Download as pdf or txt
Download as pdf or txt
You are on page 1of 145

This dissertation h a s been

m icrofilm ed exactly a s received 6 7 -6357

PIERCE, Cyril Marvin, 1939-


FORCES INVOLVED IN THE AXISYMMETRIC EXTRUSION
OF M E T A L S T H R O U G H CONICAL DIES.
The Ohio State University, Ph.D„ 1966
Engineering, metallurgy

University Microfilms, Inc., Ann Arbor, Michigan


FORCES INVOLVED IN THE AXISYMMETRIC

EXTRUSION OF METALS THROUGH

CONICAL DIES

DISSERTATION

Presented in Partial Fulfillment of the Requirements for


Degree Doctor of Philosophy in the Graduate
School of The Ohio State University

By

Cyril Marvin Pierce, S.B., M.S.

* * * * * *

The Ohio State University


1966

Approved by

Department of Metallurgical
Engineering
AC KNOWLEDGMENTS

The author wishes to express his sincere appreciation

to Dr. J.W. spretnak of Ohio State University for his

guidance, encouragement, and patience throughout this project.

Grateful acknowledgment is given to the Air Force Materials


Laboratory, and in particular to Dr. Harris Burte, Mr. I.

Perlmutter, and Mr. Vincent DePierre, for providing the

opportunity to pursue graduate studies at the Ohio State

University and to conduct the research described in this

report. Pilot plant research requires the cooperation and

perseverance of several people, and the author wishes to

recognize the many ideas and talents of the following

individuals who aided in the experimentation: Mr. D. Carnahan, •

Mr. P. DuManoir, Mr. S. Inouye, Mr. T. Jones, Mr. M. Myers,

Mr. G. Saul and Mr. R. Sweeney.


VITA

February 15, 1939 Born — Boston, Mass.

1960 ........... S.B., Mass. Inst, of Tech.,


Cambridge, Mass.

1960 - 1961 .......... Research Assistant, Metallurgy


Department, Mass. Inst, of Tech.,
Cambridge, Mass.

1961 , . . . . . . M.S., Mass. Inst, of Tech.,


Cambridge, Mass.

1961 . . . . . . Research Engineer, Watertown


Arsenal Foundry, Watertown, Mass.

1961 1964 ...........Air Force Officer, Air Force


Mat er ia1s Laboratory, Wright-
Patterson AFB, Dayton, Ohio.

1964 - 1 9 6 6 .......... Project Engineer, Air Force


Materials Laboratory, Wright-
Patterson AFB, Dayton, Ohio.

1961 - 1966 . . . . . . Graduate Student, Department of


Metallurgical Engineering, The Ohio
State University, Columbus, Ohio.
CONTENTS

Chapter Page

Acknowledgments ii
Vita iii
Contents iv
Tables vi
Illustrations vii
Symbols x

I INTRODUCTION 1

The extrusion process 1


Purpose and scope of the study 4
Extrusion forces 5

II THEORETICAL ANALYSES OF THE EXTRUSION PROCESS 7

Analytic treatment of container friction 7


Energy for uniform deformation 8
Slab method of stress analysis 10
Slip-line solutions 14
Upper-bound solutions 16
Visioplasticity 20

III FURTHER DEVELOPMENT OF EXTRUSION THEORY 22

Alternate treatment of container friction 22


Application of work hardening theory to
the slab method of stress analysis 24
Upper-bound solutions with the triangle
velocity field 26

IV EXPERIMENTAL PROCEDURE 36

Description of the extrusion equipment 36


Determination of the total applied force 37
Determination of the deformation force 38
Temperature determination by infrared
radiation measurements 43
Temperature determination by calori-
metric measurements 44
Summary of experiments 45

iv
CONTENTS

Chapter Page

V RESULTS 49

Description of thepressure-distance curves 49


Temperature measurements 50
Evaluation of the deformation pressure-flow
stress ratio 51

VI ‘ DISCUSSION 53

Effect of hillet length 53


Effect of lubrication 55
Effect of die angle 59
Comparison of results with deformation
theories 59

VII SUMMARY AND CONCLUSIONS 62

VIII APPENDIXES 64

A Analyses of slab solution as.B


approaches zero 64
B Determination of energy dissipation
on boundary surfaces 65
C Analysis of calorimetric measurements 69
D Force due to acceleration 71
E Determination of the average strain
rate 73

IX BIBLIOGRAPHY 76

V
TABLES

Table Page

3.1 Calculated maximum value of ^ 80

3.2 , Various contributions to the total P^/Y for


R = 6 and © C = 60° 81
3.3 Values of the velocity field geometry factors
which yielded minimum predicted pressures 82
4.1 Compositions of the billet materials 83
4.2 Preparation of Corning Nr. 0010 glass mixture 34

4.3 Extrusion experiments with 1018 steel at


1800°F and 45° die angle 85
4.4 Extrusion experiments with 1018 steel at
1800°F and 60° die angle 86

4.5 Extrusion experiments with OFHC copper at


800°F and 45® die angle 87
4.6 Extrusion experiments with OFHC copper at
800°F and 60° die angle 88
6.1 Effect of billet length on the pressure
required to extrude unlubricated steel at
1800°F 89
6.2 Effect of billet length on the pressure
required to extrude 00 10 glass lubricated
1018 steel at 1800°F 90
6.3 Effect of lubrication on the pressure
required to extrude OFHC copper at 800°F 91
6.4 Calculated values of yL\c.and for OFHC
copper at 800°F 92

vi
ILLUSTRATIONS

Figure Page

1.1 Photograph of 700 ton horizontal extrusion


press 93

1.2 Schematic illustration of forward extrusion


process 94

1.3 Schematic representation of forces during


extrusion 95

2.1 Idealized pressure distribution in an


extrusion billet 96

2.2 Dimensions in extrusion 97

2.3 Normal pressure on surface element 98

2.4 Frictional force on surface element 98

2.5 pd/Y vs» ln R from slab method of stress


analysis 99

2.6 Slip-line field for r less than 2 sin /


(1+2 sin oC. ) 100

2.7 Slip-line field for r greater than


2 sin o C /(1+2 sin oC ) 101

2 .8 Pd/Y vs* ln R from siip-line solution for


frictionless extrusion 102

2.9 Velocity field with cylindrical deforming


regions for straight die extrusion 103

2.10 Radial velocity field for axisymraetric


extrus ion 104

2.11 Pd/Y v s * I*1 R from upper-bound solutions 105

2.12 Green's proposed velocity field for sheet


drawing 106

3.1 Schematic relationship between the inter­


face shear stress and the normal pressure 107

3 .2 Pd/Y t vs* R for low values of the work


hardening coefficient 108

vii
ILLUSTRATIONS

Figure Pagd
3.3 vs. ln R for high values of the
work hardening coefficient 109
3.4 Triangle velocity field representation
for conical die extrusions 110
3.5 Displacement of outer region 111
3.6 Velocity diagram of the 45 line 112
3.7 from triangle velocity field 113
4.1 Schematic illustration of AFML extrusion
press with load cell 114
4.2 Modified die backer for load cell 115

4.3 Load cell dimensions 116

4.4 Equipment modifications for radiation


pyrometry 117

4.5 Photograph of extrusion calorimeter 118


4.6 Extrusion billet dimensions 119

5.1 Pressure-distance curves for OFHC copper 120

5.2 Pressure-distance curves for 1018 steel 121


5.3 Log^o strain rate vs. log^o flow stress of
copper at 800°F 122
5.4 Log-,n strain rate vs. login flow stress of
steel at 1800°F iU 123
6 .1 Measured exit temperature vs. Et/V for OFHC
copper at 800°F and c*C = 45° 124
6 .2 Photograph- of grids on lubricated and un­
lubricated steel extrusion billets 125
6.3 Ln(Ft/F(j) and (Ft - F^) vs. L for extrusion
Nr. 1720 126

6.4 Deformation pressure vs. ln R for ©<. = 45° 127

viii
ILLUSTRATIONS

Figure Page

6.5 Deformation pressure vs. ln R for «=<. = 60° 128

6.6 Experimental P^s/Y vs. ln R for lubricated


copper and steel billets 129

6.7 Experimental P(js/Y vs. ln R for unlubricated


copper and steel billets 130
SYMBOLS

height of the deformation zone

constants in the equation for


heat capacity Cp = a+bT+cT

constants

transverse area of the inside of


the container liner

transverse area of the extruded


bar

real area of contact at an inter-


face

radius of the extruded bar

f/tan oC

slip-line field constants

heat capacity of the bar in


the temperature range T.t to Tj

heat capacity of the bar in the


temperature 2 9 8 .2 *K to Tg

heat capacity of graphite

diameter of the billet at anytime


during the operation

inside diameter of the container


liner

diameter of the extruded bar

exponential

energy required to produce defor­


mation (includes energy dissipated
as friction on the die face)

total energy required for extrusion

true rate of work

x
SYMBOLS

coefficient of friction on the


die face

f(oC ) Avitzur's function of o C

*a force due to acceleration of the


billet

Fd force required to deform the


material (includes the friction
force on the die face)

friction force at the die-con-


tainer liner interface

1fb friction force at the billet-


container liner interface

'ff friction force at the follower


material-container liner inter­
face

JLc force measured by the load cell

total force required to sustain


extrusion

acceleration due to gravity

G,G‘/H/H' functions of r

nrt “ H298.2 enthalpy increment from 298.2 to


Tt
k shear stress

K constant representing the shear


stress of an interface

Ki‘
H heat loss coefficient

length of billet in the container


liner

original billet length

length of extruded bar

xx
SYMBOLS

mass of billet

mass of extruded bar

applied normal load

normal preaeura
deformation pressure (F^/Aq )

steady state value of the defor­


mation pressure

total pressure (Ft/Ac )

radius

reduction ratio (Ac/Ap)

tangential velocity discontinuity


on a boundary

surface area

time

temperature of the water after


one hour

initial temperature of the bar


when it enters the calorimeter

initial temperature of the water

highest transformation tempera­


ture below Tj

velocity component

velocity component

velocity of extruded bar

volume

ram velocity

no. of moles of extruded bar

xii
SYMBOLS

no. of moles of graphite

weight of water

distance coordinate

flow stress

flow stress at £_ = 1

average flow stress

flow stress at .2% strain •

cylindrical coordinate

half-angle of a conical die

slip-line curve designations

redundant deformation and fric­


tion factor

geometric variables in the tri­


angular velocity field

angle in the triangular velocity


field

rate of true strain

average rate of true strain

effective strain rate under the


admissible velocity fields

angle of rotation

coefficient of friction

coefficient of friction at the


billet-container liner interface

density

effective stress under the admis­


sible velocity fields

xiii
SYMBOLS

principal stresses

average flow stress of junctions


at an interface

anti-clockwise rotation angle


between the slip-line and the
chosen x-direction

effective deformation

shear strength
CHAPTER I

INTRODUCTION

The extrusion process

Compared with most metalworking processes, extrusion

is relatively new having first been commercially practiced


for the manufacture of lead pipes in the beginning of the

nineteenth century. In recent years it has found increased

utilization in the fabrication schedule of many metallic- pro­

ducts since the operation is applicable both to the "easy

to form" and the "difficult to form" materials. In the for­

mer case (Cu, Al, etc.) it is used to produce final shapes

with excellent dimensional tolerances and surface finishes,

requiring little, if any, additional machining. In the

latter instance (refractory metal alloys, superalloys, etc.),

a resulting compressive component of stress makes this pro­

cess suitable for primary breakdown, which involves the

conversion of cast or sintered billets into intermediate

shapes characterized by metallurgical structures more amenable

to further processing into wrought end products.

Industrial extrusion techniques are treated in depth

by Pearson and Parkins (reference 1), and by L.V. Prozorov

(reference 2), but Figures 1.1 and 1.2 are presented in

order to introduce some of the basic features of the direct

extrusion operation. In this process the billet is pushed

by a stem into the die, while the term "indirect extrusion"

is reserved for those processes where the billet remains


stationary and the die is continually advanced. The illus­

trated press is of the more common horizontal design and

derives its power from a high pressure hydraulic accumulator.

Although many presses utilize a high pressure fluid supplied

straight from the pumps, these generally are of low capacity

and are severely restricted to slower operating speeds.

The press can be visualized as being divided into two

sections, one to control delivery of the energy and a second

to counteract the resultant forces. Each section is individ­

ually secured in bedrock, but they are united by columns

(commonly termed tie rods). The particular press in Figure

1 .2 is actually a four-column press, only two of which can

be seen in the drawing.

The die is commonly positioned in one of two locations —

either in the container liner or in a special fixture (termed

a die holder) which is positioned directly behind the con­

tainer. The die is secured by the die backer which in turn

is supported by the bolster plate. The container is generally

resistance heated as high as 800°F for hot extrusion to avoid

the rapid chilling of billet materials when billet-tooling

contact is established. In such instances all the contact

tools are made of hot working tool steels,-usually heat

treated to a room temperature hardness of 40-60 on the

Rockwell C scale with the exact specifications depending on

the particular steel involved and the specific function of

the part.
When desired, billets are heated in auxiliary furnaces

of a variety of types before being positioned in the container

liner. When difficult to form metals are to be extruded, a

nose block of softer material shaped to the die contour is


generally placed ahead of the billet to insure a more uni­

form pressure on the die face as the force is applied. This

practice eliminates the need for machining the expensive

billet material and can lead to significant cost reductions.

A graphite follower block is then placed directly behind

the billet to force the entire billet through the die and to

thus avoid a time consuming removal operation. This technique

is essential when extruding materials which exhibit thermal

shock sensitivity since cooling rapidly from high processing

temperatures, which occurs if the billet end remains in the

die, can result in the formation of tensile stresses of

sufficient magnitude to cause crack formation within the

product. Such materials must be completely extruded, then

the bar quickly removed from the runout tube and placed in

an insulating media.

A tool steel cylinder (referred to in this country as

a dummy block and in Great Britain as a pressure pad) is

positioned behind the graphite in order to protect the stem

from the billet heat and the abrasiveness of chilled metal

and lubricants.

Ideally, the deformation process occurs in the die

while the remainder of the billet is compressively restrained


in the container liner. It is this constraint which permits

the very high reductions germane to the extrusion operation

since buckling or tensile instability cannot occur in the

billet material. For example, wire drawing (which exhibits

a similar stress field in the deformation zone) is commonly

limited to approximately sixty percent reduction of area

per pass while the area reduction ratios up to 100:1 (99 per­

cent reduction) can be obtained in a single pass during the

extrusion of soft materials such as annealed copper, aluminum

or lead.

Purpose and scope of the study

What then limits the extrudability of a given material?

In many instances, the particular material under consideration

does not exhibit sufficient deformability, even under this

ductilizing state of stress, to undergo large reductions in

area; while in other cases the strength of the material

(defined as the flow stress in uniaxial tension at the pro­

cessing temperature and strain rate) is sufficient to require

pressures above that which the tooling can sustain (usually

200,000 psi). It is the latter limitation which will be

fully explored in this dissertation; that is, the forces

involved in the extrusion operation will be measured and

analyzed as functions of process parameters in order to

minimize the total required pressure and thus extend the

extrudability of a given material. Since the limits imposed

by inadequate deformability will not be considered, the


analyses and experiments will deal with materials that can

he treated as continua during the processing operation.

