Sie sind auf Seite 1von 70

NSW 700: WELDING METALLURGY

CHAPTER 12: Mechanical Testing


Page 111
Copyright Reserved

1.
2. .
3.
4.

5.

6.

7.

8.

9.

10.
11.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.11
Copyright Reserved

12. MECHANICAL TESTING


12.1 Tensile testing
12.1.1 The engineering stress strain curve
The engineering tensile test curve is probably one of the most widely used tests to determine the mechanical
properties of metals. It is simple to perform and can provide significant engineering data on the strength of
materials. It has, however, also its limitations and the full understanding what this test can and can not, do is
essential for the discipline of mechanical metallurgy
In essence the engineering stress strain curve is obtained by pulling in tension a tensile test sample of which the
initial diameter and gauge length have been measured, until failure occurs and plotting the engineering stress
S (as the load P at any point/original cross sectional area A0) as a function of the engineering strain e which
is obtained from the relationship: extension L at any point/original gauge length L0. These relationships are
given once more below in following the accepted notation for the engineering stresses S and strains e to
distinguish them from the later true stresses and strains .
Engineering stress at any point:

P
A0

(Eq 12.1.1-1)

where P is the applied load at any point and A0 is the original cross sectional area of the test sample.
Engineering strain at any point:

L
L0

(Eq 12.1.1-2)

where L is the length increment at any point and L0 is the original gauge length of the test sample. A typical
engineering stress strain curve is shown below.

Figure 12.1(a): A typical engineering stress strain curve for a relatively ductile metal or alloy

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.12
Copyright Reserved

Note particularly the definitions of the:

Offset yield strength S0.2;


The tensile strength SUTS or Smax;
The fracture stress Sf;
The uniform strain eu;
The strain to fracture ef; and
The point where so-called necking begins at eu and Smax.

Engineering tensile strength Rm or Smax


The engineering tensile strength or the ultimate tensile strength (UTS) of a material is obtained from the
maximum load that the material can sustain under uniaxial loading conditions before the onset of plastic
instability. It is commonly quoted and has served a purpose for many decades as a design criterion where a
suitable safety factor was included in the design of a structure or a component to allow for the difference
between the point of yielding and the maximum tensile stress value. With the more advanced understanding of
the role of complex stresses within most structures that are not only loaded uniaxially, however, the more
rational use of basing the design rather on the strength at yielding has become generally accepted. In the SI
units the tensile strength is known as Rm.
Engineering yield strength Rp or S0.2
The offset yield strength ( or sometimes called the proof strength in the UK) is an engineering method to
define the practical onset of plastic yielding where a definite yield point is difficult to establish, as in the above
example for a ductile metal. Mostly the arbitrary offset of 0.2% plastic strain is used to define this point of
yielding as S0.2 or Rp0.2 according to the SI units whereas actual yielding (defined by the onset of the
movement of the first dislocations) has, of course, already started before at the so-called elastic limit.
Sometimes the offset plastic strain of 0.1% or even 0.5% are also used and the correct way of notification is to
include the offset strain as a subscript to the strength designation, i.e. Rp0.1 or Rp0.5 as the case may be.
Note that strictly in the SI units, the notation Rp is usually reserved for hardened and tempered steels whereas
ReL and ReH are reserved for unhardened steels with a clearly defined yield strength range.
The proportional limit is defined as the highest point on the stress strain curve where the stress is still directly
proportional to the strain, i.e. a linear or elastic proportion. The true point of yielding is called the elastic limit
and is defined as the greatest stress that the metal can withstand without any measurable permanent strain after
a complete release of the load. The determination of the elastic limit is a tedious method to perform and requires
a process of loading and unloading with measurements for permanent strain in between. The measurement of the
elastic limit is, of course, dependent on the sensitivity of the strain measuring equipment and at the ultimate point
of an infinite sensitivity, the elastic limit would coincide with the proportional limit (so-called because of the
linear proportion of the elastic strain and Youngs modulus from Hookes law in the elastic region, i.e. S el = e
E).
The term flow stress is also frequently encountered, particularly with high temperature deformation studies, and
this defines a stress level at any point in the plastic deformation portion of the stress strain curve where the
material flows under deformation.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.13
Copyright Reserved

Some of these yield parameters are defined in the Figure below.

Figure 12.1(b): A typical engineering stress strain curve for a ductile material with the criteria for yielding defined
as: point A: elastic limit, point A: proportional limit, point B: yield strength, line CB: offset yield strength,
OC: offset elongation

Note: Distinguishing between the strength level and a stress level needs to be understood.

A strength level, i.e the offset yield strength S0.2 or the UTS Smax, is a specific stress point (or
sometimes called a critical point) on the stress strain curve that defines a material characteristic or
constant and is independent from the test method or the laboratory where the test was performed. In fact
it is a value that can be found even outside the laboratory in hand books.
A stress level S, however, defines any point on the stress strain curve at a given test load P and is
known as a test variable. The difference between a stress (a test variable) and the strength (a material
constant) is sometimes even confused in the literature.
Finally, one also needs to distinguish between an independent test variable (for instance the load P or
diameter D0 or cross sectional area A0 of the specimen, which can mostly be freely chosen) and a
dependent test variable, i.e. the stress S or engineering strain e at any point on the S e curve that
arises from the application of the load P and the value of A0.

Engineering strain e
The onset of yielding (which is fundamentally defined by the onset of permanent dislocation movement under
stress) is most difficult to measure in practice and is a function of the sensitivity of the measuring instrument.
The terms uniform strain and necking are related, with the first term describing the engineering strain e
during which the gauge diameter of the tensile sample reduces uniformly as it elongates homogeneously. At the
point of necking, however, a so-called plastic instability sets in and further elongation is now concentrated
heterogeneously in only a small area of the gauge length and this area is called the necking area. The further
elongation of the sample with applied load is now primarily concentrated in this necked area until final fracture
occurs. In engineering terms, this stress at the maximum load at which necking begins, is called the tensile
strength or the Ultimate Tensile Strength UTS SUTS or Rm of the material.
A measure of ductility
Ductility is a qualitative term that is a subjective property of the material and the short comings of a general term
such as ductility, will be encountered once more in Chapter 5 in the topic of Fracture Mechanics where it will
be seen that a seemingly ductile material can also fracture in a seemingly brittle manner under the right
conditions. Nevertheless, the use of the general term ductility is useful as it is very widely used in practice.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.14
Copyright Reserved

Finding a quantitative expression for the ductility of a metal, is of interest in three areas:

To indicate to which extent a metal can be deformed during cold and hot working operations such as
bending, rolling etc.
To indicate to the designer to which extent a metal in a structure will deform plastically before it
fractures. A metal with a high ductility, will give ample warning before it fractures catastrophically.
Finally, the ductility of a metal can serve as an indicator of the level of impurities in the metal as the
ductility is very often sensitive to these impurities.

From the above stress strain curve, the engineering strain at fracture ef is given by:
ef

L L f L 0

L0
L0

(Eq 12.1.1-3)

where Lf is the measured gauge length at fracture and L0 is the original gauge length. Because a significant
proportion of the fracture elongation will be concentrated in the necked area, the fracture elongation ef will be
a function of the original gauge length L0 and should be quoted always with the value of ef, i.e. for instance
ef(25mm). In line with the international SI system, the fracture elongation is usually known as A 5 and would be
written as A5(25 mm).
The reduction in area in the SI units, is known as Z:
Z

A0 Af
A0

(Eq 12.1.1-4)

where A0 is the original cross sectional area of the gauge length and Af that at fracture. The reduction in area
Z does not suffer from the gauge length problem of the elongation at fracture ef and is independent from the
original gauge length L0. From this, a zero gauge length elongation value has been defined through the
constancy of volume principle, where it is assumed that practically no necking occurs because of the vanishingly
short gauge length (or conversely, that our gauge length is so short that it falls fully in the center of the necked
region):
A0L0 = AfLf
Lf A0
1

And
L 0 A f (1 Z)
and defining the zero gauge length elongation e0:
e0

Lf L0 A0
1
Z

1
1
L0
Af
(1 Z)
(1 Z)

e 0 A S( 0 )

Z
1 Z

(Eq 12.1.1-5)
(Eq 12.1.1-6)

(Eq 12.1.1-7)

(Eq 12.1.1-8)

This expression will represent the elongation at fracture of a specimen with a very short gauge length of L0 0
and e0 or A5(0) will essentially be independent of the initial gauge length.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.15
Copyright Reserved

Another method to avoid the complications of the gauge length dependence of the total strain to fracture e f or
A5, is to use the uniform strain eu until the onset of necking. This term correlates quite well with stretcher
forming operations although the engineering stress strain curve is often quite flat in this region and it is,
therefore, often difficult to find the exact point of the uniform strain on this curve.
The modulus of elasticity
The slope of the initial linear portion of the stress strain curve is the modulus of elasticity or Youngs modulus
according to Hookes law:
E

S el
e el

(Eq 12.1.1-9)

where Sel is the engineering elastic stress at any point on the linear portion of the curve and eel is the
corresponding engineering elastic strain at that same point. The elastic modulus E is also a measure of the
stiffness of the material and is very little dependent on the microstructure of the material. This is basically
because it is derived from the interatomic forces of the lattice atoms and these are not affected by the macroeffects of the microstructure, such as the presence of precipitates, the grain size etc. The elastic modulus is,
however, sensitive to the temperature and generally reduces slightly with an increase in temperature.
The modulus of resilience UR
Another parameter that is sometimes obtained from the elastic portion of the stress strain curve, is the modulus of
resilience UR of the material. This parameter is the ability of a material to absorb energy when deformed
elastically and to return this energy when it is again unloaded. The modulus of resilience UR is usually
measured as the elastic strain energy per unit volume required to stress the material to the elastic limit and is
obtained from the area under the elastic portion of the stress strain curve:
UR

S el e el S el (S el / E) Sel2

2
2
2E
UR

S el2
2E

(Eq 12.1.1-10)

The toughness of a material


The toughness of a material is its ability to absorb energy in the plastic strain range and is a parameter that is
particularly difficult to define from a stress strain curve, as will be seen later in Chapter 4 where Fracture
Mechanics will be covered. In many components such as train couplings, crane hooks, chains etc, it is of interest
to know whether a component will be able to withstand an occasional load exceeding its yield strength without
fracturing catastrophically in a rapid or brittle manner. Where pre-existing flaws in a component may be present
that provide sites for a stress concentration, fracture may even occur at a nominal stress below the yield strength.
This forms the basis for Linear Elastic Fracture Mechanics. In both cases the concept of yield before break
should apply.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.16
Copyright Reserved

The toughness of a material is actually represented by the total area under the stress strain curve and not only
the elastic portion as with the modulus of resilience. This total area is a measure of the total work per unit
volume done on the material before fracture will occur and is often known as UT. This is still a simplistic way of
measuring the fracture toughness of a material as uniaxial stress conditions very seldom exist in real life (for
instance at the tip of a crack) and that the generally accepted way is by measuring the fracture toughness as KIC
for plane strain conditions

Figure 12.1(c): Comparison of the typical stress strain curves for a steel with a high toughness (the structural steel)
and one with a low toughness (the spring steel) with an indication of the modulus of resilience UR for the two steels
also shown as the cross hatched areas under the elastic portions of the curves.

