Sie sind auf Seite 1von 17

5

Mechanical Behavior

5.1

INTRODUCTION

It is the mechanical behavior o f materials which is often the deciding factor


in designing or building the products o f our society, for example, machinery,
bridges, and aircraft. It would be funny to think o f a bridge constructed of
lead girders since we are accustomed to the use o f steel for such purposes
because o f its mechanical behavior. Similarly the use o f aluminum for the
construction o f automobile engines was at one time considered strange
today it is a fact. Just as metal has proven its value in replacing wooden
structures, so today newer metals and alloys, ceramics, and plastics are replacing steel in many applications. One o f the prime considerations in such replacement is that the mechanical behavior o f the new material be more suited
for the service conditions than the mechanical behavior o f former materials.
Steel rods may be bent and the bend retained upon release of the bending
forces. Elastic bands may be stretched and on release o f the stretching force
the band returns to its original configuration. These two phenomena are not
unrelated. Steel rods may be bent and are found to return to the unbent condition i f the bend is very small. The type o f deformation where the original
shape may be regained is called elastic deformation and occurs in metals
and ceramics as well as in plastics and rubber, but to a lesser degree. In real
materials some elastic deformation precedes the permanent or plastic deformation mentioned above for steel.
Materials that are suitable for room-temperature applications because of
their mechanical behavior often prove unsuitable at subzero or elevated temperatures. The selection of proper materials for engineering applications is
dependent on the response of the material to the forces acting upon it in service
- the mechanical response. T o understand mechanical behavior and the
factors influencing this behavior under different conditions, it is necessary for
us to discuss the various types o f deformation and service conditions and
examine the atomic mechanisms of deformation. T o accomplish this end,

'

;
i

Slrain
we must introduce certain terms which are commonly used in discussing
the mechanical behavior of matter.

5.2

STRESS

I f a load P is applied to both ends o f a bar of cross-sectional area A, it is


more convenient to express the force on the bar in terms of force per unit
area rather than total force. The force per unit area of cross section is called
the stress acting on the bar and is denoted by the symbol o-, where

(5.1)
The stress is expressed in units o f l b / i n - (psi) where the force is given in
pounds (lb) and the cross-sectional area in square inches (in-).
A force applied in such a way that it tends to extend the bar being tested
(Figure 5.1a) is said to be a tensile force and the bar is said to be in a state
of tension. I f we imagine the bar cut in half (Figure 5.1b), the bottom half o f
the bar exerts a force on the top half which is equal to the force P but in the
opposite direction. This force exerted by one part of the specimen on the
other is the product of the stress cr and the area A. The stress on the bottom
half of the bar is directed upward to oppose the force P. Even though the
stresses on the two halves are directed opposite to one another, they both
have the same sign. Tensile stresses are considered positive, o- > 0. A bar
loaded in such a way as to cause it to be compressed is said to be in compression (Figures 5.1c and d), and the compressive stresses are considered
negative, (r < 0.

1
1

The advantage o f using stress rather than load lies in the geometry o f the
bar. Bars of different cross-sectional areas under the same load do not suffer
the same stress. As we will discuss below, the behavior o f materials under
load depends on the force per unit cross-sectional area, the stress, and not on
the magnitude of the load.

r
5.3

STRAIN

The deformation of a specimen subjected to a load may take the form o f


an elongation or contraction depending on the direction of the force vector.
Specimens in tension (Figure 5.1a) will be elongated and those in compression
(Figure 5.1c) will be compressed. The over-all change in length o f a specimen 5 is equal to the difference in the length at any particular measurement
/ and the original l e n g t h / ; that is.
8= /-/.

(5.2)

131

752

Mechanical

Behavior
IP

--

t P
(a)

(b)

(c)

Figure 5.1 Tension and Compression. A force


sectional a r e a
results in a stress cr of magnitude
if the applied force is directed away from the bar
compression if the applied force is directed into the

(d)

P applied to a bar of crossP/A^. The bar is in tension


as shown in (a), and it is in
bar as shown in (c).

