Metalformingandthefinite Elementmethod Shiro Kobayashi
Metalformingandthefinite Elementmethod Shiro Kobayashi
FINITE-ELEMENT METHOD
OXFORD SERIES ON ADVANCED MANUFACTURING
SERIES EDITORS
J. R. CROOKALL
MILTON C. SHAW
SHIRO KOBAYASHI
SOO-IK OH
TAYLAN ALTAN
Kobayashi, Shiro.
Metal forming and the finite-element method /
Shiro Kobayashi, Soo-Ik Oh, Taylan Allan.
p. cm. — (Oxford series on advanced manufacturing;
4) Bibliography: p. Includes index.
ISBN 0-19-504402-9
1. Metal-work—Mathematical models. 2. Finite-element method.
I. Oh, Soo-Ik. II. Allan, Taylan. III. Title. IV. Series.
TS213.K56 1989
671'.072'4—dc!9 88-11995
CIP
135798642
Printed in the United States of America
on acid-free paper
PREFACE
Berkeley S. K.
Columbus S. O.
May 1988 T. A.
CONTENTS
Symbols, xiii
1. Introduction, 1
1.1 Process Modeling, 1
1.2 The Finite-Element Method, 3
1.3 Solid Formulation and Flow Formulation, 4
1.4 Metal Forming and the Finite-Element Method, 5
References, 6
2. Metal-Forming Processes, 8
2.1 Introduction, 8
2.2 A Metal-Forming Operation as a System, 8
2.3 Classification and Description of Metal-Forming Processes, 11
References, 24
5. Methods of Analysis, 73
5.1 Introduction, 73
5.2 Upper-Bound Method, 74
5.3 Hill's General Method, 78
5.4 The Finite-Element Method, 83
5.5 Concluding Remarks, 88
References, 88
Index, 371
This page intentionally left blank
SYMBOLS
Temperature
CONSTRAINTS!
INPUT
FIG. 1.1 Block diagram for process design and control in metal forming.
Introduction 3
References
1. Thomsen, E. G., Yang, C. T., and Bierbower, J. B., (1954), "An Experimen-
tal Investigation of the Mechanics of Plastic Deformation of Metals," Univ.
California Pub. Engg., Vol. 5.
2. Hill, R., (1963), "A General Method of Analysis of Metal-Working Proc-
esses," /. Mech. Phys. Solids, Vol. 11, p. 305.
3. Hill, R., (1950), "The Mathematical Theory of Plasticity," Oxford University
Press, London.
4. Prager, W., and Hodge, P. G., Jr., (1951), "Theory of Perfectly Plastic
Solids," Chapman and Hall, London.
5. Hoffman, O., and Sachs, G., (1953), "Introduction to the Theory of Plasticity
for Engineers," McGraw-Hill, New York.
6. Unksov, E. P., (1961), "An Engineering Theory of Plasticity," Butterworths,
London.
7. Johnson, W., and Mellor, P. B., (1973), "Engineering Plasticity," Van
Nostrand and Reinhold, London.
8. Ford, H., and Alexander, J. M., (1963), "Advanced Mechanics of Materials,"
Longmans Green, London.
9. Alexander, J. M., and Brewer, R. C., (1963), "Manufacturing Properties of
Materials," Van Nostrand, London.
10. Thomsen, E. G., Yang, C. T., and Kobayashi, S., (1963), "Mechanics of
Plastic Deformation in Metal Processing," Macmillan, New York: Macmillan-
Collier, London.
11. Kalpakjian, S., (1967), "Mechanical Processing of Materials," Van Nostrand,
Princeton, NJ.
12. Avitzur, B., (1968), "Metal Forming and Processes," McGraw-Hill, New
York.
Introduction 7
13. Johnson, W., and Kudo, H., (1962), "The Mechanics of Metal Extrusion,"
Manchester University Press, Manchester, UK.
14. Johnson, W., Sowerby, R., and Haddow, J. B., (1970), "Plane-Strain
Slip-Line Fields," American Elsevier, New York.
15. Johnson, W., Sowerby, R., and Venter, R. D., (1982), "Plane Strain Slip Line
Fields for Metal Deformation Processes," Pergamon Press, Oxford.
16. Blazynski, T. Z., (1976), "Metal Forming," Wiley, New York.
17. Rowe, G. W., (1977), "Principles of Industrial Metalworking Processes,"
Edward Arnold, London.
18. Slater, R. A. C., (1977), "Engineering Plasticity," Wiley, New York.
19. Hosford, W. F., and Caddell, R. M., (1983), "Metal Forming; Mechanics and
Metallurgy," Prentice-Hall, Englewood Cliffs, NJ.
20. Altan, T., Oh, S. I., and Gegel, H., (1983), "Metal Forming; Fundamentals
and Applications," American Society for Metals, Metals Park, Ohio.
21. Boer, C. R., Rebelo, N., Rystad, H., and Schroder, G., (1986), "Process
Modelling of Metal Forming and Thermomechanical Treatment," Springer-
Verlag, Berlin.
22. Mote, C. D., Jr., (1980), "Introduction to the Finite Element Method,"
Lecture Note, University of California at Berkeley.
23. Zienkiewicz, O. C., (1977), "The Finite Element Method," 3rd Edition,
McGraw-Hill, New York.
24. Strang, G., and Fix, G. J., (1973), "An Analysis of the Finite Element
Method," Prentice-Hall, Englewood Cliffs, NJ.
25. Huebner, K. H., (1975), "The Finite Element Method for Engineers," Wiley,
New York.
26. Zienkiewicz, O. C., (1984), Flow Formulations for Numerical Solutions of
Forming Processes, "Numerical Analysis of Forming Processes," edited by J.
F. T. Pittman et al., Wiley, New York, p. 1.
27. Nagtegaal, J. C., and Veldpaus, F. E., (1984), On the Implementation of
Finite Strain Plasticity Equations in a Numerical Model, "Numerical Analysis
of Forming Processes," edited by J. F. T. Pittman et al., Wiley, New York, p.
351.
28. Kobayashi, S., (1982), "A Review on the Finite Element Method and Metal
Forming Process Modeling," /. Appl. Metal Working, Vol. 2, p. 163.
29. Kobayashi, S., (1985), "Metal Forming and the Finite Element Method—Past
and Future," Proceedings of the 25th Int. Conf. Mach Tool Des. Res., April,
Birmingham, p. 17.
2
METAL-FORMING PROCESSES
2.1 Introduction
In metal forming, an initially simple part—a billet or sheet blank, for
example—is plastically deformed between tools (or dies) to obtain the
desired final configuration. Thus, a simple part geometry is transformed
into a complex one, in a process whereby the tools "store" the desired
geometry and impart pressure on the deforming material through the
tool-material interface.
The physical phenomena constituting a forming operation are difficult to
express with quantitative relationships. The metal flow, the friction at the
tool-material interface, the heat generation and transfer during plastic
flow, and the relationships between microstructure/properties and process
conditions are difficult to predict and analyze. Often, in producing discrete
parts, several forming operations (preforming) are required to transform
the initial "simple" geometry into a "complex" geometry, without causing
material failure or degrading material properties. Consequently, the most
significant objective of any method of analysis is to assist the forming
engineer in the design of forming and/or preforming sequences. For a
given operation (preforming or finish-forming), such design essentially
consists of (1) establishing the kinematic relationships (shape, velocities,
strain-rates, strains) between the deformed and undeformed part, i.e.,
predicting metal flow; (2) establishing the limits of formability or produci-
bility, i.e., determining whether it is possible to form the part without
surface or internal defects; and (3) predicting the forces and stresses
necessary to execute the forming operation so that tooling and equipment
can be designed or selected.
For the understanding and quantitative design and optimization of
metal-forming operations it is useful (a) to consider a metal forming
process as a system and (b) to classify these processes in a systematic way.
finally the plant environment in which the process is being conducted [1].
Such a system is illustrated in Fig. 2.1, using impression die forging as an
example [2].
The "systems approach" in metal forming allows study of the effects of
process variables on product quality and process economics. The key to a
successful metal-forming operation, i.e., to obtaining the desired shape
and properties, is the understanding and control of metal flow. The
direction of metal flow, the magnitude of deformation, and the tempera-
tures involved greatly influence the properties of the formed components.
Metal flow determines both the mechanical properties related to local
deformation and the formation of defects such as cracks or folds at or
below the surface. The local metal flow is in turn influenced by the process
variables, which are discussed below.
Material Variables
For a given material composition and deformation/heat-treatment history
(microstructure), the flow stress (or effective stress), and the workability
(or formability) in various directions (anisotropy), are the most important
material variables in the analysis of a metal-forming process.
For a given microstructure, the flow stress is expressed as a function of
strain, strain-rate, and temperature. To determine the actual functional
relationship, it is necessary to conduct torsion, plane-strain compression,
and uniform axisymmetric compression tests. Workability or formability is
the capability of a material to deform without failure; it depends on (1)
conditions existing during deformation processing (such as temperature,
rate of deformation, stresses, and strain history), and (2) material variables
10 Metal Forming and the Finite-Element Method
Friction
The mechanisms of interface friction are very complex. One way of
expressing friction quantitatively is through a friction coefficient n, or a
friction shear factor m. There are various methods of evaluating friction,
i.e., estimating the value of \n or m. Tests most commonly used are the
ring and spike tests for massive forming and the plane-strain-draw and
stretch-draw tests for sheet forming [2].
Deformation Mechanics
In forming, material is deformed plastically to generate the shape of the
desired product. Metal flow is influenced mainly by (1) tool geometry, (2)
friction conditions, (3) characteristics of the stock material, and (4)
thermal conditions existing in the deformation zone. The details of metal
flow influence the quality and properties of the formed product and the
force and energy requirements of the process. The mechanics of deforma-
tion, i.e., the metal flow, strains, strain-rates, and stresses, can be
investigated by process modeling. Some analysis methods for process
modeling are outlined in Chap. 1 (see Section 1.1), and process modeling
by the finite-element method is the main subject of this book.
Product Properties
The macro- and microgeometry of the product, i.e., its dimensions and
surface finish, are influenced by process variables. The processing condi-
tions (temperature, strain, and strain-rate) determine the microstructural
variations taking place during deformation and often influence final
product properties. Consequently, a realistic systems approach must
include consideration of (1) the relationships between properties and
microstructure of the formed material and (2) the quantitative influences
of process conditions on metal flow and resulting microstructures.
Metal-Forming Processes 11
in contact, and friction between them has a major influence on the process.
In massive forming, the input material is in billet, rod, or slab form, and a
considerable increase in the surface-to-volume ratio occurs in the formed
part. In sheet forming, a sheet blank is plastically deformed into a
three-dimensional object without any significant changes in sheet thickness
and surface characteristics.
Processes that fall under the category of massive forming processes
(Table 2.1), have the following distinguishing features:
• The workpiece undergoes large plastic deformation, resulting in an
appreciable change in shape or cross section.
• The portion of the workpiece undergoing permanent (plastic) defor-
mation is generally much larger than the portion undergoing elastic
deformation; therefore, elastic recovery after deformation is
negligible.
The characteristics of sheet-metal forming processes (Table 2.2) are:
• The workpiece is a sheet or a part fabricated from a sheet.
• The deformation usually causes significant changes in shape, but not
in cross section, of the sheet.
• In some cases, the magnitudes of permanent plastic and recoverable
elastic deformations are comparable; therefore, elastic recovery or
springback may be significant.
Most significant metal-forming processes are listed in Tables 2.1 and 2.2
[3]. Selected massive and sheet forming processes are described in Figs 2.2
through 2.29 [2-10].
FIG. 2.2 Closed-die forging without flash. A billet with carefully controlled volume is
deformed (hot or cold) by a punch in order to fill a die cavity without any loss of material.
The punch and the die may be made of one or several pieces.
FIG. 2.3 Closed-die forging with flash. A billet is formed (hot) in dies (usually with two
halves) such that the flow of metal from the die cavity is restricted. The excess material is
extruded through a restrictive narrow gap and appears as flash around the forging at the die
parting line.
FIG. 2.4 Coining opeation. Coining is a closed-die forming operation, usually performed
cold, in which all surfaces of the work are confined or restrained, resulting in a well-defined
imprint of the die on the workpiece. It is also a restriking operation used to sharpen or change
an existing radius or profile.
FIG. 2.5 Forward extrusion forging. A punch compresses a billet (hot or cold) confined in a
container so that the billet material flows through a die in the same direction as the punch.
13
FIG. 2.6 Backward extrusion forging. A moving punch applies a steady pressure to a slug
(hot or cold) confined in a closed die, and forces the metal to flow around the punch in a
direction opposite the direction of punch travel.
(A) (B)
FIG. 2.7 Robbing (A) in a container and (B) without restriction. Robbing is the process of
indenting or coining an impression into a cold or hot die block by pressing with a punch.
14
FIG. 2.8 Nosing. Nosing is a hot or cold forming process in which the open end of a shell or
tubular component is closed by axial pressing with a shaped die.
FIG. 2.9 Open-die forging. Open-die forging is a hot forming process in which metal is
shaped by hammering or pressing between flat or simple contoured dies.
15
FIG. 2.10 Various stages in orbital forging processes. Orbital forging is the process of
forming shaped parts by incrementally forging (hot or cold) a slug between an orbiting upper
die and a nonrotating lower die. The lower die is raised axially toward the upper die, which is
fixed axially but whose axis makes orbital, spiral, planetary, or straight-line motions.
FIG. 2.11 Radial forging of a shaft. This hot or cold forming process utilizes two or more
radially moving anvils, or dies, for producing solid or tubular components with constant or
varying cross sections along their lengths.
16
FIG. 2.12 Upsetting with flat-heading tool. Upsetting is the process of forming metal (hot or
cold) so that the cross-sectional area of a portion, or all, of the stock is increased.
FIG. 2.13 Schematic of the rolling process for sheet and plates. Sheet and plate rolling is a
hot or cold forming process for reducing the cross-sectional area of the stock with the use of
rotating rolls. In general, the rolled material elongates and spreads simultaneously while the
cross-sectional area is reduced.
FIG. 2.14 Roll passes for rolling an angle (L) shape. Shape rolling is a cold or hot forming
process for reducing as well as shaping the cross section of the metal stock by passing it
through a series of rotating sets of rolls with appropriately shaped grooves.
17
FIG. 2.15 Principles of ring rolling. Ring rolling is a process whereby a hollow circular blank
(cold or hot) is formed into a ring. A main roll presses on the outside diameter of the blank,
which is supported by a mandrel on the inside diameter. Shaped cross sections are obtained
by appropriate contouring of the mandrel and the roll. The height of the ring is controlled by
auxiliary rolls.
FIG. 2.16 Rotary tube piercing, a, Double-conical working rolls; b, guide roll; c, billet; d,
conical piercer point. A hollow is formed by peripherally rolling a cylindrical hot billet over a
conical piercer point. The billet is driven by a pair of cone-shape rolls, set askew to the
longitudinal axis of the billet. The frictional load between the rolls and the billet causes the
billet to rotate and forces it to advance longitudinally over the piercer point.
18
(a) (b)
FIG. 2.17 Shear forming from a plate, (a) Start, (b) partially or completely formed part.
Shear forming is a process for hot or cold seamless shaping of dished parts by the combined
forces of rotation and pressure. This process differs from spinning principally in that it
reduces the thickness of the formed part.
FIG. 2.18 Tooling and metal flow for direct and indirect extrusion process. The product of
desired cross section is obtained by forcing a heated billet through a die without lubricating
the billet, the container, or the die. In the direct extrusion process, the product is extruded in
the direction of ram movement. When the product is extruded in a direction opposite that of
ram travel, the process is called indirect extrusion.
19
FIG. 2.19 Hot extrusion setup using glass lubrication. The heated billet is forced through a
die, using some form of lubrication, to obtain a product of desired cross section. Glass is the
most widely used lubricant for extruding long lengths from steels and high-temperaure alloys
on a production basis.
(a) (b)
FIG. 2.20. Drawing of (a) rod or wire and (b) tube. Drawing is the process of reducing the
cross-sectional area and/or the shape of a rod, bar, tube or wire (cold or hot) by pulling
through a die.
20
FIG. 2.21 Schematic of ironing. Ironing is the process of smoothing or thinning the wall of a
shell or cup (cold or hot) by forcing the shell through a die with a punch.
FIG. 2.22 Tube sinking process. Tube sinking is the process of sizing the outside diameter of
a tube by drawing the tube (cold or hot) through a die without supporting the tube internally
with a mandrel.
A B C D
FIG. 2.23 Typical brake-bending operations. (A) Air bending; (B) air rounding, (C) die
bending; (D) die rounding. Brake bending is a forming operation widely used for forming flat
sheets into linear sections, such as angles, channels and hats. There are two typical
brake-forming setups: air bending and die bending. In air bending, the workpiece is
supported only at the outer edges so that the length of the ram stroke determines the bend
angle of the part. In die bending, the sheet is forced into a female die cavity of the required
part angle.
21
Workpiece
Bending roll
Driven rolls
Support rolls
FIG. 2.24 Roll bending (three-roll forming). Roll bending gives a curvature to a sheet, bar,
or shaped section by bending it between two or three cylindrical rolls that can be adjusted.
FIG. 2.25 Roll forming. This process is used to produce long components of various cross
sections. The sheet metal is formed by passing it through a succession of progressively shaped
power-driven contoured rolls.
22
1C) Id)
FIG. 2.26 Various spinning operations, (a) Hollow shape forming; (b) bulging, (c) reducing,
(d) threading. Spinning is the process of shaping seamless dished parts by the combined
forces of rotation and pressure. Spinning does not result in any change in thickness.
FIG. 2.27 Deep drawing. (A) First draw; (B) redraw; (C) reverse draw. In deep drawing, a
sheet blank (hot or cold), usually subjected to a peripheral hold-down pressure, is forced by a
punch into and through a die to form a deep recessed part having a wall thickness
substantially the same as that of the blank. This process is used to produce cylindrical or
prismatic cups with or without a flange on the open end. Cups or tubes can be sunk or
redrawn to increase their length and to reduce their lateral dimensions.
23
24 Metal Forming and the Finite-Element Method
FIG. 2.28 Rubber-pad forming, (a) With rubber punch; (b) with rubber pad. This is a
forming operation for producing shallow parts. A rubber pad is attached to the press slide
and becomes the mating die for a punch, or group of punches, which has been placed on the
press bed or plate. The rubber pad is confined in a container (pad holder), and the entire
slide with attached pad holder is forced against the tools, usually by hydraulic pressure. As
the slide descends, the pliable but virtually incompressible rubber fills the space between the
slide and the dies and forces the metal to take the exact contours of the dies.
FIG. 2.29. Rubber-diaphragm hydroforming. In this process, the blank is held between a
diaphragm, which closes the ram pressure chamber, and a blank holder. A male punch works
against the diaphragm, and the metal is shaped by balancing the pressure of the ram chamber
against the pressure of the press base chamber on which the punch is mounted.
References
1. Altan, T., Lahoti, G. D., and Nagpal, V., (1981), "Systems Approach in
Massive Forming and Application to Modeling of Forging Processes," /. Appl.
Metal Working, ASM, Vol. 1, No. 2, p. 29.
2. American Society for Metals, (1961), "Metals Handbook", Eighth Edition,
Vol. 1 (Properties and Selection of Metals) and Vol. 4 (Forming), American
society for Metals, Metals Park, OH.
3. Altan, T., Oh, S. I., and Gegel, H., (1983), "Metal Forming: Fundamentals
and Applications," ASM International, Metals Parks, OH.
4. International Institution for Production Engineering Research, (1962), Dic-
tionary of Production Engineering, Vol. 1 (Forming and Drop Forging), Vol. 3
Metal-Forming Processes 25
(Sheet Metal Forming), and Vol. 5 (Cold Extrusion and Upsetting), Verlag,
W. Girardet, Essen.
5. Lahoti, G. D., and Altan, T., (1976), "Design of Dies for Radial Forging of
Rods and Tube," Technical Paper MF76-390, Society of Manufacturing
Engineers, Dearborn, MI.
6. Aluminium, American Society for Metals, (1967), Vol. 3 (Fabrication and
Finishing), edited by K. R. Van Horn, American Society for Metals, Metals
Park, OH, p. 81.
7. Geleji, A., (1967), "Forge Equipment Rolling Mills and Accessories,"
Akademiai Kiado, Budapest.
8. Beyon, R. E., (1956), "Roll Design and Mill Layout," Association of Iron and
Steel Engineers, Pittsburgh.
9. Sachs, G., (1951), "Principles and Methods of Sheet-Metal Fabricating,"
Reinhold, New York.
10. Lange, K., (1972), "Lehrbuch der Umformtechnik/Textbook of Forming
Technology" (in German), Springer-Verlag, Berlin.
3
ANALYSIS AND TECHNOLOGY
IN METAL FORMING
3.1 Introduction
The design, control, and optimization of forming processes require (1)
analytical knowledge regarding metal flow, stresses, and heat transfer, as
well as (2) technological information related to lubrication, heating and
cooling techniques, material handling, die design and manufacture, and
forming equipment.
The purpose of using analysis in metal forming is to investigate the
mechanics of plastic deformation processes, with the following major
objectives.
• Establishing the kinematic relationships (shape, velocities, strain-
rates, and strains) between the undeformed part (billet, blank, or
preform) and the deformed part (product); i.e., predicting metal flow
during the forming operation. This objective includes the prediction of
temperatures and heat transfer, since these variables greatly influence
local metal-flow conditions.
• Establishing the limits of formability or producibility; i.e., determin-
ing whether it is possible to perform the forming operation without
causing any surface or internal defects (cracks or folds) in the
deforming material.
• Predicting the stresses, the forces, and the energy necessary to carry
out the forming operation. This information is necessary for tool
design and for selecting the appropriate equipment, with adequate
force and energy capabilities, to perform the forming operation.
Thus, the mechanics of deformation provides the means for determining
how the metal flows, how the desired geometry can be obtained by plastic
deformation, and what the expected mechanical properties of the pro-
duced part are.
For understanding the variables of a metal-forming process, it is best to
consider the process as a system, as illustrated in Fig. 2.1 in Chap. 2. The
interaction of most significant variables in metal forming are shown, in a
simplified manner, in Fig. 3.1. It is seen that for a given billet or blank
material and part geometry, the speed of deformation influences strain-
rate and flow stress. Deformation speed, part geometry, and die tempera-
ture influence the temperature distribution in the formed part. Finally,
26
Analysis and Technology in Metal Forming 27
FIG. 3.1 Simplified illustration of the interactions between major process variables in metal
forming
flow stress, friction, and part geometry determine metal flow, forming
load, and forming energy.
In steady-state flow (kinematically), the velocity field remains un-
changed, as is the case in the extrusion process (Fig. 3.2B); in nonsteady-
state flow, the velocity field changes continuously with time, as is the case
in upset forging (Fig. 3.2A) [1].
(A) (B)
FIG. 3.2 Metal flow in certain forming processes. (A) Non-steady state upset forging; (B)
steady-state extrusion [1].
28 Metal Forming and the Finite-Element Method
FIG. 3.3 Flow stress vs. strain, and strain-rate vs. strain, for type 403 stainless steel at 1800,
1950 and 2050°F (tests were conducted in a mechanical press where e was not constant) [4].
glass for temperatures up to 2300T. The fixture and the specimens are
heated to test temperature and then the test is initiated. Examples of
high-temperature uniaxial flow stress data are given in Figs. 3.3 and 3.4.
FIG. 3.4 Flow stress vs. strain, and strain-rate vs. strain, for Waspaloy at 1950, 2050 and
2100°F (tests were conducted in a mechanical press where e was not constant) [4].
k being the shear strength of the deforming material, where 0 < ra < 1.
Studies in forming mechanics indicate that eq. (3.3) adequately repre-
sents the friction condition in bulk forming processes while eq. (3.2) is
commonly used for representation of friction in sheet-metal forming. A
reason for this is that the compressive normal stress at the interface in
sheet-metal forming is much smaller in magnitude, in comparison with that
in bulk deformation processes. For various forming conditions, the values
of m vary as follows:
• m = 0.05-0.15 in cold forming of steels, aluminium alloys, and
copper, using conventional phosphate-soap lubricants or oils
• m = 0.2-0.4 in hot forming of steels, copper, and aluminum alloys
with graphite-based (graphite-water or graphite-oil) lubricants
• m= 0.1-0.3 in hot forming of titanium and high-temperature alloys
with glass lubricants
• m = 0.7-1.0 when no lubricant is used, e.g., in hot rolling of plates or
slabs and in nonlubricated extrusion of aluminium alloys
In determining the friction factor m for hot forming, in addition to
lubrication effects, the effects of die chilling or heat transfer from the hot
material to colder dies must be considered. Therefore, the lubrication tests
used for determining friction factors must include both lubrication and
die-chilling effects. Consequently, in hot forming, a good test must satisfy
as well as possible the following requirements.
• The specimen and die temperatures must be approximately the same
as those encountered in the actual hot forming operation.
• The contact time between specimen and tools under pressure must be
approximately the same as in the forming operation of interest.
• The ratio of the newly generated deformed surface area to original
surface area of the undeformed specimen must be approximately the
same as in the process investigated.
• The relative velocity between deforming metal and dies should have
approximately the same magnitude and direction as in the forming
process.
Lubricity, as defined by the friction factor m, is most commonly
Analysis and Technology in Metal Forming 33
measured by using the ring test. In the ring test, a flat ring-shaped
specimen is compressed to a known reduction. The change in internal and
external diameters of the forged ring is very much dependent on the
friction at the die-workpiece interface. If friction were zero, the ring
would deform in the same way as a solid disk, with each element flowing
radially outward at a rate proportional to its distance from the center.
With increasing deformation, the internal diameter of the ring is reduced if
friction is high, and is increased if friction is low. Thus, the change in the
internal diameter represents a simple method for evaluating interface
friction [5,6].
H Q -H. INCH
DISPLACEMENT. H Q -H, MM
FIG. 3.5 Load vs. displacement curves obtained in closed-die forging of an axisymmetric
steel part (dimensions in inches) at 2012°F in three different machines with different initial
velocities, Vpi [5].
excessive chilling of the surface layers of the formed part near the
die-material interface.
FIG. 3.6 Outline of a CAD/CAM procedure for forging die design and manufacture.
FIG. 3.7 Metal flow and load-stroke curve in impression-die forging: (A) upsetting, (B)
filling, (C) end, (D) load-stroke curve.
FIG. 3.9 Schematic illustration of a flat-face die for extrusion of "T" sections [7].
Analysis and Technology in Metal Forming 39
• Determination of die bearing lengths for balancing metal flow to avoid
the twisting and bending of the extrusion emerging from each orifice
• Determination of the extrusion load: quite difficult since the extruded
sections are usually nonsymmetrical, resulting in a complex 3D metal
flow in the deformation zone
Some semiempirical analytical techniques have been developed for
computer-aided design of extrusion dies. However, these techniques are
approximate and could be vastly improved by developing FEM-based
computer programs for metal-flow simulation.
FIG. 3.10 Schematic illustration of forming sequences in cold forging of a gear blank [8]. Left to right: sheared blank, simultaneous forward rod and
backward cup extrusion, forward extrusion, backward cup extrusion, simultaneous upsetting of flange, and coining of shoulder.
Analysis and Technology in Metal Forming 41
FIG. 3.11 Four types of metal flow in rolling: (A) strip, (B) plate, (C) simple shape, (D)
complex shape (broken and solid lines illustrate the sections before and after deformation,
respectively).
FIG. 3.12 Schematic representation of strip rolling (strip has constant width).
FIG. 3.13 Schematic illustration of five different roll-pass designs for a steel angle section
[18].
nature, and several empirical formulae have been established for estimat-
ing spread [15]. Attempts were also made to predict elongation and spread
theoretically [16, 17].
Rolling of shapes, also called caliber rolling, is one of the most complex
deformation processes (Figs. 3.11C and D). A round or round-cornered
square slab is rolled in several passes into (a) relatively simple sections
such as rounds, squares, or rectangles, or (b) complex sections such as
L,U,T,H, or other irregular shapes. For this purpose, certain intermediate
shapes or passes are used, as shown in Fig. 3.13 for rolling of angle
sections [18]. The design of these intermediate shapes, i.e. roll pass design,
is based on experience and differs from one company to another, even for
the same final rolled section geometry. Relatively few quantitative data on
roll pass design are available in the literature.
At the present stage of technology, the process variables are considered
in roll pass design by using a combination of empirical knowledge, some
calculations, and some educated guesses. A methodical way of designing
roll passes requires not only an estimate of the average elongation, but
also the variation of this elongation, within the deformation zone. The
deformation zone is limited by the entrance, where a prerolled shape
enters the rolls, and by the exit where the rolled shape leaves the rolls.
The deformation zone is cross-sectioned with several planes. The roll
position and the deformation of the incoming billet are investigated at each
of these planes. Thus, a more detailed analysis of metal flow and an
improved method for designing the configuration of the rolls are possible.
It is evident that this technique can be drastically improved and made
extremely efficient by use of computer-aided techniques and, ultimately,
by FEM.
In recent years, most companies that produce shapes have computerized
their roll pass design procedures for rolling rounds [19,20] or structural
shapes [21, 22]. In most of these applications, the elongation per pass and
the distribution of the elongation within the deformation zone for each
pass are predicted by using an empirical formula. If the elongation per pass
Analysis and Technology in Metal Forming 45
FIG. 3.14 Cold rolled (round to triangle) and cold drawn shape, requiring numerous
annealing steps [23].
depend on the flow stress of the drawn part, die geometry, and interface
friction.
Drawing with a fixed plug is widely known and used for drawing of
large-to-medium-diameter straight tubes. The plug, when pushed into the
deformation zone, is pulled forward by the frictional force created by the
sliding movement of the deforming tube. Thus, it is neccesary to hold the
plug in the correct position with a plug bar. In drawing of long and
small-diameter tubes, the plug bar may stretch and even break. In such
cases it is advantageous to use a floating plug. This process can be used to
draw any length of tubing by coiling the drawn tube at high speeds of up to
2000 ft/min. In drawing with a moving mandrel, the mandrel travels at the
speed at which the section exits from the die. This process, also called
ironing, is widely used for thinning of the walls of drawn cups or shells, for
example, in the production of beverage cans [24] or artillery shells [5].
The principle of a can ironing press is illustrated in Fig. 3.15. In the
figure, the press is horizontal, and the ram has a relatively long stroke and
is guided by the hydrostatic bushing (A). The front seal (B) prevents
mixing of the ironing lubricant with the hydrostatic bushing oil. With the
ram in the retracted position, the drawn cup is automatically fed into the
press, between the redraw die (D) and the redraw sleeve (C). The redraw
sleeve centers the cup for drawing and applies controlled pressure while
FIG. 3.15 Schematic illustration of multiple-die ironing operation for manufacturing bever-
age cans [24].