Note that the spring steel is required to have the larger modulus of resilience UR due to its function, but it has
the lower toughness if compared to the structural steel. The toughness of a material is, therefore, a parameter
that is affected by both the strength and ductility of the steel whereas the resilience is affected primarily by
only the strength (or elastic limit) of the steel.
Many attempts have been made to integrate the area underneath a stress strain curve and some approximations
have been proposed as follows:
For a ductile material such as structural steel:
UT SUTS ef

or U T

(S 0.2 S UTS )e f
2

(Eq 12.1.1-11)

where S0.2 is the 0.2% offset yield strength and SUTS the tensile strength of the material.
For a spring steel:

UT

2S UTS e f
3

(Eq 12.1.1-12)

At most these equations are only rough guidelines and attention to the published values of the fracture toughness
KIC should rather be given as the latter is based on sound fundamental principles of the stress intensity K I (or
KII or KIII as the case may be) at the tip of a notch or crack and not only the nominal uniaxial stress S applied
to the structure or component.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.17
Copyright Reserved

12.1.2 The true stress strain curve


The true stress strain curve versus the engineering stress and strain
The engineering stress strain curve is a convenient and easy measurement of some very useful engineering data
on the strength of materials. It has a major shortcoming, however, in that it is entirely based on the original
dimensions of the specimen whereas the constancy of volume principle, dictates that these dimensions must
change as the specimen elongates. The true stress strain curve overcomes this by relating the stress to the
instantaneous cross section and the strain to the instantaneous gauge length at every point during the tensile
test. In an engineering stress strain curve, the significant effect of necking on the cross sectional area is also
ignored with the consequence that the engineering stress appears to fall after necking has started. During necking,
however, the material must continue to work harden and the actual stress during this stage should continue to
rise.
A true stress strain curve is also known as the flow curve of the material as it is based on the actual plastic flow
properties of the material.
The true strain during uniform strain
Up to the point of necking during which uniform or homogeneous deformation occurs, the volume of the gauge
length may be assumed to be constant, i.e. A0 L0 = Ax Lx where Ax and Lx are the cross sectional area and the
gauge length at any point x up to the point of necking.
The true strain at any point will be given by the summation of the actual strains at every point along the stress
strain curve, as follows:
L1 L 0 L 2 L1 L 3 L 2
dL
L
(Eq 12.1.2-1)

......etc
ln
L0
L1
L2
L
L0

A
D
L
ln 0 2 ln 0
L0
A
D

(Eq 12.1.2-2)

From the engineering strain equation:


L L x L 0 L x
e

1
L0
L0
L0

(Eq 12.1.2-3)

ln

and
or

(e 1)

Lx
L0

and

ln(e 1) ln

= ln(e + 1)

Lx

L0

(Eq 12.1.2-4)
(Eq 12.1.2-5)

This equation is valid until the point of necking is reached and means that the true diameter (or the cross
sectional area Ax) need not be measured at every point but can be calculated from the engineering strain e if the
starting values of L0 and A0 are known. Beyond the point of necking where the strain becomes
inhomogeneous, however, the true diameter of the necked region needs to be measured at every point until
fracture occurs.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.18
Copyright Reserved

The true stress


The true stress at any point on the stress strain curve is given by:
P
x
Ax
Using the expression for the engineering stress:
P
S x
A0
one can rewrite the expression for the true stress as;
P A
A0 Lx
P A
and
x 0 x 0

A0Ax A0 Ax
A x L0
L
Therefore:
S x S(e 1)
L0
Therefore:

= S (e + 1)

(Eq 12.1.2-6)

(Eq 12.1.2-7)

(Eq 12.1.2-8)
(Eq 12.1.2-9)

The true stress strain curve


The true stress strain curve is shown schematically below for the case where the calculation of the true stresses
and true strains according to the above equations, were used throughout to fracture.

Figure 12.1(d): The schematic difference between the engineering and the true stress strain curves.

Note that because of the corrective factor of (e + 1) that is applied to the engineering stress S to convert it to the
true stress , the true stresses are higher than the engineering stresses. This also applies to the true strain at
fracture f or point b. The stress analysis after necking has started, is not necessarily a simple one as a necked
region introduces some elements of a mild notch with its triaxial stress distribution, into the test. Often the true
stress strain curve is simply taken as a linear extrapolation from the point a where necking has started, to point
b where fracture occurs.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.19
Copyright Reserved

The true maximum stress u at the maximum load Pmax


The true stress u at the point where true uniform elongation u ends and necking starts, corresponds to the
true tensile strength. Assume that the cross sectional area of the specimen is Au at this point and that the load
has reached the maximum value of Pmax.
The UTS in engineering terms is defined as:
P
S u max
A0
P
and
u max
Au
Eliminate Pmax from these two equations:
A
u S u 0
Au

and

Therefore:

(Eq 12.1.2-10)
(Eq 12.1.2-11)

A
ln 0
Au

(Eq 12.1.2-12)

u = Su exp(u)

(Eq 12.1.2-13)

True fracture stress f


Because the true fracture stress f should strictly be corrected for the triaxial state of the stresses in the necked
region (which is not a simple problem) the calculated true fracture stresses are generally only approximations
and are prone to error. This, therefore, is a parameter that is not used very frequently in practice.
The true fracture strain f
The true strain at fracture would be given by:
f = ln(A0/Af)
where Af is the specimen area at fracture, i.e.
after necking. This parameter represents the maximum strain that the specimen may undergo before it fractures
and should correspond to the engineering strain at fracture ef. Because the earlier equations to convert an
engineering strain to a true strain, are not valid beyond the point where necking starts, one may not use this
conversion directly. It is true, however, that the reduction in area Z is closely related to the true strain at
fracture f through the following equation:
f ln

1
(1 Z)

(Eq 12.1.2-14)

The true uniform strain u


This is the true strain based on the load Pmax until necking starts and it may be calculated either from the length
at the end of uniform strain load Lu or from the cross sectional area at this point Au as the constancy of
volume equation of A0L0 = AuLu still holds up to this point, even with subsequent necking, and:
u ln

A0
ln(e u 1)
Au

(Eq 12.1.2-15)

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.110
Copyright Reserved

The true local necking strain n


This would be the incremental strain required to deform the specimen from the onset of necking to the point of
fracture, i.e. n = (f - u) and is also given by:
A
n ln u
Af

(Eq 12.1.2-16)

where Au and Af are the respective cross sectional areas at the end of uniform strain and at fracture.
The effect of the initial gauge length L0 on ductility determinations in the tensile test
As stated before, the quantitative measurement of the ductility of a metal is quite difficult to define. A seemingly
ductile metal may also fracture rapidly in a brittle mode under certain conditions. Nevertheless, the term
ductility, even if applied rather loosely, is ingrained and has some use in engineering design. Attempts at
determining the ductility of metals via the stress strain curve have, therefore, been undertaken.
The measured elongation in a tensile test depends on the dimensions of the specimen and, in particular, on the
gauge length L0 chosen. The total strain to fracture therefore, consists of two components, i.e. the uniform
strain u and the incremental necking strain n that occurs between the onset of necking and fracture.
The uniform strain u depends to a large degree on the metallurgical microstructure of the material which is
again dependent on the strain hardening exponent n (see later) as well as the shape of the specimen gauge
length, its shoulder profiles etc. The variation of the locally measured strain at an infinitely small gauge length is
shown schematically below.

Figure 12.1(e): Schematic representation of the localized elongation with position along the gauge length of a tensile
specimen with a necked region.

This gauge length effect is shown below for a specimen where the gauge length was varied from about 25 mm to
230 mm.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.111
Copyright Reserved

Figure 12.1(f): Effect of gauge length on the measured elongation during a tensile test.

The extension of a tensile specimen at the point of fracture, can be given by:
L f L f L 0 L n e u L 0

(Eq 12.1.2-17)

where Ln is the necking extension up to fracture and euL0 is the uniform extension obtained from the
engineering strain eu.
The tensile engineering elongation at fracture ef is then given by:
ef

L f L 0 L n

eu
L0
L0

(Eq 12.1.2-18)

This clearly proves that the total engineering elongation ef at fracture, is a function of the initial gauge length L0
with a shorter gauge length L0 leading to greater total elongation ef.
12.1.3 Standardised specimen dimensions
Because of the above dependence of the engineering elongation on the gauge length L 0, specimen sizes have
been standardised through ASTM A370: Standard Methods and Definitions for Mechanical Testing of Steel
products and ASTM E8: Standard Methods of Tension Testing of Metallic Materials.
Location of specimens: The orientation and location of specimens from a product need to be defined carefully
and is covered in the above ASTM Standards. This particularly true in wrought products where rolling or forging
direction needs to be considered but also in steels that had been cooled from a higher temperature. The cooling
rate at the surface of a cooling product is always faster than in the interior and the microstructure of steels (but not
only steels) is particularly sensitive to cooling rate. For instance, specimens cut from near the surface of a casting
will be stronger than for one cut from the interior.
On long bar products of round, hexagonal or square shape, ASTM A370 recommends that the specimens be cut
from the area halfway between the surface and the center of the bar. For forgings, ASTM E8 recommends that
specimens be taken from the thickest part of the forging.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.112
Copyright Reserved

Size and shape of specimens:

Figure 12.1(g): (Top) ASTM A370 nomenclature of tension specimens and (bottom) specimen dimensions for round
and flat specimens.

Note that in round specimens the diameter to gauge length ration is maintained at 4:1 for other specimen
sizes in order to be able to compare elongation values with each other;
Specimen with a 50 mm gauge length from thin flat products. These are defined as sheet, thin plate, flat
wire and strip, band, hoop etc. with a nominal thickness of between 0.13 to 16 mm ;
Specimen with a 200mm gauge length from flat products, such as plate, shapes and other flat products.
In flat products with a thickness greater than 16 mm, round specimens are preferably machined from the
flat product.

Curved specimens: Curved specimens generally result from large pipes cut in the hoop direction or other
irregular shapes. These pose a particular challenge in the production of gas or liquid line pipe as the straightening
or flattening of the curved specimen introduces an oppositely signed strain into the specimen which introduces the
Bauschinger effect and often leads to lower measured strength values. The Bauschinger effect is found where a
specimen is partly strained in tension to a certain flow stress 1 and the stress is then relaxed. There after the
stress is reapplied but now in compression and yielding then takes place at a stress 2 which is significantly
lower than the original 1.
This is illustrated schematically below where the difference in yield strength is shown by the drop (1 2).

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.113
Copyright Reserved

Figure 12.1(h): Illustration of the Bauschinger effect. The specimen was strained in tension along line OA first and at
point A the stress was relaxed according to line AD. Upon reapplication of the tensile stress the specimen will yield
again at point A (or A) and continue along the extension of the original curve. The stress was then reversed and the
specimen restrained in compression and it now deformed along line DB and not line DAA as was the case without a
stress reversal.

In line pipe formation from flat plate, the pipe is first formed into an U and then in a second step into an O
before welding. This introduces a tensile strain after fabrication on the outer skin. After fabrication, a hoop
section is usually cut from the pipe to obtain a tensile test piece in the hoop direction (where the higher stresses
will be in operation). If this curved specimen is then flattened for tensile testing, a compressive strain is now
applied onto the outer skin and this will affect the tensile test curve through the Bauschinger effect.
Barbas law: This makes it imperative that ductility comparisons can only be done on geometrically equivalent
specimens and various equations have been proposed to find the geometrical equivalence. As a general rule one
may use Barbas law that predicts that if the ratio ( A 0 )/L0 is the same between two differently sized
specimens, then geometrical equivalence has been achieved.
Comparing elongation of a round with a flat specimen: ASTM A370 Section VI, provides an empirical
conversion equation for steels:

A
e fl e r 4.47

(Eq 12.1.3-1)

where efl is the measured engineering elongation of a flat specimen with cross sectional area A and gauge
length L, er is the engineering elongation of a standard round specimen with 50 mm gauge length and 12.5 mm
diameter and a is an empirical constant and has been found to be 0.4 for many grades of carbon and low alloy
steels and is 0.13 for annealed austenitic stainless steel. This equation may not be used for flat specimens below
0.64 mm thick and is also not valid for non-ferrous materials.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.114
Copyright Reserved

12.1.4 Tensile testing machine characteristics


In many laboratories the elastic deflection of the cross head and load cell of the testing machine is disregarded as
it often is considered to be much smaller than that of the specimen. This is, however, not always true and in many
instances the elastic stiffness K of the machine needs to be considered. The effect is shown schematically in the
Figure below.

Figure 12.1(i) Schematic sketch of the effect of elastic deformation by the cross head of the machine. (i) before loading,
(ii) after loading in a very stiff machine and (iii) actual elongation with elastic deformation by the machine of M
decreasing the actual elongation.

Note that the actual elongation L is decreased by the amount M = FK where F is the applied load and K is
the spring constant or stiffness of the machine. To model the effect of machine stiffness, consider a tensile test
as shown schematically in the Figure below.

Figure 12.1(j): A schematic diagram of a tensile test with the machines elastic behaviour represented by a spring that
deforms elastically.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.115
Copyright Reserved

Assume that the elastic deformation of the machine includes that of its load cell and cross head as represented by
an imaginary spring with a spring constant of K. Consider, furthermore, that the cross head of the machine
moves at a constant speed of s = dL/dt so that the cross head displacement in a time t is given by (s t). The
total elastic displacement of the machine is given by (K F) where F is the applied force on the specimen and
the plastic elongation of the specimen is given by (p L0) where p is the plastic strain and L0 is the original
gauge length of the specimen. The total displacement (s t) of the cross head is then given by:
s t = K F + p L0

(Eq 12.1.4-1)

The plastic strain p and the plastic strain rate p of the specimen is then given respectively by:
p

{( dL / dt ) t KF}
L0

and

d {dL / dt K(dF / dt )}

dt
L0

(Eq 12.1.4-2)

Note, therefore, that the plastic strain p and the plastic strain rate p are dependent not only on the cross head
speed dL/dt but also on the instantaneous load F and the instantaneous loading rate dF/dt respectively where
both will vary with the applied stress along the stress strain curve. Experimental values of machine stiffness have
produced values ranging from 740 kg/mm up to 2970 kg/mm.
Note from the above equations that the instantaneous loading rate dF/dt varies more sharply during elastic
loading which occurs over a smaller elongation than during the later plastic loading which occurs over a larger
elongation. The stiffness of the machine, therefore, plays a more significant role in determining the yield
strength, particularly where a yield drop occurs as in low carbon steels. It is mostly for this reason that the
determination of the elastic Youngs modulus E from a conventional stress strain curve is not very accurate
unless a machine with a very high stiffness had been used. At the UTS, however, dF/dt = 0 and the measured
plastic strain rate equals the actual strain rate p at this point.
12.1.5 An equation to describe the true stress strain curve
The Hollomon equation
Many attempts have been made throughout the years to express the true stress strain curve mathematically, with
varying success. The equation that appears, however, to possibly come the closest is the so-called Hollomon
equation:
= K n
(Eq 12.1.5-1)
where K is a constant (called the strength coefficient) and n is called the strain hardening exponent. A loglog plot of this equation should provide a straight line as shown below.

