Plastic Behavior
Plastic Behavior
At a certain value of strain (e ≅ 10−3) called the yield point, a material ceases to behave elastically
—does not return to its original condition when the load is removed but undergoes permanent
deformation. This is called plastic behavior. It is found experimentally that there is negligible
change in specific volume during plastic flow, and this suggests that more than the stretching of
atoms is involved. A suitable model to explain plastic flow and its negligible change in density
is slip in which one layer of atoms slides over another in shear. However, when the shear stress
required to cause one layer of atoms in a perfect lattice to slide over an adjacent layer is estimated,
the required value far exceeds the flow stress of ordinary materials.
45
46 METAL CUTTING PRINCIPLES
relative to each other due to the application of shear stress (τ ). The resisting stress (τ ) may be assumed
to vary linearly with displacement (x). When x = a1 (the horizontal atom spacing), the atoms will be
realigned and τ = 0. Also, when x = a1/2, a displaced atom will be equally attracted to its nearest
neighbors B in Fig. 5.1b and τ should again be zero. As a first approximation, the maximum resisting
shear stress (τ 0 = shear strength) might be expected to occur when x = a1 /4. For an assumed linear
relation between stress and displacement, it follows that
τ x
= (5.1)
τ 0 a1/4
From the definition of simple shear strain (γ ),
x
γ= (5.2)
a2
from Eqs. (5.1) and (5.2),
τ a1
τ0 = (5.3)
4 a2γ
but from Eq. (4.3),
G a1 G
τ0 = ≅ (5.4)
4 a2 4
since a1 ≅ a2.
When this value of theoretical shear strength (τ 0) is compared with values for ordinary materials,
a very large discrepancy is found. For example, mild steel has a value of G of about 11.6 × 106 p.s.i.
(80,093 MN m−2). Thus from Eq. (5.4)
G 11.6 × 106
τ0 = = = 2.9 × 106 p.s.i. (19,651 MPa)
4 4
However, the actual shear strength of such a material is only about 30,000 p.s.i. (207 MPa). This
represents a discrepancy of 2.9 × 106/0.03 × 106 ≅ 97 or two orders of magnitude.
While the assumed linear variation of stress with displacement to such a large strain (γ = –14 )
might be questioned, this is not the problem. This may be shown by substituting a sinusoidal
variation of stress with displacement for a linear one. That is, when τ = τ 0 sin x/(a1/4) (2π) instead
of τ = τ 0 x/(a1 /4) [Eq. (5.1)], the value of τ 0 is found to be G/2π instead of G/4. This changes
the discrepancy to 1.85 × 106/0.03 × 106 = 62 which is still a very large value. Thus, the assumed
relation between stress and displacement does not appear to represent the basic difficulty.
DISLOCATION THEORY
In 1934 a linear crystal imperfection called a dislocation was independently and simultaneously
proposed by three materials scientists (Orowan, Polanyi, and Taylor). Figure 5.2 shows one type
of dislocation known as an edge dislocation in which there is one more atom above the plane of
slip than below. This results in some atoms above the slip plane being attracted to their nearest
neighbors below the plane while others are resisting separation from their nearest neighbors. The
net effect of such a configuration is to require far less energy to cause displacement of the upper row
relative to the lower row than if all atoms acted in the same way as in Fig. 5.1.
PLASTIC BEHAVIOR 47
METAL WHISKERS
It was found quite by accident (Herring and Galt,
1952) that very small samples of metals free of
dislocations may be produced by electrodeposition
at very low values of potential. These “whiskers”
are produced so slowly there is plenty of time for
each atom to find its proper place and there are
no accidents of growth (dislocations). When such
whiskers are tested in bending, they
1. are elastic to the point of fracture (no plastic
flow)
2. are nonlinearly elastic, the elastic stress–
strain curve resembling the first quarter
cycle of a sine wave
3. have a very large strain at fracture approach-
ing γ = –41 suggested by the derivation for
theoretical strenth (τ 0)
From the foregoing discussion, it is evident
that metals are ductile and flow plastically due to
the presence of structural defects called dislocations.
Plastic behavior is ordinarily possible only due to
dislocations. In the absence of dislocations, metals
are perfectly brittle (yield strain p fracture strain).
