Sie sind auf Seite 1von 16

CHAPTER

NINETEEN
DRAWING OF RODS, WIRES, AND TUBES

19-1 INTRODUCTION

Drawing operations involve pulling metal through a die by means of a tensile


force applied to the exit side of the die. Most of the plastic flow is caused by
compression force which arises from the reaction of the metal with the die.
Usually the metal has a circular symmetry, but this is not an absolute require¬
ment. The reduction in diameter of a solid bar or rod by successive drawing is
known as bar , rod, or wiredrawing, depending on the diameter of the final
product. When a hollow tube is drawn through a die without any mandrel to
support the inside of the tube, this is known as tube sinking. When a mandrel or
plug is used to support the inside diameter of the tube as it is drawn through a
die, the process is called tube drawing. Bar, wire, and tube drawing are usually
carried out at room temperature. However, because large deformations are
usually involved, there is considerable temperature rise during the drawing
operation.

19-2 ROD AND WIREDRAWING

The principles involved in the drawing of bars, rod, and wire are basically the
same, although the equipment that is used is different for the different-sized
products. Rods and tubes, which cannot be coiled, are produced on drawbenches
(Fig. 19-1a). The rod is pointed with a swager, inserted through the die, and
clamped to the jaws of the drawhead. The drawhead is moved either by a chain
drive or by a hydraulic mechanism. Drawbenches with 1 MN pull and 30 m of
runout are available. Draw speeds vary from about 150 to 1500 mm s~
1

The cross section through a conical drawing die is shown in Fig. 19-1b. The
entrance of the die (the bell) is shaped so that the wire entering the die will draw

635
ÿ

636 PLASTIC FORMING OF METALS

Die holder Draw head

(a)

Approach angle
Bell
Steel casing

9 %
ÿ I
f
M
A
4
*49 /
t t M I A 4 4
% ÿ
&9
9
fc 4 4
%
* » /M
/ M
X
9
4

9
4
4

4
4
4

i M 9 4 4 4

9 A
%
m M9 4 4 4 9ÿ
<
M
f M4 9 * 4
A
% 9 \ 4 4 4 4
/ T 9 9 4 4 9 9 4 94
ÿ * 4W m
J
M9
Mk
4
*94 *
9 9 4 9 ÿ
4 4 4 * ft §
§ M 9 4 9 449499
U 9 9 4 4

99444*9% M 99 14 4 4*994
4 «
9 9 4 \ f A 9 4*
*
94 9 4 4
4 4 4 9 9 9*
/ M 9 4 9 4 4

t
9
*««>ÿ•«•
M**4*944*
4
4
4
4
9
4 4 9
4 9
9
9
9 % JW
f » 44499949
%44944949
* M 9 4 9 9

9
ÿ
4 4 9 9 4 « ft
V #|
J 9
* 9
Mm 99499444
4

ÿ
M 1 §449444499
999444*44% / |%
M M449494*49
§4 *494*944
%% *
994**44*49%
9 4 4
/
§ M 994*44944
t *4*99449*'
44949*49 4% / M4949*»494
ft / M49944994*
I §9 9 4 * *
4
*
4 9 4
§999494*4*
9 4 9 4 4
4J
*994 l| 11 M 4 4494*9449
9 4 4 9 4 9 4%W
4
4449 *4944 W
499944444
* 1 . Ja 994444949
44***44494
499*449949 4 4 4 4 4 4 9
« « 1 4 1 > 4 1 ÿ
'> ÿ
VI* « * i i « 44
» »
*9
,994499499 4 Jr §944* 9 94

4 + 4 4 4 9 4
V < V * 9* 4 ÿ 9 4 4
449994449 # V !L 9 9 4*4
9 4 4 9 *99 1#

Bearing
Carbide nib

ack relief Figure 19-1 (a) Schematic drawing


of a drawbench; ( b ) cross section
(b) of a drawing die.

lubricant with it. The shape of the bell causes the hydrostatic pressure to increase
and promotes the flow of lubricant into the die. The approach angle is the section
of the die where the actual reduction in diameter occurs. The half die angle a is
an important process parameter. The bearing region does not cause reduction but
it does produce a frictional drag on the wire. The chief function of the bearing
region is to permit the conical approach surface to be refinished (to remove
surface damage due to die wear) without changing the dimensions of the die exit.
The back relief allows the metal to expand slightly as the wire leaves the die. It
also minimizes the possibility of abrasion taking place if the drawing stops or the
die is out of alignment. Most drawing dies are cemented carbide or industrial
diamond (for fine wires). The die nib is encased for protection in a thick steel
casing.
The distinction between wire and rod is somewhat arbitrary. In general the
term wire refers to small diameter products under 5 mm which may be drawn
rapidly on multiple-die machines.
DRAWING OF RODS, WIRES, AND TUBES 637

Die
Wire

Coil
block
Side view of
Top view bull block

Figure 19-2 Wiredrawing equipment (schematic).