Analysis and Technology in Metal Forming 47
the cup is drawn through the first die (D). As the ram proceeds, the redrawn
cup is ironed by passing through the carbide dies (E), which gradually
reduce the wall thickness. The ironed can is pressed against the doming
punch (T), which forms the bottom shape of the can. When the ram starts
its return motion, the mechanical stripper (G), assisted by the air stripper
(F), removes the can from the ironing punch (H). The punch is made of
carbide or cold-forging tool steel. The stripped can is automatically
transported to the next machine for trimming of the top edge of the can
wall to a uniform height.
WRINKLES
FRACTURE
EARING
FIG. 3.16 Some defects observed in cup drawing [28].
FIG. 3.17 Forming limit diagram and deformed rectangular strip and circular blank with
circular cutoff [34].
49
50 Metal Forming and the Finite-Element Method
that will produce failures in an actual forming operation can be drawn.
Figure 3.17 shows the forming limit diagram for different materials.
Experimental methods are used to construct the diagram. The Nakajima
method [32] uses a hemispherical punch and rectangular blanks with
various widths and lubrication conditions. Analogously to the Nakajima
method, Hasek [33] used hemispherical punch stretching of circular blanks
with various circular cutoffs. The method of Hasek has an advantage of
eliminating or reducing the wrinkling phenomena that occur at the edges
of rectangular strips in the Nakajima method. In Fig. 3.17, deformed
specimens (FEM simulation [34]) in the above two methods are also
shown.
Friction at the punch-workpiece interface is an important factor for
formability, and it changes the strain-path of a critical site of the sheet in
the forming limit diagram of Fig. 3.17. Material parameters, such as
strain-hardening and strain-rate sensitivity, are important for formability.
Another important factor in sheet-metal forming is the anisotropy
(direction-dependent properties) of the sheet metal. Earing in deep
drawing of cups is due to anisotropy (see Fig. 3.16). Anisotropy also
influences formability. For example, the limiting drawing ratio in cup
drawing increases with increasing r-value, which is the ratio of the width
strain to the thickness strain in uniaxial tension and is a measure of
anisotropy in the thickness direction [26].
In sheet-metal forming, considerations of residual stresses and spring-
back are particularly important. Consequently elastic properties of sheet
metals cannot be neglected in the analysis of problems in which these
considerations are of major concern.
Although simple analytical technique considers only the stress com-
ponents in the plane of the sheet (plane-stress situation), complexity of
product shapes and deformation in forming make it difficult for research in
analytical methods to advance technology that would be useful in
metal-stamping plants [35].
Among many areas of research need, die design and process design are
of particular technological importance and two specific problems, namely,
design of drawbeads in stamping operations and design of multistage
forming, are given as examples.
In a die-formed part the presence of sloping walls that are not confined
between the die and the punch introduces the danger of buckling or
wrinkling. Generally, this tendency to wrinkle increases with increasing
unsupported area and with decreasing metal thickness. To avoid wrinkling,
large parts made from thin metal must be formed with a considerably
higher hold-down force. In addition, beads are frequently added to the
mating die and hold-down surfaces, either all around or only in the areas
where wrinkling would otherwise develop. A drawbead is shown in Fig.
3.18 in the drawing of a box-shaped part [25]. Beads retard and control the
flow of sheet metal into the die cavity. Design of drawbeads includes the
determination of their orientation, configuration, and location.
Analysis and Technology in Metal Forming 51
In most sheet-metal parts, the shape is formed almost entirely in the first
operation. Subsequent processes are generally trimming, piercing, flang-
ing, and some minor restriking of detail. There are instances, however, in
which the formability of the metal is such that it is impossible to reach the
final form in one operation, and a number of intermediate shapes are
required. An example is shown in Fig. 3.19 for forming an automobile
FIG. 3.19 Multistage forming processes for an automobile wheel center [36],
52 Metal Forming and the Finite-Element Method
References
1. Lange, K., (editor), (1972), "Study Book of Forming Technology," (in
German), Vol. I, II and III, Springer-Verlag, Berlin.
2. Johnson, W., and Sowerby, R., (1978), "Metal Forming Processes: Analysis
and Technology," ASME, AMD, Vol. 28, p. 1.
3. Allan, T., and Boulger, F. W., (1973), "Flow Stress of Metals and Its
Application in Metal Forming Analyses," Trans. ASME, J. Engr. Ind., Vol.
95, p. 1009.
4. Douglas, J. R., and Allan, T., (1975), "Flow Stress Determination for Melals
al Forging Rates and Temperalures," Trans. ASME, J. Engr. Ind., Vol. 97, p.
66.
5. Allan, T., Oh., S. I, and Gegel, H., (1983), "Melal Forming: Fundamenlals
and Applications," ASM International, Metals Park, OH.
6. Schey, J. A. (editor), (1983), "Metal Deformation Processes: Friction and
Lubrication," Marcel Dekker, New York, 1970; superseded by Schey, J. A.,
"Tribology in Metalworking: Lubrication, Friction and Wear," American
Sociely for Melals, Melals Park, OH.
7. Billhardt, C. F., Nagpal, V., and Allan, T., (1978), "A Computer Graphics
System for CAD/CAM of Aluminum Exlrusion Dies," SME Paper MS78-957.
8. Sagemuller, F., (1968), "Cold Impact Exlrusion of Large Formed Parts,"
Wire, No. 95, p. 2.
9. Dieter, G. E., (1961), "Mechanical Metallurgy," McGraw-Hill, New York,
Chapter 29, p. 488.
10. Geleji, A., (1967), "Forge Equipmenl, Rolling Mills and Accessories,"
Akademiai Kiado, Budapesl.
11. Larke, E. C., (1957), "The Rolling of Strip, Sheet and Plate," Chapman and
Hall, London.
12. Orowan, E., (1943), "The Calculation of Roll Pressure in Hoi and Cold Flal
Rolling," Proc. Inst. Mech. Eng., Vol. 150, p. 140.
13. Alexander, J. M., (1972), "On Ihe Theory of Rolling," Proc. R. Soc. London,
Series A, Vol. 326, p. 535.
14. Laholi, G. D., Shah, S. N., and Allan, T., (1978), "Computer Aided Analysis
of the Deformations and Temperalures in Strip Rolling," Trans. ASME, J.
Engr. Ind., Vol. 100, p. 159.
15. Ekelund, S., in H. Neumann, (1963), "Roll Pass Design," (in German), VEB
Deulscher Verlag, Leibzig, p. 48.
16. Oh, S. I., and Kobayashi, S., (1975), "An Approximate Melhod for
Three-Dimensional Analysis of Rolling," Int. J. Mech. Sci., Vol. 17, p. 293.
17. Neumann, H., (1969), "Design of Rolls in Shape Rolling," (in German), VEB
Deulscher Verlag, Leibzig.
18. Schulza, A., (1970), "Comparison of Roll Pass Designs Used for Rolling
Angle Sections," (in German), Stahl and Eisen, Vol. 90, p. 796.
19. Neumann, H., and Schulze, R., (1974), "Programmed Roll Pass Design for
Blocks," (in German), Neue Hutte, Vol. 19, p. 460.
20. Suppo, U., Izzo, A., and Diana, P., (1973), "Eleclronic Computer Used in
Roll Design Work for Rounds," Der Kalibreur, Vol. 19, p. 3.
Analysis and Technology in Metal Forming 53
4.1 Introduction
The theory of plasticity describes the mechanics of deformation in
plastically deforming solids, and, as applied to metals and alloys, it is
based on experimental studies of the relations between stresses and strains
under simple loading conditions. The theory described here assumes the
ideal plastic body for which the Bauschinger effect and size effects are
neglected. The theory also is valid only at temperatures for which
recovery, creep, and thermal phenomena can be neglected. The basic
theory of classical plasticity is described by Hill [1], and also in References
[2-5], in addition to the books listed in Chap. 1. A concise description of
the general plasticity theory necessary for metal forming is given in the
book by Johnson et al. [6]. In this chapter, certain important aspects of the
theory are presented in order to elucidate the developments of the
finite-element solutions of metal-forming problems discussed in this book.
First, various measures of stress and strain are introduced. Then, the
governing equations for plastic deformation and principles that are the
foundations for the analysis are described. The extension of the theory of
plasticity to time-dependent theory of viscoplasticity is outlined in Section
4.8. Particular references are made, in Sections 4.3 through 4.7, to the
books by Hill [1] and by Johnson and Mellor [7], and to the section on
general plasticity theory in the book by Johnson et al. [6].
FIG. 4.1 Uniaxial tension, (a) Tension specimen; (b) stress-strain curves.
axial direction by the force P to the length / and the cross-sectional area A
at time t, as shown in Fig. 4.1. The response of the material is recorded as
the load-displacement curve, and converted to the stress-strain curve as
shown in the figure. The deformation is assumed to be homogeneous until
necking begins.
There are two modes of description of the deformation of a continuous
medium, the Lagrangian and the Eulerian. The Lagrangian description
employs the coordinates Xt of a typical particle in the reference (or
undeformed) state as the independent variables, while in the case of
Eulerian description the independent variables are the coordinates xt of a
material point in the deformed state. When the deformation is in-
finitesimal, where products of the derivatives of the displacements can be
neglected, we need make no distinction between the two.
In the infinitesimal deformation theory, the stresses and strain-rates (or
infinitesimal strains) are expressed with respect to a fixed coordinate
system in the material configuration at a time under consideration. In
56 Metal Forming and the Finite-Element Method
uniaxial tension, they are defined by
stress
straain-rate
infinitesimal strain ds
where the dot denotes the time derivative. The stress defined in eq. (4.1) is
called true stress or Cauchy stress. The total amount of deformation is
measured by integrating infinitesimal strain as
and
Plasticity and Viscoplasticity 57
since
where ut are velocity components and y,7 are engineering shear strain-rate
components. Using suffix notation, eq. (4.11) can be written as
where Ilt I2, and 73 are quantities independent of the direction of the axes
chosen and called the three invariants of the stress tensor oti. They are
defined by the relations
The first (linear) and second (quadratic) invariants have particular physical
significance for the theory of plasticity.
From the experimental fact that the yielding of a material is, to a first
approximation, unaffected by a moderate hydrostatic pressure, or tension,
it follows that yielding depends only on the principal components
Plasticity and Viscoplasticity 59
where
Two Criteria
Two simple criteria have been in extensive use for the analysis of metal
deformation. Tresca's criterion (shear stress criterion) given in 1864 is
with o1 >CT2> o3. It may alternatively be written in the form of eq. (4.19)
in terms of J2 and J3, but the results are complicated and not useful.
Another criterion was proposed by Heuber (1904), by von Mises (1913),
and by J. C. Maxwell in a letter to Kelvin in 1856 [7]. It has been
traditionally called the von Mises criterion or the Huber-Mises criterion,
but may be appropriately called the Maxwell-Heuber-Mises criterion.
(We refer to the distortion energy criterion according to a physical
interpretation of the criterion suggested by Hencky in 1924.) The criterion
states that yielding occurs when J2 reaches a critical value, or, in other
words, that the yield function/of eq. (4.19) does not involve J3. It can be
written in the alternative forms
or
or
FIG. 4.2 Geometrical representation of a plastic state of stress in (tr1; a2, a3) space.
Plasticity and Viscoplasticity 61
FIG. 4.3 Yield locii on the Jt-plane for distortion energy criterion and maximum shear stress
criterion.
(4.22), the two criteria can be made to agree with each other and with
experiment for uniaxial tension. The two loci are shown in Fig. 4.3.
For most metals the distortion energy criterion fits the data more closely
than the shear stress criterion. Furthermore, the distortion energy criterion
is continuous and convenient to use in numerical analysis. Therefore, the
distortion energy criterion is exclusively used in this book. Other yield
criteria, such as those for anisotropic sheet materials and for porous
metals, are introduced in Chaps. 11 and 13, respectively.
In the notation used in eq. (4.24), a recurring letter suffix indicates the
sum, as explained for eq. (4.21a), and a comma denotes partial
differentiation as explained for eq. (4.12).
62 Metal Forming and the Finite-Element Method
Equilibrium with Tractions
The stress along the boundary surface S is in equilibrium with an applied
traction Ft (force per unit surface area). Equilibrium of the stresses is
written as
where «, is the unit outward normal to the surface. Writing eq. (4.25a), in
unabridged notation in the two-dimensional case (see Fig. 4.4), one has
where V is the volume of the body and S is the surface. Since otj is
where e,y is the strain-rate derivable from vv, according to eiy = 2(wy + wjti).
For proof, note that
and
or
The plastic strain-rate can be represented in the same (al, o2, o3) space by a
free vector G(e1p, e2p, £3P), where the factor G is introduced to obtain the
dimension of stress. This vector lies in the jr-plane since k-f + s2p + £3P =
0 (no volume change). Because df/do^ are the direction cosines of the
outward normal to the yield surface /(%) = C, the plastic strain-rate
vector is normal to the yield locus at point (cr/, o2'', o3) in the n plane.
The plastic strain-rate vector associated with the yield locus of the
distortion energy criterion is shown by PQ in Fig. 4.5. The concept of
plastic potential and its associated flow rule also apply to the shear stress
criterion. It will not be discussed here, however, since the shear stress
criterion is not used in this book, as stated in Section 4.3. It should be
noted that the yield locus must be convex (concave to the origin) at all
points if the stress corresponding to a given strain-rate is to be unique.
Maximum Plastic Work Principle
The maximum plastic work principle follows from the flow rule and the
convexity of the yield locus, and is
where otj is a yield state of stress, ei/> is the associated strain-rate, and a,,*
is any other stress state represented by a point either on or inside the yield
surrace.
Since £,yp is parallel to the outward normal to the yield locus at the point
Oi/, (0i/ ~ Oij*)£ijp is proportional to the scalar product of the outward
normal to the yield locus at the point a// with the chord joining a^*' to o^'
(see Fig. 4.6). Therefore, eq. (4.29) holds and the equality sign applies
when <T,/ = ai*' or cr,-, and ot* differ only by a uniform hydrostatic stress.
Drucker [11] arrived at eq. (4.29) from a definition of a stable plastic
material.
The Prandtl-Reuss Equation and the Levy-Mises Equation
With the yield function given by f{otj) =J2 = 2<V<V> (eq. 4.21a), eq.
(4.28) becomes, in the rate form,
since
noting that the repeated subscripts k and / indicate summation with respect
to these quantities.
Equation (4.30) can be written as
where a is written for Y or \/3 k, and known as the flow stress, the
effective stress, generalized stress, or equivalent stress.
Another hypothesis for strain-hardening relates a to a certain measure
of the total plastic deformation. A quantity de, known as the effective,
generalized, or equivalent infinitesimal plastic strain is defined according to
(4.38)
The derivation of eq. (4.38) is as follows. The plastic work-rate per unit
volume, using the flow rule (4.28), is written as
from which eq. (4.38) results for the yield function /(a,y) of eq. (4.21a),
and a denned by eq. (4.35). The expression for e in terms of strain-rate
components, corresponding to the effective stress o of eq. (4.35), can be
obtained by considering the inversion of the flow rule. Then
from which
Then
Equation (4.42) is the first extremum principle and it states that a lower
bound to the rate of work of the actual surface traction can be obtained
from a statically admissible stress field.
In the quasi-static flow of a rigid plastic solid, velocity discontinuity is
permissible. When the velocity field u{ contains surfaces of discontinuity,
the principle of virtual work-rate gives, instead of eq. (4.40),
Plasticity and Viscoplasticity 69
where i is the shear stress component of the stress a{j of the complete
solution along the velocity discontinuity surface 5D* with the amount of
tangential discontinuity |Aw*| in the velocity field «,*. If u* is the actual
70 Metal Forming and the Finite-Element Method
and also
from the maximum plastic work principle, where o^* satisfies the yield
criterion and is associated with Et*. Therefore, we have
4.8 Viscoplasticity
The mathematical theory of plasticitiy adequately describes the time-
independent aspect of the behavior of materials but is inadequate for
analysis of time-dependent behavior. An approach to achieving a satisfac-
tory formulation for time-dependent behavior has been to generalize
plasticity to cases within the strain-rate-sensitive range. One such generali-
zation has been provided by the theory of viscoplasticity. The history of
this theory goes back to as early as 1922 [14]. Since then, generalizations of
early versions have been achieved and various forms of the theory of
viscoplasticity have been provided (summarized, for example, by Perzyna
[15] and by Cristescu [16]). In this section, we describe an approach to the
construction of equations for rigid viscoplastic materials. Then, the
extremum principle is presented as the basis for the finite-element
formulation of viscoplastic flow analysis.
Constitutive Equations
What follows is based on work by Perzyna. We consider a rigid viscoplastic
material and apply the infinitesimal theory for each incremental deforma-
tion. Perzyna introduced a function F(<7y) such that
where k is the static yield stress in shear. Considering F(cTy) as the function
similar to the plastic potential in the theory of plasticity, the constitutive
Plasticity and Viscoplasticity 71
equation is expressed as
Then
from eq. (4.50). Equation (4.52) is a familiar rate-dependence law and the
exponent m is the strain-rate sensitivity index.
Extremum Principle
Hill [17] derived the (second) extremum principle associated with the
deformation process of the rigid-viscoplastic materials. Among all the
possible constitutive equations, we will restrict our attention to the case
where there exists the work function £(£</), sucn that
72 Metal Forming and the Finite-Element Method
where e^, «,- are actual quantities and the starred quantites are kinemati-
cally admissible ones. It can be shown that the solution is uniquely
determined at points of the body where E is strictly convex, but not
necessarily in all respects elsewhere.
References
1. Hill, R., (1950), "The Mathematical Theory of Plasticity," Oxford University
Press, London.
2. Mendelson, A., (1968), "Plasticity: Theory and Application," MacMillan, New
York.
3. Szczepinski, W., (1979), "Introduction to the Mechanic of Plastic Forming of
Metals," Sijthoff & Noordhoff International Publishers, The Netherlands.
4. Yamada, Y., (1965), "Mechanics of Plastic Deformation," Nikkan Kogio
Shinbun-Sha, Tokyo.
5. Chakrabarty, J., (1987), "Theory of Plasticity," McGraw-Hill, New York.
6. Johnson, W., Sowerby, R., and Venter, R. D., (1982), "Plane Strain Slip Line
Fields for Metal Deformation Processes," Pergamon Press, Oxford, UK.
7. Johnson, W., and Mellor, P. B., (1973), "Engineering Plasticity," Van
Nostrand Reinhold, London.
8. Sokolnikoff, I. S., (1956), "Mathematical Theory of Elasticity," McGraw-Hill,
New York.
9. Naghdi, P. M., and Trapp, J. A., (1975), "The Significance of Formulating
Plasticity Theory with Reference to Loading Surfaces in Strain Space," Int. J.
Engg. Sci., Vol. 13, p. 785.
10. Green, A. E., and Naghdi, P. M., (1965), "A General Theory of an
Elastic-Plastic Continuum," Arch. Rational Mech. Anal., Vol. 18, p. 251.
11. Drucker, D. C., (1951), "A More Fundamental Approach to Plastic Stress
Strain Relations," Proc. 1st U.S. Natl. Congress, Appl. Mech. ASME, p. 487.
12. Drucker, D. C., Greenburg, H. J., and Prager, W., (1951), "The Safety
Factor of an Elastic-Plastic Body in Plane Stress," Trans. ASME., J. Appl.
Mech., Vol. 18, p. 371.
13. Drucker, D. C., Prager, W., and Greenburg, H. J., (1952), "Extended Limit
Design Theorems for Continuous Media," Q. Appl. Math., Vol. 9, p. 381.
14. Bingham, E. C., (1922), "Fluidity and Plasticity," 1st edition, McGraw Hill,
New York, p. 125.
15. Perzyna, P., (1966), "Fundamental Problems in Viscoplasticity," Adv. App.
Mech., Vol. 9, p. 243.
16. Cristescu, N., (1967), "Dynamic Plasticity," North-Holland, Amsterdam.
17. Hill, R., (1956), "New Horizons in the Mechanics of Solids," J. Mech. Phys.
Solids, Vol. 5, p. 66.
5
METHODS OF ANALYSIS
5.1 Introduction
In Chap. 4 we described fundamentals for the analysis of metal-forming
processes and derived some useful principles. The governing equations for
the solution of the mechanics of plastic deformation of rigid-plastic and
rigid-viscoplastic materials are summarized as follows:
Equilbrium equations:
Yield critertion
Constitutive equations:
with
Compatiblity conditionms:
chapter, the two methods are briefly outlined with examples. The basic
formulations for the finite-element method are given in Section 5.4.
Discretization of these basic formulations and numerical solution tech-
niques are discussed in Chaps. 6 and 7.
FIG. 5.1 (a) Compression of a solid cylinder between flat rough dies, (b) An admissible
velocity field.
76 Metal Forming and the Finite-Element Method
velocity components in each region are assumed to be
in Region 1
and
in Region 2
In eqs. (5.3) and (5.4), a = 2H/D with dimensions as shown in Fig. 5.la.
It is to be noted that the line of discontinuity becomes straight for n = 0
and the velocity field becomes continuous throughout the body for n = 1.
An upper bound to the rate of total forming energy E is obtained by
calculating the energy rate of deformation of a continuously deforming
body and the energy-rates due to shearing along the surface of velocity
discontinuity, based on an assumed velocity field. The upper bound to the
forming pressure, then, can be found from the equation
where
FIG. 5.2 Comparison of upper bounds to the average forming pressure for the compression
of a cylinder.
78 Metal Forming and the Finite-Element Method
value of n that makes pja minimum, and is compared with the pressures
for n = 0 and n = 1.
Remarks
The technique outlined above together with the increased use of compu-
ters led to the development of generalized upper-bound approaches to a
wider class of problems. Examples of such approaches are the analysis of
axisymmetric forging of complex shapes [10], estimation of lateral spread
in plate rolling [11] and in forging [12], and upper bounds to extrusion and
drawing through dies of various shapes [13-16].
In these examples, construction of admissible velocity fields becomes
more complicated and upper-bound calculations must often resort to
numerical means. A question arises, then, whether pursuing this approach
is worth the effort, because detailed flow is much more important than
forces. Kudo [17] surveyed the development and use of the upper-bound
method and concluded that the approach should continue to be useful in
teaching, in workshop trials, and in research on bulk forming processes.
Also, it will be noted in Section 5.5 that the finite-element method
emerges from the principle involved in the upper-bound method and from
the concept of discretization.
for all virtual velocity fields w,- that are continuous and continuously
differentiable. Introducing an approximate stress field that may not
necessarily satisfy equilibrium in eq. (5.6a) and applying the divergence
theorem, we obtain
where CT,-,M/ is the traction on S computed from the considered stress field.
If eq. (5.6b) holds for all virtual velocity wjt then dOfj/dXi = 0 in V and
a,y«, = FJ on S. Equations (5.6a) and (5.6b) thus are equivalent. Equation
(5.6b) asserts that a sufficient condition for the considered stress distribu-
tion to satisfy the required statical conditions is that the divergences should
have zero net work-rate in the stated class of virtual motions, or the
divergences should be "orthogonal" to the virtual motions.
The interpretation of this theorem leads to a criterion of approximation,
namely, the static conditions for the approximating field o^ can be
regarded as closely satisfied overall when eq. (5.6a) is satisfied for a
sufficient subclass of virtual orthogonalizing motion w,, with the traction Ff
prescribed on SF and computed from the approximating field otj on the
remainder of the surface.
For metal-forming processes, the surface 5 of the deforming zone
consists of three distinct parts, S = Sc + SF + Sj, where Sc adjoins a tool or
container, SF is unconstrained, and S/ is the junction with the rigid zone;
on surface SF, ordinarily the traction is zero; the frictional constraint over
surface Sc is represented by a constant frictional stress mk (where k is the
shear yield stress of the deforming body) or a Coulomb coefficient of
friction (i.
The selection criterion for the approximating field a,-,- is then expressed
bv
for constant
frictional stress
for Coulomb
friction
{lj} = (lr, le, 4} = {-1, 0, 0}, the criterion (5.7) can be written as
and
In deriving eq. (5.14), the yield condition in the form of a'2= — §Y,
where Y denotes the yield stress in uniaxial compression, is used as a first
order of approximation.
Evaluating eq. (5.14) at z=H and using eq. (5.13), we obtain the
expression for the compression load P as
and thus
Substitution of eq. (5.16) into eq. (5.14) with the help of eq. (5.15), results
in
where
Substituting eq. (5.18) into the approximate velocity field (5.8), the
incipient of bulge formation can be described by the radial velocity
component at r = D/2. Figure 5.3 shows the profile of the radial velocity
component along the free surface for H/D = 0.5 and several values of
friction factor m. In order to trace the development of bulge, calculation
must be repeated step by step, updating the current geometrical
configuration.
To illustrate the method, the formulation given in this example uses a
simpler approximating velocity field than that given by Hill.
Remarks
It appears that the class of problems to which the Hill method can be
applied effectively is one for which the determination of the geometrical
changes and load are the main objectives. This class of problems can be
grouped into three areas: (1) simple upsetting of rings and blocks, (2)
flat-tool forging and its allied problems, (3) steady-state processes with
unknown steady-state configurations. The method is well-based on mathe-
matically sound principles and is amply flexible in introducing various
simplifying assumptions into the analysis. Actual procedure using the
method usually involves extensive calculations and sometimes encounters
overwhelming difficulties in obtaining final solutions. However, as pointed
Methods of Analysis 83
FIG. 5.3 Computed profile of the radial velocity component along the free surface in
compression.
out in Section 5.5, the basis of the finite-element method is the discrete
representation of the principles and ideas given in Hill's method.
and
Note that eq. (5.24) is the expression of the virtual work-rate principle, if
dut is considered to be the virtual motion (see eq. (4.26) in Chap. 4).
Using the symmetry of the stress tensor and the divergence theorem, eq.
(5.24) results in
Methods of Analysis 85
For the surface integral term in eq. (5.25), the boundary conditions that
diij = 0 on Su (essential boundary condition) and o^n, = Ft on SF
(suppressible boundary condition), are imposed.
Substitution of a,y = a,/ + <5,yOm, where am = \okk, the hydrostatic stress,
into eq. (5.25) gives
where x is the coordinate, u is the field variable and u' is the derivative of
u with respect to x. Then
Boundary Conditions
The traction boundary condition on SF is either zero-traction or ordinarily
at most a uniform hydrostatic pressure. However, the boundary conditions
along the die-workpiece interface are mixed. Furthermore, in general,
neither velocity nor force (including magnitude and direction) can be
prescribed completely along this interface, because the direction of the
frictional stress is opposite to the direction of the relative velocity between
the deforming workpiece and the die, and this relative velocity is not
known a priori. Situations exist, e.g., in extrusion and drawing, in which
the direction of metal flow relative to the die is known. This class of
problems can be solved if the magnitude of the frictional stress fs is given
according to the well-known Coulomb law, fs = ftp, or the friction law of
constant factor m, expressed by fs = mk (where k = Y/VJ); here, p is the
die pressure and k is the shear yield stress.
For problems such as ring compression, rolling, and forging, the
unknown direction of the relative velocity between the die-workpiece
interface makes it difficult to handle the boundary condition in a
straightforward manner. A unique feature of this type of problem is that
there exists a point (or a region) along the die-workpiece interface where
Methods of Analysis 87
the velocity of the deforming material relative to the die becomes zero (see
Section 3.8 of Chap. 3), and the location of this point (or region) depends
on the magnitude of the frictional stress itself. In order to deal with these
situations, a velocity-dependent frictional stress is used as an approxima-
tion to the condition of constant frictional stress. At the interface Sc the
velocity boundary condition is given in the direction normal to the
interface by the die velocity, and the traction boundary condition is
expressed by
where fs is the frictional stress, € is the unit vector in the opposite direction
of relative sliding, us is the sliding velocity of a material relative to the die
velocity (relative sliding velocity), and u0 is a small positive number
compared to us. The approximate expression (5.30) for a constant
frictional stress has been used for the smooth transition of the frictional
stress in the range near the neutral point [26].
Treatment of a Rigid Region
The principles described in the preceding sections apply to a domain in
which the entire body is deforming plastically. In metal-forming processes,
however, situations do arise in which rigid zones exist, and unloading
occurs. The rigid zones are characterized by a very small value of effective
strain-rate in comparison with that in the deforming body. If these portions
are included within the control volume V (the volume included in the
statement of boundary value problems), the value of the first term of the
basic equation (5.21) or (5.22) cannot be uniquely determined because of
the undefined value of the effective stress when the effective strain-rate
approaches zero.
This difficulty is removed by assuming that the stress-strain-rate
relationship in eq. (5.1c) is approximated by
where £0 takes an assigned limiting value, say 10~3 [27]. This presumed
stress-strain-rate relationship is equivalent to the assumption of a New-
tonian fluid-like material behavior for nearly-rigid regions. For these
regions, the first term of the basic equation, Svode dV, is then replaced
by
References
1. Kobayashi, S., and Thomson, E. G., (1965), "Upper and Lower Bound
Solutions to Axisymmetric Compression and Extrusion Problems," Int. J.
Mech. ScL, Vol. 7, p. 127.
2. Johnson, W., and Mellor, P. B., (1973), "Engineering Plasticity," Van
Nostrand Reinhold, London.
3. Johnson, W., and Kudo, H., (1962), "Mechanics of Metal Extrusion,"
Manchester University Press, Manchester, UK.
4. Johnson, W., and Kudo, H., (1960), "Use of Upper Bound Solutions. . .For
Axisymmetric Extrusion Processes," Int. J. Mech. Sci., Vol. 1, p. 175.
5. Kudo, H., (1960), "Some Analytical and Experimental Studies of Axisym-
metric Cold Forging and Extrusion—I," Int. J. Mech. Sci., Vol. 2, p. 102.
6. Kudo, H., (1961), "Some Analytical and Experimental Studies of Axisym-
metric Cold Forging—II," Int. J. Mech. Sci., Vol. 3, p. 91.
7. Avitzur, B., (1968), "Metal Forming: Processes and Analysis," McGraw-Hill,
New York.
8. Avitzur, B., (1983), "Handbook of Metal Forming Processes," Wiley, New
York.
9. Kobayashi, S., (1964), "Upper-Bound Solution to Axisymmetric Forming
Problems," Trans. ASME, J. Engr. Ind., Vol. 83, p. 326.