The ductility characteristic of metals is due to the Fig. 5.5 Bubble model showing grain boundary running
tendency for metals to generate large numbers of diagonally across center of field and dislocation just above
dislocations. center of field.
introduced by defining the stress in terms of the original area instead of the actual instantaneous
area. Figure 5.7 shows the curve of Fig. 5.6 replotted in terms of true stress (σ ) and true strain (ε )
where stress and strain are now defined in terms of the instantaneous area and gage length instead
of initial values (A0 and l0). A metal that is highly strain hardened before testing will yield a much
flatter curve that approaches the “ideal plastic” which has a perfectly flat σ –ε curve in the plastic
region. Since there is no change of volume in the plastic regime,
Al = A0 l0 (5.5)
and by definition
P Pl S(l + Δl )
σ= = = 0 = S(1 + e) (5.6)
A A0 l0 l0
Also,
dl
dε =
l
Integrating and evaluating the constant of integration:
l G l (1 + e) J
ε = ln = ln H o K = ln (1 + e) (5.7)
lo I lo L
Equations 5.6 and 5.7 are useful in converting S and e into σ and ε. From
Eq. (5.7) it is evident that the difference in ε and e is negligible for values of
e less than 0.1. The strain at the yield point SY is ε Y ≅ 10−3.
If Fig. 5.7 is replotted on log–log coordinates (Fig. 5.8), a straight line is
obtained having a slope n and hence beyond the yield point
σ = σ1ε n (5.8)
Fig. 5.8 Log–log true-stress–
true-strain uniaxial tensile curve. where σ1 is the stress corresponding to ε = 1 and n is called the strain
[Y = Yield point. U = Ultimate hardening index. Equation (5.8) is widely used in analytical studies involving
point. F = Fracture point.] plastic flow in the cold working regime.
PLASTIC BEHAVIOR 51
NECKING
In the initial plastic region (between Y and U in Figs. 5.6 and 5.7) the area decreases uniformly
along the gage length and the increase in load (P) required due to strain hardening exceeds the
decrease in P required due to reduction in area. However, at point U, a localized neck begins to
form. Then the change in P required due to area reduction exceeds the change in P due to strain
hardening, and what appears to be negative strain hardening occurs if stress is measured in terms
of original area Ao instead of actual area A. The necking strain may be obtained as follows:
P = σA (by definition) (5.9)
Differentiating both sides
dP = σ dA + A dσ (5.10)
At point U, dP = 0 and hence
dσ dA
=− (5.11)
σ A
Since
Al = constant (5.12)
A dl + l dA = 0 (5.13)
or
dA dl
− = = dε (5.14)
A l
Hence from Eqs. (5.11) and (5.14)
dσ
=σ (5.15)
dε
and from Eq. (5.8)
dσ
= σ1n ε n−1 = σ (5.16a)
dε
and
σ1 = σ/ε n (5.16b)
then
nσ
=σ (5.17)
ε
or at point U
εU = n (5.18)
Since the value of n is about 0.2 for a ductile metal (annealed), the natural strain at necking (ε U)
is about 0.2.
52 METAL CUTTING PRINCIPLES
When a concentrated neck forms in a tensile specimen, added complications are involved. First
of all, the strain cannot be measured in terms of length since the strain is nonuniform along the gage
length. However, due to the constance of volume,
ld 2 = lod o2 (5.19)
l Ad D 2 Ad D
ε = ln = ln o = 2 ln o (5.20)
lo CdF CdF
Thus, after necking, the strain must be measured in terms of the minimum diameter at the neck (d ).
G A a2 + 2aR − r 2 D J
σa = σc H1 + ln K (5.21)
I C 2aR FL
where a is the section radius and R is the profile radius at the neck (Fig. 5.9). Thus, the maximum
tensile stress occurs at the center of the specimen and equals
G A a DJ
σa,max = σc H1 + ln 1 + K (5.22)
I C 2R F L
The corrected uniaxial tensile stress (σc ) is equal to the stress that would
obtain in the absence of a neck and in turn is equal to the stress at the outer
periphery of the neck:
H
σc = (5.23)
A RD A a D
1+2 ln 1 +
C a F C 2R F
HYDROSTATIC STRESS
When tensile specimens are tested under different
values of hydrostatic pressure, a negligible effect
is found upon the yield stress and the post yield
shape of the true stress–strain curve. For example,
Bridgman (1952) found an increase of less than
2% in the ultimate stress (ε = 0.2) of AISI 1045
steel when subjected to a hydrostatic pressure of
100,000 p.s.i. (690 MPa). He similarly found an
Fig. 5.14 Influence of hydrostatic pressure on flow stress–
strain curve and fracture stress (x) P1 < P2 < P3. increase in Brinell hardness of less than 4% with
a hydrostatic pressure of 100,000 p.s.i. However,
hydrostatic stress has an appreciable influence on
fracture stress, a compressive hydrostatic stress increasing fracture stress and a tensile hydro-
static stress decreasing fracture stress. Figure 5.14 illustrates these effects. Bridgman expresses
the increased fracture strength (σf) with pressure in terms of a pressure coefficient of ductility
(K):
σf = σf0 + Kσ H (5.24)
this behavior. The following equation resembling Eq. (5.8) for strain hardening is frequently used
to relate flow stress and strain rate at a given strain and temperature
σ = σ1M m (5.25)
where M = dε /dt and σ1 and m are material constants. The exponent m (strain rate sensitivity)
is found to increase with temperature, especially above the strain recrystallization temperature.