Figure 12.1(k): A schematic log-log plot of the strain hardening equation.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.116
Copyright Reserved

The slope of this straight line is n and the value of the strength coefficient K is found at the interpolation of =
1, (or at Z = 0.63) as shown above.
Note: The Hollomon equation is an empirical equation and has no theoretical basis. It is also very seldom that
a metal will show a single straight log-log line over the entire plastic deformed area and it is, therefore, common
to rather find the value of n at relatively low plastic strains where a straight line on the {log vs log } plot is
usually found. In spite of these shortcomings, it still remains a very useful mathematical tool to analyse the
curve.
The value of the strain hardening coefficient n may have various values as shown below.

Figure 12.1(l): The various values that the strain hardening coefficient n may assume, as shown for the plastically
deformed region of the curve.

Where n = 0, a perfectly plastic solid is found (i.e. no work hardening) and where n = 1 a perfectly elastic
solid is found (i.e. a fully elastic material without any plastic deformation). For most metals the value of n
appears to be between 0.10 and 0.50.
The strain hardening coefficient n as a material constant
The strain hardening coefficient n is known as a material constant and is, therefore, affected by the alloy
content and the microstructure of the steel but not by the external test conditions. Values for n have been
determined extensively for steels as well as for other metals and some values are shown below in the Table:
Table: Typical values of K and n for some metals
Metal
Fe 0.05%C steel
SAE 4340
Fe 0.6%C steel
Fe 0.6%C steel
Copper
70/30 Brass

Condition
Annealed
Annealed
Quenched and tempered at 540C
Quenched and tempered at 705C
Annealed
Annealed

n
0.26
0.15
0.10
0.19
0.54
0.49

K (MPa)
77
93
228
178
47
130

The value of n appears to decrease with alloying element additions for steels and the following empirical
equation has been developed for low Carbon mild steels for this purpose:
n = 0.28 0.2%C 0.25%Mn 0.044%Si 0.039%Sn 1.2%Nf
where the element additions are in wt% and Nf = free Nitrogen content in wt%.

(Eq 12.1.5-2)

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.117
Copyright Reserved

In general it has been found that grain size has only a small effect on the value of n and the following empirical
relationship has been proposed for low Carbon mild steels as a function of the ferrite grain size:
n

5
1 / 2
(10 D GS
)

(Eq 12.1.5-3)

where n thus increases with an increase in ferrite grain size DGS where the grain size is measured by the
intercept length method. The evidence for this relationship is, however, not very conclusive and generally it is
accepted that the value of n does not vary significantly with ferrite grain size in steels.
The rate of strain hardening d/d
From the equation one may arrive at the rate of strain hardening d/d as follows:
d(log ) d(ln ) d
n


d(log ) d(ln ) d
or

n
d

(Eq 12.1.5-4)

(Eq 12.1.5-5)

Note: that the strain hardening exponent n is not the same as the rate of strain hardening d/d (also called
the work hardening rate) as the latter is given by the instantaneous slope of the curve at any point
whereas n is a constant.
Alternative true stress true strain equations
Alternative true stress true strain curve equations have been proposed from time to time as deviations from the
Hollomon equation at low strains (< 10-3) and at high strains (> 1) are often found. For instance, it is often
found that the log-log plot leads to two straight lines with different slopes and it has been proposed in such cases
that the following equation should apply:
= K (0 + )n

(Eq 12.1.5-6)

where 0 is considered to be the strain hardening that the sample has received prior to the tensile test.
Another common relationship is the so-called Ludwik equation:
= 0 + Kn

(Eq 12.1.5-7)

where 0 is the yield strength and K and n have the same meaning as with the Hollomon equation. This
equation is probably more acceptable than the original Hollomon equation as the latter implies that at zero true
strain the stress must be zero, yet this is not true. The value of 0 is, of course the yield strength and if we
accept this as the elastic limit, the value of 0 is obtained at the intercept between the elastic portion and the
plastic portion of the true stress strain curve and this has been shown to be at:

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.118
Copyright Reserved

K 1 n
0 n
E

(Eq 12.1.5-8)

In the case of austenitic stainless steels such as the 18/8 variety, the true stress strain curve has been found to
best fit the following equation:
= Kn + exp(K1)exp(n1)

(Eq 12.1.5-9)

where exp(K1) proportional limit and n1 is the slope of the deviation of stress from the Hollomon equation
plotted against .
Some authors have also argued (quite convincingly) that the true strain in any of the above equations, should be
the plastic strain only and that the elastic strain should be subtracted:
p = total el = total (0/E)

(Eq 12.1.5-10)

12.1.6 Instability in uniform strain


A condition to find the point where necking starts, was proposed by Considere and is based on the assumption
that the onset of necking is at the point of maximum load Pmax.
Pmax = u Au
and

dP = u dA + Au d = 0

(Eq 12.1.6-1)
(Eq 12.1.6-2)

as Pmax is at its maximum where dP = 0.


dA d
(Eq 12.1.6-3)

Au u
From the constancy of volume relationship Lu Au = L0A0 at the last point of uniform elongation:

or the instability condition is:

Au dL + Lu dAu = 0
dA
dL
d

Lu
Au
and from the instability condition above:

and

d
u
d

(Eq 12.1.6-4)
(Eq 12.1.6-5)

(Eq 12.1.6-6)

The point of necking at maximum load Pmax can, therefore, be found from the true stress strain curve where it has
a subtangent of unity (i.e. d/d = u/1) or where the rate of stain hardening d/d equals the stress u.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.119
Copyright Reserved

Figure 12.1(m): The geometrical representation of the necking criterion where the true strains are used.

By introducing the engineering strain e rather than the true strain the construction becomes somewhat
simpler.
d d de d (dL / L 0 ) d L u


d de d de (dL / L u ) de L 0

Therefore:

d
(1 e u ) u
de

u
d

de (1 e u )

Figure 12.1(n): Consideres construction using the engineering strain instead of the true strain

(Eq 12.1.6-7)

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.120
Copyright Reserved

Note that in this construction, the engineering strain e is plotted at the bottom against the true stress on the
vertical axis. In this construction the point of the end of uniform strain u need not be estimated from the curve
as was the case in the upper Figure where the true strain was plotted. Here all one needs to do is to find the
point A of the negative engineering strain = 1 on the horizontal axis and draw a straight line to point C
where the straight line will form a tangent to the stress strain curve. Note that this straight tangent should cross
the true stress axis at point D where OD equals the engineering stress at necking or Rm or the UTS.
Finally, one may also substitute the Hollomon equation at the point of plastic instability and find the two
equalities that were also derived above:

d
n u
d
u

Therefore:

and

d
u
d

u = n

(Eq 12.1.6-8)

(Eq 12.1.6-9)

It has also been shown that this relation ship of u = n is independent of the power law used and will also apply
if equations other than the Hollomon equation have been used.
12.1.7 The yield point phenomenon and strain ageing
The stress strain curve
The yield point phenomenon is particularly of interest in Fe C and Fe C N alloys where the interstitially
dissolved C and N atoms may segregate to dislocations, thereby effectively locking them against movement
until a certain breakaway stress is reached. Once these dislocations have torn themselves loose from the cloud
of C or N atoms, the stress required to move the dislocation is now lower than the breakaway stress and a
so-called upper yield point and lower yield point is observed in these alloys.

Figure 12.1(o): Typical yield point behaviour of a low Carbon steel

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.121
Copyright Reserved

Note the definition of the upper yield point, the lower yield point and the yield point elongation during which
so-called Lders bands at about 45 to the tensile axis of the specimen, form heterogeneously across the
specimens gauge length, progressively covering almost the entire gauge length. As is evident from the above
Figure, upon straining, these steels will elongate elastically first until a sharp yield point, the upper yield point, is
reached. Then a sudden drop in load will be observed with a subsequent fluctuation about some seemingly
constant value of the load before it rises again steadily as for a normal stress strain curve. Typically, the first
Lders band will form at a stress concentration such as an inclusion or a notch and more than one can form at the
same time. Where more than one Lders band is formed during the yield point elongation, the lower yield stress
will fluctuate, each time dropping slightly as a new band forms. The Lders bands are also sometimes called
stretcher strains or even Hartmann lines.

Figure 12.1(p): Lders bands formed on a SAE 1008 low carbon steel with a tensile elongation just beyond the upper yield point.

Static strain ageing after deformation


Strain ageing is used to describe the strengthening behaviour of a low Carbon steel that had been deformed
through the yield point (often called temper rolling or skin pass rolling) and that regains its strain ageing
strengthening at relatively low temperatures after a certain amount of time. This is shown quite clearly below for
an Fe 0 0.03%C steel that had been cold deformed by about 4% strain to take it past the initial locking of the
dislocations by the Carbon atoms and then annealed for different times at 60C where the return of the yield
point was determined as the Carbon atoms diffused back to the dislocations.

Figure 12.1(q): Strain ageing of an 0.03%C steel after straining first to about 4% and then ageing at 60C for different times.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.122
Copyright Reserved

Note the following:

The initial strain of 4% past the initial yield point had removed the yield point phenomenon completely
(see bottom curve for no ageing) and the curve continues where the initial stress strain curve had been
interrupted at 4% strain.
Ageing for only a few minutes at 60C re-introduces the yield point once more with the upper yield
strength increasing with further ageing at 60C. This is proof of the high diffusion mobility of C and N
atoms in -Fe at these relatively low temperatures.
After about 126 hours little further strain ageing occurred and basically all the dislocations had been
filled with C and N atoms.
Both the upper yield strength and the yield elongation appear to increase as the extent of dislocation
locking increases until a point of saturation is reached.

Dynamic strain ageing during deformation


Dynamic strain ageing occurs while the deformation proceeds and mostly leads to a serrated stress strain curve.
As the C and N atoms must now move with the already moving dislocations (due to the plastic strain) and
keep up with them, dynamic strain ageing is sensitive to the temperature and the rate of straining . Note
from the Figure below where straining at 25C leads only to static strain ageing even after a static annealing
treatment of 2 hours at 200C but at 200C during deformation, dynamic strain ageing has set in and causes
the serrated flow curve.

Figure 12.1(r): Curve (i) Dynamic strain ageing at 200C as compared to curve (ii) that was statically strain
aged at 25C in an Fe 0 0.03%C steel. Note that the static specimen was also statically aged at 200 C for 2 hours
before resuming deformation at 25C.

Dynamic strain ageing generally has the following characteristics:


It is accompanied by a high rate of work hardening (see Figure below);
The flow stress during dynamic strain ageing has a negative strain rate dependence, i.e. if the strain rate
increases, the flow stress will decrease;
The stress strain curve is serrated as the C atoms repeatedly capture the dislocations but are almost
immediately torn off again by the strain moving the dislocations; and
The hardness (or stress at a given plastic strain at a constant strain rate) goes through a maximum as a
function of the temperature.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.123
Copyright Reserved

As the deformation temperature is increased beyond a certain point, the serrations will disappear.

Figure 12.1(s): Stress strain curves for an Fe 0.03%C steel as function of temperature where dynamic stain ageing
occurs between about 100 and 200C

12.1.8 Effect of strain rate


The effect of rate of strain application during a tensile test may have an influence on the strength values found
and normally the strain rate is prescribed in the ASTM specifications. In most cases will the yield strength
increase if the strain rate is increased and this is something to consider in industry where rapid processing is
often necessary for purposes of productivity.
d
and is measured in s-1.
(Eq 12.1.8-1)
dt
The increase in flow or yield strength, measured at various strains on the stress strain curve, is shown below as a
function of the strain rate of testing.

The strain rate is defined as:

Figure 12.1(t): The true yield stress at various strains on the stress strain curve as a function of the testing strain
rate for a low C steel at room temperature.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.124
Copyright Reserved

Note in particular that the flow stress for the low C steel is only strain rate sensitive at strain rates in excess of
about 10-3 to 10-2 s-1 when the tests are performed at room temperature. At higher temperatures this strain rate
sensitivity is far more pronounced, as will be seen later when the topic of creep will be covered.
12.1.9 Effect of temperature
The flow stress and ductility of metals normally vary significantly with the temperature of testing although the
general characteristics of the stress strain curve remain the same. Generally the flow stress increases as the
temperature is lowered and conversely, the flow stress decreases as the temperature is increased. This is
illustrated for stainless steel AISI 304 in the Figure below.

Figure 12.1(u) Tensile stress strain curves for AISI 304 stainless steel (Left) at low temperatures and (Right) at
elevated temperatures.

The strength of metals at elevated temperature is of particular importance for the metallurgical engineer
as the strength performance may be influenced very significantly by optimising the microstructure of the
alloy. Some general yield strengths and tensile elongations, as a function of temperature, are shown
below for a number of commercial alloys designed for elevated temperature use.