10. McDermott, R. P., and Bramley, A. N., (1974), "An Elemental Upper-Bound
Technique for General Use in Forging Analysis," Proc. 15th M.T.D.R. Conf.,
p. 437.
11. Oh, S. I., and Kobayashi, S., (1975), "An Approximate Method for a
Three-Dimensional Analysis of Rolling," Int. J. Mech. ScL, Vol. 17, p. 293.
12. Juneja, B. L., (1973), "Forging of Polygonal Discs with Barrelling," Int. J.
Machine Tool Des. Res., Vol. 13, p. 87.
Methods of Analysis 89
13. Juneja, B. L., and Prakash, R., (1975), "An Analysis for Drawing and
Extrusion of Polygonal Sections," Int. J. Machine Tool Des. Res., Vol. 15, p.
1.
14. Nagpal, N., and Allan, T., (1975), "Analysis of Three-Dimensional Metal
Flow in Extrusion of Shapes with the Use of Dual Stream Functions," Proc.
3rd NAMRC, Carnegie-Mellon University, p. 26.
15. Yang, D. Y., and Lee, C. H., (1978), "Analysis of Three-Dimensional
Extrusion of Sections Through Curved Dies by Conformal Transformation,"
Int. J. Mech. ScL, Vol. 20, p. 541.
16. Kiuchi, M., (1984), "Overall Analysis of Non-Axisymmetric Extrusion and
Drawing," Proc. 12th NAMRC, Michigan Technological Univ. Houghton,
Michigan, p. 111.
17. Kudo, H., (1985), Upper Bound Approach to Metal Forming Processes—To
Date and in the Future, "Metal Forming and Impact Mechanics," edited by
S. R. Reid, Pergamon Press, Oxford, p. 19.
18. Hill, R., (1963), "A General Method of Metal-Working Processes," /. Mech.
Phys. Solids, Vol. 11, p. 305.
19. Nagamatsu, A., Murota, T., and Jimma, T., (1970), "On the Non-Uniform
Deformation of Block in Plane Strain Compression Caused by Friction," Bull.
JSME, Vol. 13, p. 1389.
20. Murota, T., Jimma, T., and Nagamatsu, T., (1966), "A Theoretical Analysis
of Circular Cylinder in Axial Compression with Friction," Proc. 16th Japan
Natl. Congr. Appl. Mech., p. 141.
21. Lahoti, G. D. and Kobayashi, S., (1974), "On Hill's General Method of
Analysis for Metal-Working Processes," Int. J. Mech. ScL, Vol. 16, p. 521.
22. Lahoti, G. D., and Kobayashi, S., (1974), "Flat-Tool Forging," Proc. 2d
North American Metal Working Res. Conf., May, University of Wisconsin-
Madison, p. 73.
23. Washizu, K., (1968), "Variational Methods in Elasticity and Plasticity,"
Pergamon Press, Oxford.
24. Lee, C. H., and Kobayashi, S., (1973), "New Solutions to Rigid-Plastic
Deformation Problems Using a Matrix Method," Trans. ASME, J. Engr. Ind.,
Vol. 95, p. 865.
25. Zienkiewicz, O. C., (1977), "The Finite Element Method," 3d Edition,
McGraw-Hill, New York.
26. Chen, C. C., and Kobayashi, S., (1978), "Rigid-Plastic Finite-Element
Analysis of Ring Compression," Application of Numerical Methods to
Forming Processes, ASME, AMD, Vol. 28, p. 163.
27. Chen, C. C., Oh, S. I., and Kobayashi, S., (1979), "Ductile Fracture in
Axisymmetric Extrusion and Drawing, Part 1: Deformation Mechanics of
Extrusion and Drawing," Trans. ASME, J. Engr. Ind., Vol. 101, p. 23.
6
THE FINITE-ELEMENT METHOD—PART I
6.1 Introduction
The concept of the finite-element procedure may be dated back to 1943
when Courant [1] approximated the warping function linearly in each of an
assemblage of triangular elements to the St. Venant torsion problem and
proceeded to formulate the problem using the principle of minimum
potential energy. Similar ideas were used later by several investigators to
obtain the approximate solutions to certain boundary-value problems. It
was Clough [2] who first introduced the term "finite elements" in the study
of plane elasticity problems. The equivalence of this method with the
well-known Ritz method was established at a later date, which made it
possible to extend the applications to a broad spectrum of problems for
which a variational formulation is possible. Since then numerous studies
have been reported on the theory and applications of the finite-element
method.
In this and next chapters the finite-element formulations necessary for
the deformation analysis of metal-forming processes are presented. For hot
forming processes, heat transfer analysis should also be carried out as well
as deformation analysis. Discretization for temperature calculations and
coupling of heat transfer and deformation are discussed in Chap. 12. More
detailed descriptions of the method in general and the solution techniques
can be found in References [3-5], in addition to the books on the
finite-element method listed in Chap. 1.
FIG. 6.1 Finite-element mesh and nodal point specifications in a forming process.
92 Metal Forming and the Finite-Element Method
assigning known values to the corresponding variables. It is to be noted
that the incompressibility condition is not required for defining a velocity
field in the formulation of eq. (5.21) or (5.22).
Equations (5.20) and (5.21) or (5.22) are now expressed in terms of
nodal point velocities v and their variations 6v. From arbitrariness of 6v{,
a set of algebraic equations (stiffness equations) are obtained as
where (/) indicates the quantity at the/th element. The capital-letter suffix
signifies that it refers to the nodal point number.
Equation (6.2) is obtained by evaluating the (dn/dvj) at the elemental
level and assembling them into the global equation under appropriate
constraints.
In metal-forming, the stiffness equation (6.2) is nonlinear and the
solution is obtained iteratively by using the Newton-Raphson method.
The method consists of linearization and application of convergence
criteria to obtain the final solution. Linearization is achieved by Taylor
expansion near an assumed solution point v = v0 (initial guess), namely,
where Au/ is the first-order correction of the velocity v0. Equation (6.3)
can be written in the form
where K is called the stiffness matrix and f is the residual of the nodal point
force vector.
Once the solution of eq. (6.4) for the velocity correction term Av is
obtained, the assumed velocity v0 is updated according to vn + <x Av, where
a is a constant between 0 and 1 called the deceleration coefficient. Iteration
is continued until the velocity correction terms become negligibly small.
The Newton-Raphson iteration process is shown schematically in Fig. 6.2.
It is seen from the figure that convergence of Newton-Raphson iterations,
the initial guess velocity should be close to the actual solution. When a
deformation process is relatively simple, the initial guess velocity can be
provided, for instance, by the upper-bound method. However, if the
process is complex and obtaining a good initial guess solution is difficult,
then the use of the direct iteration method discussed in Section 7.4 of
Chap. 7 may be appropriate.
Two convergence critera may be used. One measures the error norm of
the velocities, ||Av||/||v||, where the Euclidean vector is defined as
||v|| = (vTv)1/2, and requires such an error norm to decrease from iteration
to iteration. The other criterion requires the norm of the residual
equations, ||3jr/3v||, to decrease.
In general terms, the first criterion is most useful in the early stages of
The Finite-Element Method—Part I 93
FIG. 6.2 Schematic representation of the Newton-Raphson method: (a) convergence; (b)
divergence.
iteration, when the velocity field is still far from the solution. The second
test is most useful when slightly ill-conditioned systems reach the final
stage of iterations. The final solution is considered to be achieved when the
error norm reaches a specified small value, say 5 x 1(T5.
The finite-element method procedures outlined above are implemented
in a computer program in the following way.
1. Generate an assumed solution velocity.
2. Evaluate the elemental stiffness matrix for the velocity correction
term Av in eq. (6.4).
3. Impose velocity conditions to the elemental stiffness matrix, and
repeat step 2 over all elements defined in the workpiece.
4. Assemble elemental stiffness matrix to form a global stiffness
equation.
94 Metal Forming and the Finite-Element Method
5. Obtain the velocity correction terms by solving the global stiffness
equation.
6. Update the assumed solution velocity by adding the correctional term
to the assumed velocity. Repeat steps 2 through 6 until the velocity
solution converges.
7. When the converged velocity solution is obtained, update the
geometry of the workpiece using the velocity of nodes during a time
increment. Steps 2 through 7 are repeated until the desired degree of
deformation is achieved.
The order of the certain steps mentioned above may change depending
upon the computer programming practice.
The above procedure applies to the analysis of nonsteady-state-
processes. For steady-state processes, updating the geometry of the
workpiece is not necessary. The procedure for steady-state processes is
described in Chap. 10 for the analysis of axisymmetric extrusion and
drawing.
where fa is a function value associated with ath node, and qa(x, y) is the
shape function. It is, in general, a polynomial function of x and y defined
over the element in such a way that
The Finite-Element Method—Part I 95
where (xe, y@) is the coordinates of /3th node and 6a/i is the Kronecker
delta. Owing to the property of the shape functions given by eq. (6.6), /a
in eq. (6.5) has the value of the function / at (xa, ya) and the fa are
independent of each other.
There are various types of elements, depending upon the shape of the
element and the polynomial order of shape funtions. In the following, the
elements used in this book are discussed.
Triangular Element Family
In the triangular element family, it is convenient to define shape functions
in the area coordinate system, L1; L2, L3. The area coordinates for a
triangle, as shown in Fig. 6.3, are defined by the following linear relations:
where (xa, ya) are the coordinates of a corner of the triangle. It can be
readily shown that an alternative definition of the coordinate of a point P
can be given by the ratio of the area of the shaded triangle to that of the
total triangle as
With a linear triangular element with nodes at its corners (see Fig. 6.4a),
the shape functions qa are linear and are given by
The admissible velocity field within the element can be represented, for
both linear and quadratic elements, by
FIG. 6.4 (a) Linear triangular element; (b) quadratic triangular element.
The Finite-Element Method—Part I 97
FIG. 6.5 (a) Natural coordinate and rectangular parent element; (b) isoparametric element
(mapped on Cartesian coordinate; quadrilateral element); (c) shape function.
In this book, the linear triangular element is used for the analysis of
plate-bending under plane-strain conditions in Chap. 8. It is also used for
the analysis of sheet-metal forming in Chap. 11.
Rectangular Element Family
The shape functions of rectangular elements are, in general, denned in a
parametric form over a domain — ! < £ < ! , — l < r j < l in a natural
coordinate system (|f, r/). The element defined in the natural coordinate
system is sometimes called the parent element. The simplest of the
rectangular elements is the 4-node linear element shown in Fig. 6.5a. The
shape functions, qa, which are bilinear in £ and r/, are defined as
where (£„., r]a) are the natural coordinates of a node at one of its corners.
98 Metal Forming and the Finite-Element Method
The value of the shape function, given by eq. (6.13), is shown schemati-
cally in Fig. 6.5c. Admissible velocity fields can be defined uniquely over
the rectangular element by the nodal velocity components as
where (u^a\ u(ya)) is the velocity at the <*th node and summation is over all
four nodes.
Coordinate transformation from the natural coordinate (£, 77) to the
global coordinate (x, y) is denned by
where (xa, ya) are the global coordinates of the o-th node. Since the
coordinate transformation (6.15) uses the same shape functions (6.14), the
linear element is isoparametric and takes quadrilateral shape in the
Cartesian map, as shown in Fig. 6.5b.
Elements with shape functions of higher-order polynomials can be
defined in a similar manner. Figure 6.6 shows a quadratic rectangular
element with 8 nodes in the natural coordinate and global coordinate
systems. The shape functions are defined by
corner nodes:
mid-side nodez;
(b)
FIG. 6.6 Eight-node quadratic rectangular element: (a) natural coordinate and parent
element; (b) isoparametric element after Cartesian mapping.
the isoparametric elements are sometimes useful, since they allow more
flexibility and convenience in mesh generation. On the other hand, there
are some limitations on the element geometry. To be a valid element, the
coordinate transformation given by eq. (6.15) should be one-to-one, or the
Jacobian of the coordinate transformation (see eq. (6.29)) should be
positive at all points within the element. The validity of the element
geometry is important, since elements deform considerably in metal-
forming simulation.
Elements used for axisymmetric deformation are the same as two-
dimensional elements in terms of shape functions and coordinate transfor-
mations of the element. However, the axisymmetric element represents
the cross section of a torus (ring element), while the other two dimensional
elements represent the cross section of a straight bar. The strain-rate
definition and the volume integration procedure are therefore different.
100 Metal Forming and the Finite-Element Method
The 4-node linear isoparametric element (quadrilateral element) is used
extensively in this book for the analyses of two-dimensional and axisym-
metric deformation processes.
Three-Dimensional Brick Element
The three-dimensional brick element is a natural extension of the
two-dimensional linear element. The parent element is defined over a
domain -!<£<!, -1 < 77 < 1, — 1 < £ :£ 1, in natural coordinate system
with a node at each corner. The shape functions are defined as
In eq. (6.17), (£a, rja, £a) are the coordinates of the orth node in the
natural coordinate system. Figure 6.7 shows the brick element and its
(b)
FIG. 6.7 Three-dimensional brick element in natural and Cartesian coordinate systems.
The Finite-Element Method—Part I 101
node-ordering convention. The velocity field can be expressed by
where (u^a\ u^, u^') is the velocity of the oth node. The coordinate
transformation is given by
where (xa, ya, za) are the coordinates of the o-th node.
The use of the brick element is shown in Chap. 14 for the analysis of
three-dimensional deformation.
Note on Notation
In expressing elemental functions, we treat each velocity component as a
scalar. Sometimes it is more convenient to express a velocity field in a
vector form as
It was also shown, in Section 6.3, that the admissible velocity for all type
of elements can be expressed by
It is seen from eq. (6.22) that strain rate components can be evaluated if
dqa/dXi is known.
For the Cartesian coordinate system, we denote the coordinate jc, by
(x, y, z) for three-dimensional deformation, by (r, z, 0) for axisymmetric
deformation, and by (x, y) for two-dimensional deformation.
Let Xa, Ya, and ZK be denned as
for plane-stresss
The Finite-Element Method—Part I 103
and
for axisymmetric
deformation
Substituting eqs. (6.23) into eqs. (6.24a, 6.24b and 6.24c), the strain-rate
vectors are represented in a unified form, as
where
Note that |J| is twice of the area of the triangle. It may also be noted that
all the strain-rate components of a linear triangular element are constant
over one element, since Xa and Ya of eq. (6.33) are not functions of the
area coordinates.
The strain-rate matrix of the quadratic triangular element can be derived
in a similar manner to that shown above. However, the expressions for Xa
and Ya are much more complex, and it is easier to evaluate these
numerically following procedures similar to those used for the linear
element. It may be also mentioned that the strain-rate is not constant and
varies within the quadratic element.
Rectangular Element Family
For the rectangular family of elements, Xa and Ya in eq. (6.25) can be
written as
where |J| is the determinant of the Jacobian matrix of eq. (6.15) and is
expressed by
with
The evaluation of Xa, Ya, and Za in eq. (6.36) can be made by using eqs.
The Finite-Element Method—Part I 107
where P = BTDB.
The matrix D in eq. (6.40) takes different forms depending upon the
expression of the effective strain-rate, in terms of strain-rate components.
For example, the effective strain-rate in plane-stress problems is expressed
in a different form from that of plane-strain problems, although the
definition of the effective strain-rate is identical in both cases. The matrix
D written for plane-stress problems is not diagonal (Chap. 11). The
expression of the effective strain-rate also depends on the yield criterion.
Thus, the matrix D is different for anisotropic materials (Chap. 11) and for
porous materials (Chap. 13), and is described in corresponding chapters.
The volumetric strain-rate £„ is given by
and expressed by
where
It should be noted that the term (—djtSp/dvj) is the applied nodal point
force and that dnDl' dv, + djtp/dv, is the reaction nodal point force.
The second derivatives of n are expressed as
Evaluating stiffness matrices at the elemental level from eqs. (6.43) and
(6.44), assembling them for the whole workpiece, we obtain a set of
simultaneous linear equations (6.4).
When effective strain-rate e approaches 0, or becomes less than a
preassigned value e0, we have, by eq. (5.31) in Chap. 5,
The Finite-Element Method—Part I 109
The penalty constant K and the limiting strain rate £0 are introduced
rather arbitrarily for computational convenience. However, proper choices
of these two constants are important in successful simulation of metal-
forming processes. A large value of K is, in general, preferred to keep the
volume strain-rate close to zero. However, too large a value of K may
cause difficulties in convergence, while too small a K results in unaccept-
ably large volumetric strain. Numerical tests show that an appropriate K
value can be estimated by restricting volumetric strain rate to 0.0001—0.001
times the average effective strain-rate.
The limiting strain-rate, e0, under which the material is considered to be
rigid, has been introduced to improve the numerical behavior of the
rigid-plastic formulation [7]. Too large a value of the limiting strain-rate
results in a solution in which the strain-rate of the rigid zone becomes
unacceptably large. On the other hand, if we choose too small a value of
limiting strain-rate, then the convergence of the Newton-Raphson method
deteriorates considerably. Numerical tests show that an optimum result
can be obtained by choosing the limiting strain rate as -^ of the average
effective strain-rate.
Equation (5.21) is the basic formulation using a Lagrange multiplier in
order to remove the incompressibility constraint. It replaces dnp in eq.
(5.22) by <5jrA = J <5(Ae w ) dV. The Lagrange multiplier A is treated as an
independent variable and a function of material points denned over the
workpiece. The Lagrange multipliers are defined at points where the
volume constancy is enforced and are introduced at the reduced integra-
tion points (see Section 7.1 of Chap. 7). For example, one variable is
assigned for each linear rectangular [8] and three-dimensional brick
element, and four A are needed for a quadratic rectangular element.
Since 6A is also arbitrary in eq. (5.21), we obtain a set of simultaneous
equations given by
The first equation of (6.47) is nonlinear and eq. (6.3) applies for
linearization. The additional terms necessary for evaluation of matrices in
the stiffness equation are
Then eqs. (6.3) and (6.47) can be arranged in the system of equations
110 Metal Forming and the Finite-Element Method
References
1. Courant, R., (1943), "Variational Methods for the Solution of Problems of
Equilibrium and Vibrations," Bull. Amer. Math. Soc., Vol. 49, p. 1.
2. Clough, R. W., (1960), "The Finite Element Method in Plane Stress Analysis,"
/. Struct. Div., ASCE, Proc. 2nd Conf. Electronic Computation, p. 345.
3. Bathe, K. J., and Wilson, E. L., (1976), "Numerical Methods in Finite-Element
Analysis," Prentice-Hall, Englewood Cliffs, NJ.
4. Oden, J. T., (1972), "Finite Element of Nonlinear Continua," McGraw-Hill,
New York.
5. Desai, C. S., and Abel, J. F., (1972), "Introduction to the Finite Element
Method," Van Nostrand Reinhold, New York.
6. Hildebrand, F. B., (1974), "Introduction to Numerical Analysis," 2d Edition,
McGraw-Hill, New York.
7. Chen, C. C., Oh, S. I., and Kobayashi, S., (1979), "Ductile Fracture in
Axisymmetric Extrusion and Drawing, Part I: Deformation Mechanics of
Extrusion and Drawing," Trans. ASME, J. Engr. Ind., Vol. 101, p. 23.
8. Lee, C. H., and Kobayashi, S., (1973), "New Solutions to Rigid-Plastic
Deformation Problems Using a Matrix Method," Trans. ASME, J. Engr. Ind.,
Vol. 95, p. 865.
7
THE FINITE-ELEMENT METHOD—PART II
n X, w,
1 0 2.000000000000000
2 ±0.577350269189626 1.000000000000000
3 ±0.774 596 669 241 483 0.555 555 555 555 556
0.000000000000000 0.888888888888889
4 ±0.861 136311594053 0.347854845137454
±0.339981043584856 0.652 145 154 862 546
and, in the three-dimensional space (-1 < g < 1, - l < j j < l , -!<£<!);
where (Lu, L2I, L3f) is the area coordinate of an integration point [3].
The integration formula given by eq. (7.6) gives the exact integration of
a polynomial function of degree n. The integration points and weight
factors of the Gaussian quadrature formula for triangular elements are
listed in Table 7.2 [1].
Application of Integration Formulas
The use of Gaussian quadrature and isoparametric elements in FEM
formulation will introduce errors. Usually, it is true that a higher-order
integration is suggested in order to obtain accurate evaluations of
integrands. But a large portion of the computer execution time is spent in
performing this numerical integration. Therefore, a proper choice should
be the lowest possible order of integration that does not introduce much
error into the results.
The minimum order of integration that ensures convergence has been
established for different types of elements. For a linear rectangular
element, 2 x 2 integration points are sufficient to guarantee convergence.
For a three-dimensional brick element, it is necessary to use 2 x 2 x 2
integration points. The required integration order for quadratic elements is
3 x 3 . For the triangular element family, one-point integration is sufficient
for a linear element, while three-point integration is required for a
quadratic element. It is also known that increasing the order of integration
above the required minimum does not necessarily improve the solution
accuracy. In determining the order of integration, possible overconstraints
imposed by the FEM formulation must be also considered. For example,
the incompressibility requirement cannot be satisfied at all points in a
Number of
Integration Integration Coordinates Weight
Points Order (^,, L2, L3) Factors
1 Linear i i i 1
3> 3> 3
3 Quadratic U,o i
i
o,U
I n 1
3
1
2>«> 2 3
1 1 1 27
4 Cubic 3> 3' 3 48
25
0.6,0.2,0.2 48
25
0.2,0.6,0.2 48
25
0.2,0.2,0.6 48
114 Metal Forming and the Finite-Element Method
rectangular linear element except for uniform deformation [4]. In order to
relieve this overconstraint, one-point integration is used for the volumetric
strain-rate term in a linear rectangular element. This is known as the
reduced integration scheme. The reduced integration in a linear rectangu-
lar element, in effect, imposes the volume constancy averaged over that
element.
For higher-order elements, it is found that the integration order of the
dilatational contribution may be reduced by 1 to obtain the best results.
That is, the volumetric strain-rate term is integrated at 2x2 integration
points for a quadratic rectangular element and at one point for a quadratic
triangular element. For the linear triangular element, which gives constant
strain-rate over the element, the reduced integration technique cannot be
applied. This is why the linear triangular element is not used for the case in
which volume constancy is required. When the linear triangular element
with the regular integration order is used for forming simulation, the
model tries to achieve the volume constancy averaged over several
elements. It has been shown that if four linear triangular elements are
arranged to form a rectangle, then the volume constancy is satisfied over
the four elements within the rectangle [4]. It may be also mentioned that
the number of Lagrange multipliers necessary to enforce volume constancy
should be the same as the number of the reduced integration points of the
element used.
The Gaussian integration points are the natural locations to evaluate the
strain-rate and stress within the element, since all the necessary informa-
tion to calculate the strain-rate and stress are already evaluated at these
points. It is shown [5], however, that the calculated stress values at the
regular integration points deviate considerably from the actual ones when
volume constancy is enforced. On the other hand, stress and strain-rates
evaluated at the reduced integration points are known to be more
accurate. Consequently, throughout this book, reduced integration points
are used for stress and strain-rate evaluations.
The integrations of the boundary terms can also be evaluated by the
Gaussian quadrature. It is shown that a large number of integration points
are necessary when the integrand contains higher-order polynomials. The
friction term given by eq. (5.30) cannot be approximated by a low-order
polynomial over the usual range of the sliding velocity [6] and it is
necessary to use a higher-order integration formula. It may sometimes be
more convenient to use Simpson's formula to calculate the frictional
energy rate along the tool-workpiece interface [7], since this formula can
be applied to represent the local behavior of the integrand more accurately
by subdividing the integration range. Simpson's formula is given by
Since the stiffness matrix has been evaluated at the elemental level,
evaluated elemental stiffness matrices should be assembled as shown in eq.
(7.8).
Often the assembled stiffness equations are numbered in the order of
nodal-point numbering and the elemental stiffness equations are numbered
in the order of element connectivity, so that the connectivity of each
element can be used to correlate the corresponding equations and
variables in global and elemental stiffness equations.
It is also seen from eq. (7.8) that the stiffness equation corresponding to
node 10 contains the variables of nodes associated with the elements
surrounding node 10. That is, the equation contains the velocity com-
ponents of nodes 5, 6, 7, 9, 10, 11, 15, 16, and 17. Therefore, the global
stiffness matrix is a sparse matrix because of the limited range of influence
of a node velocity on the admissible velocity field. The sparseness of the
global stiffness matrix can be utilized to reduce the computational effort in
solving the linear equations. Thus, the stiffness matrix is arranged in a
banded matrix form, as shown in Fig. 7.2, by proper numbering of nodes.
Gaussian Elimination
Completing assemblage of elements, the global stiffness equation can be
written in the form of eq. (6.4) in Chap. 6, namely,
forms. The simplest way is to store the stiffness matrix in a banded matrix
form, and to apply Gaussian elimination over the maximum band width. In
a banded matrix solver, the same number of matrix elements is stored for
each equation. In the skyline solver, K is stored column-wise, from the
diagonal matrix element to the last nonzero element in the column [8].
The number of operations necessary to solve a linear matrix can be
estimated by ^nm\, where n is the total number of equations and mk is the
half band width of the matrix. Here, one operation consists of one
multiplication (or division), which is almost always followed by an
addition. The estimated number of operations shows that the computer
time necessary to solve the matrix equation is proportional to the square of
the half band width. It is therefore a must to number the nodes in such a
way that the assembled stiffness equation has minimum band width.
The variables of the stiffness matrix can be "eliminated" when the
stiffness matrix is assembled partially. That is, when the stiffness equation
for node / is completely assembled, then the corresponding variable can be
expressed by other variables, although the assembly is not complete. The
frontal solution technique is a linear equation solver based on the
elimination of variables during assembly. The frontal solution technique
[10] is known to be advantageous when the stiffness matrix with minimum
band width is still sparse. Such cases arise for the three-dimensional
finite-element method. In the frontal method, element numbering, but not
node numbering, influences the matrix solution efficiency; therefore, new
nodes can be added without renumbering the nodes.
where d is measured from the x axis in the global coordinate system to the
x' axis of the local coordinate system in counterclockwise direction. In the
three-dimensional coordinate system, the transformation matrix is ex-
pressed by
where T^ is a direction cosine between the coordinate axis of xt and x'j. The
transformation matrix for all nodes on the surface Sc can be constructed as
where UD is the tool velocity and n is the unit normal to the interface
surface.
In the direction of the relative sliding between the die and the
workpiece, the frictional stress /, is prescribed as the traction boundary
condition. The frictional stress is usually represented according to the
Coulomb law or as a constant frictional stress, as discussed in Chap. 3. The
friction representation by a constant friction factor m (eq. (3.2) in Chap. 3)
is approximated by eq. (5.30) in Chap. 5, in order to deal with
neutral-point problems in metal forming.
Equation (5.30) expresses that the magnitude of the frictional stress is
dependent on the magnitude of the relative sliding and that their directions
are opposite to each other. Then, the relationship can be written as
and
NODE VELOCITY
where (x,, y/) are the coordinates of node /, t0 is the time at current
configuration, and Af is the time-increment. The strain is updated in a
similar manner from the strain-rate solution. In general, the time
increment At can be determined by considering several factors, such as the
time (A£die) necessary for a next free node to contact the die surface, a
desired maximum strain-increment (Afstrain), and a maximum allowable
time-increment (A?a). The actual time-increment is determined by taking
the minimum of A£die, Afstrain, and Af a . The time necessary for a next free
node to contact the die can be determined by calculating these time
increments for all free nodes and choosing the minimum time-increment.
The time-increment required to limit the maximum strain-increment can
be readily obtained from strain-rate solutions. The maximum allowable
time-increment is given rather arbitrarily. However, consideration of the
error in the volume constancy is a factor for determining its magnitude.
During the time increment Af, elements lose volume after the geometry
is updated. Consider a two-dimensional element, as shown in Fig. 7.6,
where the element (1234) with a width of W and a height H is deformed to
the shape (12'3'4') after a time increment At. The volume constancy
requires that
or
where V and AF are the volume of the element and the amount of volume
change, respectively. Equation (7.18) suggests that updating by eq. (7.16)
always results in a volume loss during a time-increment and that the
volume loss rate is proportional to the square of (A///H). Assuming that
(A/////) is constant at all increments, we have, after n increments,
The total volume loss given by eq. (7.19) is plotted in Fig. 7.7 as a function
of total reduction in height and of &.H/H. It is seen from the figure that
0.7% volume is lost at 50% reduction in height if (A/f///) is 0.01, and the
volume loss reaches 2.3% at 90% reduction in height with (AH/H) of
0.01. Therefore, it is necessary to use (A/////) less than 0.01 in order to
maintain less than 2% volume loss at 90% reduction in height. It is also
seen that volume loss increases considerably with larger step size.
In case of axisymmetric deformation, the volume of an element (solid
cylinder) is represented by V — nR2H. The volume loss rate can be
FIG. 7.7 Volume loss during plane-strain FEM simulation as function of deformation for
various step sizes of A/////.
The Finite-Element Method—Part II 125
FIG. 7.8 Volume loss during plane-strain FEM simulation as function of deformation for
various step sizes of A///W 0 .
calculated by
7.6 Rezoning
In practical forming processes, deformation is usually very large, and it is
not uncommon to encounter effective strain with magnitudes of 2 or more.
Moreover, the relative motion between the die surface and the deforming
material is also large. Such large deformations and displacements cause the
following computational problems during FEM simulation with a Lagran-
gian mesh:
where EN is the effective strain value at node N and £7 is the effective strain
value at the center of element / that surrounds the node N. AjN is the area
contribution of ;th element to node N and is denned by
where qa is the element shape function. The interpolation from the old
mesh to the new mesh is done by evaluating the effective strain value at
the integration point locations of the new mesh system.
As can be expected, the area-weighted average method, in general,
provides very accurate interpolation results for internal nodes when the
trend of the field variable distribution is relatively well defined by the
values at surrounding elements. The method often fails at the workpiece
boundary when the field variable has large gradients, since the boundary
nodes do not have a sufficient number of surrounding elements. However,
for all practical purposes, the method provides sufficient accuracy for
metal-forming simulations. Other methods that improve the accuracy of
interpolation at the workpiece boundary have been suggested and are
discussed in Reference [13].
The rezoning procedure has been applied to simulations of various
128 Metal Forming and the Finite-Element Method
Rnoiusmm
FIG. 7.10 FEM mesh assignment for rezoning.
of degrees of freedom near the flash. The new mesh is also shown in Fig.
7.10 (bottom half). Note that the new mesh system allows far more
degrees of freedom near the flash. Interpolation of effective strain was
carried out by the area-weighted average method and the results are shown
in Fig. 7.11. It is seen from the figure that the strain distributions are
almost identical before and after rezoning.
