You are on page 1of 42

TECHNICAL ENGLISH

1. OVERVİEW
As one of their responsibilities, mechanical engineers design hardware so
that it shouldn’t break when being used and so that it can carry the
forces acting on it reliably and safely. As an example, consider Boeing’s
767 commercial airliner, which weighs up to 350,000 lb when fully
loaded. When the airplane is parked on the ground, its weight is
supported by the landing gear and wheels. During flight, the aircraft’s
wings create an upward lift force that exactly balances the weight. Each
wing, therefore, carries a force that is equal to half of the airplane’s
weight, which in this case is equivalent to some 70 family sized
automobile sedans.

2
Source: An introduction to Mechanical Engineering, Jonathan Wickert, Kemper Lewis, 2013
Subjected to the lift force, the wings bend upward, and, if the flight
happens to encounter rough weather, the wings will bend up and
down by an additional considerable amount as the plane is buffeted
by turbulence. When engineers selected the aircraft’s materials, they
took into account the facts that an aircraft’s wings are subjected to
large forces, that they sag under their self-weight, and that they bend
upward in response to the lift forces. The wings are designed to be
strong, safe, and reliable while being no heavier than necessary to
meet the design requirements.

3
Source: An introduction to Mechanical Engineering, Jonathan Wickert, Kemper Lewis, 2013
By applying the properties of force to systems as described in the
previous section, you have seen how to calculate the magnitudes and
directions of forces that act on certain structures and machines.
Knowing those forces alone, however, is not enough information to
decide whether a certain piece of hardware will be strong enough
and not fail in its task. By “fail” or “failure,” we not only mean that
the hardware will not break, but also that it will not stretch or bend
so much as to become significantly distorted. A 5-kN force, for
instance, might be large enough to break a small bolt or to bend a
shaft so much that it would wobble and not spin smoothly. A larger-
diameter shaft or one that is made from a higher-grade material, on
the other hand, might very well be able to support that force without
incurring any damage.

4
Source: An introduction to Mechanical Engineering, Jonathan Wickert, Kemper Lewis, 2013
With those thoughts in mind, you can see that the circumstances for
a mechanical component to break, stretch, or bend depend not only
on the forces applied to it, but also on its dimensions and the
properties of the material from which it is made. Those
considerations give rise to the concept of stress as a measure of the
intensity of a force applied over a certain area. Conversely, the
strength of a material describes its ability to support and withstand
the stress applied to it. Engineers compare the stress present in a
component to the strength of its material in order to determine
whether the design is satisfactory.

5
Source: An introduction to Mechanical Engineering, Jonathan Wickert, Kemper Lewis, 2013
As an example, the broken crankshaft shown in Figure 5.1 was
removed from a single-cylinder internal-combustion engine. This
failure was accelerated by the presence of sharp corners in the
shaft’s rectangular keyway, which is used to transfer torque between
the shaft and a gear or pulley. The spiral shape of the fracture surface
indicates that the shaft had been overloaded by a high torque prior
to breaking. Engineers are able to combine their knowledge of forces,
materials, and dimensions in order to learn from past failures and to
improve and evolve the design of new hardware.

6
Source: An introduction to Mechanical Engineering, Jonathan Wickert, Kemper Lewis, 2013
7
Source: An introduction to Mechanical Engineering, Jonathan Wickert, Kemper Lewis, 2013
In this section, we will discuss some of the properties of engineering
materials and examine the stresses that can develop within them.
Within these topics lie the discipline known as solid mechanics.
Tension, compression, and shear stress are quantities that engineers
calculate when they relate the dimensions of a mechanical
component to the forces acting on it. Those stresses are then
compared to the material’s physical properties to determine whether
failure is expected to occur. When the strength exceeds the stress, we
expect that the structure or machine component will be able to carry
the forces without incurring damage. Engineers conduct these types
of force, stress, materials, and failure analyses while they design
products.

8
Source: An introduction to Mechanical Engineering, Jonathan Wickert, Kemper Lewis, 2013
2. TENSION and COMPRESSION
The type of stress that is most readily visualized and useful for you to
develop intuition about materials and stresses is called tension and
compression. Figure 5.3 shows a built-in round rod that is held fixed
at its left end and then later placed in tension by the force F that pulls
on the right end. Before the force is applied, the rod has original
length L, diameter d, and cross-sectional area:

9
Source: An introduction to Mechanical Engineering, Jonathan Wickert, Kemper Lewis, 2013
10
Source: An introduction to Mechanical Engineering, Jonathan Wickert, Kemper Lewis, 2013
Engineers usually calculate the cross-sectional area of round rods,
bolts, and shafts in terms of their diameter rather than their radius r
(A = πr2) since it’s more practical to measure the diameter of a shaft
using a caliper gauge. As the force F is gradually applied, the rod
stretches along its length by the amount ΔL shown in Figure 5.3(b). In
addition, the diameter of the rod shrinks by a small amount due to an
effect known as Poisson’s contraction, which is a topic described in
the next section. In any event, the change in diameter Δd is smaller
and usually less noticeable than the lengthwise stretch ΔL.