In the hot working region, metals tend to approach the behavior of a Newtonian liquid for which
m = 1, particularly metals of very small crystal size (~ 1 μ m) which are superplastic (capable of
deformation to very large strains at high temperature and low strain rate).
In general, increased temperature and strain rate have opposite effects on the flow stress of a
material.
HARDNESS
Hardness is a term having different meaning to different people. It is resistance to penetration to
a metallurgist, resistance to wear to a lubrication engineer, a measure of flow stress to a design
engineer, resistance to scratching to a mineralogist, and resistance to cutting to a machinist. While
PLASTIC BEHAVIOR 59
these several actions appear to differ greatly in character, they are all related to the plastic flow stress
of the material (Y ).
The wide variety of hardness test procedures that have been used may be classified as follows:
1. Static indentation tests in which a ball, cone, or pyramid is forced into a surface and the load
per unit area of impression is taken as the measure of hardness. The Brinell, Vickers,
Rockwell, Monotron, and Knoop tests are of this type.
2. Scratch tests in which we merely observe whether one material is capable of scratching
another. The Mohs and file hardness tests are of this type.
3. Plowing tests in which a blunt element (usually diamond) is moved across a surface under
controlled conditions of load and geometry and the width of the groove is the measure of
hardness. The Bierbaum test is of this type.
4. Rebound tests in which an object of standard mass and dimensions is bounced from the test
surface and the height of rebound is taken as the measure of hardness. The Shore
Scleroscope is an instrument of this type.
5. Damping tests in which the change in amplitude of a pendulum having a hard pivot resting
on the test surface is the measure of hardness. The Herbert pendulum test is of this type.
6. Cutting tests in which a sharp tool of given geometry is caused to remove a chip of standard
dimensions.
7. Abrasion tests in which a specimen is loaded against a rotating disc and the rate of wear is
taken as a measure of hardness.
8. Erosion tests in which sand or abrasive grain is caused to impinge upon the test surface
under standard conditions and the loss of material in a given time is taken as the measure of
hardness. The hardness of grinding wheels is measured in this way.
The equipment and detailed test conditions for most of the hardness tests in use today may be
found in texts by O’Neill (1934), Williams (1942), Tabor (1951), and von Weingraber (1952).
SCRATCH HARDNESS
One of the oldest scales of hardness, and one that is still in use by geologists, is that due to Mohs
(1822) in which the scratch resistance of a material is compared with the scratch resistances of a
standard series of ten minerals. Each of these minerals is assigned a number from 1 to 10. A Mohs
hardness of 1 corresponds to a very soft material while 10 refers to the most scratch resistant of all
known materials—the diamond. Mohs’ scale of hardness is given in Table 5.1.
In applying the Mohs scale practically, the standard minerals are not employed, but the simple
tests given in Table 5.1 under the heading working scale are used instead. The Mohs hardness of
engineering metals falls within the narrow range from 4 to 7.
Tabor (1956) has shown that whereas a stylus must be 20% harder than a flat surface in order
to scratch the flat surface, the minerals constituting the Mohs’ scale increase in hardness in geometric
ratio by a factor of approximately 1.4. Diamond is an exception since it is approximately three times
as hard as corundum.
As can be seen, Mohs’ method of describing hardness is little more than qualitative. While
attempts have been made to put the scratch technique on a firmer quantitative basis by measuring
the width of a scratch made with a diamond under standard conditions (Bierbaum, 1930), such tests
are difficult to carry out with the necessary precision. The indentation method is by far the most
practical and useful method of measuring hardness.
60 METAL CUTTING PRINCIPLES
Vickers (Fig. 5.21b) and Knoop (Fig. 5.21c) indentors are blunt pyramids.