Wiredrawing usually starts with a coil of hot-rolled rod. The rod is first
cleaned by pickling to remove any scale which would lead to surface defects or
excessive die wear due to abrasion. The next step is to prepare the rod so that the
lubrication is effective. With high-strength materials a soft surface coating such as
copper or tin may be used as the lubricant. More typically, conversion coatings
such as sulfates or oxalates may be applied to the rod. These are used in
conjunction with a lubricant, typically soap, in dry drawing. In wet drawing the
dies and the rod are completely immersed in an oil lubricant containing an EP
additive.
When the rod diameter is sufficiently small to permit coiling, block drawing is
usually employed because it allows the generation of long lengths in a much
smaller floor space than required for a draw bench (Fig. 19-2). Because the area
reduction per drawing pass is rarely greater than 30-35 percent many reductions
are required to achieve the overall reduction. Multiple block machines with one
die and one draw block for each reduction are common. Since the wire diameter
will decrease after each pass the velocity and length of the wire will increase
proportionately. Thus, the peripheral speed of each draw block must increase in
turn if there is to be no slippage between the wire and the block. One way to
achieve this is to equip each drawing block with its own electric motor with
variable speed control. However, a more economical design is to use a single
electrical motor to drive a series of stepped cones (Fig. 19-3). The diameter of
each cone is designed to produce a peripheral speed equivalent to a certain size
reduction. When precise agreement in wire surface velocity and block peripheral
velocity is not achieved the wire slides on the blocks as they revolve, causing
friction and evolution of heat. The drawing speed in multiple-die machines may
reach 10 m s_1 for ferrous drawing, but with nonferrous drawing speeds up to
30 m s"1 are common.
Heat generation is a major concern in drawing operations. Although most rod
and wiredrawing is done cold, plastic deformation and friction can generate wire
temperatures of several hundred degrees Celsius. This heat is only partially
638 PLASTIC FORMING OF METALS

Electric motor

Blocks

Figure 19-3 Stepped multiple iredrawing. (J. N. Harris," Mechanical Working of Metals
p. 208, Pergamon Press, New York , 1983.)

removed by interpass cooling, and since the dies extract little heat, they get quite
hot.
Nonferrous wire and low-carbon steel wire are produced in a number of
tempers ranging from dead soft to full hard. Depending on the metal and the
reductions involved, intermediate anneals may be required. Steel wire with a
carbon content greater than 0.25 percent is given a special patenting heat
treatment. This consists in heating above the upper critical temperature and then
cooling at a controlled rate or transforming in a lead bath at a temperature
around 315°C to cause the formation of fine pearlite. Patenting produces the best
combination of strength and ductility for the successful drawing of high-carbon
music and spring wire.
Defects in rod and wire can result from defects in the starting rod (seams,
slivers, and pipe) or from the deformation process itself.1 The most common type
of drawing defect is center burst, or chevron cracking (Fig. 15-34c). This also is
called cupping. An upper-bound analysis2 is capable of identifying the combina¬
tions of semidie angle and reduction for which less deformation energy is required
if a hole forms at the centerline. This analysis predicts that center burst fracture
will occur for low die angles at low reductions, and as a increases the critical
reduction for freedom from center burst increases. For a given reduction and die
angle, the critical reduction to prevent fracture increases with the friction.