Extrusion forces

In order to extrude a billet at constant velocity,

the force delivered by the stem must equal the several

opposing forces illustrated in Figure 1.3 or quite simply:

(1.1)
where Ft = total force required to sustain extrusion

Fff = friction force between the container and the


follower material

F^. = friction force between the container and the


upset billet

F^ « force required for deformation (includes the


friction force between the material and the
die)

Various analytical techniques, with differing degrees of

success, have been incorporated in attempts to predict the

deformation force, Fd , for the production of solid round

bars, and new solutions to this problem will be presented

as part of this dissertation. In such an analysis, the

material's properties as well as the important process

parameters must be treated as variables and, therefore,

F^ = function (Y, oC, R, f) (1 .2 .)


where Y = property representing the strength of the
material (usually taken as the average flow
stress in uniaxial tension at the processing
temperature and strain rate)

oC= half angle of the conical die

R as area reduction ratio (a measure of the degree


of deformation)

f = coefficient of friction between the die and


the deforming material

Several proposed forms of equation (1.2) will be dis­

cussed in the following chapters and their applicability

will be determined by comparison with experimental results.

Such comparisons are possible only because of the installa­

tion of unique instrumentation (described in Chapter IV)

to directly measure and thereby determine what might be

termed "inherent extrudability," since this parameter should

be independent of frictional considerations in the container.

Heretofore, the only force measured by fabricators has been

the total force, Ft , as detected by a pressure transducer

in the hydraulic system. In addition, when values of Ft

and Fd are continuously recorded during the extrusion process,

a method for in-process lubricant evaluation is provided

since the quantity Fff + Ff^ can be calculated at any time

through use of equation (1 .1 ).


CHAPTER II

THEORETICAL ANALYSES OP THE EXTRUSION PROCESS

A number of methods are available for theoretically

analyzing a deformation process# but all of these fall into

one of two classifications (reference 4): those based on a

consideration of stresses and those which rely on a knowl­

edge of metal flow patterns. The first three techniques

discussed in this chapter are of the former type while the

remaining approaches are in the latter category.

Analytic treatment of container friction

By utilizing the method presented in Pearson and Parkins'

book (reference 3)# it is possible to relate the total force,

Ft , to the deformation force, F^, provided that the difference

between these forces is entirely ascribable to the friction

force at the billet-container interface, F ^ . Consider the

equilibrium of forces about the differential volume element

illustrated in Figure 2.1.

If the billet is assumed to be in a hydrostatic state

of stress, then the pressure against the container liner is

also equal to Pt , and the friction force on the slice dx

can be expressed as:

^ ts - TT D c / V d* (2 - 2)

7
where y U c = coefficient of friction between the billet and
the container liner

Substitution of dPf^ in equations (2.1) and (2.2)

yields the governing differential equation.

(2.3)
This expression is then integrated from the deformation

region (x = o, Pt = P^) to the end of the billet at any time

(x = L, = pt) with the assumption that is constant

throughout the processing operation. The result in terms

of force variables is

(2.4)

Energy for uniform deformation

A common approach to load determinations in deformation

processing is to first estimate the energy necessary to pro­

duce the desired shape change and to then calculate the force

required to generate this work criterion.

As early as thirty years ago# Siebel (reference 5) used

this technique to make the first calculations of the force

required to deform a metal during the extrusion operation.

By neglecting friction on the die face and by assuming a

condition of uniform deformation, he was able to equate the

input work to the energy for deformation as follows:

F,Cu)=VUH*-A«.UitY) (2.5)
where \ <, = initial length of billet

A\c.“ transverse billet area (same as~K ~ ^ “


after upsetting has occurred)

\/ = initial volume of the billet

Y = flow stress

= effective deformation

It is now well established that for cold working# Y ’

must be treated as a function of (reference 6 ), and it

is for this reason that Siebel's analysis is generally

written in terms of an average flow stress# Y * calculated

over the entire range of deformation. The effective defor­

mation was taken to be the true strain in the axial direction

or

-T^- - - J/w R (2 .6 )
l—o

where L-p = length of the extruded bar

= transverse area of the extruded bar

= area reduction ratio

Upon substituting the above expression into equation

(2.5), Siebel concluded that the deformation force could be

predicted by

r ^ = f \ CLx J u . R (2.7)

Forces calculated in this manner are much lower than

those actually required in practice (references 7# 8 # and 9)

and the discrepancy is attributed to the over-simplifying

assumptions of homogeneous flow and zero friction. It is


10

just too unrealistic to propose that all the delivered ener­

gy is purely utilized in achieving a shape change and that no

redundant deformation (a form of energy dissipation) can oc­

cur. Equation (2.7) is usually considered only after multi­

plying the right-hand side by an experimentally determined

efficiency factor, termed P e . The usefulness of this energy

approach, then, seems to be in the fact that the theory pre­

dicts a logarithmic functional relationship between the re­

quired deformation force and the reduction- ratio, which is

confirmed by experimental observations.

Slab method of stress analysis

The analysis, first proposed by Sachs (reference 10)

and later presented by Hoffman and Sachs- (reference 11), is

based on the assumption of a uniform state of stress at all

points of a plane normal to the die axis. The assumed state

of stress is such that the axial and radial directions were

both principal directions having principal stresses of (J^

a n d -^ ' , respectively.

Consider the equilibrium of a volume element bounded by

two planes perpendicular to the x axis at distances x and

x + dx from the apex of the conical die (see Figure 2.2).

The following forces are axisymmetrically distributed? hence,

their resultants are acting along the x axis. .

(a) The O'* stresses on the two transverse planes yield

the resultant, neglecting infinitesimals of higher order -(see

Figure 2.2).
11

(2 .8 )

(b) The resultant of the normal pressures on the sur­

face in contact with the die is determined by considering

a surface element, shown in Figure 2.3 obtained by inter­

secting the conical die surface with two radial planes in­

cluding the angle A ©■. The total normal pressure on the

surface element is

which yields the axial component

Integrating this expression from © — O to & - Z/tt one ob­

tains the resultant of the normal pressures on the volume

element

(2.9)
o

(c) The resultant of the frictional forces is obtained

similarly by considering a surface element with the dimen-

sions dx/cos oC and^ ^ © (see Figure. 2.4). The total

frictional force on the surface element is


12

with the axial component

( ^ d ^ ) 0 ^ A e) C a S o c =

Integrating this expression from Q - 0 to 0 = 2vr, the

resultant of the frictional forces on the bar element be­

comes

(2 .10)

The equilibrium equation expresses that the sum of the

forces given by Eqs. (2.8) and (2.9) vanishes, or

^(D-vdsC-vig-x<il5)+--^IC>AE>+. ^ ^ _ i_D i D = o

Simplifying, one obtains the following differential equation:

£u<s5-v2xncADvj4= Ai>G+-|^r) = o (2.id

In the "cylindrical" stress state assumed here, both

the maximum-shearing stress criterion and the distortion-

energy criterion of plasticity are expressed by the same

relation between the principal stresses;

<£-<35e = Y
where
O vTT.=
y as defined in equation (2.5)
Therefore, the above equation can be written as:

^> = Y - ( T x (2.12)

For simplicity, the parameter "B" is introduced and


defined as

Equation (2.11) can’now be written as

(2.13)

If work hardening is neglected and the friction condi-

tion along the die face is invariant, then \ and ^ are

constants, and equation (2.13) can be integrated by parts to

yield the following expression for the deformation force.

(2.14)

In the limiting case of B = 0, it can be shown (see

Appendix A) that this relationship reduces to Siebel's

formula, equation (2.7), thereby illustrating a similarity

between the two solutions stemming from the assumption of

uniform cross-sectional flow. The results of these analyses

are displayed in Figure 2.5 where the required deformation

pressure, P^, divided by the average flow s t r e s s , V , is

plotted as a function of the reduction ratio, R, for various

values of the "friction-angle" parameter, B.


14

SIip-line solutions

This analysis is based on the fact that in plane strain

deformation, it is possible to characterize metal flow by a

net of curves, termed slip-lines, along which the shearing

stress has its maximum value. The applicability of this

technique to axisymmetric extrusion problems stems from the

prevalent willingness of many investigators (references 1 2 ,

13, 14, 15) to extend its range of relevance to certain three

dimensional geometries. The method permits an analysis of

the nonhomogeneous nature of the deformation7 but, to date,

since the treatment has been limited to rigid, perfectly

plastic, isotropic solids, it cannot precisely deal with

materials that exhibit either strain hardening or strain rate

sensitivity.

Each slip-line belongs to one of two families of curves,

OC or p , which always intersect at right angles. The

properties of the lines are such that the stress relationships

along a line can be defined through use of equations first

derived by Hencky (reference 16)

(2_is)

= C . 2.
where = mean compressive stress

shear stress

^ = anti-clockwise rotation angle between the


oC - line and the chosen x-direction

^-1= constant along an ©<. - line

C ^ = constant along a - line


15

A similar set of equations proposed by Geiringer

(reference 17) define the velocity components and V along

slip-lines, respectively.

cl — Vci-4> =• O on an © d slip-line
d V ° on a p slip-line (2.16)

Additionally, it is possible through a proper rotation

of axes to express the plane strain equilibrium equations

Q x “ ‘"-''Ip "V* Tv
(2_17)

/b K’y c.& ■£>

By use of equations (2.15, 2.16, and 2.17), it is

possible to analyze a given slip-line field to obtain the

complete stress and velocity distribution. The major diffi­

culty in the technique lies in the construction of proper

fields which truly represent the metal flow and are consis­

tent with the imposed boundary conditions. At present, there

are three approaches employed in the construction of slip-

line fields, Hill's numerical process (reference 12),

Prager's graphical technique (reference 18), and Thomsen's

wedge method (reference 19).

Johnson (reference 20) completely analyzed the sheet

extrusion operation through wedge shaped dies and thereby

extended the solutions originally proposed by Hill in 1948

(reference 21). The fields which he examined for the case of


16

frictionless deformation are shown as Figures 2.6 and 2.7,

while the corresponding results are illustrated in Figure

2.8. The information on this graph was extracted from

Figure 4 of Johnson's referenced paper but employs the

natural logarithm of the reduction ratio as abscissa rather

than the percent reduction as in the original. It is of

interest to note that as the amount of deformation diminishes,

the predicted extrusion pressures do not extrapolate to zero

as do those from the slab analysis in Figure 2.5. Without

introducing any significant errors, all the slip line solu­

tions can be said to be of the form

(2.18)

where and A 2 are constants, although better correlation

exists if only reduction ratios greater than four are con­

sidered. In a later paper (reference 24), Johnson reported

that this form of equation accurately predicted the pressures

observed during the axisymmetric extrusion of lead and alum­

inum through square dies if A^ and A 2 were set equal to .8 .

and 1.5, respectively.

Upper-bound solutions

The concept of an upper bound is treated rigorously by

Johnson and Mellor (reference 22) and by Johnson and Kudo

(reference 23). In essence, the theorem states that the

true rate at which work is performed by an applied force is


17

smaller than the internal rate of energy dissipation derived

from an admissible velocity field in the material. There­

fore, an extrusion pressure that is calculated from a stress

system which is compatible with a valid but incorrect strain


field will always be greater than the true pressure. The

proof hinges on the principle of maximum work which predicts

that materials will flow in such a manner as to offer maxi­

mum resistance to the deforming force. This result possesses

important practical significance since fabricators may use

these solutions to schedule working operations and, because

of the above mentioned built-in safety factor, need not be

concerned with exceeding their equipment capabilities.

The internal energy can be dissipated by the steady

deformation of a volume of the plastic body as well as by

that part of the field boundary where a tangential velocity

discontinuity occurs. Mathematically the theorem can be

expressed as

(2.19)

where the true rate of work

<T,6. = effective stress and effective strain rate under


the admissible velocity fields

= volume element

= surface coefficient (■£-=■ '/PS for shear along a


boundary, and f = o for the case of frictionless
sliding),

= tangential velocity discontinuity on a boundary

d ^ = surface element along the boundary


18

In direct extrusion, the rate of work applied to the

deforming region is related to the deformation pressure,

P^, as follows:

g TCJSifa.. a (2.20)

where \ / - ram velocity

Kudo (reference 25) analytically derived an upper-

bound solution for the case of axisymmetric extrusion through

square dies (o^= 90°) by analyzing the cylindrical deforming

regions associated with the velocity field in Figure 2.9.

His results can be approximated within + 5 percent by the

following equations.

ft/s? = . 8 & +-\.-5o J U R, for smooth tools (2.21)

v . o t + V.5S J U R for rough tools (2.22)


The more complex problem (from a mathematical stand­

point) of extrusion through conical dies was treated by

Yang in 1962 (reference 26), when he attempted to extend

the plane strain concept of rigid blocks to three dimensional

extrusion geometries. With this stipulation, no energy can

be dissipated within the bounded regions of the velocity

field, and the first integral in equation (2.19) is equal to

zero. It is inconceivable, however, that a part of the metal

can exhibit a velocity component in the radial direction and

yet remain rigid. Indeed, the assumption proved to be a

poor one, since Yang's upper-bound analysis predicted lower


19

extrusion pressures than those observed by Frish and Thomsen

(reference 27) with similar process conditions.

Avitzur (references 28 and 29) proposed the radial

field illustrated in Figure 2.10 and proceeded to compute

the required extrusion pressure as a function of reduction

ratio, die angle, material flow stress, and friction condi­

tion at metal-die interface. Since he was only concerned

with a containerless operation (similar to wire drawing),

the results in the original papers are not plotted beyond

that stress required to initiate yielding in the billet

material ( P a - Y ) . The analysis-, however, can be extended

to even greater stresses if a container is- provided in order

to prevent the bulging limitation. For this condition, the

following equations by Avitzur are still valid and were used

to obtain the solutions presented in Figure 2.11.

~~ (2.23)

for perfectly smooth tools

and

^ / V \Fs (2. 24)

for shear along the die face

where is a complex trigonometric function of oC.


defined and tabulated in references 28 and 29.

In a contemporary investigation, Kobayashi (reference

30) analyzed the axisymmetric extrusion problem using the


20

velocity field illustrated in Figure 2.12, which was first —

proposed by Green (reference 31) as being applicable to

plane strain sheet drawing. Solutions were obtained for the

case of perfectly smooth tools with several die angles up to

fifty degrees, and two of these are plotted in Figure 2.11.

The extrusion pressures predicted by the upper-bound

analyses resemble those obtained from the slip-line field

solutions (Figure 2.8) in that the values do not extrapolate

to zero as the deformation decreases; however, the former

are considerably higher than either the plane strain or

homogeneous deformation predictions.