Figure 12.1(v): (i) The tensile yield strength and (ii) the elongation of various commercial alloys at elevated
temperature MAR-M200 is a Ni based alloy, MAR-M509 is a Cobalt based alloy, 304 is AISI 304 stainless steel,
7075 T6 is an Al Mg Zn alloy and 1015 is a low Carbon SAE steel..

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.125
Copyright Reserved

Note the peak in yield strength of the MAR-M200 Ni based alloy at about 700C. This is due to the
strengthening precipitate - Ni3Al. The ductility does not always behave predictably and the existence of a
ductility trough in some alloys is usually related to some microstructural effects.
12.1.10 The notched tensile test
The necked region as mild notch
It was stated before that the necked region during non-uniform strain is really a mild notch where a triaxial
stress distribution exists. A triaxial stress distribution will have both radial stresses r and tangential stresses
t which will raise the value of the longitudinal or axial stress x required to cause plastic flow. The average
stress at the neck which is found by dividing the axial tensile load by the minimum cross sectional area of the
specimen at the neck, is higher than the stress that would be required to cause flow if only simple tension would
apply as in a uniaxial stress distribution. This is shown schematically below.

Figure 12.1(w): The schematic stress distribution at the neck of a tensile tested specimen. R is the radius of the neck
with a minimum diameter of a.

Bridgeman provided a corrective equation to allow for the triaxial stress distribution in the necked region, with
the assumptions:

The contour of the neck has regular radius R and is approximated by the arc of a circle;
The cross section of the neck remains circular;
The strains are constant over the cross section of the neck; and
The Von Mises criterion for plastic yielding applies

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.126
Copyright Reserved

According to Bridgeman the uniaxial flow stress corresponding to that that would have existed in the tensile
test if no necking had occurred:

x ( ave)
[{(1 2R ) / a}{ln(1 a / 2R )}]

(Eq 12.1.10-1)

This correction generally lowers the calculated true stress at fracture f from that calculated without the
correction.
Notched tensile tests
Ductility measurements on smooth and regular tensile specimens do not always reveal a potential weakness in the
metallurgical structure because of the absence of a triaxial stress distribution in such an unnotched specimen.
Notch sensitivity tests are, therefore, often required and these are done on a pure comparison basis, i.e.
comparing a notched with an unnotched specimen tested under the same conditions and made from exactly the
same material.
Definition: The tendency for reduced ductility in the presence of a triaxial stress distribution and steep stress
gradients, is called the notch sensitivity of the material.
A notched tensile test is usually done to reveal the notch sensitivity of the material. The dimensions of the notch
are prescribed in various standards and are typically a 60 notch with a root radius of 0.025 mm or less. The
depth of the notch is usually chosen to be such that the remaining cross sectional area of the specimen after the
notch has been introduced, is about one half of the unnotched specimen. In the notched specimen the notch
strength is defined by the maximum load divided by the original cross sectional area at the notch.
The notch sensitivity is determined by the Notch Sensitivity Ratio or NSR:
NSR

S max( notched)
S max( unnotched)

(Eq 12.1.10-2)

If the NSR < 1, then the material is notch brittle. The reduction in area at the notch Znotch may also be
measured as a parameter of ductility.
In general and as a rule of thumb, the Notch Sensitivity Ratio NSR of a material will decrease (i.e. it becomes
more notch brittle) as the metallurgical structure increases in strength and hardness. This is because a harder
material will generally, restrict the flow of material at the root of a notch more than in a ductile material and
plastic flow will take place less easily and fracture will occur more easily in the harder material.
The sensitivity of the notched strength to the metallurgical structure, is shown below where it should be noted that
the conventional elongation measurement in the bottom Figure, was unable to detect any potential notch
sensitivity in the specimens that had been tempered deliberately in the temper embrittlement range, i.e in the
range 330 to 480C and it was only apparent with the notched tensile strength results in the top Figure.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.127
Copyright Reserved

Figure 12.1(x): The tensile properties of notched and unnotched specimens of a low alloy steel that was quenched and
tempered through the range that causes temper embrittlement.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.21
Copyright Reserved

12.2 The hardness test


12.2.1 Brinell hardness
The Brinell hardness test
Basically all hardness tests measure the resistance to plastic deformation of the material through an indentation
method. Because plastic or permanent deformation takes place under the indentation, the hardness test
inherently is also a measure of the work hardening of the material.
The Brinell hardness test (ASTM E 10) is a simple test in which a hardened steel ball of a pre-determined
diameter D is placed under load P for a certain time onto the surface of the material of which the hardness is
to be measured, and the size of the indentation is measured afterwards. This is shown schematically below.

Figure 12.2(a): (i) The schematic arrangement of the Brinell hardness tests with (ii) the measurement of the
projected diameter d of the indentation and (iii) the geometry of the indentation.

The Brinell hardness number is defined as:


BHN

P
{D / 2}{D D d }
2

P
Dt

(Eq 12.2.1-1)

where d is the diameter of the indentation as measured on the surface and t is the depth of the indentation,
both measured on the material after the test, as shown above. In practice, a typical load of 3000 kg with a
hardened steel ball with a minimum hardness of 850 HV (Vickers hardness) and of diameter 10 mm, is used
for medium hard materials such as most steels and cast irons and the load is normally applied for about 30
seconds. With a softer material such as Al and its alloys, the load may be decreased to 500 kg and for a
very high hardness, a Tungsten ball is usually used to avoid plastic deformation of the ball itself during the
test. The diameter of the indentation ball may also be varied typically between 5 and 10 mm although a 10
mm ball is the standard. Although the above equation may be used in arriving at a Brinell number, these
values are usually read of suitable tables based on the above equation and are given in the units kg/mm2.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.22
Copyright Reserved

It is normally recommended that the load/ball ratio be selected to give an indentation of between 2.5 to about
6 mm which will provide an indentation diameter d of between 25 and 60% of the balls diameter. The
indentation should always be measured in two perpendicular directions and the mean value taken as the reading.
The maximum range in which the Brinell test will provide reasonably reliable hardness measurements, is
normally between BHN 16 for soft Al and BHN 630 for hardened steel.
Matters to care for in the Brinell test

The indentation ball diameter: The Brinell hardness test has some shortcomings, with the main one
being that the indentation load P via the above expression does not give a mean pressure on the
indentation as the surface area of the indentation is accommodated in the equation and not the
projected area. This brings about that the Brinell number may not be fully independent of the
diameter of the indentation ball D or the load P. Converting the above equation to a geometric one in
which the included angle is used (see Figure (iii) above) the BHN is given by:
BHN

P
{( / 2)D (1 cos )}
2

(Eq 12.2.1-2)

In order to obtain a standard BHN number independent of the ball diameter D, the value of the
included angle 2 must stay constant and the ball diameter varied according to the ratio:
P1
P
(Eq 12.2.1-3)
22 cons tan t
2
D1 D 2
Maintaining this ratio may not always be convenient and the BHN from one material to another may,
therefore, vary with load and ball diameter.

Measurement of the indentation: The indentation diameter d is usually measured by an optical


microscope and should be measured to the nearest 0.01 mm. The plastic deformation around the
indentation, may also lead to some measurement errors if so-called ridging or :sinking takes
place.

Figure 12.2(b): (i) Ridging type (ii) sinking type and (iii) flat type indentation of the Brinell test.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.23
Copyright Reserved

Note that in the ridging type indentation, the ridges diameter d extends above the surface of the
work piece and in the sinking type, the edge of the indentation is below the surface of the work piece.
Cold worked metals and decarburised steels are most likely to show ridging and fully annealed and
lightly case hardened steels will often show the sinking type of indentation. Both ridging and sinking
bring some uncertainty to the measurement of the true indentation diameter d and should be avoided,
i.e. when ridging is present, the apparent diameter of the indentation is greater than the true value and
when sinking occurs, the apparent diameter is smaller than the true diameter.

Anisotropy in the material: When the material to be tested is anisotropic, i.e. its grain structure varies
with direction due to rolling, for instance, the indentation will not be round. Here the mean value of
four readings at 45 to each other on each indentation and also some readings in the various directions
of rolling must be used.

The thickness of the material: The Brinell test may not be used on very thin sheet material as the
plastic deformation under the indentation may protrude onto the work table below the specimen. A rule
of thumb is that the work piece should have a thickness of at least 10 times the depth of the
indentation.

Flatness of the surface: Surfaces with a radius of curvature of less than 25 mm should not be tested
by the Brinell method otherwise the shape of the indentation becomes uncertain.

Surface finish: The measurement of the indentation diameter d is best done on a finely machined,
ground or polished surface. As the indentation is usually visible with the naked eye, the Brinell test
should not be used where the presence of the indentations may later be aesthetically unpleasing or may
weaken the structure.

Spacing of indentations: As plastic deformation occurs at and near an indentation, they should not be
spaced closer than about three times the indentation diameter to each other and should also not be
closer than the same distance from the edge of a sample.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.24
Copyright Reserved

Brinell hardness testing machines


A wide variety of Brinell hardness testing machines are available on the market, ranging from hydraulic to
dead weight, from stationary to portable, from laboratory to automatic production machines, etc. A typical
hydraulic Brinell hardness testing machine is shown below.

Figure 12.2(c): A hydraulic Brinell hardness testing machine

12.2.2 Meyer hardness


The Meyer hardness test
The Meyer hardness test is very similar to the Brinell test and actually uses the same hardness testing machine,
with the important difference that Meyer proposed that rather than measuring the surface area the projected
area should be measured as this would accommodate the shortcoming of the traditional Brinell test on its
dependence on the load and indenter diameter. In the Meyer hardness number H M the relationship between
load P and the projected area of the indentation, therefore, rather uses the mean pressure and provides the
following relationship:
P
HM 2
(Eq 12.2.2-1)
r
or if the diameter of the indentation d is used:

HM

4P
d 2

(Eq 12.2.2-2)

The Meyer hardness HM also has the units of kg/mm2, the same as the Brinell hardness number.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.25
Copyright Reserved

Because of the difference in defining the area of the indentation, the Meyer hardness is less sensitive to the
load P and diameter D of the indenter and should be a more representative and fundamental hardness
measurement technique. For a cold worked material, the Meyer hardness is essentially constant with load
whereas the Brinell hardness decreases as the load increases. For an annealed material, the Meyer
hardness will increase continuously with load whereas the Brinell hardness will first increase and then
decrease with load.
For various reasons, however, the Meyer hardness is seldom used in industry.
Meyer did propose the following empirical rule for the choice of load and indentation diameter d:
P=kdm

(Eq 12.2.2-3)

Where d is the indentation diameter, m is a material constant related to the materials work hardening
exponent n (see earlier) and k is a material constant that expresses the materials resistance to deformation.
For fully annealed materials, the value of m is about 2.5 while m 2 for strain hardened materials and the
relationship to the Hollomon strain hardening coefficient n is given by the following empirical rule:
mn+2

(Eq 12.2.2-4)

Indentation by an indenter
It is worthwhile to understand the plastic flow of a metal more fully as it is deformed by an indentation load P.
Although the indenter deforms the material at the surface plastically and leaves an indentation behind, there is
an underlying volume of material surrounding this plastic zone that will only be stressed elastically, as shown
below.

Figure 12.4(d): (i) The elastic and plastic zones underneath an indenter and the typical deformed grid pattern in
(ii) soft clay and (iii) in steel which shows typical compressive strain field lines (contours close to each other) at
the surface of the indenter with elastically stressed contours further below..

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.26
Copyright Reserved

Note that the plastically deforming zone is not entirely free to strain in any direction and is constrained by the
surrounding elastically stressed material, almost as in a closed-die forging operation where the flow of the work
piece is constrained by the die itself. The mean compressive stress required to cause plastic flow underneath the
indenter in a hardness test is, therefore, greater than in an ordinary compression test without any constraint.
Various theories have been proposed to account for the effect of this constraint factor in the Brinell and Meyer
hardness tests with one of these plastic/elastic analyses predicting that the mean pressure between the indenter
and the indentation pm 3 0 where 0 is the flow stress of the material and the factor 3 is the constraint
factor
12.2.3 Rockwell hardness
The Rockwell hardness test
As with the Brinell and the Meyer hardness tests, the Rockwell hardness test (ASTM E18) relies on the
principle of an indentation but now under a load increment (and not an absolute or fixed load as in the
Brinell and Meyer tests) and the incremental depth of the indentation is measured as representing the
Rockwell hardness of the material.

Figure 12.2(e): The principle of the Rockwell hardness test with the 120 spherical-conical diamond indenter
shown on the right for testing the hardness of steels and other relatively hard materials.