1
R. N. Wright, Workability in Extrusion and Wire Drawing, "Workability Testing Techniques,"
chap. 9, American Society for Metals, Metals Park, Ohio, 1984.
2
B. Avitzur, "Metal Forming: Process and Analysis," pp. 172-176, 240-241, McGraw Hill Book
Company, New York, 1968.
DRAWING OF RODS, WIRES, AND TUBES 639

P
a

k
b <jx + dax -*=— - ha
\
Figure 19-4 Stresses acting on an element of wire in
P plane-strain strip drawing.
>ÿ

19-3 ANALYSIS OF WIREDRAWING

Although wiredrawing appears to be one of the simplest metalworking processes,


a complete analysis that enables calculation of the draw force to better than ± 20
percent of the observed value is a rather difficult problem. We shall use wiredraw¬
ing as an example of how to add to the model to account for realistic conditions.
We have already seen that the uniform-deformation-energy method predicts a
draw stress given by Eq. (15-13).
A 1
axa ao ln A a0 In - (19-1)
a 1 r
This equation not only neglects friction, but it neglects the influence of transverse
stresses and of redundant (shearing) deformation.
As a first step we shall consider the problem of strip drawing (see Sec. 15-2
and Fig. 15-3) where a Coulomb friction coefficient \i exists between the strip and
the die. The friction stress pp opposes the motion of the strip through the die
(Fig. 19-4). Referring to Eq. (15-6), the equilibrium of forces in the x direction is
oxdh + h dox +2p tan a dx 4- 2}ip dx = 0 (19-2)
and since dh = 2 dx tan a,
oxdh + h dax 4- p (1 + ju cot a) dh = 0 (19-3)
Since the yield condition for plane strain is ox + p Oq and B = ju cot a, the
differential equation for strip drawing is
da dh
~ (19-4)
°xB ffo(l + B) ~h
If B and are both constant, Eq. (19-4) can be integrated directly to give the
draw stress oxa.
B

oxa ao
1+ B
1-
fV 1+ B
!-(!-#ÿ) B
(19-5)
~
°0
B \ h B
However, wiredrawing is conducted with conical dies. An analysis1 following
that for strip drawing but integrating around the circumference of the die results
1
O. Hoffman and G. Sachs, "Introduction to the Theory of Plasticity for Engineers," McGraw-Hill
Book Company, pp. 176-180, 1953.
\

640 PLASTIC FORMING OF METALS

pS sin a

P.\

l±pS cos a Figure 19-5 Forces acting on a conical element

in

IB
1+ B /
Da
°xa °0 1 (19-6)
B

A similar, but slightly different analysis for wiredrawing with friction is given
by Johnson and Rowe.1 The surface area of contact between the wire and the die
(Fig. 19-5) is given by

S = (19-7)
sin a

and the mean normal pressure on this area is p. The forces acting in the axial
direction are given in Fig. 19-5. The draw force Pd is balanced by the horizontal
component of the frictional force and the horizontal component of the normal
pressure.

Pd = ppS cos a + pS sin a (19-8)

Pd — pS(fi cos a + sin a) P —:- ( jix cos a


sin a
~
+ sin a)
Pd=p(A Aj(fic ota + 1) =p(Ah - Aa){\ + B) (19-9)
In the absence of friction, B = 0 and
A
f>d = P(Ab ~
Aa) = °(Aln
Aa

which is really Eq. (19-1). Therefore, the draw stress with friction is given by
P A
axa ao ln A (1+5) (19-10)
Aa a

Example Determine the drawing stress to produce a 20-percent reduction in



a 10-mm stainless steel wire. The flow stress is given by a0o = 1300eO (MPa).
1
R. W. Johnson and G. W. Rowe, J. Inst. Met., vol. 96, p. 105, 1968.
2
J. G. Wistreich, Proc. fust. Mech. Eng. ( London ), vol. 169, pp. 654-665, 1955.
DRAWING OF RODS, WIRES, AND TUBES 641

The die angle is 12° and ju = 0.09


B = [i cot a = 0.09/tan6° = 0.8571
1 1
In = In = 0.223
1- r 1 - 0.2
0.30
Ke? 1300(0.223)
a = 637 MPa
n 4- 1 1.30 ÿ

A 10 mm Aa = Ah - rAh = Ah{ 1 - r) = 10(0.8) = 8 mm


From Eq. (19-6)
B
1+ B
°xa = 1-
° 5 Ab I
/ 1.8571
8571
= 637 [1 - 0.8° ] = 240 MPa
\ 0.8571
From Eq. (19-10)
A
axa a In (1 +B)
Aa
1.0
637 In — (1.8571) = 264 MPa
0.8

Note, that there is about a 10-percent difference between these expressions.


However, they both predict a draw stress appreciably lower than the uniaxial
flow stress.
If the wire is moving through the die at 3 m s~\ determine the power
required to produce the deformation.

distance moved
Power = force x
time
71
Drawing force Pd = oxa Aa = 240 N mm x — (8) mm 12.06 kN

Power = 12.06 kN x 3 m s -i 36.18 kW

Equation (19-10) includes a term for uniform deformation energy and a term
for frictional energy. The redundant deformation can be taken into consideration
by including a factor <j> which allows for the influence of redundant deformation
in raising the flow stress of the material.