Visioplasticitv

Visioplasticity is a technique developed by Thomsen

and co-workers (reference 32) to obtain the true strain,

strain-rate, and stress distributions in axisymmetric

deformations. The method relies on an experimental deter­


mination of the strain-rate pattern from which the stress

distribution can be obtained by the solution of the equili­

brium equations using stress-strain rate relationships. The

flow pattern is calculated from the progressive movement of

machined grid lines on the meridian plane of a split billet.

It is important to realize that this approach does not yield

a predicted extrusion pressure for various process conditions

as did the previously discussed theoretical techniques; in­

deed, this parameter must be measured since it is a necessary

boundary condition in the analysis. With respect to rod


extrusion, the visioplasticity technique has, to date, been

utilized only to study the flow behavior through square

dies (references 7 and 33), although a detailed investigation

of the stress distribution resulting from extrusion through

conical dies of various angles has recently been initiated

(reference 34).
CHAPTER III

FURTHER DEVELOPMENT OF EXTRUSION THEORY

Alternate treatment of container friction

Modern friction theory (reference 35) is based on the


observation'that metal surfaces are not smooth, and When the

normal force is relatively light, the real area of contact

is much less than the apparent area due to the impingement

of microscopic asperities. The tangential friction force

to shear these junctions must therefore be expressed by

p =■ tr (3.1)

where Ar is the real area of contact and 'b is the mean shear

stress necessary to overcome plowing and rupturing of junctions.

The real area of contact, however, wi*l depend on the applied

normal load, N, and the average yield stress value of all the

junctions, O T ; so that

a
/ V
= JtL.
^ (3.2)

and by substitution into equation (3.1)

F=4js-Nj-= y u N 0.3)

where is defined as the coefficient of friction.

At the commencement of material flow beneath the asperi­

ties, it can be shown (reference 36) that GT is equal to the

normal applied pressure, p, and that the relationship between

1T and p is variable ( is not constant) over a large range

22
23

of applied pressures. Referring to Figure 3.1 (which was

extracted from the above reference), the mechanisms for

sliding friction may be divided into the three indicated

regions. In region I, the shear stress varies linearly with

the normal pressure and the coefficient of friction is a

constant. Region II is a transition range in which plastic

flow is initiated, while in region III the real area of con­

tact approaches the apparent area, the shear stress is con­

stant, andyC( decreases with increasing pressure. A recent

investigation with hot metal sliders (reference 37) experi­

mentally confirmed this general analysis since it was found

that at the higher investigated pressures the variation of

the friction coefficients with pressure could be formulated

by

(3.4)

where K was a constant representative of the shear^stress of

the interface.

In view of the above information, it is no longer rea­

sonable to assume a constant coefficient of friction in the

container liner of an extrusion press since the pressures

generated in this area are usually greater than those which

define region I of Figure 3.1. It appears more realistic

to substitute the value from equation (3.4) into

equation (2.2). This operation results in a new differential

equation

(3.5)
24
which upon integration from (x = o, Pt = P^) to (x = L, Pt -

Pt ) yields the following expression for the total force:

4 K L (3.6)

Application of work hardening theory


to the slab method of stress analysis

There were two distinctive features associated with the

solutions obtained from the slab method of stress analysis

(see Figure 2.5). First, the predicted extrusion pressures

are quite low and second, these values always extrapolated

to zero as the reduction ratio approached one, regardless of

the friction condition or the die angle. These traits can

only be due to either the hypothesis of a homogeneous state

of deformation or the inattention paid to a real material's,

propensity for work hardening. The latter assumption can be

treated if the yield locus is allowed to expand with increas­

ing strain. The stress system is still determined by equa­


tion (2.13)

JDAcr* ■vZ<nU.D-v-2.(Y-^') = O (2.13)

but now the flow stress will be expressed as

(3.7)

where = true strain

= flow stress at ■=.\

Vr\ = work hardening coefficient


25

It can easily be shown that

Y V/ <£>•&VY\
i * = Y t e. (3 .8 )

where is tbe fl°w stress at .2% strain.


By substituting this relationship in equation (3.7)/

the instantaneous flow stress can be expressed as a function

of the tensile yield strength and the true strain.

£-2.vn rr>
= Yv e £ (3.9)

The true strain in the axial direction at any diameter/ D/

is given by

€. ^ / x j ) (3.10)

Through use of equations (3.9) and (3.10), the differential

equation (2.13) can be converted to

which, in turn, reduces to

DA'S* A D = - ; T +'c e D Vr5iIi (3-12)

In order to obtain an integral form, the above expression

must be multiplied by jj> , after which the integration

proceeds from to = D c.• aPP1Y in<3 tlie boundary

conditions t h a t G x ^ o at Cb — ? and at '£b- .


26

(3.13)
IV

(3.14)

Dp
The integral in equation (3.14) was numerically evaluated

on a high speed digital computer yielding solutions for the

required extrusion pressure (see Figures 3.2 and 3.3) as a '

function of the tensile yield strength, work hardening

exponent, reduction ratio, friction condition and die angle.

Upper-bound solutions with the


triangle velocity field

The upper-bound velocity fields solved by Avitzur

(references 28 and 29) and Kobayashi (reference 30) do not

represent the only possibilities for metal flow; as a matter

of fact several alternative proposals exist. H. Kudo

(references 38 and 39) introduced the concept of triangular

deforming regions and later (references 40 and 41) proposed

the velocity field illustrated in Figure 3.4 as being appli­

cable to the conical die axisymmetric extrusion process.

This field will be analyzed in this section for load predic­

tions because the material flow as given by the field con­

figuration compares closely with observed deformation patterns.

The method of determining the deformation pressure is

characterized by the following procedure. First, the upper-

bound rate of work must be calculated from equation (2.19)


27

which requires a knowledge of the velocity of the material

at any point within the deforming regions as well as the

tangential velocity discontinuities at the border surfaces.

The next step involves the solution of equation (2.20) for

To determine the horizontal velocity component in the

outside region, , consider the displacement of region II

that is produced by the displacement (in unit time) of the

rigid zone 135. These movements are illustrated in Figure

3.5. By calculating the volumes of displaced material, it

is found that

T r C v - ^ X 0 = *-^-vr^a

or

and

(3.15)

The vertical velocity along the discontinuity surface

45 can be calculated by utilizing the fact that a velocity

discontinuity cannot exist normal to a field boundary without

violating the condition of incompressibility. With reference

to Figure 3.6, the following relationships are established:

v — (3.16)
v -v5 - Tr
28

In a similar manner, it was determined that

v35 \ L J (3.17)

To determine the vertical velocity in the outer de­

forming region, , assume the condition that this velocity

varies linearly with Z. and therefore will have the follow­


ing form,

(3-18)

where G(\r) and H(f-) are independent functions of the radius,

For the line 35,

and

V3ET (3 -19>

For the line 45,

and

(3 .20)

Solving the set of simultaneous equations generated by

these conditions, it is found that


29

and

G - O V & = & [l- b +•2-r]

or

V 'z= (3 •21>

The vertical velocity along 34 in the outer region is found

by substituting r = b into equation (3.21):

(3-22>

Since the velocity along 34 can never change its down­

ward direction, this equation places a physical restriction

on the magnitude of V* for a given reduction in area. These

values are presented in Table 3.1.

To satisfy steady state conditions, the displacement of

each streamline across region I must be equal, or mathemati­

cally

c
\ U^cit is equal for each streamline.
54,

This condition can be fulfilled by restricting in region I

to equal a constant. Also, since the normal velocity to the

34 surface in region II must be equal to the normal velocity

to this surface in region I, it follows that

= /2.^ b (3.23)
The vertical velocity in this region, V\ is determined in a

manner analogous to that employed in region II. Let this

velocity be of the following form:

V\=■Gr Jr V \ /(.V^‘Z- (3.24)

where G 1 ( y—) and H 1 ( \r~) are independent functions of the

radius.

For the line 36,

3C,= - (3.25)

V 5I2.= ^ <3*26)

For the line 46,

-V- ciC^-V) (3 -2 7 )

(3-28)

By simultaneously solving the set of equations above:

. - 6 - > 2')CP.~>ST l-V- ~7- (3.30)


V, - L t>-2- 2- b> V- Z.zr~)=>

and along 34, where


31

With the knowledge of the velocity of the material at

every point in the field, it is now possible to refer to

equation (2.19) and consider the various terms which contri­

bute to the rate of energy dissipation. For materials which


obey the Mises yield criterion and the Levy-Mises stress-

strain rate relation, the rate of energy dissipated in the


* "fcVi
V/ deforming region can be expressed as

where

After completing the substitutions defined by equation

(3.31), the rates of energy dissipated in the inner and

outer regions are found to be

(3.32)

& 5(m

and

b z-asr

. ( |- y< -z
+x U v l z a v-x~)J t-A zA i- (3.33)
32
The solutions to the above integrals can be obtained

numerically on a high speed digital computer such as the

IBM 7094 which was made available to this investigator by

the Air Force Materials Laboratory.

Consideration must now be given to the energy that is

dissipated on the boundaries of the selected regions which

are divided into straight line segments that are examined

independently in Appendix B. The integrations in the Appendix

result in the following determinations which comprise the

second term of equation (2.19).

(3.34)
J\S

(3.35)

(3.36)
■4S —1

(3.37)

- a n r (3 .3 8 )
The contributions from various regions and field

boundaries to the total rate of energy dissipation are ex­

pressed in equations (3.32 - 3.39), where it is shown that

for given process conditions, the magnitude of each term is

dependent on either a , ^ ^ or a combination of the three

variables. Since the most accurate extrusion pressure that

can be determined from the upper bound solution is the

lowest calculated pressure, must be evaluated as a

function of the three variables and a minimum value must

be sought.

A simplification in the procedure can be made if the

metal is restricted to flow over the die surface (no buildup

of metal occurs on the die). For this case, the variables

a and ^ are no longer independent but are related by the

following expression:

-4— r ,. C\ (3.40)

In essence, for this flow condition, /<{ is a function of

only two geometric variables, V and ^ , and the minimum

of the mathematical surface formed in^/ST/ V • anc^ ^


space must be investigated. Two situations were considered— -

that of perfectly smooth tools ( = o ) and that of

perfectly rough tools ( ^ '/f3 ). Using equations

(3.32 - 3.39) and equation (2.19), the contribution of each

discontinuity surface and of each deforming region to the

total required pressure is calculated for a selected reduction


34

ratio and die angle as a function of and . A sample

of this technique is illustrated in Table 3.2 where the com­

puter results for a reduction ratio of six to one and conical

die of half angle 60° are displayed. As mentioned above,

the contributory pressures due to the 15 and 45 discontinuity

surfaces are not included in the smooth tool summation but

are added to obtain the resulting Q for the case of per­

fectly rough tools.

Those values of ft* and ^ which yielded minimum predicted

pressures are presented in Table 3.3, from which the most

suitable velocity field for a given combination of extrusion

parameters may be constructed. The predicted deformation

pressures calculated from these fields are plotted in Figure

3.7 as functions of the material's flow stress, reduction

ratio, die angle and the friction condition at the metal-die

interface. The higher curve for each die angle was calculated

from that field which resulted when a shear condition existed

along the die face and the container liner (see Table 3.3),

but does not include the pressure contribution due to the

friction generated along the 15 surface, since this term has

already been treated as part of , the total friction

force between the upset billet and the container liner.

The predicted pressure-flow stress ratios are shown to

vary almost linearly with the natural logarithm of the reduc­

tion ratio at higher deformations but do not remain linear

nor extrapolate to zero as the deformation decreases. When


35

these solutions are compared with those previously illus­

trated in Figure 2.11, the following similitudes and con­

trasts are noted:

1) As with Kobayashi's solutions, the linear dependence

of pressure with In R does not apply at small extrusion ra­

tios.

2) A maximum variation in the die friction condition

did not alter the value of tsV'Y' at zero deformation for

these solutions and those proposed by Avitzur.

3) In general, the solutions calculated from Kudo's

triangle velocity field are.very close to those calculated

by Avitzur for die half angles of 45 and 60 degrees, but

this theory predicts higher pressures for a die half angle

of 30° than those proposed by either Avitzur or Kobayashi.


CHAPTER IV

EXPERIMENTAL PROCEDURE

Description of the extrusion equipment

The main objectives of the experimental portion of

this dissertation were to first develop techniques for

measuring the forces involved in the direct extrusion

operation and to then determine the relationships between

these forces and several process parameters. This investi­

gation utilized the extrusion press shown in Figures 1.1

and 1.2, which possesses the following characteristics:

1. Water-so liable oil hydraulic system with a


3000 psi accumulator

2. Ram speed, 20 to 900 inches per minute

3. Ram stroke, 30 inches

4. Rated capacity, 600 tons

5. Peak capacity, 700 tons

6. Container liner I.D., 3.072 inches

7. Container temperature, 800°F

It is important to note that the stem speed cannot be

preselected with a great degree of accuracy nor can it be

controlled so as to be constant during the entire extrusion

operation. The rate of fluid flow between the accumulator

and the ram is not only a function of the orifice size, which

can be preset by a manually controlled valve, but also de­

pends on the total extrusion pressure and the remaining

accumulator pressure, both of which vary during the process.

36
37

Although the entire extrusion procedure was oftentimes

completed within one second, the following parameters were

continuously recorded on a direct reading oscillograph

which was selected for its very rapid (30 millisecond)

response time.

1. Force on the stem

2. Force on the columns

3. Force on the hydraulic system

4. Force on the die

5. Position of the stem

6. Velocity of the'stem
i

Determination of the total applied force

It was suspected that the total force, Ft/ as measured

by a pressure transducer, was not indicative of the true

force delivered by the stem to the system. These suspicions

developed when a surge of approximately 100 tons was noted

in the recorded values of this parameter during several

extrusion operations. Since there was no rational explana­

tion for this behavior either in terms of material deformation

or in terms of lubricity at any interface, it was concluded

that the recorded surges were characteristic of the dynamic

hydraulic system and were not manifestations of discontinui­

ties in either F^, F ^ , or F ff.

Two systems for recording the actual total transmitted

load were then developed to circumvent this difficulty.

Strain gages positioned directly on the stem constituted one


38

method while strain gages applied to the tie rods (the sup­

porting columns for the press) served as the alternate

technique. The press was brought to full capacity (a condi­

tion in extrusion termed "bottom out") and both sets of

gages were set to coincide with the static force as deter­

mined by the area of the hydraulic pistons and a calibrated

pressure gage. The response of the stem.was found to be

linear from zero to 665 tons (bottom out) at the aforemen­

tioned setting but a calibration curve had to be constructed

for the column gages. This situation was traceable to the

very small deflections in the massive support column which

dictated that a large voltage (10.7 volts) be set across

the gages' electrical bridge in order that sizable outputs

could be obtained. This large a voltage understandably pro­

duced thermal transients which in turn caused the observed

nonlinearity of the system.

Determination of the deformation force

In order to measure the force transmitted to the die,

F^, it was necessary to modify the existing extrusion tooling.