11
Source: An introduction to Mechanical Engineering, Jonathan Wickert, Kemper Lewis, 2013
If the force is not too great, the rod will return to its original diameter
and length (just like a spring) when F is removed. If the rod is not
permanently deformed after F has been applied, the stretching is said
to occur elastically. Alternatively, the force could have been great
enough that the rod would be plastically deformed, meaning that,
when the force was applied and then removed, the rod would be
longer than it was initially.

12
Source: An introduction to Mechanical Engineering, Jonathan Wickert, Kemper Lewis, 2013
Although the force is applied at only one end of the rod in our
example, its influence is felt at each cross section along the rod’s
length. As shown in Figure 5.4, imagine slicing through the rod at
some interior point. The segment that is isolated in the free body
diagram of Figure 5.4(b) shows F applied to the rod’s right-hand end
and an equivalent internal force on the segment’s left hand end that
balances F. Such must be the case; otherwise the segment shown in
the free body diagram would not be in equilibrium. The location of
our hypothetical slice through the rod is arbitrary, and we conclude
that a force of magnitude F must be carried by the rod at each of its
cross sections.

13
Source: An introduction to Mechanical Engineering, Jonathan Wickert, Kemper Lewis, 2013
14
Source: An introduction to Mechanical Engineering, Jonathan Wickert, Kemper Lewis, 2013
Since the rod is formed of a continuous solid material, we do not
realistically expect that the internal force will be concentrated at a
point as depicted by the force vector arrow in Figure 5.4(b). Instead,
the influence of the force will be spread out and smeared over the
rod’s cross section; this process is the basic idea behind stresses in
mechanical components. Stress is essentially an internal force that
has been distributed over the area of the rod’s cross section [Figure
5.4(c)], and it is defined by the equation

15
Source: An introduction to Mechanical Engineering, Jonathan Wickert, Kemper Lewis, 2013
Like the force F, the stress σ (the lowercase Greek character sigma) is
oriented perpendicularly to a hypothetical slice made through the
cross section. When the stress tends to lengthen the rod, it is called
tension, and σ > 0. On the other hand, when the rod is shortened, the
stress is called compression. In that case, the direction of σ in Figure
5.4 reverses to point inward, and σ < 0. The orientations of forces in
tension and compression are shown in Figure 5.5.

16
Source: An introduction to Mechanical Engineering, Jonathan Wickert, Kemper Lewis, 2013
Similar to the pressure within a liquid or gas, stress is also interpreted
as a force that has been distributed over an area. Therefore, stress
and pressure have the same dimensions. In the SI, the derived unit of
stress is the pascal (1 Pa = 1 N/m2), and the dimension pound-per-
square-inch (1 psi = 1 lb/in2) is used in the USCS. Because large
numerical quantities frequently arise in calculations involving stresses
and material properties, the prefixes “kilo” (k), “mega” (M), and
“giga” (G) are applied to represent factors of 103, 106, and 109,
respectively. Therefore,

17
Source: An introduction to Mechanical Engineering, Jonathan Wickert, Kemper Lewis, 2013
While stress is related to the intensity of a force’s application, the
engineering quantity called strain measures the amount that the rod
stretches. The elongation ΔL in Figure 5.3 is one way to describe how
the rod lengthens when F is applied, but it is not the only way, nor is
it necessarily the best. If a second rod has the same cross-sectional
area but is only half as long, then according to Equation (5.2), the
stress within it would be the same as in the first rod. However, we
expect intuitively that the shorter rod would stretch by a smaller
amount. To convince yourself of this principle, hang a weight from
two different lengths of rubber band, and notice how the longer
band stretches more.

18
Source: An introduction to Mechanical Engineering, Jonathan Wickert, Kemper Lewis, 2013
Just as stress is a measure of force per unit area, the quantity called
strain is defined as the amount of elongation that occurs per unit of
the rod’s original length. Strain ε (the lowercase Greek character
epsilon) is calculated from the expression

Because the length dimensions cancel out in the numerator and


denominator, ε is a dimensionless quantity. Strain is generally very
small, and it can be expressed either as a decimal quantity (for
instance, ε = 0.005) or as a percentage (ε = 0.5%).