°Xa ~ In ~(l + B) (19-11)


Aa
642 PLASTIC FORMING OF METALS

Flow curve for drown wire

Basic flow
curve

Figure 19-6 Procedure for determining redundant de


Strain formation of drawn wire.

The redundant work factor is defined by

<£> =f(a, r ) = (19-12)


where <j>= the redundant work factor
6*= the "enhanced strain" corresponding to the yield stress of the metal
which has been homogeneously deformed to a strain e

The redundant work factor may be determined in a straightforward manner for


drawn wires1 as shown in Fig. 19-6. The flow curve of a drawn wire is superim¬
posed on the flow curve for the annealed metal. The origin of the curve for the
drawn metal is displaced along the strain axis by an amount equal to the drawing
reduction, e = In(Ah/Aa) = In [1/(1 — r)]. The fact that the yield stress for the
drawn wire is above the basic flow curve for the material is due to the redundant
work that it received. To determine <#>, the flow curve for the drawn metal is
moved to the right to e* where the curves coincide. The area swept out in this
procedure is the redundant work per unit of volume. An alternative approach2 is
to use upper-bound analysis to account for the redundant deformation by
assuming simple yet reasonable deformation regions. Unfortunately, this leads to
complex equations.
Rb 2 ( a I R R L \
2/( a) In ——b —I - cot a + 2fx cot a 1 - In In +
Ra ÿ3 \ sin2 a \ Ra Ra R« /
oxa ao
1 + 2 fi(L/Ra)
(19-13)
where /(a) a complex function of semidie angle
L the length of the die land
R , the radius of the billet
R = the wire radius
1
The use of microhardness measurements to determine <f> is discussed by W. A. Backofen,
"Deformation Processing," pp. 138-140, Addison-Wesley Publishing Company, Inc., Reading, Mass.,
1972.
1
B. Avitzur, Trans. ASME J. Eng. Ind.< vol. 86, pp. 305-316, 1964.
DRAWING OF RODS, WIRES, AND TUBES 643

A less complicated, but still useful expression, that is based on an upper-bound


analysis,1 contains the redundant work term, f tan a.
*

/ m R 2
°xa ao V 1 + sin 2 a
In
Ra
H— tan a
3
(19-14)

where m = the friction factor, i.e., t; = km.


The concept of deformation-zone geometry, introduced in Sec. 15-8, provides
a convenient method for treating the redundant work in wiredrawing.2 For the ÿ
%

drawing of round wire


a
1/2 2
A = 1 + (1 -r) (19-15)
r
where a = the approach semiangle, in radians
r= the drawing reduction

Commercial wiredrawing often employs a in the range 6 to 10° and drawing


reductions of about 20 percent. Thus, the A values typically range from 2-3.
Higher values of A correspond to lower reductions and higher die angles, while
lowef values correspond to higher reductions and lower die angles. Analysis of
experimental data shows that the redundant work factor is related to A by3
A
<f> = Cl + C2A « 0.8 + — (19-16)
4.4
The expression for draw stress given in Eq. (19-11) contains terms for
homogeneous deformation, friction, and redundant work. The last two terms are
functions of semidie angle a . The effect of die angle on these components of the
total energy required to cause deformation is shown schematically in Fig. 19-7.
The ideal work of plastic deformation Up is independent of die angle. For a fixed
coefficient of friction, the work required to overcome friction Uf decreases with
increasing a as shown by Eqs. (19-7) and (19-8). On the other hand, the
redundant work Ur increases with increasing a as shown by Eq. (19-14). The
summation of these components of total energy UT leads to a curve which has a
minimum at some optimum die angle a*. Increasing the reduction and the
friction raise the optimum die angle.
The optimum value of a corresponds to a minimum in total energy of
deformation or of draw stress. This also can be expressed in terms of an optimum
value of A.
1/2
A„pt = 4.9 (19-17)
In (1/1 - r)

1
W. F. Hosford and R. M. Caddell, "Metal Forming," pp. 162-163, Prentice-Hall, Inc.,
Englewood Cliffs, N.J., 1983.
2
R. N. Wright, Wire Technology , vol. 4, pp. 57-61, 1976.
3
J. G. Wistreich, Proc. Inst. Mech. Eng., vol. 169, p. 654, 1955; R. M. Caddell and A. G. Atkins,
Trans. ASME J. Eng. Ind., vol. 90, pp. 411-419, 1968.
644 PLASTIC FORMING OF METALS

Constant ÿ and reduction

a Figure 19-7 Components of total energy of deformation.

