the operational cost for the first year.
Cm is the maintenance cost for the first year.
x is the equipment life in years.
IOC is the amount by which operational cost increases annually.

© 1999 by CRC Press LLC


IMC is the amount by which maintenance cost increases annually.
CI is the cost of investment.

Taking the derivatives of Equation (9.56) with respect to x, we get

dMC 1 CI
= (IOC + IMC) − 2 (9.57)
dx 2 x

Equating Equation (9.57) to zero and rearranging yields

12
x* =  
2 CI
(9.58)
 IOC + IMC 

where

x* is the optimum replacement time interval.

Example 9.6

Assume that the following data are specified for mechanical equipment:

IMC = $500, IOC = $600, and CI = $15, 000

Calculate the optimum time for the mechanical equipment replacement.


Using the above data in Equation (9.58) yields

 2 (15, 000) 
12

x* =  
 600 + 500 
= 5.2 years

Thus, the optimum time for the mechanical equipment replacement is 5.2 years.

9.9.2 EQUIPMENT MAINTENANCE MODEL


This mathematical model can be used to calculate the optimum number of inspec-
tions per facility per unit time. This information is quite useful to decision makers
because inspections are often disruptive but on the other hand such inspections
normally reduce the equipment downtime because of fewer breakdowns. In this
model, expression for the total equipment downtime is used to obtain the optimum
number of inspections.
The total facility downtime per unit time is given by [35]:

T D T = n (D T I) + (D T B) k n (9.59)

© 1999 by CRC Press LLC


where

TDT is the total facility downtime per unit time.


k is a constant for specific facility.
n is the number of inspections per facility per unit time.
DTI is the downtime per inspection for a facility under consideration.
DTB is the downtime per breakdown for a facility under consideration.

By differentiating Equation (9.59) with respect to n, we get

dTDT
= D T I − (D T B) k n 2 (9.60)
dn

Setting Equation (9.60) equal to zero and rearranging yields

n* = [ k (DTB) DTI]
12
(9.61)

where

n* is the optimum number of inspections per facility per unit time.

Inserting Equation (9.61) into Equation (9.59) yields

[
T D T* = 2 (DTB) (D TI) k ]
12
(9.62)

where

TDT* is the minimum total facility downtime per unit time.

Example 9.7

The following data are associated with certain mechanical equipment:

DTI = 0.02 month, DTB = 0.1 month, and k = 4. (9.63)

Calculate the optimum number of inspections per month and the minimum value of
the total equipment downtime by using Equations (9.61) and (9.62), respectively.
Thus, inserting the given data into Equation (9.61) yields

n* = [4 (0.1) 0.02]
12

= 4.5 inspections month

© 1999 by CRC Press LLC


Similarly, substituting the specified data into Equation (9.62), we have

TDT * = 2 [(0.1) (0.02) 4]


12

= 0.18 months

Thus, the optimum number of inspections per month is 4.5 and the minimum
value of the total equipment downtime is 0.18 month.

9.10 PROBLEMS

1. Give reasons for the development of the mechanical reliability field.


2. Describe the following mechanical failure modes:
• Compressive failure
• Stress concentration failure
• Shear loading failure
• Bending failure
3. What are the important causes of gear failures?
4. Give four different ways of defining a safety factor.
5. Define the terms “stress” and “strength”.
6. An item’s stress and strength are normally distributed. Develop an expres-
sion for the item reliability.
7. A component’s stress and strength are distributed exponentially with mean
values of 12,000 and 40,000 psi, respectively. Determine the item reliability.
8. Use the Mellin transform approach to estimate the item reliability in
Problem 7. Comment on the end result in comparison to the one obtained
for Problem 7.
9. List at least 10 important sources for obtaining failure data for mechanical
items.
10. A mechanical component’s stress and strength are Maxwellian distributed.
Develop an expression for the component reliability.

9.11 REFERENCES
1. Weibull, W., A statistical distribution function of wide applicability, J. Appl. Mech.,
18, 293-297, 1951.
2. Freudenthal, A.M., Safety and probability of structural failures, Trans. Am. Soc. Civil
Eng., 121, 1337-1397, 1956.
3. Freudenthal, A.M., Fatigue in Aircraft Structures, Academic Press, New York, 1957.
4. Freudental, A.M. and Gumbel, E.J., Failure and survival in fatigue, J. Appl. Phys.,
25, 110-120, 1954.
5. Dhillon, B.S. and Singh, C., Engineering Reliability: New Techniques and Applica-
tions, John Wiley & Sons, New York, 1981.
6. Redler, W.M., Mechanical reliability research in the national aeronautics and space
administration, Proc. Reliability Maintainability Conf., New York, 763-768, 1966.