Faces of the Vickers indentors are inclined at an angle of 136°, while the ridges
of the Knoop indentor have angles of 130° and 172 –12 °, respectively. Because the
Knoop indentor penetrates only about half as deeply as the Vickers for the same
load, it is frequently preferred for studies of superficial hardness. The Vickers
hardness is the load divided by the contacting surface area, while the Knoop
hardness is the load divided by the projected area, and hence corresponds to the
Meyer value.
The hardness test is very easily conducted, but not so easily interpreted.
Action beneath the indentor is complex and must be understood if full use is to Fig. 5.22 Plane strain uni-
be made of hardness values. axial compression test.
The simple compression test, Fig. 5.22, provides another measure of resist-
ance to plastic flow that is more widely used in design analysis. If friction is kept to a low value
on the die faces, a compression specimen will deform as shown by the dotted lines in Fig. 5.22
without barreling, and the uniaxial flow stress will be
P
σ1 = (5.44)
A
where A is the cross-sectional area of the specimen.
It is important to relate Meyer hardness to uniaxial flow stress, a term with which most
engineers are accustomed. The plastic zone beneath a hardness indentation is surrounded by elastic
material which acts to hinder plastic flow in a manner similar to the die surfaces in a closed die
forging. In the simple compression test the entire specimen goes plastic, and there is no resistance
to side flow because the specimen is surrounded by air. Therefore, a greater mean stress is required
to cause plastic flow in the hardness test than in the simple compression test.
The relation between the Meyer hardness and the uniaxial flow stress may be expressed as
follows:
HM = Cσ1 (5.45)
where C is called the constraint factor for the hardness test. Experimentally, C approximates three
for the Brinell, Vickers, and Knoop hardness tests. A central problem in the theory of hardness is to
explain the origin of constraint factor, C.
Fig. 5.23 Slip line field solutions for flat two-dimensional punch. (a) Prandtl (1920) SLF for punch face with
high friction. (b) Mohr’s circle for field ABC. (c) Mohr’s circle for field AEG. (d) Hill (1950) for frictionless
punch face.
The Mohr’s stress circle (Chapter 3) for all points in region ABC is shown in Fig. 5.23b. The
convention adopted in labeling the α and β directions is such that the directions of the algebraically
greatest stress (σ1) lies in the first quadrant of the α – β coordinate system. Thus the positive α and
β directions at point D are directed as shown in Fig. 5.23a according to the usual convention. The
Mohr’s circle of Fig. 5.23b will hold for all points within region ABC. The normal stress on the
shear plane for any point within region ABC will be pD = k.
In going from C to E along the curved α line there will be no change in shear stress but a change
in normal stress according to the following Hencky (1924) equation:
p + 2φ k = constant (5.46)
where φ is the angle between the α and x axes. At point C (Fig. 5.23a) φ = π /4 and at point E,
φ = −π /4. From Eq. (5.46)
π
pE = pc + 2k = k(1 + π ) (5.47)
2
The Mohr’s circle diagram for point E and all points in region AEG is shown in Fig. 5.23c. The
normal stress on the face of the punch (σ 2′ ) in Fig. 5.23c will be
A πD
σ 2′ = k + pE = 2k 1 + = 2k(2.57) (5.48)
C 2F
PLASTIC BEHAVIOR 63
equilibrium completely are simpler but more approximate than SLF solutions, and these are
generally referred to as upper bound solutions. An example of an upper bound solution for the
indentation hardness problem is given in Fig. 5.24. In this solution, deformation is assumed to
occur by slip only along the solid lines shown in Fig. 5.24a. At all other points, the material is
considered to be rigid. The shear flow stress at all points is assumed to be k (uniaxial plane strain
flow stress in shear) as in the SLF approach. The external work per unit time is then equated to
the internal work per unit time. The external work per unit time is the product of the punch force
P and the punch velocity V, while the internal work per unit time is the sum of the products of
shear force on each flow surface and the appropriate velocity. The appropriate velocity is obtained
from a hodograph (velocity diagram) that satisfies velocity continuity. Figure 5.24b shows the
synthesis of the partial hodograph for the right half of Fig. 5.24a. Here, distance 0–1 corresponds
to the absolute velocity of the punch (taken to be unity), 1–2 to the relative velocity of region 2
of Fig. 5.24a relative to region 1, and 2–3 to the velocity of region 3 relative to region 2. Velocity
vectors 0–2 and 0–3 correspond to the absolute velocities of blocks 2 and 3. Fig. 5.24b satisfies
continuity of flow to the right since the rate of flow into the deformation zone equals the rate of flow
from the deformation zone (a × 1 = 2a × –12 ).