Referring to Figures 1.2 and 4.1, the inside diameter of the

die backer was increased so as to accommodate a hollow H12

tool steel cylinder. Figure 4.2 is a drawing of the final

die backer configuration while Figure 4.3 indicates the

dimensions of the inserted cylinder. To allow clearance

for a variety of extruded products, several shells were

"manufactured having various inside diameters. Near one end


39

of the cylinder four strain gages (120 ohm, 3/8" foil type)

were positioned 90° apart alternately measuring tension and

compression on the outer surface. The slot illustrated in


»

Figure 4.2 was milled to allow an exit for the electrical

leads attached to the strain gages, while the dimensions and

tolerances for the inside diameter of the die backer and the

outside diameter of the cylinder were so designed (.003 in.

clearance) as to assure an unrestricted deformation of the

load cell's lateral surface without so loose a fit that the

cylinder could not align itself properly with respect to the

container liner and the die.

Originally it was thought that the load cells could be

calibrated in a 500 ton vertical forge press where a solid,

similarly gaged, but previously calibrated H12 cylinder was

being utilized to record upset forging loads. Experimentally,

however, this procedure proved unsatisfactory for when plac­

ing the load cell to be calibrated on top of the standardized

cylinder, the imposed end conditions on the load cell were

quite different than those encountered by its placement in

the extrusion press. In the former case one end of the load

cell would be in complete contact with the standard cell

while in the latter situation, as can be seen from Figure 4.1,

the end against the bolster is not completely covered. This

difference in end boundary conditions produced a pronounced

effect on the response of the gages, resulting in differences

of as much as 100 tons at the 500 ton level.


For this reason it v/as determined that the most accurate

method of load cell calibration would be by means of a series

of "in situ" measurements. An undersized (diameter 1/8"

smaller than the container liner) H12 tool steel block was

placed into the container liner with one end against the

stem and the other in direct contact with the load cell.

A force was increasingly applied (up to 500 tons) to the

stem and since the block did not upset to the container

liner, the known force on the stem was fully transmitted to

the load cell producing the desired calibration. Using

this technique, each load cell was calibrated in place

with the exact boundary conditions experienced during the

extrusion operation.

Yet another factor had to be reckoned v/ith before the

measurement system could be pronounced operational. From

previous observations it v/as known that as a billet extruded,

the pressure upset the die causing intimate contact between

the outside surface of the die and the inside surface of

the container liner. This was not an accidental arrangement

but rather a purposely designed method for increasing die

performance. The dies at the Air Force Materials Laboratory

are machined so that on reaching the container temperature

(800 F), the dies are expanded to the inside diameter of the

hot container liner. When the upsetting occurs during ex­

trusion, the dies are now laterally supported by the rigid

tool steel liner. Since support is provided by this means.


the yield strength of the die material need not be the sole

design factor, and therefore the tempering temperature of

the tool steel can be so selected as to provide a balance

between yield strength and toughness. It is for this reason

that the H12 dies are heat treated to a moderate (for this

material) Rockwell C40-44 rather than to the higher hard­

nesses used by other investigators. The success of this

technique is well established in terms of increased d i e '

life and freedom from die cracking.

However excellent the previously discussed procedure

was in terms of tooling utilization, it resulted in a fric­

tion force between the die and the liner which did not allow

a complete transmittal to the load cell of the axial force

placed on the die. In essence, equation (1.1) must be re­

written in terms of the measured values as follows:

F t = F ^ F j b + Ft vFLC (4.i)
where F,- is the friction force at the die-container liner
r
interface and F_„ is the force determined by the load cell

deflection.

To ascertain the magnitude of F so as to add its value

to the recorded load cell readings and obtain F , a series


d
of mild steel (1018) plugs were machined to the contour of

each extrusion die. Each die and its plug was then placed

in the container and allowed to heat for ten minutes as is

the customary pre-extrusion procedure. A five hundred ton


load was slowly applied to the plug which upset in the die

and transmitted the load to the die surface as in the extru­

sion operation. The force on the stem and the force on the

load cell were continuously recorded and their difference

was, of course/ the frictional force between the die and the

container liner. Conical dies having included angles of

both 90° and 120° were tested in this manner with a lubricant

of Fiske 604 applied to the outside surface in an effort to

minimize this frictional effect. Fiske 604 is a commercially

available lubricant of graphite suspended 'in a calcium base

grease. In general, the force value increased with increas­

ing applied load and decreasing included angle. The measured

magnitude of this force was generally less than twenty tons,

so that its addition to the load cell reading was more in the

nature of a refinement than as a major consideration.

Of prime concern in the early experimental stages was

the high temperature experienced by the load cells due to

their proximity to the 800°F container. This problem was

eliminated with the inauguration of the following procedures:

1. Provisions were made in the tooling to provide

for continuous forced air cooling of the load cell.

2. The container was not sealed against the sup­

port tooling until the last possible moment.

3. Whenever possible a reflective heat shield was

positioned between the container and the support tooling.


43

Temperature determination by in­


frared radiation measurements

Several investigators (reference 42 - 4-6) have shown

that the transient heat flow which occurs during the hot.

extrusion process may lead to either substantial increases

or decreases in the temperature of the deforming billet

depending on the material's physical properties, the degree

of deformation, the nature of the boundary surfaces and the

time required to complete the operation. Since R.M. Treco

(reference 47) was able to explain many anomalies in his

force measurements by means of these temperature deviations,

it was thought that at least some temperature determinations

should be included in the experimental procedure.

An infrared radiation detector, capable of achieving

95 percent of full scale deflection in thirty milliseconds,

was connected to the oscillograph in an attempt to measure

the surface temperature of the extruded bar along its length

as it emerged from the die exit. A 3/4" diameter hole was

drilled through the top of the bolster tooling (see Figure

4.1), and a specially prepared copper tube was inserted onto

a shoulder as illustrated in Figure 4.4. During each experi­

ment, argon gas was introduced at the top of the tube at the

rate of 15 cfh in order to purge the system of vapors or dust

which could absorb some of the infrared radiation. A quartz

lens (G.E. type 101) was used to seal the top of the tube

and a reflecting mirror was mounted above the lens to direct

the radiation to the detector. In a series of controlled


44

experiments, the lens was tested for its transparency and

the mirror for its reflectivity with respect to the energy

accepted by the detector. The lens was determined to be

94 percent transparent, while the mirror reflected 93 percent

of the considered radiation. These factors were then utilized

to correct the recorded signal.

Fifteen oxygen-free-high-conductivity (OFHC) copper

billets were extruded in the temperature range of 800-l400°F

as well as twelve mild steel and ten molybdenum alloy billets

in the temperature ranges of 1600-1950°F and 2600-3800°F, re­

spectively. In each case, the recorded temperature measure­

ments were so sporadic that the technique had to be abandoned.

There are generally two sources of error encountered in

radiation or optical pyrometry. One stems from radiation

absorbing media located between the target and the detector,

while the other arises from a nonuniform or uncalibrated

surface condition. Since careful steps had been taken to

either eliminate or correct for absorbing media, the failure

of this method is probably due to the nonuniformity of the

product surfaces.

Temperature determination bv
calorimetric measurements

In hopes of at least ascertaining the average product

temperatures, a rather inelegant calorimeter (see Figure 4.5)

was constructed to receive the extruded bars. The rectangular


box, 36" long x 4" wide x 5" deep, was fashioned from 1/8"

type 304 austenitic stainless steel and then surrounded with

2" thick cork insulation in order to minimize the heat losses

A hole was drilled in the cover, which was similarly fabri­

cated, to allow for the placement of a mercury thermometer.

In the many tests designed to calibrate the equipment as well

as during the actual experimentation, the trough was careful­

ly filled with 7500 c.c. of water at least ninety minutes

prior to placing the bar within the calorimeter. The enthal­

py balance for the system and the resulting expression for

the temperature of the extruded bars are presented in Appen­

dix C. Three 1018 steel rods and three OFHC copper rods were

heated in an electric furnace to 1800°F and 800°F, respec­

tively, in order to ascertain the accuracy and reproducibil­

ity of this technique. So much steam evolved when the steel

bars were tested that the method was abandoned for use at

this high a -temperature; however, for the copper bars, the

calorimetric measurements predicted temperatures to within

+ 8°F of the actual value. Unfortunately, since variable de­

lays were experienced in removing the extruded copper from

the runout tube during the actual experimentation, the real

accuracy must have been less than that indicated above.

Summary of experiments

Over seventy extrusions of several different alloys were

required before the force measurement system was deemed


operational. Controlled extrusion experiments were then

performed with hot-rolled OFHC copper at 800°F and 1018

mild steel at 1800°F under various process conditions.

These materials, the compositions of which are given in

Table 4.1, were selected for two main reasons. First, infor­

mation was available that described their flow stresses as

functions of strain, strain rate, and temperature in the

regime of interest (reference 48); and second, they repre­

sented two metal systems that are commonly extruded on an

industrial scale. The copper extrusion temperature was

selected so as to minimize any heat flow between the billet

and the hot (800°F) container liner, while the steel was

processed at 1800°F in order to simulate the commercial

practice of hot working in the single phase austenite region.

All the billets were machined in accordance with the

drawing shown in Figure 4.6 and were heated for one hour in

an electric furnace set at the desired extrusion temperature.

The copper billets were heated in air, but an argon atmosphere

was provided when heating the steel specimens in order to

prevent the formation of the iron oxides. This latter pro­

tection was necessary because the oxide growth in air at

1800°F is so great that a billet cannot be placed in the

container liner until these oxides are removed by the standard

technique of striking the billet a number of times with a

steel bar. This "knocking off" operation is time consuming

and allows the billet to cool appreciably.


47
The reduction ratio was, of course, an important para­

meter throughout the investigation and, as such, was varied

from two to ten. In order to test the sensitivity of the

force measurements to various friction conditions, billets

of different lengths were extruded in both the lubricated

and unlubricated conditions. To lubricate the copper ex­

trusions, Fiske 604, a commercially available suspension of

graphite in a calcium base grease, was swabbed onto both the

container liner and the die face. The lubricated steel bil­

lets were coated with Corning glass Nr. 00101 in accordance

with the procedure outlined in Table 4.2 prior to being

placed in the furnace.

Originally, conical dies with three different entrance

angles, namely, oC. = 30°, 45°, and 60° were to be used in the

investigation; however, all the 30° dies cracked during the

calibration procedure. These H12 dies were found to contain

.140% sulfur, and subsequent metallography revealed the

existence of large sulphide particles throughout the material.

Extrusion dies are normally machined in the annealed condi­

tion and then heat treated to the desired hardness; but the

supplier, when contacted, indicated that he had machined

these particular dies in the hardened condition because the

alloy had displayed such excellent machinability. The three­

fold combination of unrelieved stresses at the root of the

^-Corning glass Nr. 0010 is of the potash-soda-lead type


with an approximate composition of 63%Si02, 7 .6%Na2 0 , 6%I<2 0 ,
0.3%CaO, 3.6%MgO, 21%PbO, and 1%A1203.
48

machining grooves, the very low toughness exhibited by this

microstructure, and the demand for deformability resulted in

catastrophic tool failure at applied loads of less than fifty

tons. All the billets were therefore extruded through coni­

cal dies with half-angles of either 45 or 60 degrees.

The variations in the process parameters are summarized

in Tables 4.3 and 4.4 for the extrusion of 1018 steel at

1800°F and in Tables 4.5 and 4.6 for the extrusion of OFHC

copper at 800°F. The other values included in these tables

were derived from the experimental results and will be dis­

cussed in the next section.


Since the ram velocity was not constant during most of

the extrusions, an acceleration force was present; however,

the analysis in Appendix D clearly indicates that its mag­

nitude was inappreciable and could not alter the force balance

expressed as equation (1.1).


CHAPTER V

‘ RESULTS

Description of the pressure- ...


distance curves

The recorded values of the total pressure (Pt) and the

deformation pressure (P^) are shown as functions of the ram

travel for two copper extrusions in Figure 5.1 and for two

steel extrusions in Figure 5.2. The zero reading on the

abscissa represents that point at which extrusion commenced

(i.e., after the billet had been completely upset against

the container liner and into the die) and is characterized

by a maximum in the total pressure curve. The last point

in each curve signifies the onset of "suck in#" a condition

created when the graphite follower material entered the de­

formation zone to replace the faster moving material along

the centerline of the extrusion. Only between these points

is it safe to assume that the recorded ram travel corresponded

to the homogeneous movement of the upset billet in the con­

tainer liner.
The shape of each of the total pressure curves shown in

Figures 5.1 and 5.2 is representative of all the extrusions

performed with the identical lubrication condition. For

example, the total pressure curves for the entire series of

glass lubricated steel billets followed the trend illustrated

by extrusion number 1796 in Figure 5.2, in that Pt decreased

at first and then rose once again even though the billet was
continually being ejected from the container liner.
49
The deformation pressure curves were similar for all

but two of the extrusions (Nrs. 1706 and 1708) in that the

P^ values were initially at a maximum and then continually

decreased until about 1.0 and 1.5 inches of billet had been

extruded, after which the pressure did not vary by more than

2.7 ksi (a force of 10 tons). The larger value noted at

the onset of each extrusion is attributed to the difference

between the static and dynamic coefficients of friction at

the metal-die interface. The constant pressure value, which

was eventually attained in each extrusion except 1706 and

1708, was taken to be the steady state deformation pressure,

Pds / under the indicated process conditions.

In order to facilitate correlations between extrusion

pressure and process variables, the initial and final readings

from both the total pressure and the deformation pressure

curves as well as the value of the steady state deformation

pressure are presented for each extrusion in Tables 4.3 - 4.6.

Temperature measurements

Since approximately ninety percent of the work required

to plastically deform a metal is converted into heat (reference

49), no interpretation of the measured extrusion temperatures,

as shown in Tables 4.5 - 4.6, would be valid without a consid­

eration of the input energies. A program was therefore pre­

pared for the IBM 7094 digital computer which integrated all

the total force (F^) — distance and deformation force (F^) —

distance curves. To account for the differences in the


51

billet lengths, the results were normalized with respect to

the extruded volume and are presented in Tables 4.3 - 4.6

as Et/V, the total energy expended per unit volume, and E^/V,

the energy per unit volume required to perform the deformation

(includes the energy dissipated as friction on the die face).

Evaluation of the deformation


pressure-flow stress ratio

In order to compare the predicted steady state deforma­

tion pressure v/ith the measured values, it was first necessary

to determine the average flow stress of each material ('y' )

under the experimental conditions, in as much as the results

of all the deformation theories discussed in Chapters II and

III were obtained as functions of this parameter. Although

a virtual dearth of information exists in the properties of

materials at the temperatures, strains, and strain rates

germane to the common deformation processes, Alder and

Phillips (reference 48) had conducted laboratory compression

tests with a phosphorus deoxidized copper and a .17% carbon

mild steel to 50 percent axial reduction at various elevated

temperatures and rapid strain rates. Since these materials

were expected to behave quite similarly to those used in the

present investigation, the pressures required to upset the

cylindrical specimens to 50 percent engineering strain were

taken as the average flow stresses. By graphically inter­

polating between the test temperatures selected by these

authors, it was possible to obtain the flow stresses of copper


at 800°F and of mild steel at 1800°F as functions of the

process strain rate (see Figures 5.3 and 5.4). In actuality,

it should be noted that these flow stress measurements are

probably higher than the theoretical tensile test values

because of the friction forces which v/ere generated on the

ends of the compression rods, even though a lubricant had

been applied in order to minimize this effect.