© 1999 by CRC Press LLC


7. Lipson, C., Sheth, N.J., and Disney, R.L., Reliability Prediction-Mechanical
Stress/Strength Inference, Rome Air Development Center, Griffiss Air Force Base,
New York, 1967, Report No. RADC-TR-66-710.
8. Kececioglu, D., Reliability analysis of mechanical components and systems, Nuclear
Eng. Des., 19, 259-290, 1972.
9. Kececioglu, D. and Haugen, E.B., A unified look at design safety factors, safety
margins and measures of reliability, Proc. Annu. Reliability Maintainability Conf.,
522–530, 1968.
10. Dhillon, B.S., Mechanical Reliability: Theory, Models and Applications, American
Institute of Aeronautics and Astronautics, Washington, D.C., 1988.
11. Mechanical Reliability Concepts, American Society of Mechanical Engineers
(ASME) Design Engineering Conference, ASME, New York, 1965.
12. Grant Ireson, W., Coombs, C.F., and Moss, R.Y., Handbook of Reliability Engineering
and Management, McGraw-Hill, New York, 1996.
13. Doyle, R.L., Mechanical-system reliability, Tutorial Notes, Annu. Reliability Main-
tainability Symp., 1992.
14. Collins, J.A., Failure of Materials in Mechanical Design, John Wiley & Sons, New
York, 1981.
15. Lipson, C., Analysis and Prevention of Mechanical Failures, Course Notes No. 8007,
University of Michigan, Ann Arbor, June 1980.
16. Howell, G.M., Factors of safety, Machine Design, July 12, 76-81, 1956.
17. Schoof, R., How much safety factor?, Allis-Chalmers Elec. Rev., 21-24, 1960.
18. McCalley, R.B., Nomogram for selection of safety factors, Design News, Sept. 1957,
pp. 138-141.
19. Shigley, J.E. and Mitchell, L.D., Mechanical Engineering Design, McGraw-Hill, New
York, 610-611, 1983.
20. Bompass-Smith, J.H., Mechanical Survival: The Use of Reliability Data, McGraw-
Hill, London, 1973.
21. Juvinall, R.C., Fundamentals of Machine Component Design, John Wiley & Sons,
New York, 1983.
22. Dhillon, B.S. and Singh, C., Engineering Reliability: New Techniques and Applica-
tions, John Wiley & Sons, New York, 1981.
23. MIL-STD-1629, Procedures for Performing a Failure Mode, Effects, and Criticality
Analysis (FMECA), Department of Defense, Washington, D.C.
24. Kececioglu, D. and Li, D., Exact solutions for the predication of the reliability of
mechanical components and structural members, Proc. Failure Prevention Reliability
Conf., published by the American Society of Mechanical Engineers, ASME, New
York, 1985, pp. 115-122.
25. Rhodes, S., Nelson, J.J., Raze, J.D., and Bradley, M., Reliability models for mechan-
ical equipment, Proc. Annu. Reliability Maintainability Symp., 127-131, 1988.
26. Raze, J.D., Nelson, J.J., Simard, D.J., amd Bradley, M., Reliability models for
mechanical equipment, Proc. Annu. Reliability Maintainability Symp., 130-134, 1987.
27. Nelson, J.J., Raze, J.D., Bowman, J., Perkins, G., and Wannamaker, A., Reliability
models for mechanical equipment, Proc. Annu. Reliability Maintainability Symp.,
146-153, 1989.
28. Boone, T.D., Reliability Prediction Analysis for Mechanical Brake Systems, NAVA
IR/SYSCOM Report, August 1981, Department of Navy, Department of Defense,
Washington, D.C.

© 1999 by CRC Press LLC


29. Minegishi, H., Prediction of Brake Pad Wear/Life by Means of Brake Severity Factor
as Measured on a Data Logging System, Society of Automotive Engineers (SAE)
Paper No. 840358, 1984, SAE, Warrendale, PA.
30. Spokas, R.B., Clutch Friction Material Evaluation Procedures, SAE Paper No. 841066,
1984, SAE, Warrendale, PA.
31. Shafer, R.E., Angus, J.E., Finkelstein, J.M., Yerasi, M., and Fulton, D.W., RADC
Non-Electronic Reliability Notebook, Report No. RADC-TR-85-194, 1985, Reliabil-
ity Analysis Center (RAC), Rome Air Development Center (RADC), Griffiss Air
Force Base, Rome, NY.
32. GIDEP Operations Center, U.S. Department of Navy, Naval Weapons Station, Seal
Beach, Corona Annex, Corona, CA.
33. Earles, D.R., Handbook: Failure Rates, AVCO Corporation, Massachussetts, 1962.
34. Guth, G., Development of Nonelectronic Part Cycle Failure Rates, Rept. No. 17D/A-
050678, December 1977, Martin Marietta Corporation, Orlando, FL. Available from
the National Technical Information Service (NTIS), Springfield, VA.

© 1999 by CRC Press LLC

You might also like