<,

Just as it is convenient to express force in terms o f stress (force per unit


area), so it is convenient to express change in length in terms o f elongation
or contraction per unit length. This is called strain and is denoted by the
symbol , where
8
(5.3)
Since both 8 and / are in units o f length, strain is dimensionless.
it is expressed as inches/inch.*

Ordinarily

* Strain is commonly expressed in terms of /x-in./in. (micro-inches per inch) if it is very small,
l/i-in. = 10"' in., that is, one millionth of an inch.

Elastic Deformation
The strain imposed on a specimen by the application of a stress may be
either positise or negative (that is. the specimen may increase or decrease in
length) depending on whether a tensile or compressive stress is applied. The
strain and elongation carry the same sign. I f a specimen of original length /
(Figure .5.2a) is subjected to a tensile stress, it will be elongated, 6 > 0, and its
new length / > / (Figure 5.2b). I f a compressive stress is applied instead,
/ < ^, and 6 < 0 as shown in Figure 5.2(c). Therefore, for tension, cr > 0 and
e > 0 while, for compression, cr < 0 and e < 0.
>0

e =0

e<0

/n

/o

(a)

/o

(b)

Figure 5.2 Strain Due to Tension and Compression.

(c)
A specimen of length

lii before deformation as shown in (a) undergoes a change in length S = /


and a strain = S//,, such that for tension, as shown in (b), 8 > 0 and e > 0, and
for compression, shown in (c), 8 < 0 and e < 0.

5.4

ELASTIC

DEFORMATION

If you bend a steel rod slightly, it springs back to its original shape when
the force is removed. A rubber band will stretch and on release of the tensile
force will resume its original shape. This type of deformation which is lost
upon release of the applied stress is called elastic deformation and occurs in
metals as well as in rubberlike materials. This is the first deformation process
to occur in metals. The amount of bending that a steel rod can withstand depends on the material. After a certain amount of bending, the rod will no
longer return to its original position on the release of the applied load. It is
then said that the elastic limit has been exceeded for the material.

133

134

Mechanical
5.4.1

Behavior

Hooke's l a w

The behavior o f a material under load can best be represented by means of


a stress-strain curve, that is a graph o f stress versus strain. A stress-strain
curve of an elastic material may be linear (Figure 5.3) i f it obeys Hooke's law,
which states that stress is proportional to strain, that is
. ;

.,

or = 6 , n

4;.

(5.4)

where the constant o f proportionality E is called the Young's modulus or elastic


modulus o f the material. N o t all materials obey this relation in the elastic
region, but it is instructive to consider the case o f the ideal elastic material
which conforms to Equation (5.4). Such an ideal elastic-brittle material would
deform elastically until the stress reached a level at which rupture would
occur. This would be the case o f an elastic band stretched to breaking. A t
any point short of rupture, i f the stress is released, the material would resume
its original shape.

Figure 5.3 Stress-Strain Curve for an Ideal Elastic-Brittle Material.

An

ideal elastic-brittle material obeys Hooke's law which states that the stress is
proportional to the strain. This results in a linear stress-strain curve up to the
point where fracture occurs.
.. ,
.,
..
,
.

5.4.2

Young's modulus

^z^: ^ ' - h r

The value o f the Young's modulus varies with the material, and Table 5.1
lists the appropriate values o f this constant for several pure metals. The units
of the modulus are I b / i n ^ . From the values listed, it is evident that i f samples
of aluminum, beryllium, lead, and tungsten were subjected to a stress o f
1000 Ib/in^, then, according to Equation (5.4), the resulting strain would be
given by

Elastic Deformation
T a b l e 5.1.

ELASTIC

PROPERTIES.