Equating the external and internal work per unit time per unit distance perpendicular to the
paper for flow to the right,
P 2 1 5k(2a)
(1) = k(2a × 1) + 3k(2a × 1) =
2 √3 √3 √3
or
P p 5
= = = 2.89
(2a)(2k) 2k √3
This value is larger than that for the SLF solution [Eq. (5.49)] as it should be since it represents a
more approximate upper bound approach where force equilibrium is completely ignored. Other
upper bound solutions may be obtained by making θ in Fig. 5.24a other than 60°. The lowest
(and best) value of p/ 2k = 2.83 is obtained when θ = 54.7°.
PLASTIC BEHAVIOR 65
ELASTIC THEORY
When a large block of material having a grid applied to a central plane is loaded by a spherical
indentor, flow patterns such as those shown in Fig. 5.25 are obtained. Study of these patterns
reveals a plastic zone that passes through the edges of the punch (Fig. 5.25c). There is no evidence
of upward flow, and little resemblance to the plastic zones of Figs. 5.23 and 5.24.
The deformed grids of Fig. 5.25 clearly indicate an elastic–plastic boundary, which has a shape
resembling that of a line of constant maximum shear stress beneath a sphere pressed against a flat
surface. This elastic problem was first studied by Hertz (1895). He found that the contact stress was
distributed in a hemispherical pattern over the surface, and that lines of constant maximum shear
stress were as shown in Fig. 5.26, where
τ max
M′ = (5.52)
J
and
τ max = maximum shear stress
J = mean pressure on punch face (Meyer hardness)
In Fig. 5.26 the punch face is shown flat for simplicity, whereas in reality it is the surface of a
large-radius sphere. The elastic–plastic boundary of Fig. 5.25c closely resembles a line between
M ′ = 0.15 and 0.20.
An alternative approach to hardness has been presented by Shaw and DeSalvo (1970) in which
the material is assumed to be plastic–elastic instead of plastic–rigid.
Fig. 5.27 Photographs of grid patterns on transverse planes of plasticene indented in plane strain by flat
punch. (a) Test arrangement. Pattern (b) when h = 1.414a as in Fig. 5.23a and (c) when h = 0.707a as in
Fig. 5.23d. In both cases the specimens were supported on a steel plate (rigid relative to plasticene).
While most of the practical indentors are not conical, they may be assigned an effective cone
half angle (θe). The effective cone half angle for the sphere is the angle that the tangent to the sphere
makes with the vertical at the edge of the indentation (Fig. 5.29a). For the sphere,
2a
θ = cos−1 (5.53)
D
When 2a/D = 0.4, θe = 66.4°.
For the Vickers indentor, the cone half angle θ is 136/ 2 = 68° for flow in direction 1 in
Fig. 5.29b. For flow in a direction inclined at an angle α to direction 1,
A tan 68° D
θ = tan−1 (5.54)
C cos α F
For direction 2, this gives θ = 72.18°. The mean value of θ may be obtained as follows:
π /4
冮
4 A tan 68° D
θe = tan−1 dα = 70.3° (5.55)
π 0
C cos α F
For the Knoop indentor (Fig. 5.27c) the cone angle (θ ) is 130/2 = 65° in direction 1, and
172.5/ 2 = 86.25° in direction 2. The mean value F, found as above, is 72.12°.
These values are summarized as follows:
A sphere loaded so that 2a/D = 0.4 is seen to be less blunt than either the Vickers or Knoop
indentors.
essentially of elastic origin with a small flow component included. This flow component causes
the hardness value to correspond to the flow stress at a small plastic strain (approximately 0.05).
It also alters the constraint factor upward or downward, depending upon whether the metal is fully
strain-hardened, or annealed, and whether the indentor friction is high or low.
associated with unloading may play an important role relative to the integrity of machined surfaces,
particularly when the metal machined has a relatively low fracture strain.
4. In forming operations, such as extrusion, drawing, rolling, forging, thread-rolling, peening,
swaging, etc., the second plastic flow associated with unloading may give rise to surface cracks.