As shown above, the indenter typical for steels is a 120 spherical-conical diamond although a ball shaped
indenter may also be used for softer materials. The load is applied in two steps, i.e. first a lower or minor load
which defines the zero indentation depth and then a higher or major load is applied for about 5 to 10
seconds. Thereafter the major load is removed whilst maintaining the minor load and the incremental
indentation depth is measured. This test, therefore, differs in principle from the Brinell and the Meyers
hardness tests

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.27
Copyright Reserved

The main reason for the two step process, is to remove any surface effects of the specimen (such as a hard
oxide layer) and also back lash on the hardness testing machine itself.
The Rockwell hardness HR is read off suitable tables (or directly from a graduated dial gauge) according to
an applicable hardness scale with over 30 scales that represent different combinations of load with indenter
(diamond or steel ball) and size of the indenter. Examples here would be the Rockwell C scale with the
hardness expressed as HRC or the Rockwell B scale with the hardness expressed as HRB and many others.
The HRC and HRB are probably the most commonly used scales on most metals and alloys.
Selecting the most appropriate Rockwell scale
On steels, Brass and most other metals either the HRC (for the harder materials) or HRB (for the softer
materials) will be sufficient. On very thin sheet materials or metals other than steel or Brass, however, some of
the other Rockwell hardness scales may be more appropriate. Factors that must be considered in the selection
of the Rockwell scale, include:

Type of material;
Specimen thickness;
Test location; and
Scale limitations

As an example, if a hardened steel or a WC drill bit has to be tested for hardness, the choice is already limited
to a diamond indenter. This leaves only 6 possible scales, i.e. Rockwell C, A, D, 45N, 30N or 15N. The
next step is to examine each scale for its accuracy, its sensitivity and its repeatability for the particular
purpose in mind.
To check whether the thickness of the specimen is large enough for the particular scale to be used, the
following empirical equations may be used to estimate the depth of indentation and to keep the specimen
thickness at least 10 times above that.
For a diamond indenter:

Depth of indentation (mm) = (100 HRC) 0.002

For a ball shaped indenter: Depth of indentation (mm) = (130 HRB) 0.002
If eventually a choice between two equally appropriate Rockwell scales is arrived at, choose the one with the
heavier load as this will spread the load over a more representative area in the specimen.
Table: Selection table for the Rockwell hardness test
Scale
symbol

Indenter

Major
load
kg
60
100
150

A
B
C

Diamond
1.588 mm steel ball
Diamond

Diamond

100

E
F

3.175 mm steel ball


1.588 mm steel ball

100
60

Typical application

Cemented carbides, this steel and shallow case hardened steel


Copper alloys, soft steels, Al alloys, malleable iron
Steel, hard cast irons, pearlitic malleable iron, deep case
hardened steel, and all other materials harder than 100 HRB
This steel and medium case hardened steel and pearlitic
malleable iron
Cast iron and Al and Mg alloys, bearing metals
Annealed Cu alloys, thin soft sheet steel

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.28
Copyright Reserved

Table continued
G

1.588 mm steel ball

150

H
K

3.175 mm steel ball


3.175 mm steel ball

60
150

6.35 mm steel ball

60

6.35 mm steel ball

100

6.35 mm steel ball

150

12.70 mm steel ball

60

12.70 mm steel ball

100

12.70 mm steel ball

150

Phosphor bronze, Cu - Be alloys, Malleable irons, with an upper


limit of HRG = 92 to avoid flattening of the steel ball
Al, Zn and Pb
Bearing materials and other soft or thin metals and use the
smallest ball and heaviest load
Bearing materials and other soft or thin metals and use the
smallest ball and heaviest load
Bearing materials and other soft or thin metals and use the
smallest ball and heaviest load
Bearing materials and other soft or thin metals and use the
smallest ball and heaviest load
Bearing materials and other soft or thin metals and use the
smallest ball and heaviest load
Bearing materials and other soft or thin metals and use the
smallest ball and heaviest load
Bearing materials and other soft or thin metals and use the
smallest ball and heaviest load

A Rockwell hardness testing machine


Although a number of variations exist on the market, most use the principles demonstrated in the following cut
away view.

Figure 12.2(f): Cut away view of a typical Rockwell hardness tester

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.29
Copyright Reserved

12.2.4 Vickers hardness as well as the Vickers Microhardness test


The Vickers hardness test
This hardness test relies once again on the principle of a diamond indentation made under load and measuring
the diagonal width of the square indentation made by a square diamond indenter with an included angle of
136 between opposite faces. The included angle of 136 and size of the diamond indenter was chosen with
reference to the Brinell test from which this was the most optimal included angle to diameter ratio. Because of
the shape of the indenter, this hardness test is often called the Diamond Pyramid Hardness test or DPH
although the acronym for Vickers hardness VHN or VPH are also used. The Vickers hardness is given by:

DPH

2P sin( / 2) 1.854P

L2
L2

(Eq 12.2.4-1)

where the included angle = 136, L is the average length of the diagonals of the indentation and P is the
applied load in kg. The shape of the diamond indenter is shown below.

Figure 12.2(g): Diamond pyramid indenter used in the Vickers hardness test with the length of the diagonal D (or
L in the formula above) shown.

The HV hardness values are usually read of tables that provide a hardness factor that must then be multiplied
by the load used, for example: If the average diagonals are measures as 40.3 m, the Vickers table gives the
hardness factor as 1.142. If a load of 500 kg was used, the Vickers hardness is given by 1.142 x 500 = 571
HV.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.210
Copyright Reserved

Advantages and disadvantages of the Vickers hardness test


The Vickers hardness test has reached wide acceptance as it measures basically all hardnesses on a single
scale from DPH = 5 to DPH = 1500. With the Brinell and the Rockwell hardness tests the scale must be
changed which makes it sometimes difficult to compare different materials with each other through their
hardnesses. Furthermore, because the diamond shaped indentations are geometrically similar no matter what the
load is, the Vickers hardness should, therefore, not be as dependent on the load as with the other two tests.
One problem to avoid is the earlier ridging and sinking that may also be encountered with the Vickers
hardness test, as shown schematically below.

Figure 12.2(h): (i) Perfect indentation, (ii) sinking and (iii) ridging in the Vickers hardness test

Surface preparation of the specimen for a Vickers hardness test is more stringent than in the case of the
Rockwell test and at very low loads of about 100 grams, a metallographically polished surface needs to
prepared. Except for very low loads, the Vickers hardness value (and the Knoop hardness, used for very hard
or ceramic materials) is reasonably independent of the load, as shown below:

Figure 12.2(i): The relationship of hardness number versus load for Vickers and Knoop hardness tests

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.211
Copyright Reserved

Vickers hardness tester


A typical Vickers hardness testing machine is shown below.

Figure 12.2(j): A typical Vickers microhardness testing machine.

Vickers Microhardness
The Microhardness testing of materials is basically done via the Vickers hardness test on a laboratory bench
apparatus also equipped with an optical microscope to position the indenter very accurately within the selected
area of the microstructure. It is well suited to determine the hardness of different constituents in a
microstructure or to determine the hardness gradients in case hardened, surface hardened or nitrided steels.
Polished sample surfaces are usually necessary for microhardness testing.
If the hardness of a hard and specific micro-constituent such as a carbide particle in the example above, needs to
be determined, the Knoop hardness is appropriate as its very oblong indentation makes it possible to place two
indentations quite close to each other. The depth and area of a Knoop indentation is typically only about 15%
of the equivalent Vickers indentation and this makes the Knoop test also applicable to test the hardness of thin
surface layers such as hard Chrome plating etc.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.212
Copyright Reserved

12.2.5 Other hardness tests


A number of less common hardness test methods exist, which include:

The Scleroscope hardness tester which is a light and portable instrument in which a diamond tipped
hammer is released to drop onto the component and the height of rebound is measured as an indication
of the hardness. Although not as accurate as some of the other tests above, it is very convenient to test a
number of large components on the shop floor such as castings or forgings.
The Durometer is a hand held instrument to test the hardness of very soft materials such as rubber and
plastics and also operates on the principle of an indentation.
The scratch hardness tests are very crude but long standing comparative tests in which the well known
Moh hardness scale for minerals and the so-called file hardness test for steels, are used. The latter
consists of standard files heat treated to a hardness between 67 and 70 HRC and the comparative test
consists of drawing the file across the component. If the file does not bite the hardness of the
component is designated as file hard.
Ultrasonic hardness testing is a technique well suited to the automatic testing of a large number of
components moving past on a conveyor belt with up to 1200 parts per hour that can be tested. It
consists of a stylus with a Vickers diamond tip or a Rockwell indenter and uses a light indentation
load of maximum 800 grams. The principle of measuring the indentation rests on measuring the
natural resonant frequency of the stylus which consists of a magnetostrictive metal rod with the
indenter attached to its tip. As the indenter rests on the component, the resonant frequency changes and
this is calibrated to the specified hardness of the component being measured.

12.2.6 Hardness conversion relationships


It is often required to compare one hardness test value with another one from a different type of hardness test.
Such conversion tables are available in the ASM Metals Handbook and other reference books. It is important
to accept, however, that these conversion tables are based on purely empirical grounds, as there is no
fundamental relationship between the different tests. The most reliable conversion tables are probably those to
convert between Rockwell, Brinell and Vickers hardnesses for steels with a hardness above 240 Brinell
hardness.
12.2.7 Relationship between hardness and the tensile strength
Relationship with the UTS of a metal
As the determination of hardness of a metal requires minimal preparation of the specimen and also is a very
quick test, various attempts have been made to correlate hardness with other mechanical properties. As the
hardness test involve plastic deformation of the component by the indenter, at best it should correlate with the
ultimate tensile strength of a metal, which is measured after work hardening and not with the yield strength
which is measured before work hardening.
For heat treated plain carbon and low alloy steels, a useful correlation between the Brinell hardness and the
UTS of the steel, has empirically been determined as:
UTS (MPa) = 3.4 BHN

(Eq 12.2.7-1)

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.213
Copyright Reserved

Relationship with yield strength


In spite of the fact that a hardness test measures the resistance of a metal to plastic deformation by an indenter,
some attempts have been made to correlate hardness with the yield strength of a metal, with varying success.
This varying success is not surprising as the fundamentals of the two determinations (hardness and yield
strength) are vastly different and may be compared at best, only in an empirical way.
Here the Meyer hardness test was found to correlate best as is shown below for Meyer hardness values
obtained from different load applications compared to the flow stress obtained from compression tests.

Figure 12.2(k): Comparison of the flow stress curve from compression tests with the hardness measurements from
Meyer hardness determinations for mild steel and Cu..

The above correlation was based on an elastic-plastic analysis of an indentation that concluded that the true
strain during an indentation is given by:
0.2d
(Eq 12.2.8-2)

D
where d is the diameter of the indentation produced by an indenter with diameter D. By varying the ratio d/D
through load variations, the true stress true strain curve was reasonably approximated.
In another empirical comparison, the following expression was obtained for the correlation of the tensile flow
stress 0 and the Vickers DPH value:
DPH
m2
0
(0.1)
3

(Eq 12.2.8-3)

where 0 is the flow stress in kgf/mm2, DPH is the Vickers hardness pyramid number and m is the Meyer
exponent in Meyers law of (m = n + 2) where n is the Holloman work hardening exponent.
Although these empirical and semi-empirical correlations of hardness with flow or yield strength have provide
some acceptable correlations, other areas of failure are also numerous and it is best to approach this type of
correlation that is not based on sound fundamentals with some caution on any material where it has not
withstood the test of time.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.31
Copyright Reserved

12.3 Fracture testing


12.3.1 Charpy impact testing
The Charpy impact testing machine
Since many decades, the so-called impact test (of which the Charpy test is the most well known) has been used
to assess the propensity of a metal to brittle fracture. Although this type of three point bend test has been
superseded by the more scientifically reliable fracture mechanics approach (see later) it is still a useful test for
comparative purposes as it is cost effective and easy to perform. Because it is a comparative test, however, the
specimen size and geometry of the notch need to be standardised. For thin plate products that are thinner than 10
mm, a sub-standard Charpy specimen has also been defined in the ASTM standards.
In essence, a metal with a notched center is placed in a special impact-testing machine and is fractured by a
swinging anvil that strikes the sample on the face opposite to the notch. The energy expended on creating the
fracture is easily determined by recording the height of swing through of the anvil.

Figure 12.3(a): (Left) The standard Charpy impact testing machine and (right) the standard ASTM Charpy impact specimen.

Calibration of the Charpy pendulum machine is to be done according to ASTM E23: Standard Methods for
Notched Bar Impact Testing of Metallic Materials
This test is particularly useful in finding the transition temperature below which an impact fracture becomes
brittle from the earlier ductile mode of fracture at the higher temperatures. It is mainly the BCC metals (and
primarily ferritic steels) that suffer from this transition in mode of fracture and that may occur in the vicinity of
room temperature or even above, and causes brittle failure at temperatures lower than the transition temperature.
This was one of the main causes of the so-called Liberty ships disastrous weld failures in the cold North Sea
during the second World War.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.32
Copyright Reserved

Standard Charpy specimen dimensions


As the Charpy test is largely an indeterminate test (its sharpness of the groove, for instance, may vary slightly
from laboratory to laboratory) it becomes primarily a comparative test and this means that the specimen
dimensions must be highly standardised. This is also the reason why the Charpy impact energy is mostly quoted
only in Joules and not, for instance in J/m2 where the m2 would refer to the fracture area . The universal
standard size is a 10 mm specimen although ASTM does allow a sub-standard specimen size for product
dimensions that are already less than 10 mm.