Before the information contained in Figures 5.3 and 5.4

could be utilized, however, it was necessary to first derive

an expression for the average strain rate experienced by the

material in the extrusion operation as a function of several

process parameters. The analysis, which is presented in

Appendix E, yielded the following relationship;

(5.1)

where V' is the ram velocity which prevailed during the con-

stant portion of the deformation pressure-ram travel curve

(see Figures 5.1 and 5.2). This velocity along with the aver­

age strain rate as calculated by equation (5.1) and the aver­

age flow stress as determined from Figures 5.3 and 5.4 are

presented in Tables 4.3 - 4.6 for each of the experimental

extrusions. The last column in each of these tables contains

the resulting deformation pressure-flow stress ratio.


CHAPTER VI

DISCUSSION

Effect of billet length

The analyses in Chapters II and III indicated that

whenever a friction condition exists along the container

liner, more force must be expended to extrude longer billets

than shorter ones due to the additional lateral area avail­

able for generating increased frictional resistance. This

fact is clearly indicated in Tables 6.1 and 6.2 where that

percentage of the maximum applied pressure which was dissi­

pated in overcoming the container friction, C f r r & L x ,0 0 ,

is presented in the last column. When the lengths of the

steel billets were enlarged from 3 inches to 4.5 inches,

this average "container friction percentage" increased from

30% to 46% for the unlubricated specimens and from 16% to

21% for those billets which had received the glass coating.

Similar data was not obtained from the copper experimenta­

tion because so large a container friction force was present

during the unlubricated extrusions that it was impossible to

remain within the capabilities of the press and process bil­

lets over three inches long.


Variations in original billet length not only altered

the total pressures, Pt , required for the extrusion of both

materials, but also influenced the deformation pressures, P^,

generated by the copper specimens as can be seen in Tables

53
4.5 and 4.6. In all instances shorter copper billets ex-

hibited lower deformation pressures than longer billets

extruded under almost identical process conditions. Since

no such trend was observed with either the lubricated or


unlubricated steel billets, it was felt that this phenomenon

must be generally related to the dissimilarities which exist

in the thermal-physical characteristics of the two materials

and in particular to the differing rates at which heat can

be removed from the deformation zone.

As the extrusion operation proceeded, the heat generated

in this region was probably not extracted to the rear by the

cooler mass of undeformed steel because of its relatively

poor thermal diffusivity value of .06 crrr/sec (reference 50) .

However, since the thermal diffusivity of OFHC copper is .95

cm^/sec (reference 51), some of the heat of deformation must

have flowed into the undeformed billet and thence to the con­

tainer liner; although, admittedly, a thermal gradient did

not initially exist across the latter transfereeite. The

larger the billet, the more efficient would be the heat ex­

traction process due to both the increased mass at the lower

temperature and the expanded contact area at billet-container

liner interface. The temperature measurements presented in

Figure 6.1 support this.hypothesis since the longer copper

billets are shown to have had a lower exit temperature than

the shorter billets for the same value of_the input energy

per unit volume. Such a situation could only have occurred


55

if the heat losses were indeed amplified as the billet

lengths were increased.

Effect of lubrication
agfeata o£ iufes?A<3afcien on fcha sefeal pi?ea<sua*o t*ac­
quired to initiate the extrusion process as well as on the

percentage of this pressure dissipated by the container

friction forces (Fff + Ff^) are presented in Tables 6.1 .-

6.3. The application of 0010 glass reduced the average

"container friction percentage" from 46% to 21% for the

4-1/2 inch long steel billets and from 30% to 16% for the

3 inch long billets. The fact that almost one-half of the

delivered energy was dissipated as container friction during

the extrusion of the longer, unlubricated steel billets is

appalling, but the use of a molten glass lubricant was shown

to be an effective method for providing significant increases

in press capabilities.

An analogous situation existed with regard to the copper

extrusions as can be seen in Table 6.3, where the average

"container friction percentage" is shown to have been reduced

from a value of 42% to 26% by swabbing Fiske 604 along the

inside of the container liner. Since the friction force at

the billet-container interface (F^) has already been shown

to increase with billet length, the lessening of container

friction, just mentioned, through the use of Fiske 604, would

appear even more prominent were billets of equal lengths com­

pared.
Improved lubricity along the container liner is also

influential in creating better metallurgical products by

promoting the more uniform deformation characteristics illus­

trated in Figure 6.2. Both steel billets were split at the


meridian plane and grooves 1/16" deep were machined in to

provide a 1/2" square grid. After the newly exposed faces

had been painted with MgO to prevent sticking, the two sec­

tions were then pinned together and each billet was extruded

at 1800°F — the one on the left with the 0010 glass as a

lubricant and the one on the right in„the unlubricated condi­

tion. The lubricated billet remained rigid up to the proper

deformation zone while the unlubricated billet underwent

severe distortion well in front of the die entrance. The

deformation in the latter case was not even symmetric, since

it is obvious that the left side of extrusion number 1806

experienced a worse friction condition than the right side.

Two additional gridded billets, processed similarly to number

1806, established beyond doubt the fact that this severely

distorted side corresponded to the top of the billet as it

was positioned in the container liner. The better lubricity

noted at the bottom of the billet can only, be attributed to

the lubricating characteristics of residual glasses which

had remained in the container liner after previous extrusions.

If the coefficients of friction in the container liner

was constant during the entire operation, then equation (2.4)

would properly describe the force relationship and a plot of


57

In (F-fc/F^) versus L for each extrusion should result in a

straight line having a slope of 4 • If# on the other

hand, the shear stress at the interface was constant, then

equation (3.6) would be more appropriate and (Ft - Fd) should

be a linear function of L with a slope equal to 4 K

Neither of the above relationships was obtained with

the data from the steel extrusions, as could have been pre­

dicted merely from the representative pressure-distance'

curves shown in Figure 5.2. Instead of the total pressure

continuously dropping as the process proceeded, hardly any

change occurred over a significant distance with the unlubri­

cated billets, and the total pressure actually increased as

the glass-coated billets were extruded.

Regardless of whether one wishes to think in terms of a

coefficient of friction or in terms of a shear stress, this

data can be explained only by the fact that the lubricity at

the subject interface must have worsened as the operation

continued. This would mean that y(/(c or increased in such

a manner as to negate the effect of the decreasing billet-

container liner area. The formation of surface oxides may

have been responsible for this phenomenon in the case of the

bare billets, while a dynamic hardening of the glass could

cause such a situation with the lubricated billets.

Quite the contrary circumstance was discovered with the

copper extrusions; that is to say, for most of the billets

both In (F^/F^) and (Ft - F^) were found to be linear functions


58
of L, as is shown in Figure 5.3 for extrusion number 1720.

The lines were purposely drawn through the data points taken

toward the completion of the extrusion in order to avoid the

static friction condition which prevailed at the onset of the

operation. The calculated values of /{/{<- from equation (2.4)

and from equation (3.5) are presented in Table 6.4 along

with values of the copper shear stress as determined by the

Von Mises criterion ('fc = . 5 ' 7 7 X ). __

Both parameters, ^and ^ , reflect the change in

lubricity along the container liner due to the use of Fiske

604. Since the copper in the unlubricated condition was ob­

served to shear in the container liner, it is significant

to note the close agreement between the average K value for

the unlubricated billets (17.5 ksi) and the average shear

stress of copper (17.6 ksi). The IK parameter, as opposed

to the coefficient of friction, has the added advantage,

therefore, of being able to define the physical nature of the

interface. This argument is not limited to just the extrusion

operation but is valid for any deformation process in which

the coefficient of friction cannot be treated as a constant

because the pressures involved are in region III of Figure 3.1.

The use of Fiske 604 with copper and.0010 glass on steel

not only reduced container friction but also lowered the

measured deformation pressures, as can be seen from Figures

6.4 and 6.5. A larger than indicated difference actually

exists between the lubricated and unlubricated curves for the


59

OFHC copper. Were the unlubricated billets as long as the

lubricated ones, their deformation pressures would have been

even greater than the illustrated values because of the

billet length effect discussed earlier. Nevertheless, in

terms of increasing press capabilities, these lubricants were

obviously not as effective in the die region as in the con­

tainer liner.

Effect of die angle

By comparing the individual curves in Figure 6.4 with

their counterparts in Figure 6.5, little practical difference

in the deformation pressures is noted between conical dies

having either a 45 or 60 degree half angle. When the strain

rate of each operation was taken into account, however, and

the deformation pressures were normalized with respect to the

material's flow stress, a definite pattern (as indicated in

Figures 6.6 and 6.7) was seen to immerge. Regardless of the

particular material or the employed lubrication practice,

lower normalized deformation pressures were obtained in the

smaller reduction ratio range with a 45 degree die, whereas

a 60 degree die was better for larger deformations. This

observation will be treated in detail in the following section.

Comparison of results with defor­


mation theories

The fact that none of the lines in Figures 6.6 and 6.7

value of zero as the reduction ratio

approaches one discredits those analyses which rely on either


60

a homogeneous dissipation of energy or a uniform state of

stress at all points of a plane normal to the extrusion axis.

Siebel's energy balance/ the slab method proposed by Sachs,

and the modified slab analysis presented in this dissertation

all fall within this category.

The intersections of the o<. = 45° lines and the oC =

60° lines in the aforementioned figures are easily explained

by considering either Avitzur's upper-bound solution (Figure ■

2.11) or the upper-bound solution derived in this dissertation

from the triangle velocity field, since the two results are

quite similar for dies having half angles of 45 and 60 degrees.

With perfect lubrication, both analyses predict that no such

intersection will occur within the reduction ratio range of

interest, but an intersection is forecast for the case of

metal shear along the die face. This is explained by the

fact that as the friction force at the metal-die interface

increases, the smaller contact area provided by the 60 degree

die becomes a predominant factor in minimizing the larger

deformation pressures required at the higher reduction ratios.

All the extruded billets experienced a sufficient fric­

tion condition at the metal-die interface to cause the resul­

tant intersection of the fas lines for the 45 degree and

the 60 degree dies. The fact that the lubrication practices

were not very effective in reducing this friction has already

been discussed, and so it is no surprise that this phenomenon

was observed in the lubricated as well as in the unlubricated


61

experiments. The slight improvement in die friction is

evident, however, when the reduction ratio at which the

intersection occurred is considered. The lubricated billets

exhibited intersection at a higher reduction ratio than the

unlubricated one, indicating that the effect of decreasing

interface area was not as prominent a factor in the former

case as in the latter.

In general, the normalized pressures calculated from

the copper extrusions fall within the range predicted by

both upper-bound analyses, but the values for the steel

billets tend to be slightly lower than theoretically antici­

pated. It could be argued, of course, that the actual nor­

malized deformation pressures should be somewhat lower than

the upper-bound solutions because of the basic theoretical

inequality expressed as equation (2.19); however, an accurate

solution will minimize this inequality to the point where it

may be considered negligible. The low values are most likely

due to an erroneous selection of the flow stress ( Y ) for

1018 steel under these conditions. As previously mentioned,

little heat was conducted from the deformation zone in these

extrusions and the corresponding temperature increase would

result in an actual flow stress lower than that selected from

the literature for steel at 1800°F. Alder and Phillips

(reference 48) noticed this effect in their forging experiments

and recent work by Cook (reference 52) confirms their obser­

vations.
CHAPTER. VII

SUMMARY AND CONCLUSIONS

1. Numerous billets of 1018 steel at 1800°F and OFHC

copper at 800°F were direct extruded on a 700 ton horizontal

press in order to determine the effects of such process

variables as reduction ratio, billet length, lubrication

condition, and die angle on the forces involved in the

operation.

2. A unique measurement system was employed to quanti­

tatively determine the deformation force at the die as well

as the total required extrusion force on the stem.

3. As the billet length increased so did the total

extrusion force, but the effect was less significant with

the lubricated billets.

4. With the copper extrusions, increased billet lengths

led to higher deformation forces at the die, but no such ef­

fect was observed in the steel experimentation. This obser­

vation was explained on the basis of heat flow considerations

and the difference in the thermal diffusivities of the two

materials.

5. The container friction force was greatly reduced,

thereby extending the capabilities of the press, whenever a

steel billet was coated with Corning 0010 glass or Fiske 604

was swabbed in the container liner prior to a copper extrusion.

These same lubrication practices, however, had a much more

limited effect in lowering the deformation force at the die.


62
63

6. The container friction forces for the copper extru­

sions were found to agree with a theory which was developed

on the basis of a constant interface shear stress rather

than the usually assumed constant coefficient of friction,

but no expression described the frictional behavior of the

steel billets since the considered lubricity appeared to

undergo dynamic changes as the operation.proceeded.

7. The deformation' forces and the deformation pres'sure-

flow stress ratios increased linearly with the reduction ratio

but did not extrapolate to zero as the reduction ratio ap­

proached one. These values were found to agree reasonably

well with two upper-bound solutions, one suggested by

Avitzur from a consideration of radial velocity fields and

one developed in this paper from an analysis of Kudo's pro­

posed triangular velocity fields.