5. In shot peening, to increase fatigue life, the surface is bombarded by hard spheres which
make a large number of small Brinell impressions in the surface and leave behind residual biaxial and
triaxial compressive stresses. The second plastic flow that accompanies unloading will obviously be
involved in any rational approach to shot peening that may be developed in the future. The simple
qualitative approach presented above suggests the origin and character of these useful residual
compressive stresses. At the same time, it becomes evident why shot peening may not always be
useful and why, with somewhat brittle materials, shot peening may have a negative effect on fatigue
life. This will be the case when the plastic flow that accompanies unloading gives rise to surface
imperfections (cracks) that are more detrimental to fatigue life than the residual compressive
stresses are beneficial.
6. In burnishing and ball sizing operations, care must be taken that the plastic flow associated
with unloading does not result in surface cracks. However, a recent Russian development takes
advantage of cracks thus developed. In this case, a carbide roller is loaded against the workpiece just
ahead of the cutting tool. This roller provides sufficient pressure to produce cracks in the material
just in front of the cutting tool, substantially reduces cutting forces, and increase tool life by as much
as 50% when chilled cast-iron rolls are machined.
The best way of experimentally proving the
existence of a second plastic flow on unloading
is to trace the shape of the free surface with a
sensitive stylus instrument first with the load
applied and then after the load has been removed.
However, this cannot be done with existing tracing
instruments since it is not possible to get the
tracing point close enough to the ball when the
load is still applied.
It was therefore decided to make a replica of
the surface under load and to compare the shape of
the free surface of the replica with that of the metal
surface after the load had been removed.
A –14 inch (6.35 mm) diameter steel sphere
was loaded into a flat steel surface as shown in
Fig. 5.33 at loads varying from 1000 to 4000 kg.
In each case, a casting was made with the load on
the ball held constant throughout the period of
solidification. The load was then removed and
the surfaces of the steel and plastic were traced in
four directions oriented 90° from each other. In all
cases, there was a substantial difference between
the loaded and unloaded surfaces.
Figure 5.34 is a representative pair of tracings
showing the difference between the loaded and
unloaded surface shapes. Fig. 5.33 Test arrangement used to obtain a replica of the
It appears that a substantial reverse plastic shape of a loaded surface. (after Shaw, Hoshi and Henry,
flow occurs when a Brinell ball is unloaded. 1979)
72 METAL CUTTING PRINCIPLES
ROCKWELL HARDNESS
Rockwell hardness values (Rockwell, 1922) are also in common use. In this case the hardness value
is an arbitrary number that varies inversely with the depth of penetration under standard loading
conditions. Indentors of different geometries are used for materials of different ranges of hardness.
The B-scale (B for ball) is used for relatively soft materials while the C-scale (C for cone) and A-
scales are used for materials of increasing ranges of hardness. Rockwell hardness determinations leave
a small dent and hence represent relatively nondestructive tests (along with Vickers and Knoop)
compared with the standard Brinell test.
Fig. 5.36 Variation of Knoop microhardness (kg mm−2) with (a) load in grams and (b) depth of impression
(d) in microns produced by loads on indentor of varying magnitude. (after Thibault and Nyquist, 1947)
More ductile hard materials also contain many closely spaced imperfections and likewise
exhibit a size effect. As long as the size of the impression is large compared with the mean imper-
fection spacing, the normal hardness vale will be obtained. However, as the load is decreased to the
point where the impression size approaches imperfection spacing, hardness values begin to increase.
Values of Knoop hardness versus load and depth are shown in Fig. 5.36 for several hard materials.
The hardest material (B4C) is seen to give values which begin to increase with decreased load at
a larger value of load than the softer material (tool steel). This is because it takes a higher load to
produce the same size impression in B4C than it does in tool steel. When hardness is plotted against
depth of impression in microns (Fig. 5.36b), all curves are found to be of similar shape and to turn
upward for depths of impression less than about 2 microns.
Representative microhardness values for a variety of hard materials and metallographic con-
stituents are given in Table 5.2.
CONVERSION CURVES
The curves of Fig. 5.37 enable relative magnitudes of hardness, expressed in several systems, to be
compared. Brinell hardness has been used as the standard of comparison. The Knoop hardness is
seen to be very nearly equivalent to Brinell hardness over the entire range while the other hardness
scales are significantly nonlinear relative to the Brinell scale. The Mohs scale of hardness is seen
to be very nonlinear, particularly at the upper end of the scale. The equivalent Brinell hardness
values for Mohs hardness values of 8, 9, and 10 are as follows: 1150, 1900, and 7000. From this it
is evident that the range from 9 to 10 on the Mohs scale covers more than four times the range of
hardness corresponding to the range from 1 to 9.