References
1. Zienkiewicz, O. C., (1977), "The Finite Element Method," 3d Edition,
McGraw-Hill, New York.
2. Strang, G., and Fix, G. J., (1973), "An Analysis of the Finite Element
Method," Prentice-Hall, Englewood Cliffs, NJ.
3. Cowper, G. R., (1973), "Gaussian Quadrature Formulas for Triangles," Int. J.
Num. Meth. Engr., Vol. 7, p. 405.
4. Nagtegaal, J. C., Parks, D. M., and Rice, J. C., (1974), "On Numerically
Accurate Finite Element Solutions in the Fully Plastic Range," Comput. Meth.
Appl. Mech. Eng., Vol. 4, p. 153.
5. Zienkiewicz, O. C., and Hinton, E., (1976), "Reduced Integration, Function
Smoothing and Non-Conformity in Finite Element Analysis," /. Franklin Inst.,
Vol. 302, p. 443.
6. Oh, S. I., (1982), "Finite Element Analysis of Metal Forming Problems with
Arbitrarily Shaped Dies," Int. J. Mech. Sci., Vol. 24, p. 479.
7. Lee, G. J., and Kobayashi, S., (1982), "Spread Analysis in Rolling by the
Rigid-Plastic Finite Element Method," Numerical Method in Industrial Form-
ing Processes, p. 777.
8. Bathe, K. J., (1982), "Finite Element Procedures in Engineering Analysis,"
Prentice-Hall, Englewood Cliffs, NJ.
9. Noble, B., (1969), "Applied Linear Algebra," Prentice-Hall, Englewood
Cliffs, NJ.
10. Irons, B. M., (1970), "A Frontal Solution Program for Finite Element
Analysis," Int. J. Num. Meth. Engr., Vol. 2, p. 5.
11. Chen, C. C., and Kobayashi, S., (1978), "Rigid-plastic Finite Element Method
Analysis of Ring Compression," Application of Numerical Method to Forming
Processes, ASME, AMD, Vol. 28, p. 163.
130 Metal Forming and the Finite-Element Method
12. Lyness, J. F., Owen, D. R. J., and Zienkiewicz, O. C., (1974), "Finite
Element Analysis of Steady Flow of Non-Newtonian Fluid Through Parallel
Sided Conduits," Int. Symp. on Finite Element Method in Flow Problems,
Swansea, p. 489.
13. Oh, S. I., Tang, J. P., and Badawy, A., (1984), "Finite Element Mesh
Rezoning and its Applications to Metal Forming Analysis," Proc. of 1st ICTP
Conference, Tokyo, p. 1051.
14. Badawy, A., Oh, S. I., and Allan, T., (1983), "A Remeshing Technique for
the FEM Simulation of Metal Forming Processes," Proc. Int. Computer Engr.
Conf., ASME, Chicago, IL., p. 143.
15. Ficke, J. A., Oh, S. I., and Malas, J., (1984), "FEM Simulation of Closed Die
Forging of Titanium Disk using ALPID," Proc. of NAMRC XII, Houghton,
ML, p. 166.
8
PLANE-STRAIN PROBLEMS
8.1 Introduction
This chapter is concerned with the formulations and solutions for plane
plastic flow. In plane plastic flow, velocities of all points occur in planes
parallel to a certain plane, say the (jc, y) plane, and are independent of the
distance from that plane. The Cartesian components of the velocity vector
u are ux(x, y), uy(x, y), and uz = 0.
For analyzing the deformation of rigid-perfectly plastic and rate-
insensitive materials, a mathematically sound slip-line field theory was
established (see the books on metal forming listed in Chap. 1). The
solution techniques have been well developed, and the collection of
slip-line solutions now available is large [1]. Although these slip-line
solutions provide valuable insight into deformation modes and forming
loads, slip-line field analysis becomes unwieldy for nonsteady-state prob-
lems where the field has to be updated as deformation proceeds to account
for changes in material boundaries. Furthermore, the neglect of work-
hardening, strain-rate, and temperature effects is inappropriate for certain
types of problems. Many investigators, notably Oxley and his co-workers,
have attempted to account for some of these effects in the construction of
slip-line fields. However, by so doing, the problem becomes analytically
difficult, and recourse is made to experimental determination of velocity
fields, similarly to the visioplasticity method. Some of this work is
summarized in Reference [2]. The applications of the finite-element
method are particularly effective to the problems for which the slip-line
solutions are difficult to obtain.
131
132 Metal Forming and the Finite-Element Method
where
with
then
Plane-Strain Problems 133
where Xa and Ya are expressed by eq. (6.35) for the node numbering
shown in Fig. 8.1.
The volumetric strain-rate ev is expressed by
where
by putting P = BTDB.
The matrix D in eq. (8.5) is given by
Then the basic equation, for example, eq. (8.1b) is discretized and a set of
nonlinear simultaneous equations (stiffness equations) is obtained from the
arbitrariness of <5v as
with P, C, and N defined above. For the solution procedure, eq. (8.6) is
linearized according to eq. (6.3) in Chap. 6.
FIG. 8.3 Velocity distributions in closed-die forging: (a) FEM (sticking friction); (b)
experiment [10].
136 Metal Forming and the Finite-Element Method
FIG. 8.4 Grid patterns in closed-die forging. Upper: experiment (33rd step) [10]. Lower:
FEM (32nd step, dark portion indicates rigid elements).
FIG. 8.5 (a) Geometry and external forces in rolling, (b) Mesh systems used for
nonsteady-state and steady-state computations.
The classical slab method gives the shape of the stress distribution,
known as the "friction hill," with maximum pressure located at the neutral
point, on both sides of which the pressure decreases monotonically.
However, in some circumstances the pressure distribution curves were
observed to have double peaks. A slip-line solution [20] also predicted the
pressure drop in the middle of the arc of contact. In the computed results,
double-peak pressure distribution curves, as well as friction-hill type
Plane-Strain Problems 139
4 dec,
FIG. 8.6 Comparison of measured [19] and computed roll-workpiece contact pressure
distributions: (a) P = 1.79 and (b) P = 13.18 for aluminum; (c) P = 1.85 and (d) P = 11.57 for
copper.
distributions, were obtained. Figure 8.6 shows the pressure variations for
some typical cases, comparing the experimental and computed results.
An important parameter that determines the deformation mode is the
ratio of the roll-workpiece contact length to the initial workpiece
thickness. This ratio is equivalent to the parameter P defined by
P = R(H0-Hl)/Hl = (Reduction)(R/H0). In Fig. 8.6a, the pressure dis-
tribution curve shows maxima near entrance and exit with a pressure drop
in the middle of contact length. This is characteristic for small values of the
parameter P. For large values of the parameter P, Fig. 8.6b shows a
friction-hill type pressure distribution. There is considerable discrepancy in
magnitude between the computed values and the experimental measure-
ments. The same observations apply to the pressure distributions for a
different material given in Figs. 8.6c and d. As typified by the results
shown in Fig. 8.6, two modes of deformation are observed. In one the
contact pressure distribution shows double peaks and deformation is more
inhomogeneous, and in the other the pressure distribution is a friction-hill
type and homogeneous deformation is dominant.
For the roll separating force and the torque in rolling aluminum and
copper, the comparisons between the computed results and the ex-
perimental values [19] are shown in Fig. 8.7. The computed results for the
roll separating force are smaller than the measured values, while the
140 Metal Forming and the Finite-Element Method
Reduction
FIG. 8.7 Comparison of measured [19] and computed roll separating force and roll torque;
(a) and (b) for aluminum; (c) and (d) for copper.
Reduction
FIG. 8.8 Comparison of measured [19] and computed roll separating force and roll torque
for steel; (a) and (b) with R/Ha = 65; (c) and (d) with R/Hn = 130.
FIG. 8.9 (a) Schematic diagram of sheet bending operation used in the computation with
T = l, W = 7.5, RP = 1.8, RD = 1.2. (b) Boundary conditions in sheet bending.
and strain distributions, and the deformed geometries at each stage of the
process.
Figure 8.10 compares the grid distortion patterns obtained from rigid-
plastic and elastic-plastic formulations. The black area of the workpiece
under the punch represents the deforming zone in the rigid-plastic analysis
and the plastic zone in the elastic-plastic analysis. From the elastic-plastic
calculations, it can be seen that yielding begins at the outer fiber near the
bending axis and spreads gradually toward the rest of the sheet. The trend
Elasto-Plastic Rigid-Plastic
FIG. 8.10 Comparison of grid distortions by rigid-plastic and elastic-plastic (elasto-plastic)
formulations for punch displacements: (a) 0.6, (b) 1.2, (c) 1.8, (d) 2.25, (e) 2.8, (f) 3.2. The
darkened area denotes the plastically deforming (plastic) zone.
146 Metal Forming and the Finite-Element Method
is the same as that predicted by the elementary beam theory. As the
process continues, a well-defined wedge-shaped elastic-plastic boundary
moves away from the bending axis. But, after a certain punch displace-
ment, the boundary remains nearly stationary. The rigid-plastic calculation
does not show the well-defined continuous spread of the plastic zone. It
can be seen, however, that the size of the deforming zone determined by
the rigid-plastic analysis compares well with that of the plastic zone
obtained in the elastic-plastic analysis. It can also be seen from the figure
that unloading takes place near the bending axis during the later stages of
deformation in the elastic-plastic case, and that the rigid zone appears
during the same stage in the rigid-plastic case.
Another interesting feature in the deformation mode of sheet bending is
that the bend radius of the workpiece does not follow that of the punch;
the workpiece separates from the punch and only a small portion of the
workpiece is in contact with it, resulting in shape inaccuracies in the bent
sheet. This is shown quantitatively in Fig. 8.11, where the bend angle and
the clearance between the punch pole and the workpiece are plotted
against the punch displacement for both analyses. The figure shows that
the curves for bend angle vs. punch displacement obtained from the two
formulations are in almost perfect agreement. From the figure it can be
seen that workpiece separation from the punch pole occurs when the
punch displacement reaches 1.48 according to the elastic-plastic calculation
and 1.68 according to the rigid-plastic analysis. The clearances that are
calculated from the two formulations show the same trend, but the
rigid-plastic formulation results in slightly larger clearance than that of
elastic-plastic analysis.
PUNCH DISPLACEMENT
FIG. 8.11 Relationship of bend angle to punch displacement, and of clearance between the
punch poles and the workpiece to punch displacement.
Plane-Strain Problems 147
Elasto-Plastic Rigid-Plastic
FIG. 8.12 Comparison of distributions of normalized bending stress (as/initial yield stress)
and effective strain by rigid-plastic and elastic-plastic (elasto-plastic ) formulations at the end
of loading (punch displacement = 3.2).
Figure 8.12 shows the distributions of bending stress and the effective
strain at the end of loading, for rigid-plastic and elastic-plastic materials.
From the figure it can be seen that the bending stress increases toward the
outer surface and toward the punch axis, as expected. It can be also seen
that the neutral line (os = 0) shifts toward the punch near the punch axis
and is at about half of the thickness in other locations. The effective strain
distributions show a similar pattern. The figure also reveals that the
distributions of bending stress and effective strain for both analyses agree
with each other very well, with minor differences.
Unloading and resulting spring-back and residual stresses are discussed
in Chap. 16.
148 Metal Forming and the Finite-Element Method
(a)
(b)
FIG. 8.13 ^umparison of predicted strain-rate (per second) distributions and experimentally
determined flow localization in side pressing of Ti-6242-0.1Si at 913°C (1675T): (a) (a + ft)
microstructure; (b) (/J) microstructure [26],
Plane-Strain Problems 149
diameter of the specimen was 10.2 mm (0.40 inch) and the deformation
was assumed to be plane-strain (zero strain in the axial direction). The
temperature during deformation was 913°C (1675°F). Figure 8.13 shows
the experimentally observed transverse sections of side-pressed cylinders
of (a + /?) and (/3) microstructures and the predicted effective strain-rate
distributions for both cases. It can be seen that the method effectively
predicts, for the same alloy and forming conditions, detailed variations in
metal flow that are due to differences in microstructure and flow stress
behavior.
References
1. Johnson, W., Sowerby, R., and Venter, R. D., (1982), 'Plane-Strain Slip-Line
Fields for Metal Deformation Processes," Pergamon Press, Oxford.
2. Oxley, P. L. B., and Hastings, W. F., (1976), "Minimum Work as a Possible
Criterion for Determining the Frictional Conditions at the Tool/Chip Interface
in Machining," Phil. Trans. R. Soc., London, Series A. Vol. 282, p. 565.
3. Kobayashi, S., Herzog, R., Lapsley, J. T., Jr., and Thomsen, E. G., (1959),
"Theory and Experiment of Press Forging Axisymmetric Parts of Aluminum
and Lead," Trans. ASME, J. Engr. Ind., Vol. 81, p. 228.
4. Kobayashi, S., and Thomsen, E. G., (1959), "Approximate Solutions to a
Problem of Press Forging," Trans. ASME, J. Engr. Ind., Vol. 81, p. 217.
5. Akgerman, N., and Altan, T., (1972), "Modular Analysis of Geometry and
Stresses in Closed-Die Forging: Application to a Structural Part," Trans.
ASME, J. Engr. Ind., Vol. 94, p. 1025.
6. Johnson, W., (1958), "Over-Estimates of Load for Some Two-Dimensional
Forging Operations," Proc. 3d U.S. Congr. Appl. Mech., ASME (New York),
p. 571.
7. Kudo, H., (1958), "Studies on forging and Extrusion Process, I," Koken-
shuho, Tokyo University, Vol. 1, p. 37.
8. McDermott, R. P., and Bramley, A. N., (1974), "An Elemental Upper-Bound
Technique for General Use in Forging Analysis," Proc. 15th Int. MTDR
Conference, Birmingham, England, p. 437.
9. Kasuga, Y., Tsutumi, S., and Saiki, H., (1974), "Material Flow in Sunken
Forging Dies," /. Japan Soc. Tech. Plast., Vol. 15, p. 147.
10. Lyapunov, N. L, and Kobayashi, S., (1974), "Metal Flow in Plane-Strain
Closed-Die Forging," Proc. of 5th North Amer. Metalworking Res. Conf.
NAMRC, p. 114.
11. Chen, C. C., and Kobayashi, S., (1980), Rigid-Plastic Finite Element Analysis
of Plane-Strain Closed-Die Forging, "Process Modeling," ASM, Metals Park,
Ohio, p. 167.
12. Oh, S. I., (1982), "Finite Element Analysis of Metal Forming Processes with
Arbitrary Shaped Dies," Int. J. Mech. ScL, Vol. 24, p. 479.
13. Wu, W. T., Oh, S. L, and Altan, T., (1984), "Investigation of Defect
Formation in Rib-Web Type Forging by ALPID," Proc. 12th NAMRC, p.
159.
14. Mori, K., Osakada, K., and Fukada, M., (1983), "Simulation of Severe Plastic
Deformation by Finite Element Method with Spatially Fixed Elements," Int. J.
Mech. ScL, Vol. 25, p. 775.
15. Zienkiewicz, O. C., Jain, P. C., and Onate, E., (1978), "Flow of Solids
During Forming and Extrusion: Some Aspects of Numerical Solutions," Int. J.
Solids Structures, Vol. 14, p. 15.
150 Metal Forming and the Finite-Element Method
9.1 Introduction
According to Spies [1], the majority of forgings can be classified into three
main groups. The first group consists of compact shapes that have
approximately the same dimensions in all three directions. The second
group consists of disk shapes that have two of the three dimensions (length
and width) approximately equal and larger than the height. The third
group consists of the long shapes that have one main dimension sig-
nificantly larger than the two others. All axially symmetric forgings belong
to the second group, which includes approximately 30% of all commonly
used forgings [2]. A basic axisymmetric forging process is compression of
cylinders. It is a relatively simple operation and thus it is often used as a
property test and as a preforming operation in hot and cold forging. The
apparent simplicity, however, turns into a complex deformation when
friction is present at the die-workpiece interface. With the finite-element
method, this complex deformation mode can be examined in detail. In this
chapter, compression of cylinders and related forming operations are
discussed. Since friction at the tool-workpiece interface is an important
factor in the analysis of metal-forming processes, this aspect is also given
particular consideration. Further, applications of the FEM method for
complex-shaped dies are shown in the examples of forging and cabbaging.
151
152 Metal Forming and the Finite-Element Method
where ra, za (a = 1, . . . ,4) are the positions of the four surrounding nodal
points of an element in the global coordinate system. For an isoparametric
element of Fig. 9.1, discussed in Chap. 6, transformation functions qa in
eq. (9.1) are the same as shape functions defined by eq. (8.2b) in Chap. 8.
Then
where
Axisymmetric Isothermal Forging 153
and
and
Equations (9.4b) and (9.4c) are obtained by replacing (x, y) in eqs. (6.35a)
in Chap. 6 by (r, z).
The determinant of the Jacobian of transformation |J| is expressed by
eq. (6.35b) in Chap. 6 with (r, z) replacement. The matrix D and the
vector C in the expressions of the effective strain-rate and the volumetric
strain-rate are the same as those given by eqs. (8.4) and (8.5) in Chap. 8.
The stiffness equations based on, for example, eq. (8.la) in Chap. 8,
become
and
where the subscript j indicates the element number and M is the total
number of elements. Linearization of eq. (9.5) is made according to eq.
(6.48) in Chap. 6.
FIG. 9.2 Grid distortions in simple compression at 50% reduction in height for the two
friction conditions.
156
Axisymmetric Isothermal Forging 157
FIG. 9.4 Workability of SAE 1040 (annealed) steel at room temperature in slow-speed
upsetting.
die. Otherwise, the maximum strain is near the axis and almost at
mid-height of the upset head. Since hydrostatic pressure seems to play an
important role in ductile fracture, it is instructive to examine the
distribution of am. This distribution is plotted on the left-hand side of Fig.
9.6. Although the pattern of the effective strain distribution remains
almost unchanged with increasing reduction in height, the hydrostatic
stress distribution changes a great deal. A very important observation,
from the fracture point of view, is that with increasing reduction in height
the hydrostatic pressure increases near the axis of the specimen, whereas it
decreases near the free surface. In fact, the hydrostatic stress becomes
tensile at higher reductions. This explains why, in heading operations,
66 Percent Reduction
FIG. 9.5 Grid distortions in simple compression at 36 and 66% reductions in height for initial
height-to-diameter ratio of 2.5.
FIG. 9.6 Effective strain (e) contours and hydrostatic pressure (—a m /Y 0 ) contours in heading
at 60% reduction in height.
158
Axisymmetric Isothermal Forging 159
FIG. 9.7 Experimental and theoretical flow lines in heading at 31% reduction, 47%
reduction, and 63% reduction (60%, theory) [11].
surface cracking is the predominant failure mode rather than the formation
of internal cracks.
A significant comparison between theory and experiment can be made in
terms of the grid distortion pattern. The flow patterns obtained ex-
perimentally by etching the formed specimens are shown in Fig. 9.7 at
approximately 31%, 47%, and 63% reductions in height. The figure also
shows the grid patterns obtained theoretically at 31%, 47%, and 60%
reductions. It can be seen that the distortion of axial lines predicted by the
theory is almost identical to the flow lines observed experimentally.
(C)
FIG. 9.8 Two deformation modes in ring compression, (a) Compression geometry; (b)
situation with low interface friction; (c) situation with high interface friction.
FIG. 9.9 Comparison of the finite-element solution with the upper bound solutions for ring
compression. Experimental data: O, copper; •, copper; A, pure aluminum [17].
162 Metal Forming and the Finite-Element Method
upon the selection of the initial or current shear yield stress k. It is also
seen that the geometrical changes are affected by the material property, as
indicated by the curves calculated with ra = 1.0 for copper and for
annealed aluminum.
The problem of ring compression was also analyzed by the elastic-plastic
finite-element method. Hartley et al. [15,16] developed a method for
introducing friction into the finite-element analysis by the inclusion of a
layer of elements. The stiffness of these elements was modified by a
function of the interfacial shear factor and the method was applied to the
analysis of ring compression.
FIG. 9.10 Grid distortions in ring compression at 50% reduction in height for various
m-values: (a) m = 0.12 (pure aluminum); (b) m=0.25 (annealed Al 1100); (c) m = 0.6
(copper), (d) m - 1.0 (copper).
Axisymmetric Isothermal Forging 163
164
Axisymmetric Isothermal Forging 165
experimental data with the so-called calibration curves, as shown in Fig.
9.9. This procedure, however, raises some questions regarding the
accuracy and efficiency of this evaluation scheme, because the curves are
not unique and depend on strain, strain-rate, and the thermal characteris-
tics of the material, and also on the specimen geometry. Furthermore,
experimental data usually do not follow the trend of predicted curve for a
constant friction value. Thus, Hwang and Kobayashi [21] proposed a
method of direct evaluation of the friction value from the experimental
measurements in ring tests. The method uses the finite-element technique,
and includes fitting curves to the experimental data with an iteration
scheme for evaluating a current friction value based on measured changes
of the ring dimensions. This scheme for evaluating friction was applied to
the experimental data of DePierre and Gurney [22] for annealed aluminum
1100, for a ring geometry of 6:2:1 (outer diameter:inner diameter:thick-
ness). One set of experiments was conducted with Johnson's wax as a
lubricant and another without a lubricant.
The results for the case with a lubricant are shown in Fig. 9.12a. The
upper part shows experimental points and the fitted curve, and the lower
part shows the computed variation of the friction factor m, which only
ranges from m = 0.1 to 0.2. The results for the case without a lubricant are
shown in Fig. 9.12b. It can be seen that the variation in friction value is
large and the estimation of friction from the calibration curves generated
for constant friction values results in an erroneous interpretation of the
experimental data.
Cabbaging [25]
In the initial stage of backward extrusion, called cabbaging, a round billet
is placed in a container. A punch is used to upset and partially pierce one
(b)
FIG. 9.12 Variation of friction value determined from experimental data [22]: (a) lubricated
with wax; (b) without any lubricant.
166
FIG. 9.13 Axisymmetric spike forging: (a) undeformed grid; (b) deformation at a die stroke
of 0.58H0 for two m values (H0 is initial billet height) [24].
167
168 Metal Forming and the Finite-Element Method
end of the billet. Thus, the outside diameter of the billet is sized, and the
pierced recess on top of the billet provides punch guidance for the
subsequent backward extrusion operation. The cabbaging operation has
been simulated using the actual production conditions with minor assump-
tions. The undeformed billet had a diameter of 73.7mm (2.94 in.) and a
height of 354mm (13.85 in.) The billet material was AISI 1046 steel, and
the operation was done at 1100°C (2012°F). The frictional shear factor
used in the analysis was m = 0.3. The punch speed used for the simulation
was 59.3 mm/s (2.3 in/s).
Figure 9.14 shows the undeformed grid lines, the calculated grid
distortions at various punch displacements, and the predicted punch load
versus displacement curve. Because of the unusually high strain concentra-
tion, the element near the punch tip underwent too much distortion, as can
be seen in Fig. 9.14. A partial "remeshing" near the punch tip was done at
a punch displacement of 91.4 mm (3.6 in.).
Compressor Disk Forging [26]
A simulation of a compressor disk forging from Ti 6242 alloy is illustrated
in Fig. 9.15. Because of symmetry, only a quarter of the disk is shown. The
cylindrical preform shape used in this analysis was 158.8mm (6.5 in.) in
FIG. 9.14 Predicted grids and punch load-displacement curve for the cabbaging process [25].
Axisymmetric Isothermal Forging 169
diameter and 63.5 mm (2.5 in.) in height. The velocity of the upper die
used in the simulation was 5.1 mm/min (0.2in./min). The bottom die was
stationary.
Forging was done isothermally at 900°C (1650°F), with an average
nominal strain-rate of about 0.175mm"1. The results show that the
finite-element method can be used effectively for simulating disk-type
forgings and for predicting strains, strain-rates, and stresses for a given
preform shape.
170 Metal Forming and the Finite-Element Method
FIG. 9.16 Metal flow patterns in flashless forging of a gear blank at 0, 40, 60, and 78%
reductions in initial billet height.
Axisymmetric Isothemal Forging 171
FIG. 9.17 Distorted grid (right half) and "remeshed" grid (left half) at 61% reduction in
billet height for flashless forging.
FIG. 9.18 Grid distortions near completion of die filling for flashless forging.
172 Metal Forming and the Finite-Element Method
References
1. Spies, K., (1957), "The Preforms in Closed-Die Forging and Their Preparation
by Reducer Rolling," (in German) Doctoral Dissertation, Technical University
Hannover.
2. Lange, K., (1958), "Closed-Die Forging of Steel," (in German), Springer-
Verlag, Berlin.
3. Kudo, H., and Aoi, K., (1967), "Effect of Compression Test Condition upon
Fracturing of a Medium Carbon Steel—Study on Cold Forgeability Test: Part
II," (in Japanese), J. Japan Soc. Tech. Plast., Vol. 8, p. 17.
4. Nagamatsu, A., Murota, T., and Jimma, T., (1971), "On the Nonuniform
Deformation of Material in Axially Symmetric Compression Caused by
Friction," Bull. JSME, Vol. 14, p. 331.
5. Lee, C. H., and Kobayashi, S., (1971), "Analysis of Axisymmetric Upsetting
and Plane-strain Side-pressing of Solid Cylinders by the Finite-element
Method," Trans. ASME, J. Engr. Ind., Vol. 93, p. 445.
6. Kobayashi, S., and Lee, C. H., (1973), "Deformation Mechanics and
Workability in Upsetting Solid Circular Cylinders," Proc. North American
Metalworking Research Conference, Vol. 1, p. 185.
7. Shah, S. N., Lee, C. H., and Kobayashi, S., (1974), "Compression of Tall
Circular, Solid Cylinders Between Parallel Flat Dies," Proc. International
Conference on Production Engineering, Tokyo, p. 295.
8. Lecocq, A. G., (1971), "Stresses in the Shank of a Bolt During Cold
Heading," Wire (English version of Draft) Coburg, Vol. 115, p. 197.
9. Thomason, P. F., (1969-70), "The Effect of Heat Treatment on the Ductility
in a Cold Heading Process," Proc. Inst. Mech. Engr., Vol. 184, p. 875.
10. Gill, F. L., and Baldwin, W. M., (1964), "Proper Wire Drawing Improves
Cold Heading Process," Metal Progress, Vol. 85, p. 83.
11. Shah, S. N., and Kobayashi, S., (1974), "Rigid-Plastic Analysis in Cold
Heading by the Matrix Method," Proc. 15th International Machine Tool
Design Research Conf., p. 603.
12. Kunogi, M., (1954), "On Plastic Deformation of Hollow Cylinders under
Axial Compressive Loading,"/. Sci. Res. Inst., Tokyo, p. 2.
13. Avitzur, B., (1983), "Handbook of Metal Forming Processes," Wiley, New
York.
14. Chen, C. C., and Kobayashi, S., (1978), "Rigid-plastic Finite-element Analysis
of Ring Compression," Application of Numerical Methods to Forming
Processes, ASME, AMD, Vol. 28, p. 163.
15. Hartley, P., Sturgess, C. E. N., and Rowe, G. W., (1979), "Friction in
Finite-element Analysis of Metal-forming Processes," Int. J. Mech. Sci., Vol.
21, p. 301.
16. Hartley, P., Sturgess, C. E. N., and Rowe, G. W., (1979), "An Examination
of Frictional Boundary Conditions and Their Effect in an Elastic-plastic
Finite-element Solution," Proc. MTDR, p. 157.
17. Male, A. T., and DePierre, V., (1970), "The Validity of Mathematical
Solutions for Determining Friction from the Ring Compression Test,"
Lubrication Technology, p. 389.
18. Schey, J. A., Ed., (1970), "Metal Deformation Processes—Friction and
Lubrication," Marcel Dekker, New York.
19. Thomsen, E. G., (1969), "Friction in Forming Processes," Annals CIRP, Vol.
17, p. 149.
20. Oh, S. I., Thomsen, E. G., and Kobayashi, S., (1975), "Calculation of
Frictional Stress Distributions at the Die-Workpiece Interface in Simple
Upsetting," Proc. 3d North American Metalworking Research Conference, p.
159.
Axisymmetric Isothermal Forging 173
21. Hwang, S. M., and Kobayashi, S., (1983), "A Note on Evaluation of Interface
Friction in Ring Tests," Proc. NAMRC XI, University of Wisconsin, Madison,
WI, p. 193.
22. DePierre, V., and Gurney, F., (1972), "A Method for Determination of
Constant and Varying Friction Factors During Ring Compression Tests," Air
Force Materials Laboratory, Report AFML-TR-72-37.
23. Oh, S. I., (1982), "Finite Element Analysis of Metal Forming Problems with
Arbitrarily Shaped Dies," Int. J. Mech. Sci., Vol. 17, p. 293.
24. Oh, S. I., Lahoti, G. D., and Altan, T., (1981), "ALPID—A General Purpose
FEM Program for Metal Forming," Proc. NAMRC-IX, State College, PA, p.
83.
25. Oh, S. I., Lahoti, G. D., and Altan, T., (1982), "Analysis of Backward
Extrusion Process by the Finite Element Method," Proc. NAMRC—X,
Hamilton, Canada, p. 143.
26. Oh, S. I., Park, J. J., Kobayashi, S., and Altan, T., (1983), "Application of
FEM Modeling to Simulate Metal Flow in Forging a Titanium Alloy Engine
Disk," Trans. ASME, J. Engr. Ind., Vol. 105, p. 251.
27. Badawy, A., Oh, S. I., and Altan, T., (1983), "A Remeshing Technique for
the FEM Simulation of Metal Forming Processes," Proc. Int. Comp. Eng.
Conf., ASME, Chicago, p. 143.
10
STEADY-STATE PROCESSES
OF EXTRUSION AND DRAWING
10.1 Introduction
Except at the start and the end of the deformation, processes such as
extrusion, drawing, and rolling are kinematically steady state. Steady-state
solutions in these processes are needed for equipment design and die
design and for controlling product properties.
A variety of solutions for different conditions in extrusion and drawing
have been obtained by applying the slip-line theory and the upper-bound
theorems [1-3]. Early applications of the finite-element method to the
analysis of extrusion [4-6] have been for the loading of a workpiece that
fits the die and container, and for the extrusion of a small amount of it
rather than extruding the workpiece until a steady state is reached. An
exception is the work by Lee et al. [7] for plane-strain extrusion with
frictionless curved dies using the elastic-plastic finite-element method. In
view of the computational efficiency, various numerical procedures par-
ticularly suited for the analysis of steady-state processes have been
developed by several investigators [8-14]. Shah and Kobayashi [8]
analyzed axisymmetric extrusion through frictionless conical dies by the
rigid-plastic finite-element method. The technique involves construction of
the flow lines from velocities and integration of strain-rates numerically
along flow lines to determine the strain distributions. An improvement of
the method was made by including friction at the die-workpiece interface.