CHAPTER VIII

APPENDIX A

Analysis of slab solution as B


approaches zero

Pd = (1 + B) (RB-1) (A.l)
Y B

L'Hopital's rule states that if f(a) = g(a) = 0 and if

the limit of the ratio f'(t)/g'(t) as t approaches a exists,

then

lim f (t) =: lim f* (t) (A.2)


t— > a g(t) t — >a g 1 (t)

Let f(B) = RB-1 and g(B) = B in equation (A.l), so that


1+B

f (o) = g(o) = 0. Now

f ' (B) = RB In R

and

g* (B) = 1
(1+B)

Using L'Hopital's rule, equation (A.2), it can now be

shown that

£d = RB In R (1+B)2 = In R (A.3)
Y lim B— > 0

64
APPENDIX B

Determination of energy dissi­


pation on boundary surfaces

a) L 15

-.-■aa-sO
OS.I * J

(B.l)
Y'Z. = Z r r ^ t

b)

_^*i>— vCi^ ‘”(m.C£>sS -v-VSk'vrsS^

S\r\Ss=.^-^- > C.o'a'Sb—


-35 *5

,c - £rrM-35Li>-

& = -'5 z 2 £ '('3 * ^ L . ( \- ’t>'T■*-V 7a ^ \ (B.2)


■Vs^

65
66

c) L45 S

V'Sm fv^> — u^c_osS

>t„ S « a £ l = 2 a , c.os.S»4=i&
‘--vs L-“VS

- £ V i ^ j L a G + Y - Y i - Y } + ■ f r - S T O ^

ciS =
C\-

(B.3)
lr=Vs

d ) l36

3
n>

f
CTi1
lA V
(
#*P
/
*>• 5‘
oo
uT*
r>
5
jj
(T*
II II it
oo II u r~Q3D £
II < Ul j
a Go
i
r ir e
lA
Cn'
<
GO tA
** + 0
0>
trE? 0 2

p £ "03
CP GO
LA GO 1
0
tA 3 (T>
r II
& P
GO 04
i i ,rtr
i/ 0
II
I1 I £ \ If
oo
r\
5!^
7 n
f3 o
(A
Qo

toJ

C\
68

• „ I
-W. ^ z~v- J *.a I

^TTV~U-4eiv-
=ss - b

{^ A S > = - - ^ k n . \ c w ^ ) _ art-PY^-ft^-t-P-I?
z 1 a i-z. (B.5)
i--o u

f) L34

\ A S- ( v34 - V 3J A s
3+ .r„ 1 2. '
3^V

d S “ £.tt b d x

^ -i>d -v- Y d -v-y £ - v - p - p ^ }


z-=-a (B.6)
APPENDIX C

Analysis of calorimetric
measurements

Consider the system to be composed of the water, the

extruded bar, and any graphite from the follower block which

found its way into the trough. The heat contained in this

system immediately after the extruded bar and the graphite

enter is equal to the heat measured at a later time when

the temperature of the system is uniform plus any heat losses

which occurred during the considered intervals. For a refer­

ence temperature of 298.2°K, this argument is expressed

mathematically as

Wall ^ Cp

where W B and Wg are the weights of the bar and the


graphite (moles)

Ww = weight of water (grams)

Tj = initial temperature of the bar when it enters


the water, the extruded bar temperature ( K)

Tt = highest transformation temperature below the


suspected value of T-j- (°K)

T0 = initial temperature of the water (°K)

Tg = initial temperature of the graphite (°K) (the


graphite is heated to 800°F prior to extrusion
and since the tooling is at 800°F, this is the
value assumed for Tg)

69
70

Tf = temperature of the system after


one hour (°K)

Kh = heat loss coefficient (cal/°K)


for one hour (evaluated from cal­
ibration experiments)
B B
(Hr^t ~ ^298.2^ = heat content (cal/mole) of the
bar above the reference temperature
g
Cp = the heat capacity (cal/mole-°K)
BI
Cp = the heat capacity (cal/mole-°K)
of the bar in the temperature
range T-j- to Tj
Bf
Cp = the heat capacity (cal/mole-°K)
of the bar in the temperature
range 298.2°K to T f

By performing the integrations, combining terms and using

the relationship,

C1 = a . + b . T + c . T“2 (C.2)
P i x i
an expression for the initial bar temperature as function of

known thermodynamic parameters (reference 53) and measured

values results.

V a m CTx -T^*- - 't"Vc6x( ^ - 4^)

-a e*<t-aihs.£>- c BVC=^;-^8.^=jc\My,^K h Ktv-


t3

-4 ^ (c.3)

The value of T-j- in equation (C.3) v/as determined by means

of a reiterative process on an IBM 7094 computer for each of

the calorimetry experiments.


APPENDIX D

Force due to acceleration

From the definition of a force

di= (D.l)

But due to constancy of volume and mass

V p (D.2)
m m

i V E = (Kc Vilk (D.3)


c^- V^pr/\ d b 4.

and

am*
(D.4)
dt

dbr1 (3D.5)
d-fc’

71
72

Upon substituting equations (D.2, D.3, D.4, D.5) into equation

(D.l), the following expression for the differential of the

acceleration force results;

< » - 6>

Since the acceleration is approximately constant, the last

term in equation (D.6) is zero and upon integration;

The value of FA as given by equation.(D.7) is less than

one pound of force for the extrusions under consideration and

is therefore deemed negligible.


APPENDIX E

Determination of the average


strain rate

The true axial strain to which a material is subjected

while passing through the die is


.2- \
£= J l w , ( _ /'tf) (B.l)

The velocity of a homogeneous slice of material in the die is

given by

(E.2)
a.-t \ v
*

where is the velocity of the ram.

From geometric considerations

X L - a x_ (E.3)
c

and

(E.4)

73
Substituting equation (E.4) into equation (E.2)

<E-5)

and

J* t-

-\ VA-t r. (e .6)

"t-O

The result of this integration is

-V3

C%T=^=o-^?) (e-7)

Substituting equation (E.7) into equation (E.l) and

differentiating the resulting expression with respect to time,

one finds the instantaneous strain rate expressed as

£ - xc ^ (E-8)

The mean strain rate for a given reduction is calculated

as follows: ^

- ^ gj
r — \ _ 4- V CR “ V)
'-'5-7T r - (E.9)
75

but since

~z.± r ^ o c (E,10)

the mean strain rate as function of the process parameters


is given by

^ ^ T^c. CK-0
CHAPTER IX

BIBLIOGRAPHY

1. Pearson, C.E. and Parkins, R.N., The Extrusion of


Metals, rev. ed. (New York, 1960), 74-141.

2. Prozorov, L.V., "Extrusion of Steel," F-TS-9806/V,


Liaison Office Technical Information Center, NCLTD,
WPAFB, Ohio.

3. Pearson, C.E. and Parkins, R.N., The Extrusion of


Metals, rev. ed. (New York, 1960), p. 203.

4. Rowe, G.W., "An Introduction to the Principles of


Metalworking" (New York, 1965), p. 51.

5. Siebel, E. and Fangmeier, E., "Researches on Power


Consumption in the Extrusion and Punching of Metals,"
Mitt K.W. Inst, fur Eisenforschung, 13 (1931).

6. Sabroff, A.M., Boulger, F.W., Henning, H.J., and


Spretnak, J.W., A Manual on Fundamentals of Forging
Practice, Supplement to TDR No. ML-TDR-64-95, Manufac­
turing Technology Division, Air Force Materials Labo­
ratory, WPAFB^ Ohio (Dec. 1964).

7. Thomsen, E.G. and Frisch, J., "Stresses and Strains in


Cold-Extruding 2S-0 Aluminum," Trans. AIME, 77 (Nov.
1955), 1343-1353.

8. Frisch, J. and Thomsen, E.G., "The Effect of Process


Variables on Extrusion Pressures of Lead," Trans. AIME,
Series B.J. Eng. Ind., 81 (1959), p. 207.

9. Pugh, H.L1.D., Battelle Memorial Institute's Final


Technical Report, "Survey of Current Knowledge of the
Deformation Characteristics of Beryllium, the Refrac­
tory Metals, and the Superalloys," DDC, Cameron Station,
Alexandria, Virginia (Feb. 1966), p. 277.

10. Sachs, G . , "Beitrag zur Theorie des ziehvorganges," z.


Angew. Math. Meehanik, 1_ (1927), 235-236.

11. Hoffman, O. and Sachs, G., "Introduction to the Theory


of Plasticity for Engineers" (New York, 1953), 176-186.

12. Hill, R., The Mathematical Theory of Plasticity, Oxford


at the Clarendon Press, London (1950).

13. Jordan, T.F. and Thomsen, E.G., "Comparison of an Unsym-


metric Slip-Line Solution in Extrusion with Experiment,"
J. Mech. Phys. Solids, £ (1956), 184-190.
76
77

14. Thomsen, E.G., "Comparison of Slip-Line Solutions with


Experiment,11 Trans. ASME, J. Appl. Mech., 23 (June,
1956), 225-230.

15. Purchase, N.W. and Tupper, S.J., "Experiments with a


Laboratory Extrusion Apparatus under Criterion of Plane
Strain," J. Mech. Phys. Solids, 1_ (1952-3), 277-283.

16. Hencky, H., "Ueber einige statisch bestimmte Falle des


Gluckgewechts im plastischern Koerpern," z. Angew. Math.
Mechanik, _3 (1923) , 241-251.

17. Geiringer, H., "Fondements mathematiques de la theorie


des corps plastiques isotropes," Memorial Sci. Math.,
No. 86, Gauthiers-Villars, Paris (1937).

18. Prager, W . , "A Geometrical Discussion of the Slip Line


• Field in Plane Plastic Flow," Trans. Rov. Inst. Technol­
ogy. 65 (1953), 1-25.

19. Thomsen, E.G., "A New Method for Construction of Hencky-


Prandtle Nets," J. Appl. Mech., 24 (1957), 81-84.

20. Johnson, W . ,"Extrusion Through Wedge-Shaped Dies, Part


I.," J. Mech. Phys. Solids, 3 (1955), 218-223.

21. Hill, R., "A Theoretical Analysis of the Stresses and


Strains in Extrusion and Piercing," J. Iron and Steel
Inst., 158 (1948), 177-185.

22. Johnson, W. and Mellor, P.B., Plasticity for Mechanical


Engineers, D. Van Nostrand Co., Inc. (New York, 1962),
287-316.

23. Johnson, W. and Kudo, H., The Mechanics of Metal Extru­


sion, Manchester University Press (1962), 204-207.

24. Johnson, W . ,"The Pressure for Cold Extrusion of Lubri­


cated Rod through Square Dies of Moderate Reduction at
Slow Speeds," J. Inst. Metals, 85 (1956-57), p. 403.

25. Kudo, H., "A Computation of Required Pressure for Extru­


sion-Forging Circular Shells," Proceedings of the 7th
Japanese National Congress on Applied Mechanics (1957),
p. 57.

26. Yang, C.T., "The Upper-Bound Solution as Applied to


Three-Dimensional Extrusion and Piercing Problems," J.
Eng. Ind., Trans. ASME (Nov. 1962), 397-404.
78

27. Frisch, J. and Thomsen, E.G., "The Effect of Process


Variables on Extrusion Pressures of Lead," J. of Eng.
Ind., Trans. ASME, 81 (1959), 207-216.

28. Avitzur, B., "Analysis of Wire Drawing and Extrusion


Through Conical Dies of Small Cone Angle," J. of Eng.
Ind., Trans. ASMS, 85 (1963), 89-96.

29. Avitzur, B., "Analysis of Wire Drawing and Extrusion


Through Conical Dies of Large Cone Angle," J. of Eng.
Ind., Trans. ASME (Nov. 1964), 305-316.

30. Kobayashi, S., "Upper-Bound Solutions of Axisymmetric


Forming Problems-II," J. of Eng. Ind., Trans. ASME
(Nov. 1964), 326-332.

31. Green, A.P., "Calculations on the Theory of Sheet


' Drawing," British Iron and Research Association,
Report MW/B/7/52 (April, 1952).

32. Thomsen, E.G. and Lapsley, J.T., Jr., "Experimental -


Stress Determination Within a Metal During Plastic
Flow," Proceedings of the American Society of Experi­
mental Stress Analysis, 11 (1954), 59-68.

33. Thomsen, E.G., Yang, C.T., and Bierbower, J.B., An Ex­


perimental Investigation of the Mechanics of Plastic
Deformation of Metals, University of California Press
(Berkeley and Los Angeles, 1954).

34. kobayashi, S., personal communication (April, 1966).

35. Bowden, F.P. and Tabor, D., The Friction and Lubrica­
tion of Metals. Part I, 2nd ed. (1954)j Part II (1964),
Oxford University Press.

36. Thomsen, E.G., Yang, C.T., and Kobayashi, S., Mechanics


of Plastic Deformation in Metal Processing, MacMillan
Co. (New York, 1965), p. 221.

37. Peterson, M.B. and Ling, F.F., "Investigation of Fric­


tional Phenomena in Hot Metal Deformation," Final
Report submitted by Mechanical Technology, Inc., Latham,
New York on Navy Contract N0w65-0363-C.

38. Kudo, H., "An Upper-Bound Approach to Plane-Strain Forg­


ing and Extrusion-I," Inst. J. Mech. Sci., _1 (1960),
57-83.

39. Kudo, H., "An Upper-Bound Approach to Plane-Strain Forg­


ing and Extrusion-II," Inst. J. Mech. Sci., 3. (1960),
229-252.
79

40. Johnson, W. and Kudo, H., "The Use of Upper-Bound Solu­


tions for the Determination of Temperature Distribu­
tions in Fast Hot Rolling and Axi-Symmetric Extrusion
Processes," Inst. J. Mech. Sci., 3. (1960), 175-191.

41. Kudo, H., "Some Analytical and Experimental Studies of


Axi-Symmetric Cold Forging and Extrusion-I," Inst. J.
Mech. Sci.. 2 (1960), 102-127.

42. Singer, A.R. and Al-Samarrai, S.H., "Temperature


Changes Associated with Speed Variation During Extru­
sion, ".J1_Jrtist1_Metals, 89 (1960-61), 225-231.

43. Singer, A.R., "Temperature Effects in Extrusion,"


Metal Industry (4 May and 11 May 1962), 346-349 and
375-376.

44. Sauve, Ch., "Phenomenes Calorifiques associes au


filage," Memorires Scientifiques Rev. Metallurg., 59
(1962), 189-207.

45. Cruden, A.K., "Temperature Measurements During the Cold


Extrusion of Steel," Report to the Dept, of Scientific
and Industrial Research, National Engineering Labora­
tory.

46. Alexander, J.M., "The Application of Experimental and


Theoretical Results to the Practice of Hot Extrusion,"
J. Inst. Metals. 90 (1961-62), 193-215.

47. Treco, R.M., "Theoretical and Experimental Analysis of


the Extrusion Process for Metals," Trans. ASM, 55
(1962), 697-718.

48. Alder, J.F. and Phillips, V.A., "The Effect of Strain


Rate and Temperature on the Resistance of Aluminum,
Copper, and Steel to Compression," J. Inst. Metals,
83 (1954-55), 80-86.

49. Taylor, G.I. and Quinney, H., "The Latent Energy Remain­
ing in a Metal After Cold Working," Proc. Royal Soc.,
143 (1934), 307-326.

50. Ghiotti, P. and Carlson, O.N., USAEC Report, ISC-709


(1956), 17-22.

51. Butler, C.P. and Inn, E.C.Y., USNRDL-TR-177, DDC No.


143863, 1-27.