19
Source: An introduction to Mechanical Engineering, Jonathan Wickert, Kemper Lewis, 2013
3. MATERIAL RESPONSE
The definitions of stress and strain, in contrast to force and elongation,
are useful because they are scaled with respect to the rod’s size.
Imagine conducting a sequence of experiments with a collection of rods
made of identical material but having various diameters and lengths. As
each rod is pulled in tension, the force and elongation would be
measured. In general, for a given level of force, each rod would stretch
by a different amount because of the variations in diameter and length.
For each individual rod, however, the applied force and elongation
would be proportional to one another, following

where the parameter k is called the stiffness. This observation is the


basis of the concept known as Hooke’s law.
20
Source: An introduction to Mechanical Engineering, Jonathan Wickert, Kemper Lewis, 2013
Continuing with our hypothetical experiment, we next imagine
constructing a graph of F versus ΔL for each of the different rods. As
indicated in Figure 5.9(a), the lines on these graphs would have
different slopes (or stiffnesses) depending on the values of d and L.
For a given force, longer rods and ones with smaller cross sections
stretch more than the other rods. Conversely, shorter rods and ones
with larger cross-sectional areas stretch less. Our graph would show a
number of straight lines, each having a different slope. Despite the
fact that the rods are made from the same material, they would
appear to be quite different from one another in the context of the F-
versus-ΔL graph.

21
Source: An introduction to Mechanical Engineering, Jonathan Wickert, Kemper Lewis, 2013
22
Source: An introduction to Mechanical Engineering, Jonathan Wickert, Kemper Lewis, 2013
On the other hand, the rods would behave in an identical manner
when their stretching is instead described by stress and strain. As
shown in Figure 5.9(b), each of the F-versus-ΔL lines would collapse
onto a single line in the stress–strain diagram. Our conclusion from
such experiments is that while the stiffness depends on the rod’s
dimensions, the stress–strain relationship is a property of the
material alone and is independent of the test specimen’s size.

23
Source: An introduction to Mechanical Engineering, Jonathan Wickert, Kemper Lewis, 2013
24
Source: An introduction to Mechanical Engineering, Jonathan Wickert, Kemper Lewis, 2013
Figure 5.10 shows an idealized stress–strain curve for a typical
structural quality steel. The stress–strain diagram is broken down into
two regions: the low-strain elastic region (where no permanent set
remains after the force has been applied and removed) and the high-
strain plastic region (where the force is large enough that, upon
removal, the material has permanently elongated). For strains below
the proportional limit (point A), you can see from the diagram that
stress and strain are proportional to one another and that they
therefore satisfy the relationship

25
Source: An introduction to Mechanical Engineering, Jonathan Wickert, Kemper Lewis, 2013
The quantity E is called the elastic modulus, or Young’s modulus, and
it has the dimensions of force per unit area. In the SI, the units GPa
are typically used for the elastic modulus, and the dimension MPsi is
used in the USCS. The elastic modulus is a physical material property,
and it is simply the slope of the stress–strain curve for low strains.
In as much as metals and most other engineering materials are
combinations of various chemical elements, the elastic modulus of a
material is related to the strength of its interatomic bonds. As one
example, steel alloys contain different fractions of such elements as
carbon, molybdenum, manganese, chromium, and nickel. A common
material known as 1020 grade steel contains 0.18–0.23% carbon (by
weight), 0.30–0.60% manganese, and a maximum of 0.04%
phosphorus and 0.05% sulfur.
26
Source: An introduction to Mechanical Engineering, Jonathan Wickert, Kemper Lewis, 2013
This alloy, like all other compositions of steel, is comprised primarily
of iron, and so the value of the elastic modulus for one alloy of steel
is not much different from that for another. A similar situation exists
for various alloys of aluminum. The following numerical values for the
elastic modulus of steel and aluminum are sufficiently accurate for
most engineering calculations and design purposes:

You can see that the elastic modulus of aluminum is lower than that of
steel by a factor of three. With the same dimensions and applied force,
an aluminum rod would stretch three times as much as a steel rod.