Young's modulus
E, I b / i n -

Poisson's ratio

Material
Aluminum
Beryllium
Copper
Gold
Iron
Lead
Magnesium
Nickel
Platinum
Silver
Titanium
Tungsten

9.900.000
43.000.000
18.000.000
11.400.000
28.500.000
2,300.000
6.500,000
30.000.000
22,000,000
1 1,000.000
16.000,000
52.000,000

.34
.01

:-

S h e a r modulus
G , Ib/in^

.42
.28
.45
.33
.31
.39
.38
.34
.27

''i
' _

3,700,000
21,200,000
6,700,000
4,000,000
11,100.000
800,000
2,500,000
11,500,000
7,900,000
4,000,000
6,000,000
20,500,000

and so
1000
9.900.000

^"^

= 1.01 X 10-^ i n . / i n .

(5.6a)

1000
:0.23 X 10-^ i n . / i n .
43,000.000

(5.6b)

1000
2,300.000

(5.6c)

4.35 X 10-^ i n . / i n .

1000
= 0.19 X 10-^ i n . / i n .
52,000,000

(5.6d)

Thus at room temperature at the same stress level, lead undergoes 23 times
more strain than does tungsten.
^
5.4.3

Poisson's ratio

If a cylindrical specimen is subjected to a force acting parallel to the axis


of the cylinder, it is said to be axially loaded. Under such conditions, the
discussion above states that the cylinder will be elongated i f in tension. Can
such an extension occur with no attendant change in the cross-sectional
area? I f it can. this implies that the material may be stretched indefinitely
by applying a tensile stress and retain the same cross section. Such is not
the case in practice. As materials are stretched or compressed, there occurs
a change in cross-sectional area.
As extension occurs parallel to the cylinder axis, in the case of tension,
the diameter of the specimen is being reduced. This radial contraction is
depicted in Figure 5.4(a). where the cylinder sketched with heavy lines is

135

136

Mechanical

Behavior

(a)
/

'

(b)

Figure 5.4 Axial and Radial Strains. As a cylinder is stressed in tension or


compression [shown in (a) or (b), respectively'], the resulting strain occurs in both
the axial and radial directions. In tension, the sample is elongated and its
cross section is reduced, while in compression, the sample length is decreased
but the cross-sectional a r e a is increased. The scale of these drawings is greatly
exaggerated as we may see from the calculations of Equations (5.6) and (5.8).

the final shape assumed by the cylinder sketched with light lines after a tensile
stress <T is applied. (The scale o f these drawings is greatly exaggerated as
will be shown in the calculations below). There is a large extension in the
axial direction as predicted by Hooke's law and an attendant radial contraction. The radial strain is called Poisson contraction in this case and its magnitude and sign are given by the relation
e_, = c =

(5.7)

where
and denote the radial strains along the x- and y-axes and
is the
axial strain along the z-axis. The constant o f proportionality v is known as
Poisson's ratio and is a physical constant o f the material similar to Young's
modulus. Table 5.1 lists the values o f Poisson's ratio for various metals.
A n analogous Poisson strain occurs when a cylinder is subjected to axial

Elastic Deformation
compression (Figure 5.4b). The axial strain
< 0 according to Equation
(5.4); but the radial strains are positive according to Equation (5.7). This
means that as the cylinder is compressed, its cross-sectional area increases.
Thus the radial strains for the above materials subjected to a tensile stress
of a-= 1000 I b / i n ^ are given by Equation (5.7) as
e^^, = -(0.34)(1.01 X 10-') = - { ) . 3 4 x lO-" i n . / i n . ,

(5.7a)

e^^^ = - ( 0 . 0 0 ( 0 . 2 3 X 10-^) = - 0 . 0 0 2 x l O - ' i n . / i n . ,

(5.7b)

e^p^ = - ( 0 . 4 5 ) ( 4 . 3 5 X 10-") = - 1 . 9 6 X 10-" i n . / i n .

(5.7c)

e^^ = -(0.27)(0.19 x 10--') = - 0 . 0 5 X 1 0 - i n . / i n .

(5.7d)

The radial strains are much smaller than the axial strains calculated in Equation
(5.6) by a factor equal to the value of Poisson's ratio, and one must bear this
in mind when studying the drawings of Figure 5.4.
5.4.4

Mechanism of elastic deformation

Imagine two balls joined together by a spring as shown in Figure 5.5(a).