52. Cook, C., personal communication (July, 1966).

53. Kelley, K.K., "Contributions to the Data on Theoretical


Metallurgy," Bureau of Mines Bulletin 476 (1949).
80

TABLE 3.1

CALCULATED MAXIMUM VALUE OF ^

Max. value
r b of y*

2:1 .707 .828

4:1 .500 .667

6:1 .408 .580

8:1 .354 .523

10:1 .316 .480

12:1 . .289 .448


TA BLE 3. 2

V A R IO U S C O N T R IB U T IO N S TO T H E T O T A L P d /Y FO R R = 6 A N D CC = 6 0 ° c

CD (2) (3 ) (4 ) (5 ) (6) (7 ) (8)


P . /X Smooth T o o I b pd AB o u g h Stools
R e g io n (l ) Region (? ) L ine 36 Line 46 Line 34 Line 35 L ine 15 Xiine 45 l-fi_____ gcols. 1-8
?

y=.2. 0 .502 .if87 .231 .482 .247 .829 .098 1 .0 8 2 .7 8 3 .9 6


.1 A 97 .232 .434 .282 2,7 6 3 .9 4
.2 .if 93 .235 .392 • 318 2 .7 5 3 .9 4
.3 .lf89 .240 • 354 .353 2 .7 5 3 .93
.4 .if 88 .248 .321 .389 2 .7 6 3 .9 4
0 .505 .If87 .217 .484 .282 .789 .'131 1 .0 1 2.7 6 3 .9 0
.1 .499 .218 .433 .320 2 .7 5 3 .8 9
.2 .U9*f .221 .388 .358 2 .7 4 3 .8 8
.3 .U90 .227 .348 •395 2 .7 4 3 .8 8
.If88 .234 .313 .433 2.7 5 3 .8 9
Vs — .so 0 • 509 .if87 .202 .488 .322 .754 .168 .945 2 .7 6 3 .8 8
.1 .501 .204 .434 .362 2.7 4 3 .8 6
.2 .if 96 .207 .385 .403 2.7 3 3 .8 5
.3 .492 .213 .343 .443 2 .7 3 * 3 .8 5
.4 .If89 .221 .305 .484 2 .7 4 • 3 .8 6
Y = **35* 0 .513 .If87 .188 .496 .367 .723 .211 . 877. 2 .7 7 ! 3 .8 6
.1 .50lf .189 .437 .411 2.7 5 3 .8 4
.2 .498 .193 .385 .455 2 .7 4 3 .8 3
.3 .493 .199 .339 .499 2 .7 4 3 .8 3 *
.If .If91 .208 .299 .543 2 .7 5 3 .8 4
y*=.+o 0 .518 .487 .173 .507 .421 .698 .261 .810 2 .8 0 3 .8 8
.1 .508 .175 .444 .469 2 .7 8 3 .8 5
.2 .501 .179 .387 .516 2 .7 7 3 .8 4
.3 .495 .186 .337 .563 2 .7 7 3 .8 4
.U .If92 .195 .294 .611 2 .7 8 3 .8 5
0 .524 .487 .159 .523 .484 .681 •321 .742 2 .8 6 3 .9 2
.1 .513 .160 .454 .536 2.83 3 .8 9
.2 .50lf .165 .392 .588 2 .8 1 3 .8 7
.3 .498 .172 .337 .640 2 .8 2 3 .8 8
.if .if 98 .183 .290 .691 2.83 3 .8 9

00
H
TABLE 3.3

V A L U E S OF THE V E L O C IT Y F IE L D G E O M E T R Y F A C T O R S W HICH
Y IE L D E D M IN IM U M P R E D IC T E D P R E S S U R E S

The first term for each extrusion condition represents the ^ value
while the second term is the p factor.

PERFECTLY SMOOTH TOOLS

Reduction Ratio

Half Angle 1.25 2 4 6 8 10 12

30° .3,.25 0, .3 0, .25 0, .3 0, .3 0, .3 0, .3


45° .55,.25 •25,.3 05,.25 0, .25 0, .3 0, .3 0, .3
60° .75,.25 .55,.3 .4,.3 .3,.25 ■25,•25 .2,.25 .15,

PERFECTLY ROUGH TOOLS

Reduction Ratio

Half Angle 1.25 2 4 6 8 10 12

30° .25,.25 0,.3 0,.25 0,.3 0, .3 0..3 0, .3


45° .5,.25 .25,.3 .1,.3 0,.25 .02,.3 0, .3 0, .3
600 .65,.25 .5,.25 .4,.3 .35,.25 35,.3 .25,.25 .25,

oo
to
F ig u r e 4 . 1 S c h e m a tic I llu s t r a t io n o f A F M L
E x tr u s io n P r e s s w ith L o a d C e ll

Steel

P Si Mn Ni

.19% .032% .016% .25% .91% .016%

Balance Fe

Copper

0 P Bi Mg Ag Ni

6.4 ppm .002% 10 ppm 1 ppm 10 ppm 10 ppm

As Fe Pb Sb Sn Si

100 ppm 10 ppm 10 ppm 100 ppm 10 ppm 20 ppm

Balance Cu
TABLE 4.2

PREPARATION OP CORNING NR. 0010 GLASS MIXTURE

Heat 375 cc of distilled water in a beaker to 100°P.

Add 1 g of Carbopol 934 to heated water and mix


thoroughly.

Add NaOH solution (7 to 12 pellets or 1 g NaOH in


59 cc of distilled water) to Carbopol 934 solution
to neutralize to pH of 7.

Add 400 to 700 cc of -100 mesh glass to neutralized


solution and mix thoroughly with electric mixer.

Brush mixture on a warm billet (170 to 190°F) until


a thickness of .020 to .025 inches is attained.
TABLE 4.3

E X T R U SIO N E X P E R IM E N T S W ITH 1018 S T E E L A T 1 8 0 0 ° F A N D 4 5 ° D IE A N G L E

Steady
Billet Total Pressure Deformation Pressure State —
intrusion Reduction Length Speed (in./sec) (ksi) (ksi) Pressures Bfc/V Ed/V V e T
Fumber Ratio (in.) max. mln. start finish start finish (ksi) Xisail (ksi) (in. /sec) ( seo'l) (ksi) Prts/?

UHLUBRICATED

1967 2.12 4.5 2.0 1.75 62 35 35 32 32 44 33 1.8 4.6 21-5 1.50

1770 4.14 3.0 2-5 1.6 115 80 62 55 55 91 58 1.6 9-2 23-0 2.40

4.12 4.5 3.5 3.0 132 80 62 55 55 102 56 3.0 16.8 24-5 2.30
1779

1771 6.21 3.0 3-25 1.5 130 73 85 70 70 88 73 1.6 15.6 24.5 2.85

1780 6.19 4.5 3.5 3.3 130 77 78 72 72 90 73 3.3 31-5 26.5 2.70

1806 6.00 5.0 2.25 0.0 this extrusion vas purposely stopped, to observe the deformation pattern

1772 8.23 3.0 2.0 1.5 147 127 89 85 85 133 86 1.7 24.1 26.0 3-25

1785 8.19 4.5 5.5 3.0 165 89 96 84 84 110 86 5.5 75.8 29.0 2.90

10.14 3.0 4.0 3.25 126 89 99 86 86 99 90 3.3 62.1 28.5 3.00


1777

1786 10.11 4.5 5.5 2.5 165 96 97 88 88 112 89 5-0 92.f 30.0 2-95

0010 GLASS LUBRICATION

1.98 4.5 5.5 4.0 34 30 26 26 26 31 26 4.9 ii.o 23-5 1.10


1969
4.07 3.0 2.0 68 61 54 51 51 61 51 2.0 10.8 23-5 2.15
1794 1.9

4.09 4.5 2.7 2.6 78 62 58 54 53 66 53 2.6 14.6 24.5 2.15


1792

6.13 3.0 2.1 2.0 82 70 70 65 65 73 67 2.0 19.0 25.0 2.60


1791
6.08 4.5 2.75 2.5 88 73 72 67 67 78 68 2.5 23.3 25.5 2.65
1793

1807 6.00 5.0 2.5 0.0 this extrusion vas purposely stopped to observe the deformation pattern

1788 8.20 3.0 2.5 2.0 86 74 77 70 70 75 70 2.1 28.4 26.0 2.70

1790 8.08 2.5 2.4 96 82 78 73 73 86 74 2.4 33.1 26.5 2.75


4.5
3.0 2.75 io4 78 97 83 83 84 86 2.9 53-9 28.0 2.95
1787 10.11 3.75
1759 10.11 4.5 3.6 3-3 103 84 92 84 84 90 85 3.6 65.9 28.5 2.95
TABLE 4.4

E X T R U SIO N E X P E R IM E N T S W ITH 1018 S T E E L A T 1 8 0 0 ° F A N D 6 0 ° D IE A N G L E

Billet Total Pressure Deformation Pressure State


Extrusion Seduction Length Speed (in./sec) (ksi) (ksi) Pressure V 4
Et/V
Number Ratio fin.) max. mln. start finish start finish (ksi) (ksi) (in./sec) (sec ^) ( L i)

UKLUHRICATED

1767 4.10 3.0 2.0 2.0 84 58 69 57 57 63 59 1.5 14.2 24.0 2.35


1781 4.06 4.5 3.25 2.75 132 97 69 57 57 112 57 3.2 30.2 26.5 2.15
1768 6.1k 3.0 2.5 2.25 108 70 82 70 70 81 7k 2.2 36.0 27.0 2.60
1782 6.1k k.5 2.5 2.0 154 123 76 69 69 131 70 2.6 41.5 27.5 2.50

1769 8.22 3.0 2.5 1.5 Ik k 92 92 78 78 116 83 1.5 36.3 27.0 2.90
1783 8.28 4.5 5-0 2.0 174 105 92 81 81 131 83 4.2 102 30.0 2.70
1778 10.18 3.0 3.25 3.0 136 95 112 92 91 103 95 3.3 106 30.5 3.00
1781* 9-77 4.5 5.0 1.5 190 117 107 90 90 l4l 93 k .5 138 31.5 2.85

0010 GLASS LUBRICATION

1801 4.09 3.0 2.0 1.5 81 62 65 57 57 66 59 1.9 18.1 25.O 2 .3 0


1800 6.13 3.0 2.25 1-9 99 76 82 70 70 81 72 1.9 31.2 26.5 2.65

179T 6.10 4.5 2.5 2.5 92 80 77 70 69 81 70 2.4 38.7 27.O 2.55


1796 8.17 ^.5 3.5 3.5 12k 95 90 78 77 98 79 3.3 77.8 29.5 2.60

1795 10.18 3.0 3.5 3.5 108 90 88 80 80 9k 81 3.2 104 30.5 2.65
1798 10.09 k.5 3.5 3.5 , 122 96 93 82 82 101 83 3.4 108 30.5 2.70

00
CTi
TABLE 4.5

E X T R U SIO N E X P E R IM E N T S W IT H O FH C C O P P E R A T 8 0 0 ° F A N D 4 5 ° D IE A N G L E

Steady-
Billet Total Pressure Deformation Pressure State T-
Speed (in../sec.) (ksi) (ksi) Pressure Temp. %/v V
& 1
Extrusion Seduction Length V* *ds/?
WiiwhpT Ratio max. mln. start finish start finish (ksi) CD Xksii (ksi) (in./dec.) (sec"l) Iteil

UHLUBRICATED

1966 2.12 3.0 2.5 1.75 8k 57 63 57 57 - 67 59 2.2 5.8 28.5 2.00

1678 k.03 3-0 2.0 1.5 15k 127 92 89 90 - lko 91 1.8 9.9 29.5 3.05

1677 6.10 3-0 1.5 .5 186 135 101 96 97 - 161 100 l.lt- 12.5 29.5 3.30

1679 8.10 2.5 3.0 .7 189 159 122 117 117 - 17U 119 2.7 36.2 31.0 3.75
1680 10.00 2.8 STOCK - - - - - - - - - - -

tTSKT? 6ok LQBKtCKEItHI

1
1968 1.97 5.0 2.0 1.5 81 k9 53 k9 k9 - 62 50 1.7 3.9 28.5 1.70

1720 k.Ok k.5 3.25 2.5 87 69 68 62 63 810 76 6k 2.5 13.8 30.0 2.10

1725 6.07 2.88 2.0 1.75 100 88 81 78 78 938 95 79 1-9 17.9 30.0 2.60

1721 6.06 3.38 2.0 2.0 99 86 85 80 80 897 91 82 2.0 18.k 30.0 2.65

1718 6.07 5.0 3.25 2.k 119 99 92 93 92 - 118 98 2-5 22.7 30.5 3.00

1717 8.17 5.0 5.0 3.5 159 115 109 103 10k - 135 106 k.k 60.0 31.5 3.30

1722 10.03 2.75 3.0 3.0 109 10k 97 90 90 951 108 93 3.0 55.7 31.5 2.85

1726 10.03 3.38 k.o 3.0 123 109 111 105 105 881 115 1CT 3.0 5k-9 31-5 3-35

1719 10.07 5-0 k.75 k.O 153 12k 111 107 108 - 137 109 k.k 80.7 32.0 3.k0

00
TABLE 4. 6

E X T R U SIO N E X P E R IM E N T S W IT H OFHC C O P P E R A T 8 0 0 ° F A N D 6 0 ° D IE A N G L E

Steady
Billet Total Pressure Deformation Pressure State
intrusion Seduction Length Speed (in./sec.) 1
(ksi) (ksi) Pressure Temp. V £ Y Pds/Y
hnnber Batio max. min^ start finish start finish (ksi) L h i ML (in./sec) (see ^) (ksi)

UHLUBRICA2ED

1706 U.05 2.6 2-75 1.25 175 117 80 77 lk9 81 2.3 22.0 30.5
1681 k.03 3-0 2.0 1.5 1U7 12k 88 85 86 - 138 86 2.1 20.0 30.5 2.80
1682 6.0k 3.0 1.5 1.0 175 lk8 103 100 100 - 163 101 1.5 23.2 30.5 3.30
1708 6.05 3.0 2.5 .6 190 128 86 92 - - 17k 92 1-9 30.7 31.0 -
1683 8.09 2.5 3.0 1.5 188 165 107 10k 10k - 177 105 2.6 61.7 31.5 3.30
1707 8.06 2.6 STOCK - - - - - - - - - - - -
168k 10.0 2.5 3.0 1.0 190 163 112 109 109 - 179 Ill 2.8 88.k 32.0 3.kO

FISKE 6(A IJIBRI/cAEIQN

171k 3.96 5-0 2.75 2.0 138 86 81 76 76 817 112 78 2.k 22.1 30.5 2.50

1723 6.0k 3-38 2.25 1.9 108 93 89 86 86 915 99 88 2.0 31.k 31.0 2.80
1716 6.02 5.0 3.25 2.5 123 95 99 92 93 - 1CTT 95 3.0 k7.k 31.5 2.95
1715 8.06 5.0 5.25 k.5 130 109 100 97 97 - 119 99 k.9 Ilk 32.5 3.00
172k 10.03 5.0 5.0 k.o 127 113 112 109 109' - 117 109 k.5 lk3 32.5 3.35

oo
oo
TABLE 6.1

E F F E C T OF B IL L E T L E N G T H ON T H E P R E S S U R E R E Q U IR E D TO E X T R U D E
U N L U B R IC A T E D 1018 S T E E L A T 1 8 0 0 ° F

Lo = 3"

Initial Initial
Total Deformation Percent
Extrusion oC Reduction Pressure Pressure Container
Number (deqrees) Ratio (ksi) (ksi) Friction

1770 45 4.14 115 62 46


1771 45 6.21 130 85 35
1772 45 8.23 147 89 39
1777 45 10.14 126 99; 21
1767 60 4.10 84 69 18
1768 60 6.14 108 82 24
1769 60 8.22 144 92 36
1778 60 10.18 136 112 18

Average = 30%

Lo = 4-1/2"