The steady-state deformation characteristics in extrusion and drawing were
obtained as functions of material property, die-workpiece interface
friction, die angle, and reduction [15].
FIG. 10.1 Boundary conditions and mesh system for steady state axisymmetric extrusion
analysis.
Steady-State Processes of Extrusion and Drawing 177
Results
The computations were carried out under various extrusion process
conditions for SAE 1112 steel and aluminum alloy 2024-T351. Since the
solution with nonwork-hardening material was used as the initial guess in
analyzing the extrusion of work-hardening materials, the results for
nonwork-hardening materials are also discussed. The computation was
performed for each solution until the error norm of ||Av||/||v|| = 0.000 08
was reached. The number of iterations to reach the above convergence
depends on the initial guess, but the average number of iterations required
for the final solution was about 25-30.
The predicted results were obtained for average extrusion pressure,
normal pressure distribution on the die, grid distortions, and for velocity,
stress, and strain-rate distributions. Some of these are discussed below.
Detailed differences in deformation and f}ow behavior, due to material
properties and friction at the die-workpiece interface, are clearly indicated
in calculated grid distortions in Fig. 10.2. The steady-state grid distortion
patterns are compared for nonwork-hardening and work-hardening (SAE
1112 steel) cases for two friction conditions. Figure 10.2a shows the
distortion of the grid lines that are originally perpendicular to the axis of
the workpiece. It is seen that there is a cusp or double peak on the axis in
the extruded part for both the work-hardening and nonwork-hardening
materials. However, the magnitude of the cusp is smaller with the former
material. With increasing friction at the die-workpiece interface, the
FIG. 10.2 Grid distortion patterns for the extrusion of nonwork-hardening (right halves) and
work-hardening (left halves) materials with (a) frictional stress fs = 0, and (b) fs = 0.4Y0: die
semi cone angle, 30°, R0/R,, = 2.37.
178 Metal Forming and the Finite-Element Method
180
Steady-State Processes of Extrusion and Drawing 181
FIG. 10.5 The mean stress distribution in extrusion through a frictionless die at R0/RC =
1.25.
FIG. 10.6 The die pressure distributions in drawing for various reductions and for the two die
friction conditions (die semi-cone angle = 6°).
FIG. 10.7 Comparison of grid distortions in extrusion of pure aluminum for the first and
second passes (/, = 0).
184 Metal Forming and the Finite-Element Method
FIG. 10.8 Stress and srain histories of a material point along the axis of (frictionless)
extrusion.
185
186 Metal Forming and the Finite-Element Method
FIG. 10.10 Stress and strain histories of a material point along the axis of drawing.
References
1. Johnson, W., Sowerby, R., and Venter, R. D., (1982), "Plane-Strain Slip Line
Fields for Metal Deformation Processes," Pergamon Press, Oxford.
2. Johnson, W., and Kudo, H., (1962), "The Mechanics of Metal Extrusion,"
Manchester University Press, Manchester, UK,
3. Avitzur, B., (1983), "Handbook of Metal Forming Processes," Wiley, New
York.
4. Murota, T., Jimma, T., and Kato, K., (1970), "Analysis of Axisymmetric
Extrusion," Bull. JSME, Vol. 13, p. 1366.
5. Iwata, K., Osakada, K., and Fujino, S., (1972), "Analysis of Hydrostatic
Extrusion by the Finite Element Method," Trans. ASME, J. Engr. Ind., Vol.
94, p. 697.
6. Lee, C. H., Iwasaki, H., and Kobayashi, S., (1973), "Calculation of Residual
Stresses in Plastic Deformation Processes," Trans. ASME, J. Engr. Ind., Vol.
95, p. 283.
7. Lee, E. H., Mallet, R. L., and Yang, W. H., (1976), "Stress and Deformation
Analysis of the Metal Extrusion Process," SUDAM No. 76-2, Stanford
University, June.
8. Shah, S. N., and Kobayashi, S., (1977), "A Theory on Metal Flow in
Axisymmetric Piercing and Extrusion," J. Prod. Engr., Vol. 1, p. 73.
9. Zienkiewicz, O. C., and Godbole, P. N., (1974), "Flow of Plastic and
Viscoplastic Solids with Special Reference to Extrusion and Forming Proc-
esses," Int. J. Num. Meth. Engr., Vol. 8, p. 3.
10. Zienkiewicz, O. C., Jain, P. C., and Onate, E., (1978), "Flow of Solids
During Forming and Extrusion; Some Aspects of Numerical Solutions," Int. J.
Solids Structures, Vol. 14, p. 15.
11. Zienkiewicz, O. C., Onate, E., and Heinrich, J. C., (1978), "Plastic Flow in
Metal Forming, I. Coupled Thermal Behavior in Extrusion. II Thin Sheet
Forming," Applications of Numerical Methods to Forming Processes, ASME,
AMD, Vol. 28, p. 107.
12. Dawson, P. R., (1978), "Viscoplastic Finite Element Analysis of Steady State
Forming Processes including Strain History and Stress Flux Dependence,"
Applications of Numerical Methods to Forming Processes, ASME, AMD,
Vol. 28, p. 55.
13. Dawson, P. R., and Thompson, E. G., (1978), "Finite Element Analysis of
Steady-State Elasto-Visco-Plastic Flow by the Initial Stress-Rate Method," Int.
J. Num. Meth. Engr., Vol. 12, p. 47.
14. Derbalian, K. A., Lee, E. H., Mallet, R. L., and McMeeking, R. M., (1978),
"Finite Element Metal Forming Analysis with Spacially Fixed Mesh," Appli-
cations of Numerical Methods to Forming Processes, ASME, AMD, Vol. 28,
p. 39.
15. Chen, C. C., Oh, S. I., and Kobayashi, S., (1979), "Ductile Fracture in
188 Metal Forming and the Finite-Element Method
11.1 Introduction
The stress-state is said to be plane when the direction normal to the plane
is a principal stress direction and the magnitude of the stress in this
direction is zero. This situation occurs when a sheet is loaded along its
edges in the plane of the sheet. In-plane deformation of sheet metal, such
as bore expanding and flange-drawing, is an example of plane-stress
problems. For out-of-plane deformation of sheet metals, such as punch
stretching, sheet bending, and cup drawing, a simple analytical method is
the use of membrane theory. This theory neglects stress variations in the
thickness direction of a sheet and considers the distribution of stress
components only in the plane of the sheet. Thus, the basic formulations for
the analysis of both in-plane and out-of-plane deformations contain only
the stress components acting in the plane of the sheet. However, the
analysis of out-of-plane deformation requires consideration of large
deformation, while the infinitesimal theory is applicable for in-plane
deformation analysis.
Many materials employed in engineering applications possess mechani-
cal properties that are direction-dependent. This property, termed
anisotropy, stems from the metallurgical structure of the material, which
depends on the nature of alloying elements and the conditions of
mechanical and thermal treatments. Metal sheets are usually cold-rolled
and possess different properties in the rolled and transverse directions.
Therefore, in sheet-metal forming in particular, the effect of anisotropy on
the deformation characteristics may be quite appreciable and important.
In the past the calculation of the detailed mechanics of large plastic
deformation of metal sheets has been achieved with some success by
numerical methods. However, without exception, these studies have dealt
with deformations that possess a high degree of symmetry, and were
concerned with the anisotropy existing only in the direction of sheet
thickness (normal anisotropy). Methods that are capable of solving
nonaxisymmetric problems in forming of anisotropic sheet metal are still
being sought. The finite-element method is one of those methods. It was
applied to the elastic-plastic analysis of nonaxisymmetric configurations of
sheet stretching with normal anisotropy by Mehta and Kobayashi [1].
Yamada [2] presented a stress-strain matrix for a material that is
189
190 Metal Forming and the Finite-Element Method
elastically isotropic, and that obeys Hill's anisotropic yield criterion in the
plastic range. He treated the incipient deformation of a circular blank of
anisotropic material in the flange drawing process. For the deformation
analysis of a metal sheet having three mutually perpendicular axes of
anisotropy (orthotropic material), the rigid-plastic finite-element method
has been applied to plane-stress bore expanding and flange drawing [3,4].
To analyze out-of-plane forming processes, Wang [5] proposed a method
of solution in which two spatially independent variables are required to
define the geometry. The method is based on a variational procedure and
assumes that the material of the sheet is rigid-plastic. On the same
principle, Kim et al. [6, 7] analyzed three processes, namely, the bulging of
a sheet subjected to hydrostatic pressure, the stretching of a sheet with a
hemispherical punch, and the deep drawing of a sheet with a hemispherical
punch.
Using the elastic-plastic approach, complete solutions of stretch-forming
and deep-drawing problems, taking into account the contact problem at
the blank holder, die, die profile, and punch head, were obtained by Win
[8]. On the basis of the nonlinear theory of membrane shells, Wang and
Budiansky [9] developed a procedure for calculating the deformations in
the stamping of sheet metal by arbitrarily shaped punches and dies.
Onate and Zienkiewicz [10] presented a finite-element formulation
based on an extension of the general viscoplastic flow theory for
continuum problems to deal with thin shells.
Toh and Kobayashi [11,12] analyzed sheet-metal forming processes,
axially symmetric and nonsymmetric, by the finite-element method based
on the membrane theory. The finite-element model takes into account the
rigid-plastic material characteristics and includes the normal anisotropy of
the sheet metal as well as the finite deformation that occurs during the
sheet-forming process.
In this chapter, first the yield function for anisotropic materials is
defined and in-plane deformation processes are analyzed using the
infinitesimal theory. Then the large-strain formulation using membrane
theory is described for the analysis of out-of-plane deformation processes.
The formulation is extended to sheet-metal forming of general shapes and
applied to square-cup drawing. Finally, in Section 11.8 we discuss the
nonquadratic yield function for anisotropic materials.
where the orthotropic axes are taken as the coordinate axes (x, y, z) and
F, G, H, L, M, N are anisotropy parameters. The parameters in eq. (11.1)
are not definite but their ratios are for defining the behavior of a given
material. Thus, eq. (11.1) is rewritten as
then
Under plane-stress conditions (oz = TZX = ryz = 0), eq. (11.2) reduces to
where o is the tensile yield stress. The ratio of the transverse to the
through-thickness strain (or strain-rate) is
Substitution of eq. (11.5), using the stresses given by eq. (11.8), into (11.9)
results in
where rx, r45 and ry are the /--values for a = 0° (x direction), 45°, and 90° (y
direction), respectively. For the anisotropy to be rotationally symmetric
about the z axis, f=g and n=f + 2h (and / = m) in addition to the
condition given by eq. (11.3b). If there is complete spherical symmetry, or
isotropy,
differs for anisotropic materials from that for isotropic materials. With
reference to eq. (11.7), the matrix D for orthotropic materials, becomes
and
194 Metal Forming and the Finite-Element Method
and
In Fig. 11.2, the strain distributions for normal anisotropy and those in
the direction a = 45° for planar anisotropy are compared (ah the angle
from the principal axis of anisotropy). The distributions in the directions of
a = 5° and 85° are almost identical to those obtained for normal aniso-
tropy. The formation of "ears" and "hollows" in flange drawing is shown
in Fig. 11.3a and the corresponding strain contours are given in Fig. 11.3b.
The ears appear along the axes of anisotropy and a hollow is formed at
approximately a = 47° at an early stage and changes its location slightly
near the end of the process. According to Hill [13], the principal axes of
stress and strain-increment coincide for the orientations a = 0°, a ( = 49°),
and 90° for the present planar anisotropy; the ears are formed at a = 0°
and 90°, and a hollow at a = (90 - a)° = 41°. The location of the hollow,
shown in Fig. 11.3, is not in agreement with Hill's prediction, but is closer
to that for minimum r-value (a = 51°).
R A D I U S / R0
FIG. 11.2 Strain distributions at various deformation stages of bore expanding.
where «0 is a constant.
It is easily shown that all the components of strain rate, denned by
£ij = 2(MI,/ + "/,;)> become zero, even though the length of the sheet would
change after a finite time interval A?. This suggests that there is at least
one deformation mode that cannot be determined by the variational
principle formulated by eq. (11.11). The zero-energy deformation mode
tFrom Section 11.4 through Section 11.7, where finite strain formulation is utilized, u, is
used as a component of displacement, ut for a velocity component, and v, for a nodal point
displacement.
FIG. 11.3 (a) Deformed grid pattern in flange drawing, (b) Effective strain contours in flange
drawing.
197
198 Metal Forming and the Finite-Element Method
o R
FIG. 11.4 Out-of-plane deformation of a ring element.
In eq. (11.16), A*0 and As are the lengths of an incremental line element
in the meridian direction before and after the time increment, respectively,
and R0 and R are the radial positions of a material point before and after
the time increment, respectively. The components of the strain-rate can be
expressed by
and
Note that the components of strain-rate and stress are functions of time,
and that the coordinate axis with which these quantities are denned rotates
with the material fiber. Owing to the work-hardening, a in eq. (11.18b) is
also a function of time and is expressed as
where H' = do/de and is the slope of the true stress vs. true strain curve of
the material.
Substituting eqs. (11.17) and (11.19) into eq. (11.15), we can readily
obtain the variational functional integrated over a time-increment At as
where T is the thickness of the sheet, A is the surface area of the sheet,
and the effective strain increment AE is expressed as
where u^a) and u^ are the radial and axial displacements at the ath node,
respectively, and the summation is performed for the two nodes. The
shape function qa is written in the natural coordinates as
The radial positions of a material point before and after the deformation,
,/?0 and R, can be expressed by
where
Sheet-Metal Forming 201
where
Assembling the eqs. (11.24) and (11.25) for all elements in the finite-
element scheme, we obtain the linear simultaneous equations
where r0 and z0 are radial and axial positions of a nodal point at the
present configuration, ur and uz are the increments of radial and axial
202 Metal Forming and the Finite-Element Method
z
where starred (*) quantities are initial guesses and Aw r , A«2 are perturba-
tions. By rearranging eq. (11.28) we have
where
and
FIG. 11.7 Comparison of the circumferential strain distributions between computed and
experimental [17] results, at three punch positions.
204 Metal Forming and the Finite-Element Method
FIG. 11.8 Comparison of the computed results with experiment [17] for thickness strain
distribution, at three punch positions.
and thickness strain. The agreement between the experimental data and
the numerical solution is reasonable, considering the fact that the exact
friction condition is not known.
2. Effect of friction. The parameters used for computation are:
• Stress-strain characteristics: a = Ce° 23 (C a constant)
• Normal anisotropy: r = 1.27
• Punch radius: Rp = 1.0
• Radius of the blank: Rb = 1.0
FIG. 11.9 Punch load vs. punch depth in deep drawing with hemispherical punch for
different coefficients of friction.
or more often, by the ratio of the blank diameter to the punch diameter.
This ratio is called the limiting drawing ratio and the test used to determine
the limiting drawing ratio is called the Swift test.
Deep drawing is not only a useful method of material testing, but also
one of the basic operations in sheet-metal stamping. In practice, various
shapes may be used for the bottom of the punch; however, most past
FIG. 11.11 Distribution of thickness strains for [ip=Q.Q4, ;irf = 0.04, and comparison with
experiment [18].
functional integrated over the finite time duration At. It has been shown in
Section 11.4 that the variational functional dn is integrated over the finite
time-increment A/ can be expressed in terms of the logarithmic strain-
increment, provided that the directions of principal strains are known in
advance. For a general out-of-plane deformation, the principal stretch
direction is not known in advance owing to the shear strain component in
the plane of the sheet.
FIG. 11.12 Distribution of circumferential strains for ft p =0.04, jUrf = 0.04, and comparison
with experiment [18].
208 Metal Forming and the Finite-Element Method
Punch Depth
FIG. 11.13 Punch load vs. punch depth, comparison of predictions with experiment [18].
where the dot indicates the time derivative. The Piola-Kirchhoff stress sap
is related to the Cauchy stress a,-,- according to
Noting that o,-, = s/,- at t = t0, the variational functional integrated over a
time-increment At is obtained, similarly to eq. (11.20), as
Sheet-Metal Forming 209
where h' = ds/dE, and s is defined in terms of s,y in the same form for a
given by eq. (11.13c).
The slope h' is related to H'( = dolde) according to h' =H' -2o(t0).
The derivation of this relationship can be found in References [11] and
[19].
The effective strain-increment can be written in the matrix form as
where
The matrix D in eq. (11.35) is identical to eq. (11.14c). Once the solution
of the displacement increment is obtained, the geometry is updated. The
updated geometry becomes the reference state for the next time increment
Af.
Finite-Element Formulation
Let the sheet-metal domain be decomposed into an assemblage of linear
triangular elements. A local Cartesian coordinate system is assigned to
each element in such a way that the (x-y) coordinate plane coincides with
that of the element plane and the z axis is normal to the element plane.
The displacement field within the element can be described by
where qa is the shape function of the linear triangular element given in eq.
(6.10) in Chap. 6. Equation (11.36) can be rewritten in a matrix form as
where vT={41), u«\ «<'>, u™, uf\ u(?, u?\ uf\ u< 3 >}.
The component of displacement gradient can be written as
The first derivative of the functional, eq. (11.34), with respect to the
nodal displacement can be written
where
Here, we have
etc.
Note that 3AE/3u7 is also a function of the nodal displacements,
because of the nonlinear terms involved in the expression of AE. The
solution of eq. (11.39) is obtained iteratively by the Newton-Raphson
method described in Chap. 7. The details of the stiffness matrix evaluation
can be found in Reference [19].
Since the element stiffness equations are evaluated based on the local
coordinate system, the elemental stiffness matrices should be transformed
into the global coordinate system before they are assembled.
FIG. 11.17 Comparison of the numerical solutions with the experimental data [20] for
thickness strain distributions in the transverse direction of a drawn cup.
213
214 Metal Forming and the Finite-Element Method
1/4 OF SQUARE CUP
FIG. 11.18 The square cup formed by the FEM simulation [19].
simulations are:
• Sheet material: AISI 304 stainless steel
• Stress-strain characteristic,CT= 219.6£°'43ksi
= 1514e° 43 MN/m 2
• r-value: 1.025
• Sheet thickness: 0.76 mm (0.03 in.)
• Coefficient of friction: \ip = \ID = 0.04
• Other dimensions for the punch and die are the same as those used in
drawing with square blanks.
Figure 11.19 shows the blank shapes and corresponding finite-element
meshes used in the process simulations. The same coefficient of friction
was used between the punch head, the die surface, and the sheet metal.
The deformation is controlled by incremental punch advancement and the
blank holding force is set equal to zero. The results of the computation are
compared with experiments [12].
The punch load versus displacement curves are shown in Fig. 11.20 for
various blank shapes. The finite-element solutions and experimental data
FIG. 11.19 Blank shapes and finite-element domains used in the square-cup drawing
simulations: (a) square; (b) octagonal; (c) circular.
215
216 Metal Forming and the Finite-Element Method
Punch Depth, mm
FIG. 11.20 Load-displacement curves in square-cup drawing processes with various initial
blank shapes.
FIG. 11.21 Thickness strain distributions in square cup. (a) Transverse direction; (b)
diagonal direction [12].
218
Sheet-Metal Forming 219
Under uniaxial tension (a1, o2) = (ou, 0), then eqs. (11.41) and (11.42)
give, respectively,
with the ordering of o-i > o2. In eq. (11.50) the effective strain-rate e is
introduced from
oj = 0^1 +CT2e2= 1(0! + o2)(i-l + e2) + 1(0! - CT^E! - £2).
The explicit formula for e can be obtained by eliminating ol and o2 from
eqs. (11.48) and (11.50) as
Wang [28] implemented the yield criterion, eq. (11.46), into the finite-
element procedure and applied the method for hemispherical punch
stretching to assess the effect of m-value on the resulting strain
distributions.
References
1. Mehta, H. S., and Kobayashi, S., (1971), "Finite-element Analysis and
Experimental Investigation of Sheet-metal Stretching," Rep. No. MD 71-2,
University of California.
2. Yamada, Y., (1969), "Recent Japanese Developments in Matrix Displacement
Method for Elasto-Plastic Problems," Paper presented at Japan-U.S. Seminar
on Matrix Methods of Structural Analysis and Design, Tokyo, Japan.
3. Lee, C. H., and Kobayashi, S., (1973), "New Solutions to Rigid-Plastic
Deformation Problems using a Matrix Method," Trans. ASME, J. Engr. Ind.,
Vol. 95, p. 865.
4. Lee, S. H., and Kobayashi, S., (1975), "Rigid-Plastic Analysis of Bore
Expanding and Flange Drawing with Anisotropic Sheet Metals by the Matrix
Method," Proc. 15th Int. Mach. Tool Des. Res. Conf. p. 561.
5. Wang, N. M., (1970), "A Variational Method for Problems of Large Plastic
Deformation of Metal Sheets," General Motors Research Publication GMR-
1038.
6. Kim, J. H., and Kobayashi, S., (1978), Deformation Analysis of Axisymmetric
Sheet Metal Forming Processes by the Rigid-Plastic Finite Element Method,
in, "Mechanics of Sheet Metal Forming," (Edited by D. P. Koistinen, and
N. M. Wang), Plenum Press, New York, p. 341.
7. Kim, J. H., Oh, S. I., and Kobayashi, S., (1978), "Analysis of Stretching of
Sheet Metals with Hemispherical Punch," Int. J. Machine Tool Des. Res., Vol
18, p. 209.
8. Win, S. A., (1976), "An Incremental Complete Solution of the Stretch-
forming and Deep-drawing of a Circular Blank Using a Hemispherical Punch,"
Int. J. Mech. ScL, Vol. 18, p. 23.
Sheet-Metal Forming 221
9. Wang, N. M., and Budiansky, B., (1978), "Analysis of Sheet Metal Stamping
by a Finite-element Method," General Motors Research Publication GMR-
2423.
10. Onate, E., and Zienkiewicz, O. C., (1983), "A Viscous Shell Formulation for
the Analysis of Thin Sheet Metal Forming," Int. J. Mech. ScL, Vol. 25, p. 305.
11. Toh, C. H., and Kobayashi, S., (1983), "Finite-element Process Modeling of
Sheet Metal Forming of General Shapes," Grundlagen der Umformtechnik I
Symposium, Stuttgart, p. 39.
12. Toh, C. H., and Kobayashi, S., (1985), "Deformation Analysis and Blank
Design in Square Cup Drawing," Int. J. Machine Tool Des. Res, Vol. 25, No.
1, p. 15.
13. Hill, R., (1950), "Mathematical Theory of Plasticity," Oxford University
Press, London.
14. Rebelo, N., and Kobayashi, S., (1980), "Axisymmetric Punch Stretching of
Strain Rate Sensitive Sheet Metals," Proc. 8th NAMRC, University of
Missouri, Rolla, MO., p. 235.
15. Park, J. J., Oh, S. I., and Altan, T., (1987), "Analyses of Axisymmetric Sheet
Forming Processes by Rigid-Viscoplastic Finite Element Method," Trans.
ASME, J. Engr. Ind., Vol. 109, p. 347.
16. Kim, J. H., (1977), "Analysis of Sheet Metal Forming by the Finite-Element
Method," Ph.D. Dissertation, University of California, Berkeley.
17. Kaftanoglu, B., and Alexander, J. M., (1970), "On Quasistatic Axisymmetri-
cal Stretch Forming," Int. J. Mech. Sci. Vol. 12, p. 1065.
18. Wood, D. M., (1968), "On the Complete Solutions of the Deep Drawing
Problem," Int. J. Mech. Sci. Vol. 10, p. 83.
19. Toh, C. H., (1983), "Process Modeling of Sheet Forming of General Shapes
by the Finite Element Method Based on Large Strain formulation," Ph.D.
Dissertation, University of California, Berkeley.
20. Thomson, T. R., (1975), "Influence of Material Properties in the Forming of
Square Shells," /. Australian Inst. Metals, Vol. 20, No. 2, p. 106.
21. Yoshida, K., and Miyauchi, K., (1978), Experimental Studies of Material
Behavior as Related to Sheet Metal Forming, in "Mechanics of Sheet Metal
Forming" (Edited by D. P. Koistinen, and N. M. Wang), Plenum Press, New
York, p. 19.
22. El-Walkil, D., Kamal, M. N. E., and Darwish, A. H., (1980), "Mechanics of
the Square Box Drawing Operation of Aluminum Blanks," Sheet Metal
Industry, August, p. 679.
23. Woodthorpe, J., and Pearce, R., (1970), "The Anomalous Behavior of
Aluminum Sheet under Balanced Biaxial Tension." Int. J. Mech. Sci., Vol. 12,
p. 341.
24. Hill, R., (1979), "Theoretical Plasticity of Textured Aggregates," Math. Proc.
Camb. Phil. Soc., Vol. 85, p. 179.
25. Parmar, A., and Mellor, P. B., (1978), "Predictions of Limit Strains in Sheet
Metal Using a More General Yield Criterion," Int. J. Mech. Sci., Vol. 20, p.
385.
26. Parmar, A., and Mellor, P. B., (1978), "Plastic Expansion of a Circular Hole
in Sheet Metal Subjected to Biaxial Tensile Stress," Int. J. Mech. Sci., Vol. 20,
p. 707.
27. Kobayashi, S., Caddell, R. M., and Hosford, W. F., (1985), "Examination of
Hill's Latest Yield Criterion Using Experimental Data for Various Anisotropic
Sheet Metals," Int. J. Mech. Sci., Vol. 27, p. 509.
28. Wang, N. M., (1984), A Rigid-Plastic Rate-sensitive Finite Element Method
for Modeling Sheet Metal Forming Processes, in "Numerical Analysis of
Forming Processes," (Edited by J. F. T. Pittman et al.), Wiley, New York, p.
117.
12
THERMO-VISCOPLASTIC ANALYSIS
12.1 Introduction
The main concern here is the analysis of plastic deformation processes in
the warm and hot forming regimes. When deformation takes place at high
temperatures, material properties can vary considerably with temperature.
Heat is generated during a metal-forming process, and if dies are at a
considerably lower temperature than the workpiece, the heat loss by
conduction to the dies and by radiation and convection to the environment
can result in severe temperature gradients within the workpiece. Thus, the
consideration of temperature effects in the analysis of metal-forming
problems is very important. Furthermore, at elevated temperatures, plastic
deformation can induce phase transformations and alterations in grain
structures that, in turn, can modify the flow stress of the workpiece
material as well as other mechanical properties.
Since materials at elevated temperatures are usually rate-sensitive, a
complete analysis of hot forming requires two considerations—the effect of
the rate-sensitivity of materials and the coupling of the metal flow and heat
transfer analyses.
A material behavior that exhibits rate sensitivity is called viscoplastic. A
theory that deals with viscoplasticity was described in Chap. 4. It was
shown that the governing equations for deformation of viscoplastic
materials are formally identical to those of plastic materials, except that
the effective stress is a function of strain, strain-rate, and temperature. The
application of the finite-element method to the analysis of metal-forming
processes using rigid-plastic materials leads to a simple extension of the
method to rigid-viscoplastic materials [1].
The importance of temperature calculations during a metal-forming
process has been recognized for a long time. Until recently, the majority of
the work had been based on procedures that uncouple the problem of heat
transfer from the metal deformation problem. Several researchers have
used the following approach. They determined the flow velocity fields in
the problem either experimentally or by calculations, and they then used
these fields to calculate heat generation. Examples of this approach are the
works of Johnson and Kudo [2] on extrusion, and of Tay et al. [3] on
machining. Another approach [4] uses Bishop's numerical method in which
222
Thermo-Viscoplastic Analysis 223
FIG. 12.1 Comparison of the finite-element rigid-plastic analysis (•) with dynamic analysis
( ) [18] aid experiment (—) [18].
Thermo-Viscoplastic Analysis 225
FIG. 12.2 Comparison of the finite element rigid-viscoplastic analysis (•) with dynamic
analysis ( ) [18] and experiment (—) [18].
where qn is the heat flux across the boundary surface 5g, n denotes the unit
normal to the boundary surface, and
and
The first term on the right is the heat, generated by plastic deformation
inside the deforming body. The second term defines the contribution of the
heat radiated from the environment to the element, where a is the
Stefan-Boltzman constant, e is the emissivity, and Te and Ts are environ-
ment and surface temperatures, respectively. The third term describes the
heat convected from the body surface to the environment with heat
convection coefficient h. The fourth term represents the contribution of the
heat transferred from the workpiece to the die through their interface. Td
and Tw are die and workpiece temperatures, respectively, and /ilub is the
heat transfer coefficient for the lubricant. The last term is the contribution
of the heat generated by friction along the die-workpiece interface, qf
being the surface heat generation rate due to friction.
The theory necessary to integrate (12.10) can be found in numerical
analysis books; see for instance Dahlquist and Bjorck, [21]. The conver-
gence of a scheme requires consistency and stability. Consistency is
satisfied by an approximation of the type
where fs is the friction stress, and \us\ the relative velocity between die and
workpiece. The heat qf is evenly distributed between the die and
deforming material.
It should be noted that the nodes on the die do not generally coincide
with those on the deforming material along the interface, and that the
calculation of nodal point temperatues requires interpolation. When
boundary conditions such as the convection term in eq. (12.11) apply, they
are split into two parts, wherein one containing the unknown temperatures
is added to the matrix Kc. Boundary conditions such as the radiation term
in eq. (12.11) are applied using previous iteration values for body
temperatures.
The equations for the flow analysis and the temperature calculation are
strongly coupled, making a simultaneous solution of their finite-element
counterparts necessary.
Considering Tt+&t as a primary dependent variable, we have, from eq.
(12.12), with t = 0 initially,
The coupling procedure makes use of eq. (12.15) through the following
sequence.
1. Assume the initial temperature field T0.
2. Calculate the initial velocity field u corresponding to the tempera-
ture field T0.
3. Calculate the initial temperature-rate field T0 from eq. (12.10) using
values from (1) and (2).
4. Calculate the quantity T.
5. Update the nodal point positions and the effective strain of elements
for the next step.
6. Use the velocity field at the previous step to calculate the first
approximate temperature T such as
Thermo-Viscoplastic Analysis 229
9. Repeat steps for (7) and (8) until both have converged.
10. Calculate the new temperature rate field TA,.
11. Repeat steps from (5) to (10) until the desired deformation state is
reached.