Flow curve

Sad*

'XQ

Figure 19-8 Development of limit on drawability. (After


max Backofen . )

We now shall examine the limit of drawability for steady-state wiredrawing.1


Equation (19-11) can be expressed most simply by
1 f
oxa = -lode (19-18)
V J

where t] = Up/UT is the efficiency of the deformation process. Figure 19-8 plots
the flow curve of the material, fade for an ideal process without friction or
redundant work, and 1/tjfo de for the actual drawing process. For a given strain
e = In(Ah/Aa) produced by the die the values of draw stress oxa and flow stress
oE are as indicated. The weakest link is the wire which has exited from the die. On
being deformed through the die the material is strain-hardened. For a severely
strain-hardened material the formation of a neck is quickly followed by fracture.
Therefore, the drawing limit is reached when od = o£ as shown in Fig. 19-8. If the
1
W. A. Backofen, op. cit., pp. 227-230; R. M. Caddell and A. G. Atkins, Trans. ASME, Ser. B.
J. Eng. fuel., vol. 91, pp. 664-672, 1969.
DRAWING OF RODS, WIRES, AND TUBES 645

material follows a power-law hardening relationship oE = Ken, then Eq. (19-18)


becomes
n+ 1
Ke °eE
O (19-19)
T](n + 1) ](n
t 4- 1)
Substituting the criterion for the maximum drawing strain in a single pass, that is,
ad

emax = V(" + 1) 5 (19-20)


Since e = In ( A h/A a ),
IAj,\ i)(n + 1)
(19-21)
\ / max

and by definition the reduction is r = 1 — ( Aa/Ah )


-tj(«+ 1)
rmax — 1 (19-22)
For wire subjected to repeated reductions through a series of dies, n will decrease
toward zero and the allowable reduction will decrease. A comparison of the
maximum strain in drawing with that in stretching (tension) shows the important
roleÿplayed by the strain-hardening exponent.
e(drawing) \ 7](n -f 1)
(19-23)
e(stretch) / max n
Typically the limiting strain in drawing is at least twice that in pure tension.

Example For the material drawn in the previous example, what is the largest
possible reduction?
To a first approximation the limit on drawing reduction occurs when
°xa = a
1+ B\ B
°xa a0 1 "(1 ~r)
B 1

/ 1.8571 0.8571
637 = 637 1 - (1 -r)
\ 0.8571
0.8571 1
1 = 2.167 1 - (1 -r) r = 0.51 or e — In 0.71
1
30
a0 = 1,300c 1,300(0.7)° = 1,173 MPa
0.30

A better estimate is to let axa oQ at e = 0.71, i.e., oxa 1173 MPa


0.8571
1173 = 637(2.167)1 - (1 - r)
r = 0.89
Note, that when there is no friction or redundant work (rj = 1), even if there
is no strain hardening (n = 0), Eq. (19-22) predicts that rmax = 1 - l/e =
0.63.
646 PLASTIC FORMING OF METALS

19-4 TUBE-DRAWING PROCESSES

Hollow cylinders, or tubes, which are made by hoi-forming processes such as


extrusion or piercing and rolling (see Chap. 18), often are cold finished by
drawing. Cold-drawing is used to obtain closer dimensional tolerances, to produce
better surface finishes, to increase the mechanical properties of the tube by strain
hardening, to produce tubes with thinner walls or smaller diameters than can be
obtained with hot-forming methods, and to produce tubes of irregular shapes.
The three basic types of tube-drawing processes are sinking, plug drawing,
and mandrel drawing. Since the inside of the tube is not supported in tube sinking
(Fig. 19-9a), the wall thickens slightly and the internal surface becomes uneven.
Because the shearing at the entry and exit of the die is large, the redundant strain
is higher for sinking and the limiting deformation is lower than for other
tube-producing processes. Both the inner and outer diameters of the tube are
controlled in drawing on a plug (Fig. 19-96). The plug may be either cylindrical
or conical. The plug controls the size and shape of the inside diameter and
produces tubing of greater dimensional accuracy than in tube sinking. Because of
the increased friction from the plug, the reduction in area seldom exceeds 30
percent. The situation where a carefully matched plug floats in the die throat is
shown in Fig. 19-9c. Properly designed floating plugs can give a reduction in area
of 45 percent, and for the same reduction the drawing loads are lower than for
drawing with a fixed plug. An important feature of this design is that it is possible
to draw and coil long lengths of tubing. However, tool design and lubrication can
be very critical. Problems with friction in tube drawing are minimized in drawing
with a long mandrel (Fig. 19-9d). The mandrel consists of a long hard rod or wire
that extends over the entire length of the tube and is drawn through the die with
the tube. In tube drawing with a moving mandrel the draw force is transmitted to
the metal partly by the pull on the exit section and partly by the friction forces
acting along the tube-mandrel interface. Since the mandrel is moving with a