74 METAL CUTTING PRINCIPLES
TORSION
The torsion test is an additional method of characterizing the plastic performance of materials. This
test is particularly useful for studies involving large plastic strains since neither necking (with all its
complications in the tensile test) nor barrelling (compression test) occurs. However, there is a large
strain gradient from the center to the outside diameter of the specimen which makes it somewhat
awkward to convert measured quantities (torque, θ, and angle of twist, MT) into shear stress (τ ) and
shear strain (γ ).
Three of the important applications of the torsion test are
1. to directly measure the shear modulus G in the elastic region
2. to evaluate the strain rate sensitivity of materials in the hot working region (homologous
temperature > 0.5) at high rates of strain (Gleeble test)
3. to evaluate the fracture stress of quasi-brittle materials such as quenched and tempered tool
steels. Since the mean principal stress (σ H) in the torsion test is zero (Fig. 4.9e) instead of
being tensile (σ H = σ 1/3) as in the uniaxial tensile test, there is a greater spread of fracture
values in torsion for specimens of different quality than in tension
The stress–strain curve in torsion has the same appearance as a true-stress–true-strain tensile
curve (with yield point, strain hardening, and finally fracture).
In both the elastic and plastic regimes, the shear strain increases linearly with radius, being zero
at the center of the specimen and a maximum (γm) at the outside diameter (o.d.) where
rθ
γm = (5.58)
L
where
r = radius at o.d.
θ = angle of twist of one reference plane relative to the other
L = axial distance between reference planes
In the elastic regime the shear stress at the o.d. is
MT r
τ= (5.59)
J
where
MT = twisting moment
r = radius at o.d.
J = polar moment of inertia
(J = 16π r 4/32 for cylindrical specimen)
In the plastic regime it may be readily shown as Nadai (1950)
has done that the shear stress at the o.d. (τ m ) may be obtained from
the moment of twist (MT) versus angle of twist (θ ) curve as follows
(Fig. 5.38): Fig. 5.38 Construction for converting
torsional moment vs angle of twist curve
1
τm = (AB + 3AD) (5.60) to shear-stress–shear-strain curve. (after
2 πr 3 Nadai, 1950)
76 METAL CUTTING PRINCIPLES
The effective stress σe and effective strain ε e for a torsion test may be readily shown to be
σe = √3 τm (5.61)
γm
εe = (5.62)
√3
When such values are compared with σe and εe values for a uniaxial tensile test on the same
material (σ e = σ 1 and ε e = ε 1 in uniaxial tension), the same curve is obtained to a good approximation.
While this is frequently cited as evidence for the general applicability of the concept of equivalent
stress and strain, it should be kept in mind that there is no support for the equivalent strain concept
in dislocation theory.
ε 3 ~ σ 3 − v(σ 1 + σ 2)
As previously mentioned, it is found experimentally that the change in volume in the plastic
region is essentially zero and therefore
ε1 + ε 2 + ε 3 = 0 (5.64)
When Eqs. (5.63) are substituted into Eq. (5.64), it is found that
v = –12 (5.65)
Plane stress and plane strain are important states of stress in plasticity and Fig. 5.39 gives
the Mohr’s circle diagrams for these two cases. In the example of Fig. 5.39, σ2 has been taken to
be σ 1/3 and Eqs. (5.63) and (5.65) have been used to determine the unknown strain and stress,
respectively.
As mentioned in the previous section, another relatively crude approximation in plasticity is
to use an effective flow strain [Eq. (5.39)] as well as an effective flow stress [Eq. (5.38)]. When
this is done, it is equivalent to assuming that strain hardening is a scalar type effect. In reality, strain
hardening is strongly dependent on the direction of the strain. The Bauschinger (1881) effect in
which the flow stress in compression following plastic flow in tension is found to be substantially
reduced (and vice versa) is direct evidence that strain hardening not only depends upon the
magnitude of the strain but also the direction of the strain. Considering all strain to be equivalent
relative to strain hardening is equivalent to assuming the Bauschinger effect does not exist.
PLASTIC BEHAVIOR 77
Fig. 5.39 Mohr’s circle diagrams for (a) plane stress and (b) plane strain (σ2 = σ1/3).
REFERENCES
Backofen, W. A. (1972). Deformation Processing. Addison Wesley, Reading, Mass.
Bauschinger, J. (1881). Zivilingur 27, 289.
Bierbaum, C. H. (1930). Trans. Am. Soc. Steel Treat. 18, 1009.