The iteration process for temperature calculations is not likely to require
much computing time, because only the heat input vector Q is changed
during iterations and, as a result triangularization of the matrix is
necessary only once. Moreover, additional iterations necessary to obtain a
velocity field after a new temperature field is obtained should be relatively
few, because the velocity field does not show much sensitivity to small
variations of the temperature field.
12.5 Applications
Applications of the thermo-viscoplastic analysis include compression of a
solid steel cylinder [11,22], hot nosing of a steel shells [23], and forging of
titanium alloys [24, 25].
Compression of Steel Cylinder [11]
Pohl [26] conducted temperature measurements in order to test his
uncoupled analysis, in which approximate stream functions were used for
the deformation, and finite differences for the heat balance. A solid
cylinder of a carbon steel AISI 1015 was compressed between flat dies at
room temperature. Thermocouples were inserted in the cylinder at
different locations. Upon deformation, their readings indicated the tem-
perature increases due to the heat generation. Figure 12.3 shows the
dimensions and locations of measuring points. The conditions used in
computations with the finite-element method [11] were as follows. The
deformation took place at room temperature, until a reduction of 33% in
height was achieved. The finite-element grid was composed of 132
four-node quadrilateral elements in the workpiece, and 119 in the die.
Because of symmetry, only one-quarter of the cylinder needed to be
analyzed. The friction factor m was taken as 0.65.
The die velocity was changed at each time-step to simulate a mechanical
press. Each step corresponded to 1% reduction in height, which was
equivalent to time-steps of up to 0.03 s. The flow stress was considered to
be independent of strain-rate and temperature and its values were given by
Pohl. The heat transfer characteristics, other than the thermal conductivity
and the heat capacity of the AISI 1015, which were given by Pohl, were
taken from handbooks.
FIG. 12.3 Compression of a steel cylinder: specimen geometry and thermocouple locations, and comparison of temperature distribution between theory (—)
and experiments [26].
Thermo-Viscoplastic Analysis 231
The dimensions of the die were such that at the outside boundaries a
constant temperature was imposed. The temperature values measured by
Pohl are compared with the calculated values in Fig. 12.3. The agreements
are excellent for the internal points. For the three points near the billet
surface, however, the computed results indicate that the temperature
difference for these points is small, while experiments show larger
differences. This discrepancy may be attributed to inaccuracies in the
material constants used in computations or to inaccuracies in experimental
measurements.
Hot Compression of Steel Cylinder
Wu and Oh [22] developed an FEM-based computer program (ALPIDT)
that is capable of simulating nonisothermal forming operations with
arbitrarily shaped workpieces and dies. This program consists of two
independent FEM programs, ALPID for viscoplastic deformation analysis
[27] and the program for the heat transfer analysis. They are coupled in an
efficient manner in the program ALPIDT for simulation of thermo-
viscoplastic deformation. To demonstrate the capability of ALPIDT, a hot
compression process was simulated. The temperature changes during the
initial resting and dwelling periods were included in the simulation. The
simulation also accounted for the changes in the heat transfer coefficients
between the workpiece and the die during the process.
A cylindrical billet of AISI 1020 steel was compressed between two flat
dies. The initial billet temperature was 1232°C (2250°F) and the initial die
temperature was 204°C (400°F). In order to estimate the workpiece
temperature accurately, simulation was performed in the process consisting
of three periods as follows:
1. The heated billet is placed on the lower die for 6 seconds without
deformation (free resting period).
2. The workpiece is compressed to 67% in height (deformation period),
1.5 seconds).
3. The deformed workpiece stays on the lower die for 3 seconds after
the upper die is retracted (dwelling period).
Detailed computational conditions are given in Reference [22].
In Fig. 12.4 photographs of grid-distortions and temperature distribu-
tions are shown at various stages of the process. The temperature scale
(°F) is also shown in the figure.
The predicted temperatures at the end of the resting period are shown in
Fig. 12.4a. Owing to relatively small loss of heat to the environment, the
temperatures of the upper die and of the upper portion of the workpiece
remained almost unchanged. However, heat loss from the bottom portion
of the billet to the lower die was considerable. The temperature at the
bottom of the workpiece dropped by 280°C (536°F) during the free resting
period of 6 s. At the same time, the surface of the lower die was heated to
nearly 600°C (1112°F). It is to be noted that the flow stress of the
232 Metal Forming and the Finite-Element Method
workpiece material was doubled when the temperature was reduced from
1230° to 950°C (2246° to 1742°F).
The predicted temperature distributions and grid distortions during
the deformation period are shown in Figs. 12.4b to e. At the beginning of
the deformation period, the temperature of the workpiece at the upper die
interface was higher than that at the lower die interface. However, the
temperature of the top surface dropped quickly after the upper die
contacted the workpiece. The bottom and top surface temperatures of the
workpiece become almost the same, reaching 700-800°C (1292-1472°F) at
0.375 s after deformation started. However, it is noted that the tempera-
ture gradient at the top was higher than at the bottom of the workpiece.
These differences between the temperature distributions in the upper and
lower regions of the workpiece became less as the deformation proceeded
further. The temperature distributions shown in Figs 12.4b to e suggest
that the flow stress of workpiece near the die workpiece interface could
have been three times that in the mid-height region.
The predicted grid distortions reflect the effect of temperature on the
flow stress of the workpiece, influencing metal flow. During the early stage
of deformation (see Figs. 12.4b and c), the barreling near the top surface
was more pronounced than near the bottom surface. It is also noticed that
the upper surface moved radially more than the bottom surface, suggesting
that the average flow stress was higher near the bottom surface. It can be
also seen that, throughout the deformation period, the deformation
occurred mainly in the mid-height region, while the chilled regions near
the die workpiece interfaces remained almost rigid (Figs. 12.4d and e).
Figure 12.4f shows the temperature distributions at 3 s after the upper
die was retracted from the workpiece at the end of deformation. It is seen
from the figure that the top surface of the workpiece has warmed up again,
while the average temperature of the workpiece had dropped. This was
due to the redistribution of heat within the workpiece, without die chilling
at the top surface of the workpiece. The predictions shown in Fig. 12.4
agree with general observations made in nonisothermal forging of cylindri-
cal billets.
Forging of Titanium Alloy 776242
The preform considered is a cylindrical composite material with a central
core of (a + ft} phase and an outer ring of (/^-transformed phase, the two
diffusion-bonded together. Its dimensions are given in Fig. 12.5a. The
preform was forged isothermally (dies and workpiece at the same initial
temperature) at 1227 K (1750°F) at a constant ram speed of 5.08mm/min.
(0.2in./min). The total reduction in height was 60%. At these high
temperatures, a glass-type lubricant is very effective; a friction factor of
m = 0.2 was used in computer simulations.
In isothermal forging, very slow speeds are usually employed in order to
(1) avoid increasing the flow stress of the material, which is strain-rate
dependent, and (2) allow the heat generated during deformation to spread
FIG. 12.4 Predicted temperature distributions and grid distortions at various stages of hot
compression process [22], (a) At the end of free resting (elapsed time t ~ 6s); (b) 16.67%
reduction in height (/ - 6,375s); (c) 33,34% reduction in height (/ - 6.750s); (d) 50.00%
reduction in height (i = 7.125s); (e) 66.67% reduction in height (/ = 7.500s); (f) at the end
of 3s free resting after deformation (/ - 10.5s).
This page intentionally left blank
Thermo-Viscoplastic Analysis 235
FIG. 12.5 Die and workpiece temperatures in forging a titanium 6242 alloy composite billet
at 60% reduction in height with different ram speeds: (a) preform dimensions; (b) ram
speed = 0.2 in./min.; (c) ram speed = 2.0 in./min.
FIG. 12.6 Local strain, strain-rate, and temperature variations during (a + f})/(i composite
forging and corresponding microstructures after forging. (Microstructures courtesy of C. C.
Chen, Chen Tech. Industries.)
Thermo-Viscoplastic Analysis 237
238
Thermo-Viscoplastic Analysis 239
FIG. 12.9 Temperature distributions obtained by induction heating for a shell [28].
FIG. 12.10 Load-displacement curves for (a) various friction conditions and (b) different
initial tip temperatures, and comparison with experiment [28].
that develops during hot forging. Furthermore, Gegel et al. [31] proposed
a new method of modeling the dynamic material behavior that explicitly
described the dynamic metallurgical processes occurring during hot
deformation.
Recently, Dawson [32] developed the numerical solution formulation for
242 Metal Forming and the Finite-Element Method
simulation of hot or warm metal forming under steady-state conditions. Of
particular importance was the incorporation of a methodology for using
particle stress-temperature trajectories in conjunction with deformation
mechanism maps. Thus, the assumptions made regarding the assumed
constitutive equations could be evaluated. The analysis of slab rolling of
aluminum was given as an example of possible applications. A significant
advance, made by Dewhurst and Dawson [33], is the development of a
finite-element program that models steady-state viscoplastic flow and heat
transfer in three dimensions.
References
1. Oh, S. I., Rebelo, N. M., and Kobayashi, S., (1978), "Finite-Element
Formulation for the Analysis of Plastic Deformation of Rate-Sensitive Mate-
rials in Metal Forming," IUTAM Symposium, Tutzing/Germany, p. 273.
2. Johnson, W., and Kudo, H., (1960), "The Use of Upper-Bound Solutions for
the Determination of Temperature Distributions in Fast Hot Rolling and
Axisymmetric Extrusion Processes," Int. J. Mech. Sci., Vol. 1, p. 175.
3. Tay, A. O., Stevenson, M. G., and Davis, G. V., (1974), "Using the
Finite-Element Method to Determine Temperature Distributions in Orthogo-
nal Machining," Proc. Inst. Mech. Engr. Vol. 188, p. 627.
4. Bishop, J. F. W., (1956), "An Approximate Method for Determining the
Temperature Reached in Steady-State Motion Problems of Plane Plastic
Strain," Q. J. Mech. Appl. Math., Vol. 9, p. 236.
5. Altan, T., and Kobayashi, S., (1968), "A Numerical Method for Estimating
the Temperature Distributions in Extrusion Through Conical Dies," Trans.
ASME, J. Engr. Ind., Vol. 90, p. 107.
6. Lahoti, G., and Altan, T., (1975), "Prediction of Temperature Distributions in
Axisymmetric Compression and Torsion," Trans. ASME, J. Engr. Materials
Technology, Vol. 97, p. 113.
7. Nagpal, V., Lahoti, G. D., and Altan, T., (1978), "A Numerical Method for
Simultaneous Prediction of Metal Flow and Temperatures in Upset Forging of
Rings," Trans. ASME, J. Engr. Ind. Vol. 109, p. 413.
8. Zienkiewicz, O. C., Jain, P. C., and Onate, E., (1978), "Flow of Solids During
Forming and Extrusion: Some aspects of numerical solutions," Int. J. Solids
Structures Vol. 14, p. 15.
9. Zienkiewicz, O. C., Onate, E., and Heinrich, J. C., (1978), "Plastic Flow in
Metal Forming—I. Coupled Thermal Behavior in Extrusion—II. Thin Sheet
Forming." Applications of Numerical Methods to Forming Processes, ASME,
AMD, Vol. 28, p. 107.
10. Rebelo, N., and Kobayashi, S., (1980), "A Coupled Analysis of Viscoplastic
Deformation and Heat Transfer—I. Theoretical Considerations," Int. J. Mech.
Sci., Vol. 22, p. 699.
11. Rebelo, N., and Kobayashi, S., (1980), "A Coupled Analysis of Viscoplastic
Deformation and Heat Transfer—II. Applications." Int. J. Mech. Sci., Vol.
22, p. 707.
12. Cristescu, N., (1975), "Plastic Flow Through Conical Converging Dies, Using
a Viscoplastic Constitutive Equation," Int. J. Mech. Sci., Vol. 17, p. 425.
13. Cristescu, N., (1976), "Drawing Through Conical Dies—An Analysis Com-
pared with Experiments," Int. J. Mech. Sci., Vol. 18, p. 45.
14. Zienkiewicz, O. C., and Godbole, P. N., (1975), Viscous, Incompressible
Flow with Special Reference to Non-Newtonian (plastic) Fluids, Chap. 2 in
"Finite Elements in Fluids" (Edited by R. H. Gallagher et al), Wiley, New
York.
Thermo-Viscoplastic Analysis 243
15. Zienkiewicz, O. C., and Godbole, P. N., (1974), "Flow of Plastic and
Viscoplastic Solids with Special Reference to Extrusion and Forming Proc-
esses," Int. J. Num. Meth. Eng., Vol. 8, p. 3.
16. Price, J. W. H., and Alexander, J. M., (1976), "A Study of Isothermal
Forming or Creep Forming of a Titanium Alloy," Proc. of the 4th NAMRC,
Columbus, Ohio, p. 46.
17. Price, J. W. H., and Alexander, J. M., (1976), "The Finite-Element Analysis
of Two High-Temperature Metal Deformation Processes," 2d Int. Symposium
on FEM in Flow Problems, p. 717.
18. Klemz, F. B., and Hashmi, S. J., (1977), "Simple Upsetting of Cylindrical
Billets: Experimental Investigation and Theoretical Prediction," 18th MTDR
Conference, London, p. 323.
19. Loizou, N., and Sims, R. B., (1953), "The Yield Stress of Pure Lead in
Compression," J. Mech. Phys. Solids, Vol. 1, p. 234.
20. Lippman, H., (1966), "On the Dynamics of Forging," Proc. 7th Int. MTDR
Conference, Birmingham, England, p. 53.
21. Dahlquist, G., and Bjorck, A., (1974), "Numerical Methods," Prentice-Hall,
Englewood Cliffs, NJ.
22. Wu, W. T., and Oh, S. I., (1985), "ALPIDT: A General Purpose FEM Code
for Simulation of Non-Isothermal Forming Processes," Proc. NAMRI—XIII,
Berkeley, California, p. 449.
23. Tang, M.-C., and Kobayashi, S., (1982), "An Investigation of the Shell Nosing
Process by the Finite Element Method. Part 2: Nosing at Elevated Tempera-
tures," Trans. ASME, J. Engr. Ind., Vol. 104, p. 312.
24. Rebelo, N., and Kobayashi, S., (1981), "Thermo-Viscoplastic Analysis of
Titanium Alloy Forging," ASME Publications FED—Vol. 3, Manufacturing
Solutions Based on Engineering Sciences, p. 151.
25. Oh, S. I., Park, J. J., Kobayashi, S., and Allan, T., (1983), "Application of
FEM Modeling to Simulate Metal Flow in Forging a Titanium Alloy Engine
Disk," Trans. ASME, J. Engr. Ind., Vol. 105, p. 251.
26. Pohl, W., (1972), "A Method for Approximate Calculation of Heat Genera-
tion and Transfer in Cold Upsetting of Metals," Doctoral Dissertation,
University of Stuttgart.
27. Oh, S. I., (1982), "Finite Element Analysis of Metal Forming Process with
Arbitrarily Shaped Dies," Int. J. Mech. ScL, Vol. 24, p. 479.
28. Carlson, R. K., (1943), "An Experimental Investigation of the Nosing of
Shells," Forging of Steel Shells, presented at the Winter Annual Meeting of
ASME, New York.
29. Altan, T., and Boulger, F. W., (1973), "Flow Stress of Metals and its
Applications in Metal Forming Analysis," Trans. ASME J. Engr. Ind., Vol.
95, No. 4, p. 1009.
30. Gegel, H. L., Nadiv, S., Malas, J. C., and Morgan, J. T., (1980), "Application
of Process Modeling to Analysis of Microstructural Changes During the Hot
Working of a Two-phase Titanium Alloy," Appendix K, AFWAL-TR-80-
4162, p. 403.
31. Gegel, H. L., Prasad, Y. V. R. K., Malas, J. C., Morgan, J. T., and Lark, K.
A., (1984), "Computer Simulations for Controlling Microstructure During Hot
Working of Ti-6242," ASME, PVP, vol. 87, p. 101.
32. Dawson, P. R., (1984), "A Model for the Hot or Warm Forming of Metals
with Special Use of Deformation Mechanism Maps," Int. J. Mech. Sci., Vol.
26, p. 227.
33. Dewhurst, T. B., and Dawson, P. R., (1984), "Analysis of Large Plastic
Deformation at Elevated Temperatures Using State Variable Models for
Viscoplastic Flow," Proc. Symp. Constitutive Equations: Micro, Macro, and
Computational Aspects, ASME., p. 149.
13
COMPACTION AND FORGING
OF POROUS METALS
13.1 Introduction
Powder forming, once considered a laboratory curiosity, has evolved into a
manufacturing technique for producing high-performance components
economically in the metal-working industry because of its low manufactur-
ing cost compared with conventional metal-forming processes [1,2].
Generally, the powder-forming process consists of three steps: (1) com-
pacting a precise weight of metal powder into a "green" preform with
10-30% porosity (defined by the ratio of void volume to total volume of
the preform); (2) sintering the preform to reduce the metal oxides and
form strong metallurgical structures; (3) forming the preform by repressing
or upsetting in a closed die to less than 1% residual porosity.
Powder forming has disadvantages in that the preform exhibits porosity.
Because of this porosity, the ductility of the sintered preform is low in
comparison with wrought materials [3]. In forging compacted and sintered
powdered-metal (P/M) preforms, where large amount of deformation and
shear is involved, pores collapse and align in the direction perpendicular to
that of forging and result in anisotropy. However, repressing-type defor-
mation, where very little deformation and shear are present, does not lead
to marked anisotropy [4]. A low-density preform will result in more local
flow and a higher degree of anisotropy than will a preform of high initial
density [5]. These anisotropic structures can lead to nonuniform impact
resistances of the forged P/M parts. Also, in forming of sintered preforms,
materials are more susceptible to fracture than in forming of solid
materials, and the analysis is of particular importance in producing
defect-free components by determining the effect of various parameters
(preform and die geometries, sintering conditions, and the friction
conditions) on the detailed metal flow. In this chapter, the plasticity theory
for solid materials is extended to porous materials, applicable to the
deformation analysis of sintered powdered-metal preforms.
In characterizing the mechanical response of porous materials, a
phenomeno'ogical approach (introducing a homogeneous continuum mo-
del) is employed. For the finite-element formulations of the equilibrium
and energy equations based on the infinitesimal theory, the following
assumptions are made: the elastic portion of deformation is neglected
244
Compaction and Forging of Porous Metals 245
because the practical forming process involves very large amounts of
plastic deformation; the normality of the plastic strain-rates to the yield
surface holds; anisotropy that occurs during deformation is negligible; and
thermal properties of the porous materials are independent of the
temperatures.
Starting with eq. (13.1), Oyane and his colleagues [8,9,10] derived the
plasticity equations. On the basis of this theory, they derived the slip-line
field equations and the upper-bound theorem applicable to porous metals.
The yield surface can be defined by
where o,, and e,y are apparent stresses and strain-rates, respectively,
considering a porous metal to be a continuum. The proportionality factor A
in the flow rule, eq. (13.3), is given by
where y,-,- is the engineering shear strain rate and ev is the volumetric strain
rate.
The apparent yield stress YR depends on the property of the base metal
and the relative density R (ratio of the volume of base metal to the total
volume of porous metal) [13] according to
where Yb is the yield stress of the base metal and 17 is a function of relative
density. The effects of strain, strain-rate, and temperature on the yield
stress are included in Yb = Yb(eb, eb, Tb), where eb, eb, and Tb are strain,
strain-rate, and temperature of the base metal. The relationship between
the apparent strain and strain-rate and those of the base metal are given by
and
Discretization
The element used for discretization is an isoparametric quadrilateral
element with bilinear shape function (see Fig. 8.1 and Fig. 9.1 for the
two-dimensional and the axisymmetric deformations, respectively). The
elemental velocity field is approximated by
The matrices N and B have been defined for a quadrilateral element in the
previous chapters. Substitution of eq. (13.11) into eq. (13.5) leads to
Substituting eqs. (13.10) and (13.13) into eq. (13.9) at the elemental level
and assembling the element equations with global constraints, we obtain
where (;') indicates the y'th element. Because the variation <5v is arbitrary,
eq. (13.14) results in the stiffness equations. The matrix P in eq. (13.12) is
obtained as follows. The inversion of the flow rules (13.3) can be expressed
by
with Vb being volume of the base metal and Vv as the volume of void.
Integrating eq. (13.16), we have
In eq. (13.17), R0 is the current relative density and Aey is the change of
volumetric strain in one deformation step. The average relative density Ra
is defined by
where Rt and Vt are the relative density and the volume of an element,
respectively.
Fully Dense Materials
For fully dense materials, R = 1.0 and A = 3. Then, the matrix D of eq.
(13.15) becomes infinity. Consequently, convergence behavior for the
solution becomes erratic when the relative density in the element
approaches unity.
A constraint was incorporated in the numerical procedures such that an
element with R = 0.9990 was considered as a fully dense element. This
constraint was helpful in obtaining well-behaved convergence for the
solution. Osakada et al. [16] and Mori et al. [17], also using this theory,
developed the finite-element method and applied it to the rigid-plastic
deformation analysis of fully dense materials.
Volume Integration
The program has been tested by analyzing compression of cylinders and
rings of porous materials [18]. During the test it was found that the
reduced integration must be applied to the terms involving the volumetric
strain-rate. Such a reduced integration strategy is straightforward in the
formulation of fully dense material, since the term of the volumetric strain
Compaction and Forging of Porous Metals 249
and
FIG. 13.1 Simple compression of a cylindrical sintered P/M preform. The predicted grid
distortions during simple compression at 20, 50, and 70% reductions in height. The dashed
lines are the predicted boundaries with fully dense initial preform. Initial relative density =
0.8; friction factor = 0.5.
figure that the predicted grid distortions resemble those expected in the
compression of a fully dense preform. The predicted boundaries of the
deformed workpiece for a fully dense initial preform are shown in dashed
lines for comparison. It is seen that the porous preform changes its volume
during the compression process. The average density change as a function
of height strain is shown in Fig. 13.2. The height strain is defined by
\n(H/H0) where H0 is the height of the undeformed preform and H is the
deformed height. It is interesting to note that the average density varies
very little with differerent friction conditions.
Compaction and Forging of Porous Metals 251
FIG. 13.2 The predicted average relative density changes as a function of height strain in
simple compression.
observed that the density near the free surface is lower with higher
friction, owing to the larger barrelling. At 70% reduction in height, it is
noted that the size of the fully dense region near the axis is larger with
higher friction, while the density near the free surface decreases con-
siderably as the workpiece undergoes deformation from 50 to 70%
reduction in height. In practice, fracturing is often observed at the
equatorial surface because of this density reduction.
Figure 13.4 shows the predicted load-displacement curves for two
different friction conditions. For comparison, the predicted load-
displacement curve for the fully dense preform with friction factor m = 0.5
is also shown. The load is lower with the porous preform than that for the
fully dense preform. It can be seen from the figure that the difference
between the forming loads with different frictions is very small during the
early stage of deformation. However, during the later stage, the load
begins to increase faster in the case of higher friction.
FIG. 13.4 The predicted load versus die stroke curves in simple compression. Ra indicates
the initial relative density and m indicates the friction factor.
254 Metal Forming and the Finite-Element Method
for the base metal (OFHC copper). The initial relative density was
R0 = 0.800. Y0 is the initial yield stress of the base metal and was given as
42061 psi (290MN/m 2 ). The frictional condition is given by mky (ky
represents the apparent yield stress in shear) and the friction factor m was
assumed to be 0.1.
The results, illustrating the deformation zone and the extent of
densification, are given for two preforms in Fig. 13.6, where the
distributions of apparent effective strain and relative density are shown at
24% reduction in height. In this figure, the pattern of equistrain and
density contours are remarkably similar. This is because the effective strain
contains not only the term of distortion but also the volume change that is
associated with the relative density. The effective strains are largest at the
center of the forging and at the edge of the die-specimen interface, as are
the relative densities.
For preform I, the central region under the die shows no deformation
and no change in relative density from the initial value (R0 = 0.800) for
FIG. 13.6 Effective strain (right) and relative density (left) distributions at 24% reduction in
height in forging (a) preform I and (b) preform II.
FIG. 13.7 (a) Mesh systems and (b) relative density contours before (left) and after (right)
remeshing in flange-hub forging.
FIG. 13.9 Preform shapes and mesh systems in axisymmetric closed-die forging of a pulley
blank.
257
258 Metal Forming and the Finite-Element Method
aluminum powders. The yield stress of the preform was determined from
that of the base metal of commercially pure aluminum and the constitutive
eq. (13.8). The rate-sensitivity of the material property and the tempera-
ture effect on deformation were not included in this simulation. The initial
relative density was assumed to be uniform and to be 0.780 for both
preforms. Two values of m were assumed for friction, m = 0-l and
m = 0-5.
The nonsteady-state forging process was simulated by a step-by-step
method with increments of 2% of the initial height for preform A and 5%
of the initial height for preform B. In order to complete the calculation
successfully, four and ten remeshings were required for preforms A and B,
respectively.
Preform A. The calculation was stopped at 38% and 43% reductions in
height for m = 0.1 and m = 0.5, respectively, in forging preform A. The
program is designed to automatically discontinue the calculation if the
relative density of any element falls below the limiting value of the relative
density for which the apparent yield stress becomes zero, because this
indicates the initiation of fracture.
In Figs. 13.10a and b, the distributions of relative density and hydro-
static stress at the final stage of forging are shown for two friction
conditions. The patterns of densification and hydrostatic stress are similar
to each other. For the two friction conditions, the relative density
distributions differ somewhat in the rim area, as do the hydrostatic stress
distributions. It was found that the effect of friction on the densification of
the materials was not evident at an early stage of forging. As deformation
continues, however, the densification and the compressive hydrostatic
stress are greater with higher friction. It can also be seen that the locations
of the large and small values of relative density are the same for both
friction conditions and lie at the flange part around the die corners and
near the free surfaces, respectively.
Preform B. Contrary to preform A, complete forging was possible with
preform B. This was due to the differences in material flow in the two
preforms.
The distributions of relative density and hydrostatic stress with low
friction were compared with those with high friction at the final stage of
forging in Fig. 13.11. The distribution patterns are similar to each other for
the two friction conditions at the same stage of forging except that for
higher friction, the densification and the compressive hydrostatic stress
were greater. Because of this effect of friction on densification, the final
forged part in preform B with high friction did not completely fill the
cavity at 60% reduction in height due to the volume change, while
complete filling was achieved for low friction at the same reduction in
height (see Figs. 13.Ha and b).
As in forging of preform A, densification and the compressive hydro-
static stress are concentrated at the flange part near the die corners. The
figure also shows that the weakest mechanical properties will be near the
Compaction and Forging of Porous Metals 259
FIG. 13.10 Relative density (left) and hydrostatic stress (3am, units of ksi) (right)
distributions at the final stage for the two friction conditions in forging preform A (darkened
areas indicate the possible fracture sites): (a)m = 0.1; (b)m=0.5.
tips of the rim section. However, the relative density in most of the forged
hub section reached full density, R = 0.999, in the final stage.
The experiments [19] also showed that preform B produced defect-free
pulley blanks. This agreement allows the prediction of the probable
locations of fracture during forging using the existing finite-element
program.
FIG. 13.11 Relative density (left) and hydrostatic stress, (3am units of ksi) (right)
distributions at the final stage for the two friction conditions in forging preform B: (a)
m=Q.\, (b) m=0.5.
where the subscripts R, b, and v denote the quantity related to the total
porous materials, base metals, and voids, respectively, t is the time, and p,
c, and V are the density, the specific heat, and the volume, respectively.
The temperature-rate of voids is at most of the same order as the
262 Metal Forming and the Finite-Element Method
temperature-rate of the base metal, since the heat generation due to plastic
deformation is limited to that of the base metal. The change in the internal
energy of voids can also be neglected, since the thermal capacity of the
voids is much smaller than that of the base metal. Therefore, eq. (13.23)
can be reduced to
where the apparent specific heat of the porous material, CR, can be
determined by
Combining this equation with eq. (13.24), noting that pRVR = pbVb, gives
FIG. 13.12 Schematic of plane-strain compression: (a) geometry of the die and workpiece;
(b) boundary conditions. (Superscripts c, r, f on the heat flow qn refer to convection,
radiation, and friction, respectively.)
CT£ = 3.6xlON/(mm 2 sK 4 ),
/zc = 5.5N/(mmsK), /z00 = 0.01 N/(mms K),
K = 0.85
264 Metal Forming and the Finite-Element Method
True S t r a i n in y - d i r e c t i o n
FIG. 13.13 Comparison of variations in lateral flow between theory and experiment [28].
FIG. 13.14 Variations of average relative density during compression, and comparison with
experiment [28].
(b)
FIG. 13.15 (a) Experimental local hardness distribution in the center cross-section of a bar
[28]. (b) Relative density distributions at 40% reduction in height for Ra = 0.743 with two
friction conditions: left (m = 0.5); right (m = 1.0).
265
266 Metal Forming and the Finite-Element Method
FIG. 13.16 Relative density distributions at the final stages for Ra = 0.743 with two friction
conditions: (left) 70% reduction and m = 0.5; (right) 58% reduction and m = 1.0.
13.9 Compaction
In the process of compaction, a P/M preform is placed in a cylindrical
container and pressed in a double-action press, as shown in Fig. 13.18. For
the simulation [14], the container was 50.8 mm (2.0 in.) in diameter and
152.4 mm (6.0 in.) in height. Because of symmetry, the analysis was
performed for the half-height. The flow stress behavior of the matrix
material was assumed to be Yb =0.1. The constant-friction factor law was
FIG. 13.17 (a) Temperature distributions for R0 = 0.743 at 40% reduction in height with two
friction conditions: (left) m=0.5; (right) m = 1.0. (b) Temperature distributions for
R0 = 0.743 at the final stages with two friction conditions: (left) m = 0.5; (right) m = 1.0.
267
268 Metal Forming and the Finite-Element Method
assumed, and the friction factors used in simulation were m = 0.1, 0.2, and
0.5. The punch velocities were 76.2 mm/s (3.0 in./s) for both the upper
and lower punches. The initial relative density in the simulation was taken
to be 0.8. It was also assumed that the air and the lubricant trapped in the
pores did not affect the compaction process.
Figures 13.19a and b show the predicted relative density distributions for
the friction facors m =0.2 and 0.5, respectively. It is seen that the density
is greatest near the outside radius of the moving punch and decreases in
regions remote from the punch, particularly near the container die wall.
As the deformation progresses towards full compaction, the density
distribution becomes more uniform; that is, the difference between the
highest and the lowest density becomes smaller. It can be seen that the
trend of density distributions for both friction conditions are similar, but
the density is more uniform with lower friction.