I f a tensile stress is applied, the balls move apart and the spring is extended

(c)
Figure 5.5 Spring Analogy of Interatomic Bonds.

During elastic deforma-

tion, the interatomic bonds are stretched or compressed and they behave like
springs. The bond between two atoms is represented in (a) by an unstretched
spring. The spring is shown after a tensile strain in (b) and after compression
in (c).

137

138

Mechanical

Behavior

(Figure 5.5b), while i f a compressive stress is applied, the balls move closer
together and the spring is compressed (Figure 5.5c). This is but a simple
idea o f the action o f the interatomic bonds under stress. A s a tensile stress
is applied and the specimen undergoes a tensile strain, the interatomic bonds
parallel to the tensile axis are extended while the bonds perpendicular to the
axis undergo Poisson contraction. The opposite occurs if a compressive stress
is applied in place o f a tensile stress. I n this case the bonds parallel to the
compression direction are compressed and the bonds perpendicular to this
axis are extended.

I f we consider polymers we see that the long-chain molecules may exist in


a tangled arrangement and, on the application o f a tensile stress, these chains
tend to be straightened and aligned parallel to the direction of the stress. The
stretching o f interatomic bonds occurs after the alignment o f the molecules.
A simple demonstration o f this can be performed with a band o f natural (unvulcanized) rubber. A s the band is stretched, the molecules should be untangled. I f the stretched band is run across a sponge wet with cold water,
the cooling is sufficient to freeze the chains in the parallel orientation. When
the stress is released, the band does not recover its original position immediately but requires some exposure to room temperature to unfreeze the
molecules and allow the tangled arrangement to be reinstated.
.-

:-n:-:,;i3a:.w,i;,^:i;,-;: -^'O

5.4.5

Jfe'5i>

^iM.-j-Sis.. ^i-cisi-iqitv

, - ,

Volume changes during elastic deformation

Let us consider a cube o f material subjected to a stress as shown in Figure


5.6(a). The cube suffers strains in the x, y , and ^-directions, and the changes
in length are
.
^
,
'

and
8=8=-,
The resultant tetragonal shape is shown in Figure 5.6(b). The original volume
-

^0 = a\

(5.10)

and the final volume V is


K = ( a + S_,)(a + S)(a + 8^).

\)

On multiplying out Equation (5.11), we obtain


V=a^

+ a-8^ + 28^ + a^S, + a8^8^ + a8^8^ + a8^8^ + 8^8/,.

(5.12)

Since the change in length 8 is small compared with the length a, those terms
containing the product o f two or three 6-terms will be very small and may be
dropped to give

Elastic Deformation

Figure 5.6 Volume Change During Elastic Deformation.

A cubic specimen

whose edge length is a is subjected to a uniaxial tensile stress (T^. The original
cubic shape pictured in (a) is transformed into a tetragonal cell of sides a + 8^.,
a -+
and a + 8,, as shown in (b), by the action of this stress. The specimen ii
extended along the z-axis (8^ > 0) and undergoes Poisson contraction along the
X - and y-axes (8^. < 0, 8^^ < 0).

K -

+ a% + a \ a \

(5.13)

and the change of volume A K is found to be


^V=y-y^^^aH8^

+ 8^ + 8^l

(5,4)

I f we substitute into Equation (5.14) the values of the length changes in terms
of strains as given by Equations (5.8) and (5.9), we find that
AV ^ a\{l-2v).

(5.15)

In order for the volume to remain constant during deformation. Equation (5.15)
may be set equal to zero, and we find that
:

- v = o . 5 .

(5.16)

According to the data listed in Table 5.1, Poisson's ratio is normally less than
0.5, and therefore there is a volume change during elastic deformation. There
is a certain type of elastic deformation called shear deformation which does

139

140

Mechanical

Behavior

not involve a volume change but only a shape change as we will discuss later
in this chapter.