L draft

m,

{a) {c (d)

Figure 19-9 Methods of tube drawing, (a) Sinking; (b) fixed plug; (c) floating plug; ( d ) moving
mandrel.
DRAWING OF RODS, WIRES, AND TUBES 647

velocity equal to that of the tube on exiting from the die, and since this is higher
than the velocity of the metal confined in the die channel, there is a forward
frictional drag at the interface between the mandrel and the tube which tends to
* *

cancel the backward frictional drag between the stationary die and the tube.
However, after drawing, the mandrel must be removed from the tube by rolling
(reeling), which increases the tube diameter slightly and disturbs the dimensional
tolerances.

ÿ
v

19-5 ANALYSIS OF TUBE DRAWING

In the tube drawing with a plug or mandrel the greatest part of the deformation
occurs as a reduction in wall thickness. Usually the inside diameter is reduced by
a small amount equal to that needed to allow for insertion of the plug or mandrel
before drawing. Thus, there is no hoop strain and the analysis1 can be based on
plane-strain conditions.
For tube drawing with a plug, the draw stress, by analogy with Eq. (19-5), can
be expressed by

=
1+ B
1-
'ha\B' (19-24)
°xa °o B' \hbi
ix i 4- ii2
where B'
tan a — tan

and fi1 = coefficient of friction between tube and die wall


ft 2 = coefficient of friction between tube and plug
a = semicone angle of die
/?= semicone angle of the plug; for a cylindrical plug, =0

In tube drawing with a floating plug the question of proper design of the plug
is more important than the magnitude of the drawing stress. This has been
discussed by Blazynski.2
In tube drawing with a moving mandrel, the friction forces at the stationary
die-tube interface are directed toward the die entrance as for wiredrawing and
tube drawing with a plug. However, as discussed earlier, the friction forces at the
mandrel-tube interface are directed toward the exit of the die. Therefore, for a
moving mandrel B' must be expressed as

g,_
tan a - tan j8

The draw stress is given by Eq. (19-24) with the proper value of B'. If — ju2, as

1
Hoffman and Sachs, op. cit., pp. 190-195.
2
T. Z. Blazynski, Metall. Rev., no. 142, April, 1970; D. J. Smith and A. N. Bramley. Proc. J 3th
Int. Machine Tool Design and Research Con]., Birmingham, England. 1972.
648 PLASTIC FORMING OF METALS

may often be the case, then B' = 0. The differential equation of equilibrium for
this simple case is

hdox 4- {ox + p) dh = 0
which integrates directly to the equation for ideal homogeneous deformation Eq.
(19-1). However, it is entirely possible for the coefficient of friction on the
mandrel ju2 to exceed that at the die jliÿ so that B is negative. This situation
results in a draw stress less than required by frictionless ideal deformation.
Tube sinking is the process by which the inside diameter of the tube is
decreased. Although we have neglected this aspect of the deformation until now,
some reduction in inside diameter is part of the tube-drawing process.
The stresses involved in tube sinking have been analyzed by Sachs and
Baldwin1 on the assumption that the wall thickness of the tube remains constant.
The equation or the draw stress at the die exit is analogous with the equation
describing the draw stress in wiredrawing. The cross-sectional area of the tube is
related to the midradius r and the wall thickness h by A « lirrh.

axa O0
ft
1 + B
1 ' f\B
A
(19-25)
B Abj
The yield stress Oq is taken equal to 1.1a0 to account for the complex stresses in
tube sinking. A more complete analysis of tube sinking has been given by Swift.2