From the predicted density distribution (which is a function of radial as
well as height location), the average density was calculated as a function of
height. The average density was obtained by integrating the relative
density distribution over the cross-sectional area and dividing by that
cross-sectional area at a given height.
Figures 13.20a and b show the predicted density variation as a function
of height for different friction factors, m = 0.2 and m = 0.5, respectively. It
can be seen from the figures that the density is lowest at the mid-height,
and that it increases toward the punch. It is also seen that the distribution
becomes more uniform with increasing deformation and lower friction.
Compaction and Forging of Porous Metals 269
FIG. 13.19 (a) Predicted relative density distribution during die pressing; initial relative
density = 0.8; friction factor = 0.2. (b) Predicted relative density distribution during die
pressing; initial relative density = 0.8; friction factor = 0.5.
D E N S I T Y D I S M=0.5
FIG. 13.20 (a) Predicted average density as function of height; initial relative density = 0.8;
friction factor = 0.2. (b) Predicted average density as function of height; initial relative
density = 0.8; friction factor = 0.5.
FIG. 13.21 Measured average relative densities obtained in compaction of powder in closed
dies and by hydrostatic pressing [29].
271
272 Metal Forming and the Finite-Element Method
References
1. Wisker, J. W., and Jones, P. K., (1974), The Economics of Powder Forging
Relative to Competing Processes—Present and Future, in "Modern Develop-
ments in Powder Metallurgy," (Edited by H. H. Hausner, and W. E. Smith),
Vol. 7, American Powder Metallurgy Institute, Princeton, NJ, p. 33.
2. Jones, P. K., (1973), "New Perspectives in Powder Metallurgy," Vol. 6,
Plenum Press, New York.
3. Kaufman, S. M., and Mocarski, S., (1971), "Effect of Small Amount of
Residual Porosity on the Mechanical Properties of P/M Forgings," Int. J.
Powder Metal., Vol. 7, p. 19.
4. Dieter, G. E., (1976), "Mechanical Metallurgy," 2d Ed., McGraw Hill, New
York.
5. Moyer, K. H., (1974), A Comparison of Deformed Iron-Carbon Alloy Powder
Preforms with Commercial Iron-Carbon Alloys, in "Modern Developments in
Compaction and Forging of Porous Metals 273
14.1 Introduction
A majority of the finished products made by metal forming are geometri-
cally complex and the metal flow involved is of a three-dimensional nature.
Thus, any analysis technique will become more useful in industrial
applications if it is capable of solving three-dimensional metal-flow
problems.
Nagpal and Altan [1,2] introduced dual-stream functions for describing
metal flow in three dimensions. This work showed that the proper
selection of a flow function makes the incompressibility requirement
automatically satisfied and provides general kinematically admissible
velocity fields. Yang and Lee [3] utilized the conformal transformation of a
unit circle onto a cross-section in the analysis of curved die extrusion. They
derived the stream-line equation from which a kinematically admissible
velocity field was determined. The upper-bound method was then applied
to determine the extrusion pressure for a rigid-perfectly plastic material.
An important aspect of three-dimensional plastic deformation is the
analysis of spread in metal-forming operations, such as spread in rolling or
in flat tool forging, and spread in compression of noncircular disks.
Solutions to such problems have been obtained by the use of Hill's general
method [4] and the upper-bound method [5-7].
The extension of the finite-element method to solve three-dimensional
problems is natural and not new, particularly in the area of elasticity [8].
However, the simulation of three-dimensional forming operations by the
finite-element method is relatively recent. Park and Kobayashi [9] de-
scribed the formulation for the three-dimensional rigid-plastic finite-
element method and the implementation of the boundary conditions. They
applied this technique to the analysis of block compression between two
parallel flat platens. For certain forming problems, such as those involving
lateral spread, the use of a simplified three-dimensional element is efficient
and some examples can be found for analysis of spread in rolling and flat
tool forging [10, 11].
275
276 Metal Forming and the Finite-Element Method
14.2 Finite-Element Formulation
Element Equations
The matrices for evaluation of elemental stiffness equations are defined for
a three-dimensional brick element in Chap. 6 and some of them are
recapitulated in this section. A three-dimensional brick element used for
the analysis is an eight-node hexahedral isoparametric element. As shown
in Fig. 14.1 the global coordinate system (x, y, z) is transformed into the
natural coordinate system (£, rj, £). The natural coordinate system is
defined such that £, r\, and £ vary from —1 to 1 within each element. The
velocity field within an element is expressed as
where
with
For eqs. (14.1), (14.2), and (14.3) above, reference can be made to eqs.
(6.17), (6.18), and (6.19) in Chap. 6.
The strain-rate vector E T = {ex, ey, ez, jxy, yyz, YZ*} is expressed by
where
and
where
and
FIG. 14.3 Comparison of the computed (broken curves) and experimental (solid curves) [13] spread contours at the equatorial plane in rectangular block
compression.
Three-Dimensional Problems 281
actual friction value that is present in the experiment. On the other hand,
in the simulation for WJB0 = I, m = 0.5 is a proper value to assume for
the dry condition. The simulation, for W0/B0 = 2 with m = 0.2 for the
lubricated condition, gives an excellent agreement with the experimental
result. Some discrepancy around the corner in simulation for the dry
condition may indicate that friction is not uniform over the entire
die-workpiece interface.
Compression of Wedge-Shaped Blocks
Compression of a wedge-shaped block (Fig. 14.4) between two flat parallel
dies is used in practice for workability and microstructural studies in
forging. The boundary conditions for the analysis are more complex in
comparison with those for rectangular block compression owing to the
lower degree of symmetry in the geometry. The determination of the
neutral regions at the contact surfaces, with both upper and lower dies,
represents an important problem. In particular, along the die-workpiece
contact surface, at the lower die the extent of the actual contact is not
known and varies during deformation. The finite-element code for this
problem includes additional schemes in order to take into account these
boundary conditions, as well as folding, that occur in rectangular block
compression.
The workpiece dimensions used in the simulations are shown in Fig.
14.4. The workpiece geometry and deformation are symmetrical with
respect to the (x, z) plane, so that one-half of the workpiece is taken as the
control volume for simulation. The flow stress of the material, Al-2024
(annealed), is represented by a = 225 (1 + e/1.6147) (MN/m2). Two
simulations were performed under different friction conditions: dry (m =
0.4) and lubricated (m = 0.1). A step size of 2% in height reduction was
FIG. 14.5 Grid distortion of the wedge-shaped workpiece at several height reductions (20,
40, and 60%).
NJ FIG. 14.6 Effective strain distributions and location of neutral zones (indicated by arrows) in the plane of symmetry at several height reductions of
w compression of a wedge-shaped block.
284 Metal Forming and the Finite-Element Method
lower center of the block portion and that the strain decreases monotoni-
cally toward the end of the wedge. The strains are, in general, uniform in
the height direction, particularly in the wedge portion and for low friction.
Locations of the neutral zones are symmetrical and approximately posi-
tioned at the mid-point of the contact surface at larger reductions in
height. However, at 20% reduction, the interfacial friction condition
appears to have a significant influence on the location of the neutral zones.
In using the wedge compression for a workability test, fracture is usually
observed on the side surface. The present analysis provides some
information concerning the occurrence of fracture, such as the strain path
and the corresponding stress variations at critical sites.
FIG. 14.7 Schematic diagram of the square ring used in simulation. The thick lines indicate
the volume included in the analysis.
Three-Dimensional Problems 285
lines in Fig. 14.7. A total of 250 elements with 396 nodes were used
for the analysis. Simulation was done in a step-by-step manner and 2%
of the undeformed height was used as the step size.
The results show that the buckling-type deformation mode takes place at
the beginning of deformation. That is, the mid-height of the side wall
moves outward while the die contact surface moves very little in the
horizontal direction. As a result, nodes at the die contact surface near the
inner edge separate from the die at the initial stage of the deformation.
The separated nodes touch the die shortly afterwards. Figure 14.8 shows
the perspective view of one-quarter of the workpiece with the predicted
mesh distortions of the cross section on the YZ plane at steps of 20% in
height. It is seen in the figure that at 20% reduction in height, the
mid-plane moves outward while the top and bottom surfaces change their
positions very little. This deformation mode causes buckling. At 40%
reduction in height, the outside wall forms a considerable bulge, while the
inner wall forms a dimple. The dimple at the inner wall becomes sharper
as deformation progresses and, finally, it collapses, forming a fold at 54%
reduction in height. This inner surface folding takes place first at the YZ
cross section and propagates towards the corner of the hole. The folding of
the inner surface can be seen clearly from the predicted grid distortion at
60% reduction in height.
In Fig. 14.8, locations of the neutral points are indicated by arrows. It is
seen that the spatial position of the neutral point on the YZ cross section
changes very little with deformation. However, its location relative to a
material point changes considerably.
Figure 14.9 shows the top view of the distorted mesh at steps of 20%
reduction in height. In the figure, the neutral lines are indicated by broken
lines. It is seen from the figure that the material points at the mid-side of
the inner surface move inward, while points at the corner of the inner
surface move outward until 40% reduction in height is reached; the corner
points then start to move inward. Because of the different movements of
the material points on the die contact surface, the inner surface, looking
from the top, becomes inwardly convex and the angle between the two
inner surfaces becomes less than 90°.
Aku et al. [16] carried out square-ring compression tests by using
plasticine as a model material. Though the detailed measurements of the
process have not been reported, they include the sketch of the top view of
the deformed workpiece at 30, 50 and 77% reductions in height. By
comparing their sketch with Fig. 14.8, it is found that the present
predictions agree qualitatively very well with the experimental observa-
tions. Seen from the top, the inner hole (which is inwardly convex), the
bulge at the inner surface (which is not noticeable), and the large amount
of bulge at the outside surface all indicate reasonable agreement between
the predictions and experiment. These observations are only qualitative
and further critical comparison may be necessary to fully verify the results
of the simulations.
FIG. 14.8 Perspective views of one-quarter of the workpiece for square-ring compression,
with the predicted mesh distortions on the YZ cross section at 0, 20, 40 and 60% reductions
in height (units of mm). (Arrows indicate the position of the neutral surface.) [15].
286
Three-Dimensional Problems 287
FIG. 14.9 Top views of grid distortions predicted by the analysis at 0, 20, 40, and 60%
reductions in height during square-ring compression. The broken lines indicate the neutral
lines (units of mm) [15].
and
We also have
and
In eq. (14.14), the directions of «,<3) and w,(4) are tangential to the contact
surface at points 4 and 3, respectively. Thus, the distribution of u, given in
eq. (14.14) is the result of an approximation usually involved in
discretization.
Three-Dimensional Problems 289
FIG. 14.12 The profiles of a rolled square bar for 20% reduction: W0/HQ= I.
FIG. 14.13. (a) Comparison between computed and measured [30] lateral spread for
room-temperature rolling of mild-steel plates, (b) Reduction in cross-sectional area vs.
reduction in height for room-temperature rolling of mild-steel plates.
294
Three-Dimensional Problems 295
Stroke
FIG. 14.15 Comparison between theory and experiment for spread (AWmax/W0) and
elongation (AL/L0) in three-bite flat-bar forging (Bar and tool geometry are shown in Fig.
14.14). [11]
References
1. Nagpal, V., (1977), "On the Solution of Three-Dimensional Metal Forming
Processes," Trans. ASME, J. Engr. Ind., Vol. 99, p. 624.
2. Nagpal, V., and Allan, T., (1975), "Analysis of the Three-Dimensional Metal
Flow in Extrusion of Shapes with the Use of Dual Stream Functions," Proc. 3d
NAMRC., Pittsburgh, PA, p. 26.
3. Yang, D. Y., and Lee, C. H., (1978), "Analysis of Three-Dimensional
Extrusion of Sections Through Curved Dies by Conformal Transformation,"
Int. J. Mech. ScL, Vol. 20, p. 541.
4. Lahoti, G. D., and Kobayashi, S., (1974), "Flat-tool Forging," Proc. 2d
NAMRC, University of Wisconsin, Madison, p. 73.
5. Oh, S. I., and Kobayashi, S., (1975), "An Approximate Method for
Three-Dimensional Analysis of Rolling," Int. J. Mech. ScL, Vol. 17, p. 293.
6. Sagar, R., and Juneja, B. L., (1979), "An Upper Bound Solution for Flat Tool
Forging Taking into Account the Bulging of Sides," Int. J. Machine Tool. Des.
Res., Vol. 19, p. 253.
7. Braun-Angott, P., and Berger, B., (1982), An Upper Bound Approximation
for Spread and Pressure in Flat Tool Forging, in "Numerical Method in
Industrial Forming Processes," (edited by J. F. T. Pittman, R. D. Wood, J. M.
Alexander and O. C. Zienkiewicz) Pineridge Press, Swansea UK, p. 165.
8. Zienkiewicz, O. C., (1977), "The Finite Element Method," 3d Edn. McGraw-
Hill, New York.
9. Park, J. J., and Kobayashi, S., (1984), "Three-Dimensional Finite-Element
Analysis of Block Compression," Int. J. Mech. Sci., Vol. 26, p. 165.
10. Li, G.-J., and Kobayashi, S., (1982), Spread Analysis in Rolling by the
Rigid-Plastic Finite Element Method, in "Numerical Method in Industrial
Forming Processes," (edited by J. F. T. Pittman, R. D. Wood, J. M.
Alexander and O. C. Zienkiewicz) Pineridge Press, Swansea UK, p. 777.
11. Sun, Jie-Xian, Li, G.-J., and Kobayashi, S., (1983), "Analysis of Spread in
Flat-Tool Forging by the Finite Element Method," Proc. llth North American
Manufacturing Research Conf., May 1983, Madison, Wisconsin, p. 224.
12. Park, J. J., (1982), "Applications of the Finite Element Method to Metal
Forming Problems," Ph.D. Dissertation, Department of Mechanical Engi-
neering, University of California, Berkeley.
13. Kanacri. F., Lee, C. H., Beck, L. R., and Kobayashi, S., (1972), "Plastic
Compression of Rectangular Blocks Between Two Parallel Platens," Proc.
13th. Int. Mach. Tool Des. Res. Conference, Birmingham, p. 481.
Three-Dimensional Problems 297
14. Sun, Jie-Xian, and Kobayashi, S., (1984), "Analysis of Block Compression
with Simplified Three-Dimensional Element," Proc. 1st Int. Conf. on Tech-
nology of Plasticity, Tokyo, p. 3.
15. Park, J. J., and Oh, S. I., (1987), "Application of Three-Dimensional Finite
Element Analysis to Metal Forming Processes," Proc. NAMRC XV, Beth-
lehem, PA, p. 296.
16. Aku, S. T., Slater, A. C., and Johnson, W., (1967), "The Use of Plasticine to
Simulate the Dynamic Compression of Prismatic Blocks of Hot Metal," Int. J.
Mech. ScL, Vol. 9, p. 495.
17. Chitkara, N. R., and Johnson, W., (1966), "Some Experimental Results
Concerning Spread in the Rolling of Lead," Trans. ASME, J. Basic Engr.,
Vol. 88, p. 489.
18. Helmi, A., and Alexander, J. M., (1968), "Geometric Factors Affecting
Spread in Hot Flat Rolling of Steel," J. Iron Steel Inst., Vol. 206, p. 1110.
19. Gokyu, I., Kiharai, J., and Mae, Y., (1970), "Study on the Width Spread in
Flat Rolling," /. Japan Soc. Tech. Plast, Vol. 11, p. 11.
20. Lahoti, G. D., and Kobayashi, S., (1974), "On Hill's General Method of
Analysis for Metal-Working Processes," Int. J. Mech. ScL, Vol. 16, p. 521.
21. Kummerling, R., and Lippmann, H., (1975), "On Spread in Rolling," Mech.
Res. Commun., Vol. 2, p. 113.
22. Tamano, T., (1973), "Finite Element Analysis of Steady Flow in Metal
Processing," J. Japan Soc. Tech. Plast., Vol. 14, p. 766.
23. Zienkiewicz, O. C., Jain, P. C., and Onate, E. (1978), "Flow of Solids During
Forming and Extrusion: Some Aspects of Numerical Solutions," Int. J. Solids
Structures, Vol. 14, p. 15.
24. Zienkiewicz, O. C., Onate, E., and Heinrich, J. C., (1981), "A General
Formulation for Coupled Thermal Flow of Metals using Finite Elements," Int.
J. Num. Meth. Eng., Vol. 17, p. 1497.
25. Dawson, P. R., (1978), "Viscoplastic Finite Element Analysis of Steady-state
Forming Processes Including Strain History and Stress Flux Dependence,"
Applications of Numerical Methods to Forming Processes, ASME, AMD,
Vol. 28, p. 55.
26. Shima, S., Mori, K., Oda, T., and Osakada, K., (1980), "Rigid-Plastic Finite
Element Analysis of Strip Rolling," Proc. 4th Int. Conf. on Prod. Eng.,
Tokyo, Japan, p. 82.
27. Li, G.-J., and Kobayashi, S., (1982), "Rigid-Plastic Finite Element Analysis of
Plane Strain Rolling," ASME Trans., J. Engr. Ind., Vol. 104, p. 55.
28. Kanazawa, K., and Marcal, P. V., (1982), "Finite Element Analysis of Plane
Strain Rolling," ASME Trans., J. Engr. Ind., Vol. 104, p. 55.
29. Mori, K., and Osakada, K., (1982), "Simulation of Three-Dimensional Rolling
by the Rigid-Plastic Finite Element Method," Proc. Num. Methods Ind.
Forming Processes, Swansea, p. 747.
30. Lahoti, G. D., Akgerman, N., Oh, S. I., and Altan, T., "Computer-aided
Analysis of Metal Flow and Stresses in Plate Rolling," /. Mech. Work. Tech.,
Vol. 4, p. 105.
31. Tomlinson, A., and Stringer, J. D., (1959), "SPread and Elongation in
Flat-tool Forging," /. Iron Steel Inst., Vol. 193, p. 157.
32. Baraya, G. L., and Johnson, W., (1964), "Flat-bar Forging," Proc. 5th Int.
Conf. MTDR, p. 449.
33. Kudo, H., and Nagahama, T., (1969), "Experimental Results for Upsetting
Pressure and Material Spread—Study in Transverse Upsetting Process of
Circular Rods, 1st Report," /. Japan Soc. Tech. Plast., Vol. 10, No. 106, p.
827.
15
PREFORM DESIGN IN METAL FORMING
15.1 Introduction
Preform design in metal forming refers to the design of an initial shape of
the workpiece that, when it has undergone an associated forming process,
forms the required product shape with desired property successfully
without formation of defects and without excessive waste of materials. A
carefully selected preform can contribute significantly to the reduction of
the production costs. Preform design problems are encountered in various
metal-forming processes, such as closed-die forging, shell nosing, rolling,
and sheet-metal forming. Design of an optimal preform shape requires
simultaneous determination of optimal process conditions. However, we
are here concerned with the determination of the best preform shape
under a given set of process conditions.
In this chapter, a new method of "backward tracing" is introduced as an
alternative approach to the solution of preform design [1], and the
applications of this method to some specific processes are discussed.
FIG. 15.1 Concept of the backward tracing method (geometrical configuration only).
solution for loading at P(2) is then u,^, and the second estimate of the
configuration Q(2) = F (2) + u,(,221 Af can be made. The iteration is carried
out until Q(n) = P(n> + v$2i Af becomes sufficiently close to Q.
The formulation of the concept described above will further clarify the
backward tracing technique. After finite-element discretization, the solu-
tion of the boundary value problem associated with isothermal viscoplastic
deformation satisfies a system of nonlinear, coupled algebraic equations,
expressed in the form
FIG. 15.2 Flow chart for the backward tracing procedure implemented into the finite-
element flow analysis.
For a given time-increment, the substitution of eq. (15.3) into eq. (15.1) at
time t0-i results in
Preform Design in Metal Forming 301
FIG. 15.3 Die dimensions and nosed shell configurations. L0 = 47.26, Lf = 48.84, R =
21.526, Rt = 14.648, Bf = 14.47, Hf = 6.88 (uniform), a0 = 237, b0 = 18.75 (units of mm).
For the preform shape with constant outer diameter (Type-0 preform),
RO = R0 for all n, where RQ is the radius of a nodal point located along the
outer surface of the shell and R0 is the radius at the die entrance. The
boundary condition for this type of preform is controlled in such a way that
nodes that are in contact with the die (7?o < RO) become force free when
the condition that RQ = R0 is reached during backward tracing.
For the preform shape with constant inner diameter (Type-I preform),
the criterion for controlling the boundary condition is that as soon as
R? = Rf for the nodes Rpt < Rh the condition of the corresponding outside
node is changed from that of die-contact to the force free condition, where
Rpt is the radius of a nodal point along the inner surface of the shell.
Backward tracing was carried out computationally by taking 2 mm
penetration (in loading) as one step. Thus, a total of 22 steps was
necessary to complete the calculation. The solution convergence needed
about six to seven iterations. For completing one step of backward tracing,
it was necessary to have one to two iterations for the case where the
boundary condition does not change, and two to three iterations were
required for the case where a change in the boundary condition occurs.
Convergence for backward tracing was determined by the closeness of the
solution (between Q and <2(M) described in Section 15.2) within a limit.
The two types of preforms determined by this method are compared with
the final nose configuration in Fig. 15.4. In the figure, the lengths £° and
£o, for Tpye-0 and Type-I preforms, respectively, indicate the portion of
the preform that becomes the nosed shell length Lf after nosing. Because
Preform Design in Metal Forming 303
FIG. 15.4 Nosed shell configuration and two types of preform shapes for nonwork-hardening
materials [1].
the shell elongates during nosing, the preforms are shorter than the nosed
shell. The shell material was assumed to be rigid-perfectly plastic
(nonwork-hardening).
In order to include the work-hardening effect in preform design, it is
necessary to know the strain distributions in the nosed shell, in addition to
the geometrical configurations. However, the strain distributions in the
nosed shell depend on the preform shapes and are not known (or cannot
be specified). In Fig. 15.5 the procedure used to take into account the
work-hardening effect is illustrated.
By applying the backward tracing scheme to nonwork-hardening mate-
rials, the two types of preforms can be determined. With these preforms
the solutions for loading are obtained for a work-hardening material. The
results provide strain (effective strain) distributions and the nosed shell
configurations that are very close to but not identical to the specified
geometries. These configurations are then corrected exactly according to
the specifications, maintaining the same effective strain distributions. With
FIG. 15.5 Schematic illustration of the procedure to include the work-hardening effect in prefr»rn design.
Preform Design in Metal Forming 305
306
FIG. 15.8 Backward tracing applied to a preform design of (a) the front end and (b) the back
end in rolling (conditions are the same as those of Fig. 15.7).
307
308 Metal Forming and the Finite-Element Method
are shown in Fig. 15.8a. The preform front end obtained by this method
has a circular convex contour. The figure also shows that during rolling the
curvature of the front end contour is reduced and becomes slightly concave
before the front end becomes flat at the exit. Similarly, in order to apply
the backward tracing scheme to the determination of back end preform
shapes, the back end contour obtained by the approximate solution was
modified to a flat end, and the rolling process was then traced backward.
Variations of the back end contour at several stages are shown in Fig.
15.8b. The resulting preform shape is a circular arc, similar to the front
end preform, but with a smaller curvature.
For work-hardening materials, the procedure illustrated in Fig. 15.5 was
used. With preform shapes of a plate obtained for nonwork-hardening
materials (Fig. 15.8), the rolling simulation is performed for a work-
hardening material. The results provide the strain distributions and the end
contours. The end shapes are not exactly flat. The ends are then corrected
geometrically to become flat ends, maintaining the same strain distribu-
tions. With these strain distributions, the application of the backward
tracing scheme results in preform shapes for a work-hardening material.
The results for an aluminum alloy are shown in Fig. 15.9. Comparing the
preform shapes for a work-hardening material with those for nonwork-
hardening materials, it is seen in Fig. 15.9 that at both front and back ends
FIG. 15.9 Comparison of preform end shapes for nonwork-hardening and work-hardening
materials in rolling.
Preform Design in Metal Forming 309
the deviations of the contours from the flat ends are smaller for a
work-hardening material. The end shapes obtained by the backward
tracing scheme for a work-hardening material are more complex than
those for nonwork-hardening materials. This indicates that during
nonsteady-state rolling, the deformation is influenced by the material
property.
Preliminaries
For developing design procedures, it would be helpful if we could obtain
some knowledge of the metal flow involved in forging. Thus, two
simulations by the finite-element method were performed as part of a
preliminary investigation.
One was the simulation of the simple compression of a cylinder between
two parallel flat dies (with m = 1.0). The grid patterns before and after
compression are shown in Fig. 15.11. Also shown in the figure are the
effective strain distributions after 50% reduction.
From the results of Fig. 15.11, it is seen that in the core region, a nearly
dead-metal zone is formed beneath the die, resulting in a large strain
gradient in the thickness direction. In the peripheral region, strains are less
than the average strain value within the main body, but they are greater
than average near the contact surface, because of distortion due to friction.
The strain gradient in the core region may be reduced if the size of the
dead zone can be reduced. One may achieve this by applying indentation
at the core region, followed by flat die compression. A symmetric
indentation was simulated by a hemispherical punch with a radius of 1.5 to
the depth of z = 0.35 (the amount of punch displacement = 0.15). Simula-
tion of flat-die compression was then performed until the final disk
configuration was reached. The grid distortions in the workpiece after
indentation, and at the final stage, are shown in Fig. 15.12 along with
resulting strain distributions. It is seen that the strain distribution in the
Preform Design in Metal Forming 311
FIG. 15.11 The grid patterns before and after deformation and the effective strain
distributions after 50% reduction in height in disk forging (black zone: e > l . l ; white zone:
1.1 >e>0.8; gray zone: e<0.8).
core region was altered and some improvement toward uniform strains, in
general, was obtained by indentation and compression.
Design Procedures
In the present problem, and in preform design in forging in general, the
boundary conditions and their variations during the process depend on the
preform shape itself. During the actual forming operation, a change in the
boundary conditions occurs when a nodal point on the free surface comes
into contact with the die. In backward tracing, nodal points along the
die-workpiece interface in the final configuration should be detached from
the die at some stage of deformation and all the nodal points must become
free at the completion of backward tracing.
In developing design procedures for these cases, it is necessary to
consider criteria for controlling when and where and in what sequence a
boundary node should be detached from the die. Two criteria are
conceivable at this point—one for controlling the workpiece geometry and
the other for controlling deformation.
For the first backward tracing step, the boundary conditions were
312 Metal Forming and the Finite-Element Method
FIG. 15.12 The grid patterns after identation and in the finish-forged disk, and the effective
strain distributions in the finished disk (black zone: £>!.!, white zone: 1.1>E>0.8; gray
zone: e<0.8).
Final Design
Following the design procedures discussed above, the shapes of the
preform and the stock and the resulting strain distributions in the finished
disk were obtained, and are shown in Fig. 15.13a. In Fig. 15.13a it is seen
that improvement has been achieved for the strain distributions in the core
region, but not in the peripheral region. Further improvement in design
appears to be possible only by increasing the amount of distortion in the
peripheral region with an additional preforming operation. A radial
forging operation with circular-contoured dies is considered as a process
for this purpose. In forward simulation as well as in backward tracing, the
radial forging process is replaced by a fictitious process in which the
diameter of a ring-die decreases or increases radially. During backward
tracing of this process, the boundary conditions were regulated in such a
way that the side-surface of the stock became as close to a circular cylinder
as possible. The resulting shape of the stock and the corresponding strain
distributions in the finished disk are shown in Fig. 15.13b. It can be seen
FIG. 15.13 Shapes of preforms and stocks and the effective strain distributions in the finished
disk: (a) design I (black zone: e > l . l ; white zone: 1.1>£S0.8; gray zone: esO.8); (b)
design II (black zone: e a 1.3, white zone: 1.3 > £ > 1.0; gray zone: e < 1.0).
314 Metal Forming and the Finite-Element Method
that improvement in uniformity of strain distributions in the peripheral
region was achieved.
In the design procedures discussed above, no particular attention was
paid to irregular shapes of designed stocks and preforms. The reason is
that the shapes and contours can always be smoothed out or modified to
obtain realistic and simple configurations. The final design can then be
evaluated by simulating the operation sequence with modified stock and
preform shapes. The modified stock shape and the radial forging processes
are shown in Fig. 15.14a. The initial stock is a cylindrical solid with a
hemispherical dimple at the center. Each die for radial forging has a
circular contour, and moves radially inward.
The second preforming operation is forging with a die that has the
FIG. 15.14 (a) Simulation of first preforming operation—radial forging, (b) Simulation of
second preforming operation—forging, (c) Simulation of finishing operation—flat-tool
compression.
Preform Design in Metal Forming 315
(see eq. (12.12) in Chap. 12). Substitution of eqs. (15.3) into eq. (15.5) at
time t0-i, gives
In eq. (15.7), x0, £ 0 > VQ-I, EO-I are supposed to be known or obtained from
the solution of the velocity equations. The temperature T
is a function of T 0 _i according to eq. (15.6). Since T0 and T0 are known,
eq. (15.7) can be solved for T0-i by assuming the boundary conditions at
time t0_i. Once T0_! is obtained, then T0_! can be calculated and
compared with the assumed boundary values. Iterative procedures will
result in backward tracing of temperature calculations. Details of the
calculation procedure involved in backward tracing of a thermo-
viscoplastic deformation process is given in Reference [15].
To illustrate the design method, a preform design in shell nosing at
elevated temperatures is considered. The specifications described in the
study by Lahoti et al. [16] and by Tang and Kobayashi [17] were used.
They were: dimensions of the nosed shell and die; interface friction
(frictionless); material (AISI 1045 steel); initial temperature distributions;
and nosing speed (270 mm/s). The preform design procedure consists of
the following three steps.
Step (1)—Loading simulation. Backward tracing begins with the strain and
the temperature fields in the specified configuration and in the die. In
order to know these fields, loading is simulated with a trial or "guessed"
preform, using a given initial temperature distribution. A trial preform
may be designed by considering that the material volume of the preform
must be the same as that of the shell to be formed, and by assuming that
the inner profile of the preform is a straight line. The geometry of the trial
preform is then determined by the requirement of straight outer profile.
The trial preform shape and the initial temperature distributions in the
preform are shown in Fig. 15.15.
Step (2)—Preparation for backward tracing. The final deformed shape,
obtained from the simulation in Step (1) is not the same as that specified.
The nodal point positions are adjusted so that the specified shell
configuration results and strain and temperature distributions in this
configuration are obtained. Figure 15.16 shows these distributions.