Figure 12.3(b): Standard Charpy 10 mm specimen with details of the V-groove and the anvil hammer.

The standard Charpy curve as a function of test temperature


From such an Energy absorbed (measured in Joules or Nm) versus Test temperature plot, a number of
definitions may be used to describe the impact fracture behaviour of the metal.

Figure 12.3(c): Schematic sketch of an Energy Absorbed curve as a function of the Test Temperature

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.33
Copyright Reserved

Note the following definitions:

The three zones of brittle, transition and ductile fracture appearance;

The type of fracture associated with each zone as a function of the test temperature and the temperature
where this is in transition, is often defined as the FATT (Fracture Appearance Transition Temperature),
i.e. the temperature where the transition occurs;

Level of energy absorbed is very low during brittle fracture but is high during ductile fracture, with the
latter defined as the upper shelf energy level;

A certain transition temperature at about 50% transition where the fracture transforms from brittle to
ductile fracture and is often defined as the DBTT (Ductile-to-Brittle Transition Temperature).

By definition, the
quantitatively.

FATT DBTT

although care must be taken in estimating the fracture appearance

The DBTT and design of components: Note that for design purposes of a component that may be subject to
impact stresses, any material choice with a specific DBTT must ensure that the operating temperature will
always be above the DBTT, preferably in the upper shelf energy region of its Charpy test curve. This means
that metallurgically, low DBTTs are to be sought through correct alloy choice and heat treatments. For example,
a choice of steel for use outside in winter in South Africa is probably safe if its DBTT is about 20C or lower
but this same component to be used in the Antarctica where outside temperatures may reach 50C or less, a
DBTT much lower than 50C would have to be found. If the component designed for South African
conditions were to be used wrongly in the Antarctica, it would fracture easily in the brittle zone with almost no
fracture strength.
That the various definitions of estimating this transition from brittle to ductile fracture, at least agree qualitatively,
is shown in the Figure below where the %Shear Failure, the % Lateral Expansion during ductile failure and
the Energy Absorbed are compared.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.34
Copyright Reserved

Figure 12.3(d): Fracture transition data from Charpy V-notch impact tests on ASTM A533 type B structural steel
used for nuclear pressure vessels. The % Lateral Expansion is measured from the extent of sideways shear lips which
are ductile in nature.

The following typical alternative transition temperature definitions are also used from time to time in
specifications:
Table: Example of transition temperatures obtained from a Charpy type impact test on ASTM A533 type B
steel for pressure vessels
Specified temperature
definition
Upper shelf energy
100% shear
50% shear
68 J energy
40 J energy
0% shear

Operating temperature C
90
65
25
15
-15
-55

This table immediately points to one of the difficulties of the impact test where a separate definition to suit a
particular design criterion is often used.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.35
Copyright Reserved

Shortcomings of the simple impact type of test


Comparing the simpler Charpy impact test with the much more elaborate fracture toughness test for the
determination of KIC, it may quite rightly be asked what the major and also fundamental differences are between
the two systems of understanding the brittle fracture of metals and alloys. Can Charpy impact measurements,
for instance, be used to estimate the fracture toughness properties of a metal? This question is particularly
relevant as the Charpy impact value is, in a certain sense, some measure of the toughness of a material as is also
the case with the fracture toughness value KIC. There are, however, some important differences in the two tests
and these should be noted.

First, a Charpy specimen contains a machined notch whereas the KIC test contains a much sharper
fatigue pre-crack. This leads to a significant difference in the stress field distribution at the head (or tip)
of the crack if compared to the root of a machined notch. As was seen earlier, the state of the stress field
at the tip of an advancing crack, has a considerable influence on the criterion for unstable crack
propagation. In theory this question could have been addressed by also introducing a fatigued pre-crack
into the Charpy specimen. Practically all Charpy data is, however, for machined notch specimens.

Secondly, the Charpy test measures the total energy required to fracture the test specimen, i.e. it
includes the plastic deformation energy to fold the specimen, particularly in the upper shelf area
where plastic deformation is usually the only energy measured if no fracture takes place. In the plane
strain fracture toughness KIC test, the plastic zone size ahead of the crack tip is kept to a minimum
through the plane strain conditions and the energy to advance the crack in an unstable condition, is
approached more closely.

Thirdly, in the Charpy test the total energy is measured to fracture the entire uncracked ligament of the
specimen whereas in the KIC test one rather measures the energy required for a small extension of the
crack by a. This approaches the situation in practice far more closely and provides more firm design
information to predict the possible catastrophic failure of a structure.

Fourthly, the Charpy test is, by definition, a high strain rate test and it can not readily be used to
translate its impact data to structures where the loading is more gradual yet rapid unstable crack
propagation may still occur.

Finally, the Charpy standard test makes use of 10 mm square specimens whereas true fracture
toughness measurements generally have to make use of far thicker specimens according to the
experimentally determined equation given earlier for KIC tests.

Effect of specimen thickness


It will be seen later that the specimen thickness needs to be greater than that given by the expression below for
true plane strain fracture testing:
2

K
b1 2.5 IC
(Eq 12.3.1-1)
YS
For most materials the standard Charpy specimen size of 10 mm is well below that predicted by the above
equation and true plane strain conditions will, therefore, not prevail in practically all Charpy tests. In fact, it is
probable that mixed mode or type II mode of fracture may prevail in the Charpy test. The possible thickness
effect between the Charpy test and a true fracture toughness test is shown schematically below.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.36
Copyright Reserved

Figure 12.3(e): Schematic difference between a standard Charpy test curve as a function of temperature versus that for
the true fracture toughness specimens.

In fact this thickness effect is even present where sub-standard Charpy specimens are used, for instance, where
the impact fracture energy needs to be found from plate or sheet materials that are already thinner than the
required 10 mm. Comparison of CVN impact values from differently sized Charpy specimens must, therefore,
also be treated with great caution. The thickness effect on Charpy impact energy values for plate and sheet
thicknesses below 15 mm for steel A283 is shown below.
Note that the DBTT decreases as the specimen thickness is decreased, leading to more ductile Upper Shelf
energy values at small thicknesses if compared to the more brittle energy values for the thicker specimens, both
measured at the same temperature. Note also from the top half of the Figure that the % shear failure also
decreases as the specimen thickness is increased, indicating that the mixed mode tends towards greater type I
mode (plane strain fracture) and less from type II mode as the specimen size increases.

Figure 12.3(f): Charpy impact energy curves for different specimen thicknesses for steel A283 shown as a function of the
test temperature. The impact values have been normalised as Joules per 2.5 mm specimen thickness. (Reference: RC
McNicol, Weld Res Suppl: Sept 1985, p385s)

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.37
Copyright Reserved

In fact the effect of specimen thickness on the DBTT from Charpy CVN tests has been studied in many
instances and it is common to find that the thickness effect is also different from steel to steel, as shown below for
three steels with different strengths. Note that steel A shows a shift in the DBTT by as much as 80C (from
about 50 to 30C) by an increase in thickness from about 1 mm to 16 mm.

Figure 12.3(g): Effect of specimen thickness on the Charpy DBTT for three steels with different strength levels.
(Reference: RC McNicol, Weld Res Suppl: Sept 1985, p385s)

Effect of strain rate


As stated before, the Charpy CVN test is actually a high strain rate test whereas most K IC tests are carried out
at slower strain rates. This difference may also have a significant effect on the conclusions one may reach on
toughness performance, particularly arising from the sensitivity of the yield strength of the material on the strain
rate. The effects of this on a measured DBTT in a Charpy test is shown schematically below.

Figure 12.3(h): Schematic effect of strain rate of testing on the transition temperature shift in the Charpy type impact
test. (Reference: JM Barson and ST Rolfe, ASTM STP No 466 (1970)p281)

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.38
Copyright Reserved

Note that the strain rate of testing has a significant effect on both the DBTT as well as on the Upper Shelf
impact energy. The high strain rate Charpy type of impact test may, therefore, lead to a perceived high DBTT
whereas in reality the actual DBTT at a slower strain rate may be significantly lower. This will, of course,
generally lead towards a more conservative situation in practice if slow strain rates are expected rather than fast
strain rates although the Upper Shelf impact energy will now be less conservative.
The above does, of course, not imply that true KIC values are not also strain rate dependent. In fact they are very
much so and this will certainly lead to different KIC values at the same temperature depending on the strain rate
or alternatively, will allow a lower operating temperature for the same K IC design value at slower strain rates
than at higher strain rates.

Figure 12.3(i): Effect of strain rate on the KIC fracture toughness of steel A517-F. (Reference: JM Barson and ST
Rolfe, ASTM STP No 466 (1970)p281)

It is further interesting to note that the measured shift in transition temperature with a change from a slow CVN
bend test to a high strain rate impact test, was more pronounced in lower strength steels with about no shift in
transition temperature found at strength levels above 1000 MPa.

Figure 12.3(j): Shift in transition temperature as a function of the yield strength between impact and slow bend CVN
tests. (Reference: JM Barson and ST Rolfe, ASTM STP No 466 (1970)p281

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.39
Copyright Reserved

12.3.2 The instrumented Charpy impact test


Reference: Instrumented Impact Testing, ASTM Spec Tech Publ No 563 (US) (1973)
B Augland, Brit Weld Journ: 9(1962)p434
In order to overcome some of the limitations of the Charpy test, an instrumented Charpy test has been
introduced. Whereas the normal Charpy test measures the total energy absorbed in fracturing the specimen, the
instrumented Charpy test obtains additional information by recording the load-time history of the specimen
during the test as shown schematically below. The instrumentation includes an instrumented striker fitted with
strain gauges, a dynamic transducer amplifier, a signal-recording and display system and a velocity measuring
device.

Figure 12.3(k): A typical load time history plot from an instrumented Charpy impact test.

Note that the curve allows one to distinguish between the energy required for initiating the crack and the energy
to propagate the crack until fracture as well as information on the load for general yielding P y, the maximum
load Pmax and the fracture load Pf. The fracture energy is normally computed from a loaddisplacement curve
whereas an instrumented test usually measures the loadtime curve. The fracture energy can, however, also be
calculated from such a loadtime curve by the following calculations.
If the velocity of the impacting pendulum v0 is assumed to be constant throughout the test, then one may find the
fracture energy E from the following expression:
t

E v 0 Pdt

(Eq 12.3.2-1)

where P is the instantaneous load and t is the time. In reality, however, the velocity of the pendulum is not
constant during the test as this velocity v will decrease in proportion to the load on the specimen as the fracture
process proceeds. Augland (1962) has proposed that the following correction needs to be introduced to account
for the varying velocity of the pendulum:

E t E(1 )

v 0 Pdt t

E t v 0 1 0
Pdt
4E 0 0

or

(Eq 12.3.2-2)

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.310
Copyright Reserved

E
where E0 is the initial energy of the
4E 0
pendulum. This equation has shown a very good correlation between the calculated value of the total fracture
energy Et from a loadtime instrumented test and the CVN value read off the Charpy impact tester directly.

where Et is the total fracture energy and is a correction factor

Figure 12.3(l): Comparison between the fracture impact energy from the Charpy test and that calculated from a
loadtime curve from an instrumented test. (Reference: GD Fearnehough and CJ Hoy, Journ Iron and Steel Inst:
202(1964)p912)

The standard Charpy impact test is particularly useful in determining the effects of temperature on the absorbed
impact energy, and the curves from instrumented Charpy tests will appear typically as follows for different test
temperatures that span the DBTT.

Figure 12.3(m): The load-time response for a medium Carbon steel impact tested at different temperatures that span
the DBTT. PM = maximum load, PGY = general yield load, PF = fast fracture load, PA = arrest load after fast fracture,
tM = time to maximum load, tGY = time to general yield load, WM = energy absorbed up to maximum load. (Reference:
Metals Handbook, vol 8, 9 th edition, Publ ASM (US) (1985)p266)

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.311
Copyright Reserved

Note from the Figure above that both the crack initiating energy as well as the crack propagating energy
increase as the temperature increases, i.e. as the ductility of the steel increases. Secondly, at very low
temperatures where almost no ductility is present, that the loads PM PGY and that practically no general
yielding takes place within the crack area of the steel.
12.3.3 Pre-cracked Charpy tests
To overcome the limitation introduced through the bluntness and possible variability of the Charpy V-notch,
fatigue pre-cracked Charpy specimens have been tested where the tip of the crack reaches dimensions close to the
atomic level. In un-instrumented pre-cracked Charpy tests, the impact energy was normalised by the surface area
of the uncracked ligament after the fatigue but before the impact test, i.e. CIE/A where A is the uncracked area
of the specimen before the Charpy test.
Initially these pre-cracked Charpy tests were tested by slow bending but correlation with standard CIE values was
poor due to the large differences in loading rate between a standard Charpy test and a slow bend test. A far better
correlation was reached by using pre-cracked Charpy specimens that had been instrumented as referred to above.
The fact that the loading rate could now be measured, brought the pre-cracked Charpy specimen closer to a trui
fracture toughness test.