5.5

PLASTIC

DEFORMATION

According to the stress-strain curve o f Figure 5.3, as greater stresses are


applied to a material, the strains increase due to elastic deformation. Practical
experience tells us that i f we bend a bar of metal far enough, a permanent
deformation is left in the bar. This is called plastic deformation and is distinguished from elastic deformation in that it is nonrecoverable on release of
the load. I t is the nonrecoverable plastic deformation which allows the many
shape changes we introduce in the fabrication of the products of our civilization.
5.5.1

Ideal rigid-plastic material

Before considering real materials, it is instructive to discuss two ideal materials which are extreme examples of the behavior of real materials under
stress. We will show later that real materials fall between these two extremes.
The two idealized cases are the rigid-plastic and the elastic-plastic materials.
In the rigid plastic (Figure 5.7) there is no elastic deformation. The sample
remains rigid until a definite value o f stress, the yield stress t r ^ , is reached.
A t this level o f stress, plastic deformation begins and continues with no change
in the stress. A bar of such an ideal rigid-plastic material could be stretched
into a very fine wire by just applying and maintaining the yield stress. We will
show that real materials lack this feature except for a few special cases.

O
Figure 5.7 Stress-Strain Curve of on Ideal Rigid-Plastic Material.

An ideal

rigid-plastic material undergoes no elastic deformation. It begins to deform


only when the applied stress becomes equal to the yield stress cr^^ and continues
to flow plastically at this stress level.

Plastic Deformation
5.5.2

Ideal elastic-plastic material

The elastic behavior discussed in Section 5.4 usually occurs in a material


before the onset of plastic deformation, and the ideal elastic-plastic material
(Figure 5.8) is an example of this. As the applied stress increases, the sample
is deformed elastically as is indicated by the linear segment of the stress-strain
curve. A t the yield stress cr^^, the material deforms plastically with no increase
in stress required for continued plastic deformation. The sample at any point
P has a total strain represented by the line OR (Figure 5.8). Part of this total
strain is elastic and part is plastic. The elastic strain is represented by OA
and the plastic by AR.

O A

Figure 5.8 Stress-Strain Curve of an Ideal Elastic-Plastic Material.

An

Ideal elastic-plastic material first deforms elastically in accordance with Hooke's


law until the yield point o"^^^ is reached. At this point the material flows plastically with no increase in stress. At any point P, the sample has a total strain OR
of which OA is the elastic and AR the plastic strain.

5.5.3

Recovery of elastic strain after plastic deformation

Let us consider a sample of an ideal elastic-plastic material which has been


subjected to plastic deformation. The stress-strain curve of such a sample is
illustrated in Figure 5.9. As the stress is increased from 0 to a^^, the sample
undergoes elastic deformation OA. A t the yield stress, point B on the stressstrain curve, plastic deformation begins with no further increase in stress.
When point C of the stress-strain curve is reached, the sample has undergone
a total strain OE. Part of the total strain is plastic, AE, and part is elastic,
OA. Let us now consider what will happen to the sample strained to point C
if the load is released. The elastic deformation is recovered and the stress
decreases linearly with decreasing strain from C to D. The total strain in
the sample is now OD and is entirely plastic. The elastic strain DE was removed during the removal of the load. Applying geometry to the stress-strain

Mechanical

Behavior

Figure 5.9 Recovery of Elastic Strain.

After straining into the plastic re-

gion, the elastic strain is recoverable on release of the applied stress. If the sample is strained elastically from O to 8 and plastically from B to C , the stressstrain curve on the release of the load follows the line CD parallel to OB, and the
total strain O f is reduced by the amount DB which is equal to the elastic strain OA.

curve, we can show that the strain OD is equal to the plastic strain AE. I f
we look at the parallelogram OBCD, we see that
(5.17)

OD = BC,

since they are opposite sides o f a parallelogram. For the same reason, we have
BC = AE.