Bragg, Sir Lawrence, Lomer, Wm., and Nye, J. F. (1954). Experiments with the Bubble Model of a Metal
Structure (16-mm silent film). Brent Laboratories Ltd., London.
Bragg, Sir Lawrence, and Nye, J. F. (1947). Proc. R. Soc. A190, 474.
Bridgman, P. (1944). Trans. Am. Soc. Metals 32, 553.
Bridgman, P. (1952). Studies in Large Plastic Flow and Fracture. McGraw-Hill, New York.
Brinell, J. (1901). J. Iron Steel Inst. 51, 243.
Cottrell, A. H. (1953). Dislocations and Plastic Flow in Crystals. Oxford University Press, New York.
Dugdale, D. S. (1954). J. Mech. Phys. Solids 2, 267.
Dugdale, D. S. (1955). J. Mech. Phys. Solids 3, 197 and 206.
Hencky, H. (1924). Z. für Angew. Math. u Mech. 4, 323.
Herring, C., and Galt, J. (1952). Phys. Rev. 85, 1060.
Hertz, H. (1895). Gesammelte Werke (Leipzig) 1, 156.
Hill, R. (1950). The Mathematical Theory of Plasticity. Clarendon Press, Oxford.
Johnson, W., and Mellor, P. B. (1962). Plasticity for Mechanical Engineers. D. Van Nostrand Co., Inc.,
Princeton, N.J.
Knoop, F., Peters, G., and Emerson, W. B. (1939). J. Res. Natn. Bur. Stands. 23, 39.
MacGregor, C. W. (1944). J. Franklin Inst. 238, 111.
Marshall, E. R., and Shaw, M. C. (1952). Trans. Am. Soc. Metals 44, 7.
Maxwell, C. (1856). Letter to W. Thomson, 18 Dec. 1856.
Meyer, E. (1908). Z. Ver. Dtsch. Ing. 52, 645.
Mohr, O. (1914). Abhandlung aus dem Gebiet der Technischen Mechnik, 2nd ed. W. Ernst & Sohn, Berlin.
Mohs, F. (1822). Grundriss der Mineralogie. Dresden.
Nabarro, F. R. N. (1967). Theory of Crystal Dislocations. Oxford University Press, New York.
78 METAL CUTTING PRINCIPLES
Nadai, A. (1950). Theory of Flow and Fracture of Solids, Vol. I. McGraw-Hill, New York.
Nadai, A. (1963). Theory of Flow and Fracture of Solids, Vol. II. McGraw-Hill, New York.
O’Neill, H. (1934). The Hardness of Metals and Its Measurement. Chapman and Hall, London.
Orowan, E. (1934). Z. Phys. 89, 605.
Polanyi, M. (1934). Z. Phys. 89, 660.
Prandtl, L. (1920). Nachr. Ges. Wiss. Gottingen, Math. Phys. K1., 74.
Rockwell, S. R. (1922). Trans. Am. Soc. Steel Treat. 2, 1013.
Rowe, G. W. (1977). Principles of Industrial Metal Working, 2nd ed. E. Arnold, London.
Shaw, M. C. (1954). Proc. Natn. Acad. Sci. U.S.A. 46, 394.
Shaw, M. C., and DeSalvo, G. J. (1970). J. Engng. Ind. 92, 469 (Part I) and 480 (Part II).
Shaw, M. C., Hoshi, T., and Henry, D. (1979). J. Engng. Ind. 101, 104.
Smith, R., and Sandland, G. (1922). Proc. Inst. Mech. Engrs. (London) 623.
Smith, R., and Sandland, G. (1925). J. Iron Steel Inst. 111, 285.
Tabor, D. (1951). The Hardness of Metals. Clarendon Press, Oxford.
Tabor, D. (1956). Br. J. Appl. Phys. 7, 159.
Taylor, G. I. (1934). Proc. R. Soc. A145, 362.
Thibault, N. W., and Nyquist, H. S. (1947). Trans. Am. Soc. Metals 38, 271.
Tresca, H. (1864). C. r. hebd. Séanc. A cad. Sci. Paris 59, 754.
Tresca, H. (1867). C. r. hebd. Séanc. A cad. Sci. Paris 64, 809.
von Mises, R. (1913). Nachr. Ges. Wiss. Göttingen, Math. Phys. K1., 582.
von Weingraber, N. (1952). Technische Hartemessung. Carl Hauser Verlag, Munich.
Williams, S. R. (1942). Hardness and Hardness Measurement. Am. Soc. Met., Cleveland.