Figure 12.3(n): (Left) Schematic load response curves from a pre-cracked Charpy instrumented test and (right) some
actual load response curves from steel specimens tested at different impact velocities. m 0 -= mass of hammer, s =
distance between supports. In the right hand figure the initial crack length a was 5 mm and a/W = 0.5 where W =
width (10 mm) of theC harpy specimen.

The dynamic fracture toughness KId is found from the condition:

K Id K dyn
I (t t f )
where tf is the time to fracture obtained from the load curve.

(Eq 12.3.3-1)

These instrumented Charpy tests on pre-cracked specimens still suffer from the one limitation, however, viz. the
Charpy specimen size of 10 mm is still mostly less than the required minimum thickness for a true plane strain
fracture toughness test.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.312
Copyright Reserved

12.3.4 Fracture toughness testing through Linear Elastic Fracture Mechanics


The modes of loading
Although the principles of LEFM apply equally to the three different modes of load application on the crack
advancement, some of the equations need to be adjusted specifically for each of the modes of loading. The three
basic modes of loading are shown below.

Figure 12.3(o): The three basic modes of loading in fracture mechanics with the crack advancement in Mode I defined by
the crack surfaces that are displaced normal to themselves, In Mode II the crack surfaces are sheared relative to each
other in a direction normal to the direction of the front edge of the crack and Mode III where the shearing action is
parallel to the front edge of the crack.

Mode I failure is by far the most common type in most engineering structures or machine components and has,
therefore, received the most attention in developing the stresscrack length relationships. In some cases where
the pre-existing crack is not fully transverse to the applied stress, Mode II loading may be operative but even
here, when the angle between the stress axis and the direction of the crack is greater than 60, Mode I loading
will still be the applicable one to consider. Finally, Mode III would typically be applicable to a notched torsion
bar where the notched area is subject to a pure torsional loading.
In cases where combined loading of two or more of the basic modes takes place, the situation may become more
complex and the concepts for analyzing the stress fields at the tip of the crack have not been developed
successfully yet. If the different modes are in phase, however, and the crack is still in its early development
stage, it will generally turn in a direction in which Mode I applies unless geometrical factors force it into another
direction. Consequently most of the fracture cases found in practice are confined to Mode I loading (or tensile
type of fracture). The principles and concepts used here apply, however, equally to the other modes of fracture.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.313
Copyright Reserved

The elastic stress intensity factor KI


a) The general relationship
The elastic stress field at the tip of a crack is a complex mathematical function of the exact point at which it is to
be determined.

Figure 12.3(p): The co-ordinate system for defining the elastic stress field at the tip of a crack. The stress field may be
defined in either linear (xyz) or polar (r, ) co-ordinates.

The elastic stress field in the vicinity of a crack may be calculated from equations originally developed by Irwin
for Mode I fracture. These are more conveniently given in polar co-ordinates of (r and ) and are
approximated to an area immediately ahead of the crack tip by neglecting higher order components of the
equation. This is justified in most cases as one is really only interested in the direct area where the cracking and
fracturing itself take place, i.e. at the crack front. Should the elastic stress field be sought far ahead of the crack,
the more complete equations will have to be used. The equation for the Mode I elastic stress at any point r and
just ahead of the crack tip, is given by the general equation:
ij

KI
2r

f ()

(Eq 12.3.4-1)

where KI is defined as the stress intensity factor for Mode I fracture. This is a general equation that applies
for any geometry and loading of Mode I, irrespective of the crack size.
The stress intensity factor KI is given by the limiting case of the crack radius r being substituted by an
expression for the half-crack length a {where a = f(r)} and for 0 i.e. for the stress in the y-direction
along the plane where = 0. In this case the relationship between a and r is a simple one of a = r:
KI =

with the units MPa m

(Eq 12.3.4-2)

Where the width W of the plate is significantly larger than the half-crack length a, the constant
= 1 and the expression for the elastic stress intensity factor KI becomes simply:
KI =

(Eq 12.3.4-3)

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.314
Copyright Reserved

This expression of Eq 12.3.4-2 is valid for all Mode I fracture loading cases where the width of the plate W is
significantly larger than the half-crack length a, i.e. W >> a and for the direction in the plane of the crack where
= 0. In this expression, the substitution of the crack radius r with an expression for the half-crack length a,
the crack length convention for center cracks and edge cracks as shown in the Figure below, should be
followed as 1 with any other convention. This convention is based on the rule that for any crack with two
sides, the crack length is defined as 2a whereas if the crack has only one tip, the definition of a is used for the
crack length. In practice, therefore, a typifies the half-crack length of a full crack (although the other half is
imaginary for an edge crack). Expressions for have been developed for other crack definitions and
arrangements (for instance a crack emanating from a holes inside surface in a plate), and may be typically found
in the following references:
References: H Tada, PC Paris and GR Irwin, Stress Analysis of Cracks Handbook, Del
Research Corp., Bethelem, PA (1973)
GC Sih, Handbook of Stress Intensity Factors, Lehigh Univ., Bethelem, PA (1973)
DP Rooke and DJ Cartwright, Compendium of Stress Intensity Factors,
HMSO, UK (1976)
V Kumar et al, An Engineering Approach for elastic-Plastic Fracture Analysis,
EPRI NP-1931, Electric Power Research Institute, CA, US (1981)

Figure 12.3(q): The convention used to describe the crack half length a for (left) a center crack and (right) an edge
crack, both under a uniform tensile stress.

Likewise expressions can be found for the stress intensity factors K II and KIII for Modes II and Modes III
fracture loading respectively.
To determine the fracture toughness KIC of a material, the critical stress c required to propagate a given sharp
fatigue crack with half length ac in a brittle manner before general yielding will take place, is found through the
condition:
K I c a c K IC
(Eq 12.3.4-4)

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.315
Copyright Reserved

Effect of plate thickness b


The effects of plate thickness on the existence of either plane strain (in a thick plate) and plane stress (in a thin
plate) conditions at the advancing crack tip, makes it clear that the value of the facture toughness K IC
determined in a test, will also depend on the thickness of the specimen used. In practice the following schematic
relationship is found upon measuring the fracture toughness of a material.

b1

Figure 12.3(r): (Left) Schematic effect of plate thickness on the fracture toughness K IC. At a plate thickness of b1 a stage
is reached where the fracture toughness KIC becomes independent of plate thickness and here true plane strain conditions
have been reached.. (Right) Effect of plate thickness in an Fe 18%Ni Maraging steel on the measured fracture
toughness.

The value of b1 the plate thickness where the fracture toughness KIC becomes independent from plate thickness,
has been determined experimentally and is given by:

K
b1 2.5 IC
YS

(Eq 12.3.4-5)

This is an important expression to be used both in the design of a laboratory fracture test as well as in the design
of a working component. At thicknesses below b1 plane strain conditions are not present anymore and K
values of type II (or even mixed mode of I and II) need to be considered. Note, however, that KIC is
generally always lower than KIIC and rather using KIC fracture toughness values (which are also more readily
available) will be conservative in the safety design of a component, even if K IIC conditions would be more
applicable.
This relationship also takes into account the size of the plastic zone r p ahead of the crack tip in normally ductile
materials and it is for this reason that the yield strength of the material YS appears in this empirical equation.
The factor 2.5 arises from the experimental observation that the thickness of the plate b 1 needs to be at least 47
times the size of the plastic zone rp for true plane strain conditions to apply.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.316
Copyright Reserved

The above equation, of course, implies that one needs a value for KIC before being able to calculate the value b1.
In practice this is normally done by estimating K IC from similar materials or from other tests but for precise
determinations some tests at different plate thicknesses probably need to be done to find the minimum value for
b1. A table in the ASTM E 399 Standard may also assist in an initial estimate of the value of KIC.
Experimental determinations have likewise shown that the minimum remaining ligament length (W a), i.e. that
portion of the potential crack plane that still remains uncracked at the start of a crack test, and the plate width W
also need to comply with minimum dimensions as follows:

K
a 2.5 IC
YS

and

K
w 5.0 IC
YS

(Eq 12.3.4-6)

Specimen dimensions for the ASTM standard test


A major goal of fracture mechanics is to determine values of the fracture toughness KC that are truly material
constants and that may be used in the design of any structure made from that material. Because of the earlier
observation that the fracture toughness KIC depends on the plate thickness B and that a minimum value of B 1
as well as the remaining ligament length (W a) and the plate width W should be used, the ASTM Standard E
399 specifies the following two specimen dimensions for a three point bend test as well as a tensile crack
opening test.

Figure 12.3(s): Some standardised specimens for the determination of the fracture toughness. Dimensions are in inches.
The following specimens are (i) the ASTM tensile test, (ii) the ASTM three point bend test, (iii) standard edge crack
thin plate test, (iv) center crack thin plate test, (v and vi) the Dynamic Drop Weight Tear (DDWT) test of the Naval
Research Laboratory of the US.

It was learnt earlier that the stress distribution at the crack tip is approximated through a set of stress field
equations that are strictly applicable at only very sharp crack radii. A machined notch is, therefore, not suitable
for these tests and a sharper pre-crack is normally introduced first via a machined notch through prior fatigue
testing. This pre-crack must have a length of at least 0.05W where W is the width of the plate. This places the
tip of the pre-crack beyond any influence from the machined notch. The ASTM standard also specifies the
fatigue pre-crack forming process because a too large plastic zone at the head of the pre-crack must be avoided.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.317
Copyright Reserved

Limits on the maximum alternating stress during this pre-operation are, therefore, set. Typically a low stress high
cycle test is specified, for instance, 1000 cycles at a cyclic strain of 0.03. To distinguish the fatigued pre-crack
from the later fracture crack it is usual practice to colour the pre-crack either by a dye or through some light
oxidation before the fracture test.

Figure 12.3(t): An ASTM compact plane strain fracture test showing the distinction between the pre-cracked fatigue
fracture and the later plane strain fracture upon tensile testing.

12.3.5 The loading fracture test


After the introduction of a fatigued pre-crack into the standard specimen, it is deformed, for instance, in a tensile
test machine for the tensile test specimen and the load-versus-displacement curve is recorded. The displacement
is typically measured by a double cantilever clip gauge that is inserted between the two machined faces of the
initial machined notch in the tensile type test specimen. One of three types of load/displacement curves may be
found, as shown below:

Figure 12.3(u): Typical load/displacement curves found during fracture toughness testing. Note that the initial elastic
deflection has been exaggerated for clarity purposes.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.318
Copyright Reserved

Note that in all three cases the specimen initially deforms elastically until the crack starts to propagate at an
increased displacement, either due to a ductile or a brittle mode. The following distinguishes the three curves
from each other:

Type I: This type of load/displacement curve represents the behavior of a wide variety of ductile metals
in which the crack propagates by a ductile tearing mode with increasing load. As may be seen, this
curve does not present any measurable features that indicate the onset of unstable crack propagation.
The ASTM procedure in this case is then to first draw the tangent line OA and, secondly, to draw the
tangent line OPs at a slope of 5% less than that of line OA, intersecting the curve at Ps. Next draw a
horisontal line at a load of 80% of Ps and measure the distance x1 as shown above. If x1 > 0.25(xs)
(where xs is measured as the distance between OA and Ps) the material is then too ductile and no
reliable fracture toughness value may be determined from this test. If
x1 < 0.25(xs) then the load Ps is designated as PQ and used in the appropriate equations to find the
fracture toughness.

Type II: The Type II load/displacement curve has point PQ where a sharp drop in load to Ps takes
place followed by a slower recovery in load with further displacement. The load drop P Q to Ps is
known as a pop-in and arises from a sudden unstable and rapid crack propagation before the crack
slows down once more and advances further by a ductile tearing mode. The same construction as in
Type I load/displacement curves is carried out here but in this case PQ is simply the maximum
recorded load.

Type III: The Type III load/displacement curve shows complete pop-in and an unstable crack
propagates rapidly through the entire remaining width of the plate with a corresponding decrease in load.
This type of curve is typical of a completely brittle material that fractures elastically with no plastic zone
ahead of the crack tip.