(5.18)

OD=AE,

(5.19)

Therefore we see that

and the remaining strain is entirely plastic strain. The recovered strain DE
is equal to AE and is just the original elastic strain. This example serves to
illustrate that even after the onset of plastic deformation, the elastic strain
may be recovered on the release o f the load.
5.5.4

Strain hardening

> ^ ;>

. > u

A n ordinary material does not exhibit the plastic flow at constant stress
which is characteristic of the rigid-plastic or elastic-plastic materials. Instead,
the stress increases during plastic deformation (Figure 5.10) and more stress
is required for each further increment o f strain. This phenomenon is referred
to as strain hardening or work hardening. As an illustration of work hardening,
take a piece o f metal wire or rod and bend it. N o w try to unbend it. Notice
that it does not unbend as easily as it was bent. This is the result of work
hardening. Often the wire will bend at a different location on unbending because the material at that point was not strain hardened as was the material

Enfiineering Versus True Stress-Strain

Curves

Figure 5.10

Stress-Strain Curve of a Typical Metal.

A typical stress-strain

curve siiows a metal deforming elastically until the yield point fr^^^^ is reached and
then flowing plastically. As plastic flow occurs, strain hardening sets in, and the
increased stress required for further plastic flow is shown by the steadily increasing stress-strain curve.

at the site of the original bend. The mechanism of strain hardening involves
the interaction of dislocations and will be discussed in Section 5.12.
5.5.5

Necking and fracture

In the early stages of plastic deformation, the entire sample undergoes the
same plastic strain. Because of strain hardening, the stress for plastic flow
increases. This stress eventually reaches a maximum value known as the
ultimate tensile strength (Figure 5.1 la). A t this point in the straining of the
specimen, uniform deformation ceases and localized plastic flow occurs. This
means that at some point along the length of the specimen, plastic deformation
begins to take place at an accelerated rate. This leads to a high local reduction
in cioss section and the formation of a neck. Due to the characteristic appearance of the localized plastic strain (Figure 5.12a). this part of the plastic
deformation process is known as necking. The cross-sectional area of the
specimen becomes very small in this vicinity and eventually cannot support
the applied load. A t this point, fracture occurs in the necked region (Figure
5.12b). Beyond the ultimate tensile strength, the load and stress decrease
with increasing elongation and strain until fracture occurs.
- - < =

5.6

ENGINEERING

VERSUS

TRUE

STRESS-STRAIN

CURVES

The decrease of stress during necking is characteristic of the stress-strain


curve based on the definitions of stress and strain in Sections 5.2 and 5.3,

143

t
(a)

(b)

Figure 5.11 Engineering and True Stress-Strain Curves, (a) An engineering stress-strain curve showing the yield point where plastic deformation commences, the ultimate tensile strength where necking begins, and fracture. The
ultimate tensile strength is the maximum stress recorded, (b) A true stressstrain curve showing the yield point and fracture. This curve does not have a
maximum because necking is accounted for in calculating the true stress and
true strain.

Figure 5.12 Necking and Fracture, (a) The sample of steel stressed beyond
the ultimate tensile stress has developed a necked region, (b) The same steel
stressed to fracture.

Resolution of Forces
where stress and strain are based on the original area and length. During
necking, the cross-sectional area A decreases rapidly and a smaller load is
required to maintain the same or an increased level of stress. Since A^ is
used in Equation (5.1) instead o f A. the stress level falls instead of rising in
the region from the ultimate tensile strength to fracture. Stress and strain
based on Equations (5.1) and (5.3) are known as engineering stress and engineering strain, respectively.
I f the values o f the cross-sectional area A and the length / of the specimen
at any point during testing are used instead o f the original values o f these two
quantities, true stress a and true strain e values may be calculated according
to the following equations:

^ = ^
and

\<
:

- .

? = ln(f)-*

(5.20)

^'i : :4'->

V :: ^

-ii,;'

'

(5.21)

The true stress-strain curve is illustrated in Figure 5.11(b). I t does not exhibit a decrease o f stress at the onset o f necking but rather a continual strain
hardening from the yield point to fracture.
The engineering stress-strain curve provides sufficient information for many
practical applications. The true stress-strain curve provides information o f
a less practical nature but is very important in research investigations since
it provides a more correct picture o f the testing process.