27
Source: An introduction to Mechanical Engineering, Jonathan Wickert, Kemper Lewis, 2013
Returning to our discussion of the stress–strain diagram of Figure
5.10, point B is called the elastic limit. For loading between points A
and B, the material continues to behave elastically, and it will spring
back upon removal of the force, but the stress and strain are no
longer proportional. As the load increases beyond B, the material
begins to show a permanent set. Yielding starts to occur in the
region between B and C; that is, even for small changes in stress, the
rod experiences a large change in strain. In the yielding region, the
rod stretches substantially even as the force grows only slightly
because of the shallow slope in the stress–strain curve. For that
reason, the onset of yielding is often taken by engineers as an
indication of failure.

28
Source: An introduction to Mechanical Engineering, Jonathan Wickert, Kemper Lewis, 2013
The value of stress in region B–C defines the material property called
the yield strength, Sy. As the load is even further increased beyond
point C, the stress grows to the ultimate tensile strength, Su at point
D. That value represents the largest stress that the material is capable
of sustaining. As the test continues, the stress in the figure actually
decreases, owing to a reduction in the rod’s cross-sectional area, until
the sample eventually fractures at point E.

29
Source: An introduction to Mechanical Engineering, Jonathan Wickert, Kemper Lewis, 2013
Stress–strain curves are measured on a device called a materials
testing machine. Figure 5.11 shows an example of equipment in
which a computer both controls the test stand and records the
experimental data. During tension testing, a specimen such as a rod
of steel is clamped between two jaws that are gradually pulled apart,
placing the specimen in tension. A load cell that measures force is
attached to one of the jaws, and a second sensor (called an
extensometer) measures how much the sample is stretched. A
computer records the force and elongation data during the
experiment, and those values are then converted to stress and strain
by using Equations (5.2) and (5.3). When the stress–strain data is
drawn on a graph, the slope in the low-strain region is measured to
determine E, and the value of the stress Sy at the yield point is read
off the curve.

30
Source: An introduction to Mechanical Engineering, Jonathan Wickert, Kemper Lewis, 2013
31
Source: An introduction to Mechanical Engineering, Jonathan Wickert, Kemper Lewis, 2013
4. FACTOR OF SAFETY
Mechanical engineers determine the shape and dimensions and
select materials for a wide range of products and types of hardware.
The analysis that supports their design decisions takes into account
the tensile, compressive, and shear stresses that are present, the
material properties, and other factors that you will encounter later in
your study of mechanical engineering. Designers are aware that a
mechanical component can break or otherwise be rendered useless
by a variety of causes. For instance, it could yield and be permanently
deformed, fracture suddenly into many pieces because the material
is brittle, or become damaged through corrosion.

32
Source: An introduction to Mechanical Engineering, Jonathan Wickert, Kemper Lewis, 2013
In this section, we introduce a simple model that engineers use to
determine whether a mechanical component is expected to yield
because of either tension or shear stress. This analysis predicts the
onset of yielding in ductile materials, and it is a useful tool that
enables mechanical engineers to prevent the material in a piece of
hardware from being loaded to, or above, its yield stress. Yielding is
only one of many possible failure mechanisms, however, and so our
analysis in this section obviously will not offer predictions regarding
other kinds of failure.

33
Source: An introduction to Mechanical Engineering, Jonathan Wickert, Kemper Lewis, 2013
From a practical standpoint, engineers recognize that, despite their analysis,
experiments, experience, and design codes, nothing can be built to perfection.
In addition, no matter how carefully they might try to estimate the forces that
are going to act on a structure or machine, the mechanical component could be
accidentally overloaded or misused. For those reasons, a catchall parameter
called the factor of safety is generally introduced to account for unexpected
effects, imprecision, uncertainty, potential assembly flaws, and material
degradation. The factor of safety is defined as the ratio of the stress at failure
to the stress during ordinary use. The factor of safety to guard against ductile
yielding is often chosen to fall in the range of 1.5–4.0; that is, the design is
intended to be between 150% and 400% as strong as it needs to be for ordinary
use. For engineering materials with better-than-average reliability or for well
controlled operating conditions, the lower values for the factor of safety are
appropriate. When new or untried materials will be used or other uncertainties
are present, the larger factors will result in a safer design.
34
Source: An introduction to Mechanical Engineering, Jonathan Wickert, Kemper Lewis, 2013
When a straight rod is subjected to tension as in Figure 5.3, the
possibility of it yielding is assessed by comparing the stress σ to the
material’s yield strength Sy. Failure due to ductile yielding is predicted
if σ > Sy. Engineers define the tensile-stress factor of safety as

If the factor of safety is greater than one, this viewpoint predicts that
the component will not yield, and if it is less than one, failure is
expected to occur.