19-6 RESIDUAL STRESSES IN ROD, WIRE, AND TUBES

Two distinct types of residual-stress patterns are found in cold-drawn rod and
wire, depending on the amount of reduction. For reductions per pass of less than
about 1 percent the longitudinal residual stresses are compressive at the surface
and tensile at the axis, the radial stresses are tensile at the axis and drop off to
zero at the free surface, while circumferential stresses follow the same trend as the
longitudinal residual stresses. For larger reductions of commercial significance the
residual-stress distribution is completely reversed from the first type of stress
pattern. In this case the longitudinal stresses are tensile at the surface and
compressive at the axis of the rod, the radial stresses are compressive at the axis,
and the circumferential stresses follow the same pattern as the longitudinal
stresses. The first type of residual-stress pattern is characteristic of forming
operations where the deformation is localized in the surface layers.
The effect of die angle and the amount of reduction per pass on the
longitudinal residual stress in cold-drawn brass wire was investigated by Linicus

1
G. Sachs and W. M. Baldwin, Jr., Trans. ASME, vol. 68, pp. 655-662, 1946; also see Hoffman
and Sachs, op. cit., pp. 252-255.
2
H. W. Swift, Phiios. Mag,, vol. 40, ser. 7, pp. 883-902, 1949.
DRAWING OF RODS, WIRES, AND TUBES 649

500

Holt die angle


400
o - 8"
x -
16°
+ -32°
CD
|300
c/>
9S>
4>

O
c 200
3

100

Figure 19-10 Longitudinal residual


20 30 40 50 60 70 stress in cold-drawn brass wire. (After
Reduction in area by' drawing, % Linicus and Sachs.)

and Sachs.1 Figure 19-10 shows that for a given reduction the longitudinal
residual stress increases with the half-die angle. Maximum values of residual
stress are obtained for reductions in the region of 15 to 35 percent.
For tubes produced by tube sinking, under conditions where the deformation
is relatively uniform throughout the tube wall, the longitudinal residual stresses
are tensile on the outer surface and compressive on the inner surface of the tube.
The residual stresses in the circumferential direction follow the same pattern,
while the stresses in the radial direction are negligible. Approximate measure¬
ments2 of the circumferential stresses on the outer surface of sunk tubes indicate
that the stresses increase with increasing reduction in diameter at the same rate at
which the yield stress is increased by the cold-work.
Studies3 of residual stresses in tubes produced by drawing over a plug and
mandrel showed the same distribution of residual stresses as for tube sinking. An
important result was that a substantial reduction in the level of residual stress
could be produced by tandem drawing, whereby a small reduction (2 percent) was
produced by a second die immediately following the main reduction. A systematic
approach to developing tube-drawing schedules has been proposed4 based on the
criterion of developing zero circumferential residual stress.

W. Lincius and G. Sachs, Mitt. Dtsch. Materialprufungsanst vol. 16, pp. 38-67, 1932; R. N.
1

Wright, Wire Tech., vol. 6, pp. 131-135, 1978.


2
D. K. Crampton, Trans. Metall. Soc. AIME, vol. 89, pp. 233-255, 1930.
3
S. K. Misra and N. H. Polakowski, Trans. ASME, Ser. D: J. Basic Eng. vol. 91, pp. 810-815,
1969.
4
B. J. Meadows, and A. G. Lawrence, J. Inst. Met., vol. 98, pp. 102-105, 1970.
650 PLASTIC FORMING OF METALS

BIBLIOGRAPHY

Bernhoeft, C. P.: "The Fundamentals of Wire Drawing," The Wire Industry Ltd., London, 1962.
Blazynski, T. Z.: "Metal Forming: Tool Profiles and Flow," Chaps. 4-6, John Wiley-Halsted Press,
New York, 1976.
Collins, L. W., J. G. Dunleavy, and O. J. Tassi, feds.): "Nonferrous Wire Handbook," vol. 1, 1977,
vol. 2, 1981, Wire Association, Inc., Branford, Conn.
Dove, A. B. (ed.): "Steel Wire Handbook," vol. 1, 1968, vol. 2, 1969, vol. 3, 1972, vol. 4, 1980, Wire
Association, Inc., Branford, Conn.
Rowe, G. W.: Wire Manufacture, Int. Met. Rev., vol. 22, pp. 341-354, 1977.
Wistreich, J. G.: The Fundamentals of Wire Drawing, Metall. Rev., vol. 3, pp. 97-142, 1958.

Das könnte Ihnen auch gefallen