1967 45 2.12 62 35 44
1779 45 4.12 132 62 53
1780 45 6.19 130 78 40
1785 45 8.19 165 96 42
1786 45 10.11 165 97 41
1781 60 4.06 132 69 48
1782 60 6.14 154 76 51
1783 60 | 8.28 174 92 47
1784 60 9.77 190 107 44
oo
vo
Average = 46%
TABLE 6.2

E F F E C T O F B IL L E T L E N G T H ON T H E P R E S S U R E R E Q U IR E D TO E X T R U D E
0 0 1 0 G L A SS L U B R IC A T E D 1018 S T E E L A T 1 8 0 0 ° F

Lo = 3"

Initial Initial
Total Deformation Percent
Extrusion oc Reduction Pressure Pressure Container
Number (degrees) Ratio (fcsi) (ks i) Friction

1794 45 4.07 68 54 21
1791 45 6.13 82 70 15
1788 45 8.20 86 77 10
1787 45 10.11 104 97 7
1801 60 4.09 81 65 20
1800 60 6.13 99 82 17
1795 60 10.18 108 88 19

Average = 16%

Lo = 4-1/2"

1969 45 1.98 34 26 24
1792 45 4.09 78 58 26
1793 45 6.08 88 72 18
1790 45 8.08 96 78 19
1789 45 10.11 103 92 11
1797 60 6.10 92 77 16
1796 60 8.17 124 90 27
1798 60 10.09 122 93 24

Average = 2 1 %
TABLE 6.3

E F F E C T O F L U B R IC A T IO N ON T H E P R E S S U R E R E Q U IR E D
TO E X T R U D E OFHC C O P P E R A T 8 0 0 ° F

Unlubricated
Initial Initial
Total Deformation Percent
:trus Length ©C Reduction Pressure Pressure Container
iiriber (inQ„ (degrees) Ratio (ksi) J M i ) _______ Friction
1966 3.0 45 2.12 84 63 25
1678 3.0 45 4.03 154 92 40
1677 3.0 45 6.10 186 101 46
1679 2.5 45 8.10 189 122 35
1706 2.6 60 4.05 175 80 54
1681 3.0 60 4.03 147 88 40
1682 3.0 60 6.04 175 103 41
1708 3.0 60 6.05 190 86 55
1683 2.5 60 8.09 188 107 43
1684 2.5 60 10.0 190 112 41
Avg. = 42%

Fiske 604 Lubrication


1968 5.0 45 1.97 81 53 35
1720 4.5 45 4.04 87 68 22
1718 5.0 45 6.07 119 92 23
1717 5.0 45 8.17 159 109 31
1719 5.0 45 10.07 153 111 27
1714 5.0 60 3.96 138 81 41
1716 5.0 60 6.02 123 99 20
1715 5.0 60 8.06 130 100 23
1724 5.0 60 10.03 127 112 12
VQ
H
Avg. = 26%
TABLE 6.4

C A L C U L A T E D V A L U E S OF AND F O R OFHC
CO PPER AT 800°F

Unlubricated
•usio K (ksi)
1677 .128 19.4
1678 .116 17.0
1679 .086 14.4
1681 .090 15.3
1682 .091 18.2
1683 .100 19.0
1684 .097 19.4
.101 Avg. = 17.5 Avg.

Fiske 604 Lubrication

us ioi yUc— K (ks

1715 .031 4.46


1716 .035 3.48
1717 .050 6.89
1719 .033 3.94
1720 .033 3.05
1723 .035 3.05
1725 .059 5.94
1726 .013 1.22
Avg. = .036 Avg. = 4.00 Avg.
u>
F ig u r e 1. 1 P h o to g r a p h o f 700 T o n H o r iz o n ta l E x tr u s io n P r e s s
Container Liner
Die
Die Backer
H eating Element

Platen
Dummy Block Bolster
from High Pressure Bottle

Pressure Transducer
Main Cylinder
Bumper Plate Columns ( Tie Rods) Runout Tube

Nose Block

Graphite Follower Block Billet

F ig u r e 1 .2 S c h e m a tic I llu s t r a t io n of F o r w a r d E x tr u s io n P r o c e s s
STEM
(RAM) DUMMY BLOCK
FOLLOWER MATERIAL
(GRAPHITE)
.BILLET

E Z 2 3 2 2 X
&
DIE BACKER

CONTAINER LINER

F ig u r e 1 .3 S c h e m a tic R e p r e s e n t a t io n o f F o r c e s D u r in g E x tr u s io n
L = Li L =L L= 0

F ig u r e 2. 1 I d e a liz e d P r e s s u r e D is tr ib u t io n In an
E x tr u s io n B i l l e t
STRESSES ACTING ON BIL L E T E LE M E N T
fp p

cos a

0
1 fP P

Do < f D+dD

F ig u r e 2 . 2 D im e n s io n s in E x tr u s io n

vo
F ig u r e 2 . 3 N o r m a l P r e s s u r e on F ig u r e 2 . 4 F r ic t io n a l F o r c e on
S u r fa c e B i l l e t S u r fa c e E le m e n t
R
l>l 21 8:1
8

l>
4

*0
2

0
1.0 1.5 2.0 2.5
InR-

F ig u r e 2. 5 P , /Y v s . In R f r o m S lab M eth o d o f S t r e s s A n a ly s is
100

45 45

F ig u r e 2 . 6 S li p - lin e F i e ld fo r r L e s s T han
2 s in GC/ ( l+2sinOC)
45'

45

F ig u r e 2 . 7 S lip - lin e F i e ld fo r r G r e a te r th an 2 s in GC/( 1 + 2 s in OC)

101
R —>
6:1 12:1

l>
P /Y = InR

=40

2.0 2.5
In R

102
F ig u r e 2 . 8 /Y v s . In R f r o m S lip - lin e S o lu tio n fo r F r i c t i o n l e s s
E x tr u s io n
BILLET EXTRUDED BAR

\\\\\\\\\\\\\\mmmv

F ig u r e 2 . 9 V e lo c it y F i e ld w ith C y lin d r ic a l D e fo r m in g
R e g io n s fo r S tr a ig h t D ie E x tr u s io n

103
DEFORMATION
- V O LU M E

E X T R U D E D BAR

BILLET
DIE

F ig u r e 2 . 10 R a d ia l V e l o c i t y F i e ld f o r A x is y m m e t r ic E x tr u s io n

104
105

6:1

AVITZUR
KOBAYASHI

SHEAR ON DIE FACE

NO FRICTION
Ph / Y

1.0 2.0 2 .5
InR

F ig u r e 2 . 11 P ^ /Y v s . In R f r o m U p p e r -B o u n d S o lu tio n s
D IE

B IL L E T EXTRUDED BAR

DIE

F ig u r e 2 . 12 G r e e n 's P r o p o s e d V e lo c it y F ie ld fo r S h e e t D r a w in g
l-*
o
o\
(r)
STRESS
SHEAR

REGION I REGION II REGION III

NORMAL PRESSURE(p)

F ig u r e 3. 1 S c h e m a tic R e la tio n s h ip B e t w e e n th e I n te r fa c e
S h ea r S t r e s s and th e N o r m a l P r e s s u r e

107
4-1
8

m=0
7
V

5
P^/Y

*0
3
,»0

0
1.0 1.5 2.0 2.5
InR-

108
F ig u r e 3 . 2 P /Y . v s . In R fo r L ow V a lu e o f th e W o r k H a r d e n in g
_n .>
C o e f f ic ie n t
InR-*'

109
F ig u r e 3. 3 /Y . v s . In H fo r H ig h V a lu e s o f th e W o rk H a r d e n in g
C o e f f ic ie n t
110

RIGID
EXTRUSION
RIGID
DIRECTION

RIGID

POSITION DIAGRAM VELOCITY DIAGRAM

F ig u r e 3 . 4 T r ia n g le V e lo c it y F i e ld R e p r e s e n t a t io n
fo r C o n ic a l D ie E x t r u s io n s
RI GI D

F ig u r e 3. 5 D is p la c e m e n t ,of O u ter R e g io n
1

111
5

F ig u r e 3. 6 V e lo c it y D ia g r a m o f th e 45 L in e

112
113

1.25: | 2:1 4:1 611 8:1 10:1 I2:|

8.5
HIGHER CURVES ARE FOR SHEAR ALONG DIE FACE AND CONTAINER
LOWER CURVES REPRESENT PERFECTLY SMOOTH TOOLING
O a • 30°
X Ct ■45°
D a « 60° .
4.5

SHEAR ON DIE FACE

3 .5

IV
s NO FRICTION

2.5

0 .5 1.0 1.5 2.0 2.5


In R-

F ig u r e 3 . 7 P ^ /Y f r o m T r ia n g le V e lo c it y F ie ld
Container Liner
Die
H eating Element
Load Cell
Platen
Dummy Block Bolster
from High Pressure Bottle

to Reservoir

3 Y7777,

Pressure Transducer
Main Cylinder
Bumper Plate Columns (Tie Rods) Runout Tube

Nose Block

Graphite Follower Block Billet

F ig u r e 4 . 1 S c h e m a tic I llu s t r a t io n o f A F M L
E x tr u s io n P r e s s w ith L o a d C e ll
H
H
I" DEPTH
l / 2 " x 1/2" SLOT THREADS/INCH
H

OOO

CHAMFER

1 .6 8 0

8 .4 6 9 3 .1 8 8
+ OOO + 000
-.0 0 5 -.0 0 5

115
Figure 4.2 Modified Die Backer for Load Cell
GROUND FINISH
32

o o 0>
roo o DIE RATIO Vs D
.11

4:| 1.70
6 :| 1.50
8 :1 I. 35
10:1 1.25
3.440 ±.01

F ig u r e 4 . 3 L o a d C e ll D im e n s io n s

116
117

REFLECTING MIRROR
DETECTOR

LENS RETAINING QUARTZ LENS


C AP —

ARGON

BO LSTER

F ig u r e 4 . 4 E q u ip m e n t M o d ific a tio n f o r R a d ia tio n


P y r o m e tr y
118
119

L±.05

D ±.002

D a 2 .9 6 0 FOR OFHC COPPER


D * 2 .9 5 0 FOR 1018 STEEL

F ig u r e 4 . 6 E x tr u s io n B i l l e t D im e n s io n s
200
#1717 a » 4 5 ° ,R « 8 .1 7 , L = 5 \ FISKE 6 0 4
* 1 6 7 9 a » 4 5 ° ,R » 8 .I O , L »2.5", UNLUBRICATED

180
650
PRESSURE (K S I)-*

600
160

TO N S-#-
550

140
500
■X-----

450
120

400

100
350
0 5 1.0 1.5 2.0 2.5 3 .0
RAM TRAVELtinJ

H
Figure 5.1 Pressure-Distance Curves for OFHC Copper
650
# 1 7 9 6 a * 6 0 ° , R * 8 .I7 , L * 4 1/2",0010 GLASS
# 1 7 8 3 a * 6 0 °, R * 8 .2 8 ,L * 4 1 /2 " ,UNLUBRICATED
600
160

550

140
500
PRESSURE (KSI)

TONS
— O ----- O
\ 450
120

^ 400

100
350

300
80

250
0 .5 1.0 1.5 2.0 2 .5
RAM TRAVEL (In.)
H
fO
Figure 5„Z Pressure-Distance Curves for 1018 Steel
122

FLOW STRESS(KSI)
26.0 27.0 28.0 29.0 30.0 31.0 32.0 33.0
1000

100
(sec"
RATE
STRAIN

I.Ol-----
1.40 1.42 1.44 1.46 1.48 1.50 1.52
LOG FLOW STRESS(KSI)->

F ig u r e 5 .3 L o g ^ S tr a in R a te v s . L o g ^ F lo w S t r e s s of
C o p p er a t 8 0 0 ° F
123

FLOW STRESS(KSI)
20 24 25 26 27 26 30
1000

100
RATE (sec1)
STRAIN

1.30 1.32 1.34 1.36 1.38 1.40 1.42 1.44 1.46 1.48 1.50
LOG|Q FLOW STRESS (K S I) ►

F ig u r e 5 . 4 Log^Q S tr a in R a te v s . L o g F lo w S t r e s s of
S teel at 1800°F
1000

Lo = 2 .7 5
| 950
TEMPERATURE (°F)

Lo = 3 .3 8
900

Lo = 3 . 3 8

850

Lo= 4 .5
800
60 70 80 90 100 110 120 130
Et / V ( k s i )

F ig u r e 6. 1 M e a s u r e d E x it T e m p e r a tu r e v s . E /V fo r OFHC

124
C o p p er a t 8 0 0 ° F and OC = 4 5 °
F ig u r e 6 . 2 P h o to g r a p h o f G r id s on L u b r ic a te d and
U n lu b r ic a te d S t e e l E x t r u s io n B i l l e t s
126

4 65

60

3 55

50

(Ft -Fd) TONS


2 45

40

35

30

EXTRUSION DIRECTION
25
0 1.0 2.0 3.0
L-»

F ig u r e 6 . 3 Ln (F /F ^ ) and (F t - F^) v s . L fo r
E x tr u s io n N r . 1720
127

R—►

130

450
120 X REPRESENTS UNLUBRICATED CONDITION
O REPRESENTS LUBRICATED CONDITION

400

00
(KSI)

350

90
PRESSURE

OFHC COPPER 300


80
AT 8 0 0 °F

TONS
DEFORMATION

70
250

1018 STEEL
60
' AT I800°F

2 00
50

50

30
00

20
0 .5 1.0 1.5 2.0 2.5
In R — ►

F ig u r e 6 . 4 D e fo r m a tio n P r e s s u r e v s . In R f o r GC = 45
R—
4:| 6:l 8=1 10:1 I2M

X REPRESENTS UNLUBRICATED CONDITION


450
120 O REPRESENTS LUBRICATED CONDITION
PRESSURE (KSI)

OFHC COPPER
AT 8 0 0 * F 400

100

350
DEFORMATION

TONS
1018 STEEL
90 AT I8 0 0 * F

300
80

70
250

60

200
501—
1.0 1.2 1 .4 1.6 1.8 2.0 2.2 2 .4
In R — >

, o to
F ig u r e 6 . 5 D e fo r m a tio n P r e s s u r e v s . In R fo r CC = 60 00
10:1 12:1

Ct =4 5 ’ STEEL
Ct= 6 0 ° STEEL
a s 4 5 ’ COPPER
a = 60* COPPER
OHFC COPPER4 L*5
— 1018 STEEL

« 2 .0 -

In R

129
F ig u r e 6. 6 E x p e r im e n t a l P ^ g /Y v s . In E fo r L u b r ic a te d
C o p p er and S t e e l B i l l e t s
R-»
6:l
0
O fl* 45°STEEL
A a * 4 5 ° COPPER
X a « 6 0 ° STEEL
□ a « 6 0 ° COPPER
— OFHC COPPER L« 2 .5 - 3 .0 '
— 1018 STEEL a*
o

0
0 5 1.0 1.5 2.0 2 .5
In R—>

130
F ig u r e 6. 7 E x p e r im e n t a l P ^ s /Y v s . In R fo r U n lu b r ic a te d C o p p er
and S t e e l B i l l e t s

You might also like