35
Source: An introduction to Mechanical Engineering, Jonathan Wickert, Kemper Lewis, 2013
The numerical value for the factor of safety that an engineer chooses for a
particular design will depend on many parameters, including the designer’s own
background, experience with components similar to the one being analyzed, the
amount of testing that will be done, the material’s reliability, the consequences
of failure, maintenance and inspection procedures, and cost. Certain spacecraft
components might be designed with the factor of safety being only slightly
greater than one in order to reduce weight, which is at a premium in aerospace
applications. To counterbalance that seemingly small margin for error, those
components will be extensively analyzed and tested, and they will be developed
and reviewed by a team of engineers having a great deal of collective experience.
When forces and load conditions are not known with certainty or when the
consequences of a component’s failure would be significant or endanger life,
large values for the factor of safety are appropriate. Engineering handbooks and
design codes generally recommend ranges for safety factors, and those
references should be used whenever possible.
36
Source: An introduction to Mechanical Engineering, Jonathan Wickert, Kemper Lewis, 2013
VOCABULARY Significantly: önemli derecede
Reliably: Güvenilir bir şekilde Distorted: Çarpılmış
Safely: Emniyetli Shaft: mil, şaft
Commercial: ticari Incur: uğramak
Airliner: büyük yolcu uçağı Stretch: uzamak, uzatmak
Buffet: sarsmak Intensity: yoğunluk, şiddet
Be subjected to: maruz kalmak Stress: gerilme
Sag: çökmek, bükülmek Strength: dayanım
Bend: eğilmek Satisfactory: yeterli
However: Bununla birlikte, yinede, ancak Keyway: kama yuvası
Failure: Hasar Pulley: kasnak, Makara

37
Fracture: kırılma Straight: düz, doğru
Crankshaft: krank mili Gradually: yavaş yavaş, derece derece, git gide
Tension: çekme Shrink: daralmak, büzüşmek
Compression: Basma Plastic deformation: plastic deformasyon, kalıcı
Shear: Kayma, kesme şekil değişimi
Length: uzunluk Elastic deformation: elastik deformasyon
Diameter: çap Equivalent: eşdeğer, denk
Pull: çekmek Distribute: dağıtmak, dağılmak
Rod: çubuk Spread out: yayılmak
Built-in: ankastre Smear: yayma
Fixed: sabit Lengthen: uzatmak, uzamak
Shorten: kısaltmak, kısalmak
38
Orientation: yön Slope: eğim
Reverse: ters, karşıt Interatomic bonds: atomlar arası bağlar
Elongation: uzama Molybdenum: molibden
Strain: birim şekil değiştirme, genleme, Manganese: mangan
gerinme Chromium: krom
Stiffness: rijitlik, elastik deformasyona karşı Nickel: nikel
direnç
Phosphorus: fosfor
Hypothetical: düşünsel, farazi
Sulfur: kükürt
Apply: uygulamak
Alloy: Alaşım
Stress-strain curve: gerilme-birim şekil
Steel: çelik
değiştirme eğrisi
Iron: demir
Proportional limit: orantı sınırı
39
Accurate: kesin load cell: yük hücresi
Contain: içermek extensometer: ekstansometre,
Comprised of: oluşmak uzama ölçer
Yielding: akma curve: eğri
Yield strength: akma dayanımı otherwise: aksi takdirde
Ultimate tensile strength: çekme dayanımı prevent: önlemek
Spring back: geri yaylanma prediction: tahmin
Substantially: büyük ölçüde, oldukça, factor of safety: emniyet katsayısı
önemli miktarda experiment: deney
Onset of: başlangıç estimate: tahmin etmek,
Jaw: çene değerlendirmek
Specimen: numune unexpected: beklenmeyen
40
Uncertainty: belirsizlik Slightly: çok az
imprecision: kesin olmama Margin: pay
Flaw: kusur, hata Considerable: hatırı sayılır derecede,
Degradation: giderek bozunma, özelliklerini önemli, oldukça büyük
kaybetme
Caliper gauge: kumpas
Reliability: güvenirlilik
Exceed: aşmak
Consequence: sonuç, netice, önem
Fracture: kırılma
Maintenance: bakım, tamir
Failure: Hasar
Inspection: muayene, denetim, teftiş,
Numerator: pay
kontrol
Denominator: payda
Procedure: prosedür, yöntem, izlenecek yol
Observation: gözlem
Cost: maliyet

41
Pressure: basınç

42

You might also like