.7

RESOLUTION

OF

FORCES

A force of magnitude P applied to a specimen causes a stress a- on the plane


of cross section as shown in Figure 5.1. The stress cr is a normal stress because it is directed normal to the plane under consideration. I f the plane o f
interest is not normal to the specimen axis, the force vector P acting on this
plane may be resolved into two component vectors: P, normal to the plane,
and P^, parallel to it. Figure 5.13 shows the resolution o f the load into its
normal and parallel components. I f we assume that the plane of interest has
an area A and the vector n normal to this plane makes an angle 0 with the
loading axis, then the load may be resolved into the two components whose
* The term In denotes the natural logarithm to the base e and not the common logarithm (log)
to the base 10. Values of both may be found in various mathematical tables.

SECTION 1.3

o- (ksi)
80
D
60
40
20
0

0.05 0.10 0.15 0.20 0.25 0.30

FIG. 1-12 Stress-strain diagram for a


typical structural steel in tension (drawn
to scale)

o-(ksi)
40
30

20
10
0

0.05 0.10 0.15 0.20 0.25

RG.1-13 Typical stress-strain diagram


for an aluminum alloy

FIG. 1-14 Arbitrary yield stress


determined by the offset method

Mechanical Properties of Materials

21

The diagram o f Fig. 1-10 shows the general characteristics o f the


stress-strain curve for mild steel, but its proportions are not realistic
because, as already mentioned, the strain that occurs from B to C may be
more than ten times the strain occurring from O to A. Furthermore,
the strains from C to are many times greater than those from B to C.
The correct relationships are portrayed in Fig. 1-12, which shows a
stress-strain diagram for mild steel drawn to scale. I n this figure, the
strains from the zero point to point A are so small in comparison to the
strains from point A to point E that they carmot be seen, and the initial
part of the diagram appears to be a vertical line.
The presence o f a clearly defined yield point followed by large
plastic strains is an important characteristic o f structural steel that is
sometimes utilized in practical design (see, for instance, the discussions
of elastoplastic behavior in Sections 2.12 and 6.10). Metals such as
structural steel that undergo large permanent strains before failure are
classified as ductile. For instance, ductility is the property that enables a
bar o f steel to be bent into a circular arc or drawn into a wire without
breaking. A desirable feature o f ductile materials is that visible distortions occur i f the loads become too large, thus providing an opportunity
to take remedial action before an actual fracture occurs. Also, materials
exhibiting ductile behavior are capable of absorbing large amounts of
strain energy prior to fracture.
Structural steel is an alloy o f iron containing about 0.2% carbon,
and therefore it is classified as a low-carbon steel. With increasing
carbon content, steel becomes less ductile but stronger (higher yield
stress and higher ultimate stress). The physical properties o f steel are
also affected by heat treatment, the presence of other metals, and manufacturing processes such as rolling. Other materials that behave in a
ductile manner (under certain conditions) include aluminum, copper,
magnesium, lead, molybdenum, nickel, brass, bronze, monel metal,
nylon, and teflon.
Although they may have considerable ductility, aluminum alloys
typically do not have a clearly definable yield point, as shown by the
stress-strain diagram o f Fig. 1-13. However, they do have an initial
linear region with a recognizable proportional limit. Alloys produced for
structural purposes have proportional limits in the range 10 to 60 ksi
(70 to 410 MPa) and ultimate stresses in the range 20 to 80 ksi (140 to
550 MPa).
When a material such as aluminum does not have an obvious yield
point and yet undergoes large strains after the proportional limit is
exceeded, an arbitrary yield stress may be determined by the offset
method. A straight line is drawn on the stress-strain diagram parallel to
the initial linear part of the curve (Fig. 1-14) but offset by some standard
strain, such as 0.002 (or 0.2%). The intersection o f the offset line and
the stress-strain curve (point A in the figure) defines the yield stress.
Because this stress is determined by an arbitrary rule and is not an
inherent physical property of the material, it should be distinguished
from a true yield stress by referring to it as the offset yield stress. For a

Das könnte Ihnen auch gefallen