The value found from the load/displacement curve is used to first calculate a conditional fracture toughness
called KQ from the following equation for the compact tension specimen for an initial edge half-crack length of
a and a width of W and a plate thickness of B (other equations are valid, for instance, for the three point
bending equation, etc):
3/ 2
3/ 2
5/ 2
7/2
9/ 2
PQ

a
a
a
a
a
KQ
29.6 185.5 655.7 1017 683.9
W
W
W
W
W
B W

(Eq 12.3.4-7)

The fatigued half-crack length a is measured after the fracture of the specimen. Next calculate the ratio
2.5(KQ/YS)2. If this quantity is less than both the thickness and crack length a of the specimen, then K Q KIC
and the test is valid. If this is not the case, then a thicker specimen will have to be used to find KIC
From all of the above it has become clear that the determination of fracture toughness data is no small task and
can become quite expensive (because of large specimens that have to be prepared by fatiguing after machining)
and also time consuming. It is for this reason that many data books on fracture toughness determinations have
been established which may be consulted to reduce the amount of testing, if own testing is found to be necessary.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.319
Copyright Reserved

The Dynamic Drop Weight Tear (DDWT) test


Reference: Annual book of ASTM Standards, part 10, E208-69 (1981)p416
In many ductile metals fracture occurs by considerable plastic deformation at the crack tip and it becomes
virtually impossible to obtain plane strain (or Mode I) conditions during a normal fracture toughness test unless
very thick specimens are available. This is often not the case for most structural materials and the Naval
Research Laboratory of the US has consequently developed the so-called Dynamic Drop Weight Tear
(DDWT) test for such materials. The test is intended to evaluate metals and alloys over a wider range of
toughness values than would be the case with the standard ASTM fracture toughness test.
In principle, the DDWT is carried out by placing the notched sample that had been preheated to the required
tests temperature, into a holder and dropping a weight onto the sample. Alternatively, the test has also proven to
be useful for testing samples with a welded bead into which a notch had been machined. This test has also
allowed the concept of the NDT (Nil Ductility Temperature) to be adopted, i.e. that temperature above which
the plate will not break. A break is defined as the condition where a crack grows to one or both edges of
the specimen on its tension surface. The NDT is, therefore, virtually a go no-go type of result.
The DT specimen sizes are given in the above Figures 7.10(a)(v) and (vi) for standard specimen dimensions.
Note that a fatigued pre-crack is not introduced in these more ductile materials but rather a sharp knife is used to
create a starter crack within the machined notch. The DT test specimen is fractured by a drop weight test very
similar to the Charpy impact test and the energy not absorbed is measured on a swing pendulum or by the
deformation of some Al or Pb plates in a true drop weight testing machine.
The crack arrest test
Reference: TS Robertson, Engineering: 172(1951)p445
For many applications it is important to know the temperature at which an advancing crack will be arrested and
Robertson (1951) has proposed the following test in which an internally notched sample is placed under a
temperature gradient and crack growth is initiated from the cold end by an impact. Note that sample is also
subjected to a crack-opening load. After the initiating impact, the crack will advance until it reaches a point of
sufficient ductility where it will be arrested and from a pre-calibrated temperature profile, one may determine the
crack arrest temperature TCA.

Figure 12.3(v): Schematic sample and test arrangement for the Robertson crack arrest test. Note that the sample is
subjected to a temperature gradient from left (cold) to the right (hot) as well as to a crack-opening stress. (Reference:
TS Robertson, Engineering: 172(1951)p445)

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.320
Copyright Reserved

The short rod or short bar specimen


Reference: ASTM Standard E 1304-89, Annual Book of ASTM Standards,
Publ ASTM (US) (1989)p962
LM Barker, Eng Fract Mech: 9(1977)p361
LM Barker and FI Baratta, Journ Test Eval: 8(1980)p97
The short rod or short bar specimen has been standardised to determine the KIC value by machining a deep
chevron shaped notch into the rod or bar as shown below.

Figure 12.3(w): Standardised ASTM specimen for determining KIC from a short rod or short bar. The crack formed is
shown as the shaded area. (Reference: ASTM Standard E 1304-89, Annual Book of ASTM Standards, Publ ASTM
(US) (1989)p962)

During the fracture test a wedge is pushed into the mouth of the machined chevron and both the loads P and the
crack-mouth opening are measured. A crack initiates at the tip of the chevron notch with increasing load and
grows to some length a at the peak load. KIC is then determined from the expression developed by Barker for
an ideal elastic material.
K IC

C1 Pmax
B3

(Eq 12.3.4-8)

where C1 22 and is a calibration constant, Pmax is the maximum load and B is the rod diameter or bar width.
The advantage of this standardised test is that it requires relatively small samples and does not require a fatigued
pre-crack. The latter requirement is particularly relevant to very brittle materials such as ceramics where a partthrough crack is quite difficult to introduce.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.321
Copyright Reserved

KIC determinations in brittle ceramics through microhardness testing


Reference: GR Anstis, P Chantikul, BR Lawn and DB Marshall,
Journ Amer Ceramic Soc: 64(1981)p533
AG Evans and EA Charles, Journ Amer Ceramic Soc: 59(1976)p371
BR Lawn, Fracture Mechanics of Ceramics, Ed RC Bradt, AG Evans,
DPH Hasselman and FF Lange, Publ Plenum (NY) vol 5(1983)p1
BR Lawn, SW Frieman, TL Baker, DB Cobb and AC Gonzales,
Journ Amer Ceramic Soc: 67(1984)pC67
Toughness levels in brittle ceramics have been measured through either Vickers or Knoop hardness
indentations in which microcracks are formed in the ceramic at the extremities of the hardness indentations.

Figure 12.3(x): Microcracks formed in brittle ceramics through the Vickers (right) or the Knoop (left) hardness
indentation method..

From the measurements of the crack lengths a the critical stress intensity K C may be found from the expression:
E P
K C d
(Eq 12.3.4-9)
H a 3
where d is a constant that depends on the geometry of the indenter, E is the elastic modulus of the material, H
is the hardness indentation value, P the load and a is the microcrack length.
A good correlation has been found between the KC measured by this Vickers indentation method and KC that
was determined conventionally on 12 different ceramic materials.

Figure 12.3(y): Comparison of fracture toughness KC values determined on 12 different ceramic materials by the
conventional method and by the Vickers indentation method. (Reference: GR Anstis, P Chantikul, BR Lawn and DB
Marshall, Journ Amer Ceramic Soc: 64(1981)p533)

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.322
Copyright Reserved

KIC determinations using a torsional bar


Reference: JA Wang, KC Liu, E McCabe and SA David, Fatigue Fract Eng Matls: 23(2000)p917
These authors have shown through Finite Element Analysis that V-groove that is cut spirally at a pitch of 45
around a round torsion specimen, will approach KIC testing with some confidence as the full length of the Vgroove represents the effective thickness of crack propagation. In general their K IC values determined by this
method agreed to within 6% of the KIC determined on the Al alloy 7475-T7351 on compact 0.5T specimens
and was within 2% of that for A320B steel.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.41
Copyright Reserved

12.4 Creep testing


12.4.1 The three stages of creep
If a specimen is loaded at an elevated temperature with a stress that is above the flow strength of the material
at that temperature, it will first deform instantaneously and then deform further plastically with time and the
strain-time curve will have typically one or more of the three stages of creep in the Figure below. Note that, as
was the case in the engineering stress strain test in Chapter 3, the creep test is traditionally done at constant load
and the apparent strain accelerates during Stage III while the true strain would continue at a steady rate until
fracture takes place.
The creep rate is obtained from the strain versus time curve by the slope (d/dt) at any relevant point on the
curve. Note also that not necessarily in all cases will all three stages be found in one test. At lower
temperatures, for instance, where recovery of dislocations is slow or non-existent, the specimen may undergo
only Stage I creep in which it strain hardens with time and its deformation rate will slow down and later
become almost zero with probably no ultimate fracture of Stage III.

Figure 12.4(a): The three Stages of creep with curve A at constant load and curve B at constant stress.

Reference: B Wilshire and AL Battenbough, Matls Science and Eng A: A443(2007)p156


The secondary creep stage is not always a true constant creep rate regime but it does approach constancy to some
degree.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.42
Copyright Reserved

Figure 12.4(b): Creep (strain) curves and the creep rate curves derived from the strain curves, showing the concept of a
minimum creep rate in the secondary stage of creep. Curves are for fine grained Cu. (Reference: B Wilshire and AL
Battenbough, Matls Science and Eng A: A443(2007)p156

Note the minimum creep rates in the secondary stage of creep but without true linear creep or a constant
creep rate. Because of this predomination of tertiary creep, some instances a log-log plot is often used for design
purposes as shown below for both a linear and a log-log plot of a Ni-based creep resistant alloy.

Figure 12.4(c): (i) Linearly plotted creep curve and (ii) a log-log plotted creep curve of the Ni-base alloy NASAIR (Ni
5.5Al 8.5Cr 0.7Mo 3Ta 1Ti 10W) single crystal tested in tension to rupture at 1000 C in air.

Note in Figure (i) that the Stages I and II are de-emphasized whereas the design engineer is precisely interested
in those areas. Curve (ii)s log-log plot places the Stages I and II in greater perspective.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.43
Copyright Reserved

12.4.2 Constant load testing


The tensile test (at both room and high temperature) is usually carried out by applying a constant extension
rate dL/dt (which does not really lead to a constant d/dt or , as was shown earlier) through a constant
cross head velocity and measuring the load Px. Thus, the extension L/dt is an independent experimental
variable and the measured load Px is a dependent experimental variable. In creep, however, it is more common
to apply a constant load P (now an independent experimental variable) and to measure the extension L as a
function of time where L has now become a dependent variable. The creep curve of true strain as a
function of time t is then plotted from which the steady state creep rate may be obtained.
In most industrial creep experiments, the load P is kept constant but this does not mean that the stress will be
constant as the diameter or cross sectional area A of the specimen keeps on reducing from the starting value A0
over time.

Figure 12.4(d): (Left) A typical constant load creep rig. Note the split furnace and the weights to provide a constant
load. (middle) Thermocouples attached to a specimen and (right) a typical system to measure the extension through a
linear differential transformer.

Temperature control needs to be quite accurate and typical limits would be 1.7C up to 982C and
2.8C above 982C.
Most creep rigs have an automatic device that determines the time when the specimen finally breaks.
This is known as the creep rupture time or life tr or tf.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.44
Copyright Reserved

12.4.3 Constant stress testing


Introduction
The difference between a creep curve at constant load P and one at constant stress is shown schematically
below while some actual constant stress creep curves for AISI 316L stainless steel are shown below on the right.

Figure 12.4(e): (Left) Schematic difference in the creep curve from a constant load test versus that of a constant stress
test and (right) some constant stress creep curves for AISI 316L stainless steel carried out different constant stresses.

Note in particular that the constant stress curves do not have the sharply upward sweep of the curve near the final
rupture. This is due to the load P on the specimen being systematically reduced as the cross section of the
specimen A reduces to keep the stress constant.
The Hyperbolic-weight system
Reference: EN Da Andrade, Proc Royal Soc London: 84(1910-11)p1.
This system was originally proposed by Andrade and consists of a hyperbolically shaped weight that is gradually
submersed into a liquid, thereby reducing the load on the creep specimen as it elongates. The required shape of
the weight is given by:
ML 0 1
(Eq 12.4.3-1)
y

L x
where M is the mass of the load, L0 the initial gauge length of the creep specimen, is the density of the
liquid and x and y are the Cartesian coordinates of the hyperbolic function describing the shape of the mass.

Figure 12.4(f): Schematic arrangement of the hyperbolic weight system of Andrade

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.45
Copyright Reserved

Constant stress system with balanced cams for relatively high loads
A more convenient method is to use a cam-lever arrangement to systematically reduce the load on the specimen
as its diameter reduces by creep. Such an arrangement is shown schematically below.

Figure 12.4(g): (Left) Schematic arrangement of a single cam system and (right) the normal double cam arrangement for
constant stress creep tests.

Note from the above left hand figure that the specimen is loaded through a moment arm R over a circular
loading wheel while the weight system W applies a reducing load through an initial moment arm of r 0. The
initial mechanical advantage is, therefore, r0/R.
To maintain a constant stress on the specimen the load P on the specimen must reduce by keeping the ratio P/Ax
constant where Ax is the instantaneous cross sectional area of the specimen. During uniform elongation, the
volume remains constant and, therefore, the ratio P/Lx must also remain constant. From the figure above:]
Wrx
(Eq 12.4.3-2)
R
where rx is now the instantaneous moment arm fro the weight W. This means that the following condition must
be satisfied:
rxLx = C1 r0Lo
(Eq 12.4.3-3)
P

where C1 is a constant. The instantaneous specimen length may be written as:


L x L 0 ( x 0 )R

(Eq 12.4.3-4)

where 0 is the original angle for the original position of the cam. Eliminating L x from the two above
equations, leads to the constant stress equation for the design of the cam:

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.46
Copyright Reserved

rx

r0 L 0
L 0 R ( x 0 )

(Eq 12.4.3-5)

In practice the two variables rx and x are converted to x and y coordinates by transforming the above
equation into Cartesian coordinates. Note that normally a balancing cam is added to stabilise the application of
load onto the specimen.

Figure 12.4(h): Balancing cam constant stress creep machines.

Constant stress system for low loads


Reference: MF Holmes and PJ Wray, Journ Phys E: Sciences Instr.: 3(1970)
LMT Hopkin, Proc Phys Soc B: 63(1950)p346-349
The above cam arrangement is only suitable for relatively high load applications and a somewhat similar system
may be designed for the application of low creep loads.

NSW 700: WELDING METALLURGY


CHAPTER 12: Mechanical Testing
Page 12.41
Copyright Reserved

Das könnte Ihnen auch gefallen