Step (3)—Backward tracing and control of boundary conditions. Starting
from the specified shell configuration and corresponding strain and
temperature fields, backward tracing is performed. In order to obtain a
Preform Design in Metal Forming 317
FIG. 15.15 (a) The trial preform shape and finite-element layout for loading simulation of
nosing, (b) The initial temperature distributions in the preform.
FIG. 15.16 (a) The strain distributions and (b) the temperature distributions in the specified
shell configuration and in the die.
FIG. 15.17 (a) The preform shape and (b) the initial temperature distributions obtained at
the completion of backward tracing for shell nosing.
References
1. Park, J. J., Rebelo, N., and Kobayashi, S., (1983), "A New Approach to
Preform Design in Metal Forming with the Finite Element Method," Int. J.
Machine Tool Des. Res., Vol. 23, p. 71.
2. Nadai, A., (1944), "Plastic State of Stress in Curved Shells: The Forces
Required for Forging of the Nose of High-Explosive Shells," Forging of Steel
Shells, ASME Trans., p. 31.
3. Carlson, R. K., (1944), "An Experimental Investigation of the Nosing of
Shells," Forging of Steel Shells, ASME Trans., p. 45.
4. Lahoti, G. D., Subramanian, T. L., and Allan, T., (1978), "Development of a
Computerized Mathematical Model for the Hot/Cold Nosing of Shells,"
Report ARSCD-CR-78019 to U.S. Army Research and Development
Command.
5. Kobayashi, S., (1983), "Approximate Solutions for Preform Design in Shell
Nosing," Int. J. Machine Tool Des. Res., Vol. 23, p. 111.
6. Johnson, W., (1980), The Mechanics of Some Industrial Pressing, Rolling, and
Forging Processes, in "Mechanics of Solids," (Edited by H. G. Hopkins and
M. J. Sewell), Pergamon Press, Oxford, p. 303.
7. Hwang, S. M., and Kobayashi, S., (1984), "Preform Design in Plane-Strain
Rolling by the Finite Element Method." Int. J. Machine Tool Des. Res., Vol.
24, p. 253.
8. Kobayashi, S., (1984), "Approximate Solutions for Preform Design in Roll-
ing," Int. J. Machine Tool Des. Res., Vol. 24, p. 215.
9. Biswas, S. K., and Knight, W. A., (1974), "Computer-Aided Design of
Axisymmetric Hot Forging Dies," Proc. 15th Int. MTDR Conf., p. 135.
10. Spies, K., (1957), "The Preforms in Closed-Die Forging and Their Preparation
by Reducer Rolling" (in German), Doctoral Dissertation Technical University,
Hannover.
11. Chamouard, A., (1964), "Estampage et Forge." Dunod, Paris.
12. Akgerman, N., Becker, J. R., and Allan, T., (1973), "Preform Design in
Closed Die Forging," Metal Forming, Vol. 40, p. 135.
13. Yu, G. B., and Dean, T. A., (1985), "A Practical Computer-Aided Approach
to Mould Design for Axisymmetric Forging Die Cavities," Int. J. Machine
Tool Des. Res., Vol. 25, p. 1.
14. Hwang, S. M., and Kobayashi, S., (1986), "Preform Design in Disk Forging,"
Int. J. Machine Tool Des. Res., Vol. 26, p. 231.
15. Hwang, S. M., and Kobayashi, S., (1987), "Preform Design in Shell Nosing al
Elevaled Temperatures," Int. J. Machine Tool Manufacture, Vol. 27, p. 1.
16. Lahoti, G. D., Oh, S. I., and Allan, T., (1981), "Development and
Confirmation of a Series of Mathematical Models for the Blocking, Cabbaging,
Piercing, and Nosing Operations Involved in Shell Manufacluring," Report
ARSCD-CR-81010, to U.S. Army Research and Development Command.
17. Tang, M. C., and Kobayashi, S., (1982), "An Investigation of the Shell Nosing
Process by the Finite Element Method, Part 2. Nosing at Elevated Tempera-
tures (Hot Nosing)," ASME Trans., J. Engr. Ind., Vol. 104, p. 312.
16
SOLID FORMULATION, COMPARISON
OF TWO FORMULATIONS, AND
CONCLUDING REMARKS
16.1 Introduction
In the previous chapters we have discussed only the applications of flow
formulation to the analysis of metal-forming processes. Lately, elastic-
plastic (solid) formulations have evolved to produce techniques suitable for
metal-forming analysis. This evolution is the result of developments
achieved in large-strain formulation, beginning from the infinitesimal
approach based on the Prandtl-Reuss equation.
A question always arises as to the selection of the approach—"flow"
approach or "solid" approach. A significant contribution to the solution of
this question was made through a project in 1978, coordinated by Kudo
[1], in which an attempt was made to examine the comparative merits of
various numerical methods. The results were compiled for upsetting of
circular solid cylinders under specific conditions, and revealed the impor-
tance of certain parameters used in computation, such as mesh systems and
the size of an increment in displacement. This project also showed that the
solid formulation needed improvement, particularly in terms of predicting
the phenomenon of folding. For elastic-plastic materials, the constitutive
equations relate strain-rate to stress—rates, instead of to stresses. Conse-
quently, it is convenient to write the field equation in the boundary-value
problem for elastic-plastic materials in terms of the equilibrium of stress
rates.
In this chapter, the basic equations for the finite-element discretization
involved in solid formulations are outlined both for the infinitesimal
approach and for large-strain theory. Further, the solutions obtained by
the solid formulation are compared with those obtained by the flow
formulation for the problems of plate bending and ring compression. A
discussion is also given concerning the selection of the approach for the
analysis.
In conclusion, significant recent developments in the role of the
finite-element method in metal-forming technology are summarized.
where dtj is the Kronecker delta. The first term of eq. (16.2) can be
rewritten, using the strain-hardening relationship in the form of eq. (4.37),
as
where H' is the slope of the stress-plastic strain curve H' — da/de. In
deriving eq. (16.3), note that
oa = \ o'mnd'mn = | o'mn(amn - &mn\akk)
where o'mn5mn = 0. Then, eq. (16.2) can be written in the form
For discretization of eq. (16.1), we need the inversion of eq. (16.4). The
inversion was obtained by Yamada et al. [2] as
For the last term in eq. (16.5), the coefficient a is added such that a = 0
for elastic elements and a = 1 for plastic elements. After an incremental
Solid Formulation, Comparison of Two Formulations 323
REFERENCE STATE
where p0 and p are the material densities at the reference and the
deformed states, respectively. The time derivative of the stress measure is
the Jauman derivative (DTij/Dt), where time differentiation is carried out
with the coordinate system that rotates but does not deform with the
material. It is expressed by
Then
Using the symmetry of Dr^lDf, e,7 and a^, eq. (16.13) can be rewritten in
the form
Assuming that the Prandtl-Reuss equation holds for the Jauman deriva-
tive of the Kirchhoff stress and strain rates, the constitutive relation is
given by
Equations (16.14) and (16.15) are the bases for the finite-element
discretization.
In the updated Lagrangian description, the reference state is assumed to
coincide with the current state at time t. The procedures used to integrate
the constitutive equations for a time-step Af are based on the mean normal
method of Rice and Tracey [9] and the radial return method of Krieg and
Krieg [10]. These two methods are further explained in Reference [28].
The formulations described above have been used for the analyses of
various forming processes. A complete stress and deformation analysis has
been given for plane-strain and axisymmetric extrusion until the steady
state is reached [11,12]. In applying the finite-element method to
can-extrusion the main problem was to simulate the metal flow around a
sharp corner [13]. The use of a special element to overcome this problem
was discussed.
Complete solutions of stretch-forming and deep drawing problems,
taking into account the contact problem at blank holder, die, die profile,
and punch head, were obtained by Wifi [14]. On the basis of the nonlinear
theory of membrane shells, Wang and Budiansky [15] developed the
procedure for calculating the deformations in the stamping of sheet metal
by arbitrarily shaped punches and dies.
Plane-strain rolling was analyzed by Rao and Kumar [16]. Key et al. [17]
obtained the solutions for rolling, extrusion, and sheet stretching. Hartley
et al. [18] developed a method for introducing friction into the finite-
element analysis [19].
Hartley et al. [20] also included the effects of strain-rate and tempera-
ture variations within the billet in the analysis of axial compression of a
solid cylinder. The fixed-mesh updated-Lagrangian technique was pro-
326 Metal Forming and the Finite-Element Method
At the end of the increment At, the spatial position of the body x, is
described by
(16.18)
with Aw, being the incremental displacement. Then eq. (16.16b) implies
that the virtual work-rate principle, applicable at the end of the increment,
t Note that Uj has been used for velocity throughout this book but is used for displacement
in Sections 11.4 through 11.7 of Chap. 11 and in this section.
Solid Formulation, Comparison of Two Formulations 327
is written as
where &Ekl is the increment of Lagrangian strain and <£ikjl are the
constitutive moduli. The increment of the Lagrangian strain A£w is related
to the displacement increment according to
The moduli J£ijk, are not the small-strain moduli Lijkl given by eq. (16.5),
and are given by
FIG. 16.2 Finite-element mesh system of the undeformed workpiece for (a) rigid-plastic and
(b) elastic-plastic analyses in plate bending.
FIG. 16.3 Variations of normalized bending stress along the thickness directions at various
cross sections (as indicated in Fig. 16.2), for punch displacements of (a) 2.25 and (b) 3.2.
330 Metal Forming and the Finite-Element Method
ELASTO-PLASTIC RIGID-PLASTIC
FIG. 16.4 Variation of effective strain along the thickness direction at various cross sections
(as indicated in Fig. 16.2), for punch displacements of (a) 2.25 and (b) 3.2.
strains near the outer fiber and lower values near mid-surface, than the
corresponding values predicted by the elastic-plastic analysis.
One of the main disadvantages of the rigid-plastic formulation is that it
cannot predict unloading behavior. However, in the case where the
unloading conditions can be well-defined, such as in sheet bending, it is
possible to calculate springback and the residual stress distribution by
performing the elastic-plastic unloading calculations using the results of the
rigid-plastic loading solution.
The springback angle calculated with the elastic-plastic solution is 3.03°,
and 2.94° when obtained from the other solution. Figure 16.5 shows the
distributions of the bending component of residual stress along the
thickness of several selected cross-sectional areas for both cases. The
results show that the residual stresses obtained from the elastic-plastic
solution are somewhat higher along the cross sections near the bending
axis. Comparison of the two solutions shows excellent agreement between
them, not only in overall quantities but also in detail.
The amount of springback was obtained by performing unloading
calculations. Two methods were used, namely, elastic-plastic loading and
unloading, and rigid-plastic loading and elastic-plastic unloading. The
results show that the springback angle, as well as the residual stress
distributions, calculated by both methods are in excellent agreement.
Solid Formulation, Comparison of Two Formulations 331
FIG. 16.5 Normalized residual stress distributions in plate bending: (a) elastic-plastic loading
and unloading; (b) rigid-plastic loading and elastic-plastic unloading.
FIG. 16.6 Deformed mesh in ring compression (shown for upper half of the ring
cross-section) after 50% reduction in height [29]: (a) elastic-plastic analysis; (b) rigid-plastic
analysis.
16.6. It is clear that the two solutions give almost identical results in detail.
Figure 16.8 shows plots of nodal-point loads at the die-workpiece
interface. The closeness of the two distributions confirms that both
formulations are physically consistent.
Comments on Comparison
The comparison shown above indicates that both the flow and solid
formulations are capable of predicting certain physical phenomena in
metal forming. A variety of formulations and algorithms are now available
for simulation of forming processes. The following remarks may be helpful
for the selection of an approach to a specific forming problem.
When we consider simulation of forming processes, it is important for an
engineer to be aware of the objective of simulation. Specific information
will be sought by simulation, depending on the objective. Each problem
has its peculiarities and specific requirements, and the selected method
must have the capability of delivering the required information. The
FIG. 16.7 Equivalent plastic strain contours in ring compression at 50% reduction in height
[29]: (a) elastic-plastic analysis; (b) rigid-plastic analysis.
Solid Formulation, Comparison of Two Formulations 333
FIG. 16.8 Axial nodal force distributions along the die-workpiece contact surface in ring
compression [29]: (a) elastic-plastic analysis: (b) rigid-plastic analysis.
Tehran! [40]. Lush and Anand [41] proposed implicit time integration
procedures for internal variables of constitutive equations for large-
deformation elasto-viscoplasticity.
Pecherski [42] presented the phenomenological model for large plastic
strains and the onset of deformation instability. For calculation of the
flatness defect of rolled strip, Yukawa et al. [43] solved the eigenvalue
problem for the buckling limit and determined the shape after buckling by
the large-displacement finite-element method. Mori et al. [44], in the
coupled analysis of a forming process with elastic tools, treated the plastic
deformation of the workpiece by the rigid-plastic finite-element method
and the elastic deflection by the boundary-element method.
These investigations provide some indication for ever-increasing ap-
plications of the finite element method in metal forming.
References
1. Kudo, H., and Matsubara, S., (1979), Joint Examination Project of Validity of
Various Numerical Methods for the Analysis of Metal Forming Processes, in
"Metal Forming Plasticity," (Edited by H. Lippmann), Springer-Verlag,
Berlin, p. 378.
2. Yamada, Y., Yoshimura, H., and Sakurai, T., (1968), "Plastic Stress-Strain
Matrix and its Application for the Solution of the Elastic-Plastic Problem by
the Finite Element Method," Int. J. Mech. Sci., Vol. 10, p. 343.
3. Yamada, Y., and Takatsuka, K,., (1973), "Finite-Element Analysis of
Nonlinear Problems," J. Japan Soc. Tech. Plast., Vol. 14, p. 758.
4. Larsen, P. K., (1971), "Large Displacement Analysis of Shells of Revolution,
Including Creep, Plasticity and Viscoplasticity," Ph.D. Thesis, University of
California, Berkeley.
5. Needleman, A., (1970), "Void Growth in an Elastic-Plastic Medium," Ph.D.
Thesis, Harvard University.
6. McMeeking, R. M., and Rice, J. R., (1975), "Finite-Element Formulation for
Problems of Large Elastic-Plastic Deformation," Int. J. Solids Structures, Vol.
11, p. 601.
7. Hill, R., (1958), "A General Theory of Uniqueness and Stability in Elastic-
Plastic Solids," /. Mech. Phys. Solids, Vol. 6, p. 231.
8. Hill, R., (1959), "Some Basic Principles in the Mechanics of Solids Without a
Natural Time," /. Mech. Phys. Solids, Vol. 7, p. 209.
9. Rice, J. R., and Tracy, D. M., (1983), "Computational Fracture Mechanics,"
Proc. Symp. on Numerical and Computer Methods in Structure Mechanics,
Urbana IL, Academic Press, Orlando, p. 585.
10. Krieg, R. D., and Krieg, D. B., (1977), "Accuracies of Numerical Solution
Methods for the Elastic-Perfectly Plastic Model," /. Pressure Vessel Tech.,
ASME, Vol. 90, p. 510.
11. Lee, E. H., Mallett, R. L., and Yang, W. H., (1977), "Stress and
Deformation Analysis of the Metal Extrusion Process," Computer Methods
Appl. Mech. Engr., Vol. 10, p. 339.
12. Lee, E. H., Mallett, R. L., and McMeeking, R. M., (1977), "Stress and
Deformation Analysis of Metal Forming Processes," Numerical Modeling of
Manufacturing Processes, ASME Special Publication PVP-PB-025, p. 19.
13. Hartley, P., Sturgess, C. E. N., and Rowe, G. W., (1978), "A Finite Element
Analysis of Extrusion-Forging," Proc. VI North American Metalworking
Research Conference, SME, p. 212.
336 Metal Forming and the Finite-Element Method
14. Wifi, A. S., (1976), "An Incremental Complete Solution of the Stretch-
Forming and Deep-Drawing of a Circular Blank Using a Hemispherical
Punch," Int. J. Mech. ScL, Vol. 18, p. 23.
15. Wang, N. M., and Budiansky, B., (1978), "Analysis of Sheet Metal Stamping
by a Finite Element Method," ASME Trans. J. Appl. Mech., vol. 45, p. 73.
16. Rao, S. S., and Kumar, A., (1977), "Finite Element Analysis of Cold Strip
Rolling," Int. J. Machine Tool Des. Res., Vol. 17, p. 159.
17. Key, S. W., Krieg, R. D., and Bathe, K. J., (1979), "On the Application of
the Finite Element Method to Metal Forming Processes—Part 1," Computer
Methods Appl. Mech. Engr., Vol. 17/18, p. 597.
18. Hartley, P., Sturgess, C. E. N., and Rowe, G. W., (1979), "Friction in Finite
Element Analysis of Metal Forming Processes," Int. J. Mech. ScL, Vol. 21, p.
301.
19. Hartley, P., Sturgess, C. E. N., and Rowe, G. W., (1979), "A Prediction of
the Influence of Friction in the Ring Test by the Finite-Element Method,"
Proc. 7th NAMRC, Ann Arbor, Michigan, p. 1.
20. Hartley, P., Sturgess, C. E. N., and Rowe, G. W., (1980), "Finite Element
Predictions of the Influence of Strain-rate and Temperature Variations on the
Properties of Forged Products," Proc. 8th NAMRC, University of Missouri,
Rolla, MO., p. 121.
21. Derbalian, K. A., Lee, E. H., Mallett, R. L., and McMeeking, R. M., (1978),
"Finite Element Metal Forming Analysis with Spatially Fixed Mesh," Appli-
cations of Numerical Methods to Forming Processes, ASME, AMD, Vol. 28,
p. 39.
22. Wang, N. M., (1980), "On Analytical Formulations of Computational Sheet
Metal Forming," Proc. 4th International Conference on Production Engineer-
ing, Tokyo, p. 28.
23. Oh, S. I., and Kobayashi, S., (1980), "Finite Element Analysis of Plane-Strain
Sheet Bending," Int. J. Mech. ScL, Vol. 22, p. 583.
24. Nagtegaal, J. C., and DeJong, J. E., (1981), "Some Computational Aspects of
Elastic-Plastic Large Strain Analysis," Int. J. Num. Methods Engr., Vol. 17, p.
15.
25. Nagtegaal, J. C., (1982), "On the Implementation of Inelastic Constitutive
Equations with Special Reference to Large Deformation Problems," Computer
Methods Appl. Mech. Engr., Vol. 33, p. 469.
26. Nagtegaal, J. C., and Veldpaus, F. E., (1984), On the Implementation of
Finite Strain Plasticity Equations in a Numerical Model, in "Numerical
Analysis of Forming Processes", (Edited by J. F. T. Pittman, O. C.
Zienkiewicz, R. D. Wood and J. M. Alexander), Wiley, New York, p. 351.
27. Rebelo, N., and Wertheimer, T. B., (1986), "General Purpose Procedures for
Elastic-Plastic Analysis of Metal Forming Processes," Proc. 14th NAMRC,
Minneapolis, MN., p. 414.
28. Nagtegaal, J. C., and Rebelo, N., (1986), On the Development of a General
Purpose Finite Element Program for Analysis of Forming Processes, in
"NUMIFORM 86," Balkema, Rotterdam, p. 41 [see Ref. 32].
29. Rebelo, N., Nakazawa, S., Wertheimer, T. B., and Nagtegaal, J. C., (1986),
"A Comparative Study of Algorithms Applied in Finite Element Analysis of
Metal Forming Problems," Special Publication ASME, PED-Vol. 20, p. 17.
30. Tomita, Y., (1982), "A Rigid Plastic Finite Element Method for the Effective
Prediction of the Influence of the Change in Parameters in the Constitutive
Equation upon Deformation Behavior," Int. J. Mech. ScL, Vol. 24, p. 711.
31. Eggert, G. M., and Dawson, P. R., (1987), "On the Use of Internal Variable
Constitutive Equations in Transient Forming Processes," Int. J. Mech. Sci.,
Vol. 29, p. 95.
32. NUMIFORM 86 (1986), "Numerical Methods in Industrial Forming Proc-
Solid Formulation, Comparison of Two Formulations 337
A.I Introduction
The development and testing of the computer code is an important part of
the finite-element method. In programming a practical finite-element
method code, one should consider various factors, such as generality,
computational efficiency, and pre- and postprocessing. However, such
considerations are beyond the scope of this book. Instead, a simple
two-dimensional FEM code, called SPID, written for metal-forming
simulation is presented as an example and discussed in this appendix. The
main purpose of the code is to illustrate the programming of key steps
explained in this book.
The program SPID, Simple Plastic /ncremental Deformation, is based
on rigid viscoplastic finite-element formulation. SPID is capable of
handling only simple forming processes, such as simple compression and
ring compression. Some important features and limitations of SPID are
summarized below.
• SPID is valid for rigid plastic as well as rigid viscoplastic materials.
• A constant shear factor friction law is used.
• No heat transfer simulation is included.
• It can handle one flat die only.
• It cannot handle free surface folding to the die.
• The initial guess is generated automatically.
• A banded matrix solver is used.
• SPID can handle a finite-element model with up to 100 nodes.
SPID is written in FORTRAN 77 standard, with special consideration of
portability. The program was tested on two computer systems. One was
VAX 750 with VMS version 4.4 operating system. The other was IBM
PC/AT with a math coprocessor, where the professional FORTRAN compiler
of Ryan-McFarland Corp. and DOS 3.10 operating system of Micro Soft
Inc. were used. The required minimum memory size for the IBM PC to
run SPID is 256 K Bytes.
338
Appendix The FEM Code, SPID 339
A.2 Program Structure
The program structure of SPID in terms of subroutine calling sequence is
shown schematically in Fig. A.I. Brief descriptions of each subroutine are
given as comment lines in the program listing in Section A.6. The
functional procedure of SPID is described below. Subroutine names
corresponding to each logical step are also given in parenthesis.
1. Program starts. (SPID)
2. Read input from input file SPID.DAT. (INPRED)
3. Print input information. (PRTINP)
4. Determine the maximum half-bandwidth and number of stiffness
equations. (BAND)
5. If initial guess is required, then set up for the direct iteration
procedure. If not, then use Newton-Raphson iteration method.
(NONLIN)
6. Evaluate elemental strain-rate matrix. (STRMTX)
7. Evaluate elemental stiffness matrix. (ELSHLF, VSPLON,
VSPLST)
FIG. A.I Subroutine calling sequence of SPID. The names enclosed in boxes are the
subroutine names and the double-line box indicates the continuation of the diagram.
340 Metal Forming and the Finite-Element Method
C
C METAL FORMING AND THE FINITE ELEMENT METHOD
C
C BY SHIRO KOBAYASHI, SOO-IK OH AND TAYLAN ALTAN
C
C OXFORD UNIVERSITY PRESS, 1988
C
C THIS IS MAIN PROGRAM OF RIGID VISCOPLASTIC FINITE
C ELEMENT METHOD FOR SIMPLE FORMING PROCESSES.
C
IMPLICIT INTEGERS (I-N), REAL*8 (A-H, 0-Z)
CHARACTER TITLE*70
COMMON /TITL/ TITLE
COMMON /RVA2/ EPS(5,100), STS(5,100), TEPS(IOO)
COMMON /INOT/ INPT,MSSG,IUNIT,IUNI2
COMMON /MSTR/ NUMNP.NUMEL.IPLNAX
COMMON /INVR/ NOD(4,100),NBCD(2,100)
COMMON /TSTP/ NINI,NCUR,NSEND,NITR,DTMAX
COMMON /ITRC/ ITYP.ICONV
INPT - 5
MSSG - 6
IUNIT = 3
IUNI2 = 4
C
C READ INPUT
C
CALL INPRED
OPEN(IUNIT,FILE='SPID.OUT',STATUS='UNKNOWN',FORM='FORMATTED')
OPEN(MSSG, FILE='SPID.MSG',STATUS='UNKNOWN',FORM='FORMATTED')
WRITE(MSSG,1020) TITLE
CALL PRTINP
CALL BAND(NOD,NUMEL,NUMNP)
C
C STEP SOLUTIONS
C
NINI = NINI + 1
C
DO 300 N = NINI, NSEND
NCUR - N
C
WRITE(MSSG,1050) N
IF(N .NE. NINI) GO TO 80
ICOUNT = 0
50 ITYP = 2
CALL NONLIN
ICOUNT - ICOUNT + 1
80 ITYP - 1
CALL NONLIN
IF(ICONV .EQ. 2 .AND. ICOUNT .GT. 3) GO TO 900
Appendix The FEM Code, SPID 345
IF(ICONV .EQ. 2) GO TO 50
C
CALL POTSOL
CALL PRTSOL
CALL RSTFIL
300 CONTINUE
C
CLOSE(IUNIT)
CLOSE(MSSG)
STOP
C
900 CONTINUE
WRITE(MSSG,1070)
C
STOP
1020 FORMAT(1H1,//,5X,'OUTPUT OF S P I D',//,
+ 5X,' MESSAGE FILE FOR '/,5X,A,//)
1050 FORMAT(///,' ITERATION PROCESS FOR STEP ',I5,//)
1070 FORMAT{/,' STOP BECAUSE SOLUTION DOES NOT CONVERGE. ')
END
c
SUBROUTINE ADDBAN(B,A,NQ,LM,Qq,PP)
C
C CHAPTER 7.2,
C EQUATION (7.8)
C
C ASSEMBLE GLOBAL STIFFNESS MATRIX FROM ELEMENTAL STIFFNESS MATRIX
C
IMPLICIT REAL*8 (A-H, 0-Z), INTEGERM (I-N)
DIMENSION B(l), A(NQ,1), QQ(1), PP(8,8), LM(1)
C
DO 100 I = 1, 8
II = LM(I)
DO 50 J - 1, 8
JJ = LM(J) - LM(I) + 1
IF(JJ .LE. 0) GO TO 50
A(II.JJ) = A(II,JJ) + PP(I.J)
50 CONTINUE
B(II) = B(II) + QQ(I)
100 CONTINUE
RETURN
END
c
SUBROUTINE BAND(NOD,NUMEL,NUMNP)
C
C CHAPTER 7.2
C DETERMINE MAXIMUM HALF BANDWIDTH, MBAND AND
C TOTAL NUMBER OF EQUATIONS, NEQ
C
IMPLICIT REAL*8 (A-H,0-Z), INTEGERM (I-N)
COMMON /CNEQ/ NEQ,MBAND
DIMENSION NOD(4,1)
C
MBAND = 0
346 Metal Forming and the Finite-Element Method
DO 100 N - 1, NUMEL
NMIN - NOD(1,N)
NHAX - NOD(l.N)
DO 50 I - 2,4
IF(NMIN .GT. NOD(I.N)) NMIN - NOD(I.N)
IF(NHAX .LT. NOD(I,N)) NMAX = NOD(I.N)
50 CONTINUE
MB - (NMAX - NMIN + 1) * 2
IF(MBAND .LT. MB) MBAND = MB
100 CONTINUE
NEQ - NUHNP * 2
RETURN
END
c
SUBROUTINE BANSOL(B,A,NQ,MM)
C
C CHAPTER 7.2
C THIS SUBROUTINE SOLVES THE BANDED SYMMETRIC MATRIX EQUATIONS BY
C THE GAUSSIAN ELIMINATION.
C B LOAD VECTOR
C A SYMMETRIC MATRIX IN BANDED FORM
C MM HALF BANDWIDTH
C NQ NUMBER OF EQUATIONS
C
IMPLICIT REAL*8 (A-H, 0-Z), INTEGERS (I-N)
COMMON /INOT/ INOT,MSSG,IUNIT,IUNI2
DIMENSION B(l), A(NQ,1)
C
DO 200 N = 1, NQ
IF(A(N,1) .LE. 0.) GO TO 800
DO 150 L = 2, MM
IF(A(N,L) .EQ. 0.) GO TO 150
C = A(N,L) / A(N,1)
I = N + L -1
J - 0
DO 100 K = L, MM
J - J +1
100 A(I,J) - A(I,J) - C*A(N,K)
A(N,L) = C
150 CONTINUE
200 CONTINUE
C
C LOAD VECTOR REDUCTION
C
DO 300 N - 1, NQ
DO 250 L = 2, MM
I = N + L - 1
IF(I .GT. NQ) GO TO 250
B(I) = B(I) - A(N,L) * B(N)
250 CONTINUE
B(N) = B(N) / A(N,1)
300 CONTINUE
C
C BACK SUBSTITUTION
Appendix The FEM Code, SPID 347
r
DO 400 M - 1,NQ
N = NQ + 1 - M
DO 350 K - 2, MM
L - N + K -1
350 B(N) - B(N) - A(N,K)*B(L)
400 CONTINUE
RETURN
C
800 CONTINUE
WRITE(MSSG,1020) N
STOP
1020 FORMAT(/,' NEGATIVE PIVOT AT EQUATION NO. ',15)
END
c
SUBROUTINE DISBDY(URZ,NBCD,B,A,NEQ,MBAND,ITYP)
C
C CHAPTER 7.3
C APPLY DISPLACEMENT BOUNDARY CONDITION
C
IMPLICIT REAL*8 (A-H.O-Z), INTEGERM (I-N)
DIMENSION B(1),A(NEQ,1),NBCD(1),URZ(1)
C
IF(ITYP .EQ. 2) GO TO 120
DO 100 N > 1, NEQ
IF(NBCD(N) .EQ. 0) GO TO 100
DO 70 I = 2, MBAND
II - N - I + 1
IF(II .LE. 0) GO TO 50
A(II,I) = 0.
50 CONTINUE
II - N + I - 1
IF(II .GT. NEQ) GO TO 70
A(N,I) - 0.
70 CONTINUE
B(N) - 0.
A(N,1) = 1.
100 CONTINUE
RETURN
C
120 CONTINUE
DO 200 N = 1, NEQ
IF(NBCD(N) .EQ. 0) GO TO 200
DO 170 I = 2, MBAND
II - N - I + 1
IF(II .LE. 0) GO TO 150
B(II) = B(II) - A(II,I) * URZ(N)
A(II,I) - 0.
150 CONTINUE
II - N + I - 1
IF(II .GT. NEQ) GO TO 170
B(II) = B(II) - A(N,I) * URZ(N)
A(N,I) = 0.
170 CONTINUE
348 Metal Forming and the Finite-Element Method
B(N) * URZ(N)
A(N,1) - 1.
200 CONTINUE
END
c
SUBROUTINE ELSHLFfPP.QQ.RZ.URZ.EPS.TEPS.IPLNAX.IDREC.NEL)
C
C CHAPTER 6.5
C EQUATIONS (6.43) AND (6.44)
C EVALUATION OF ELEMENTAL STIFFNESS MATRIX
C
C IDREC : IF = 1, NEWTON-RAPHSON ITERATION
C = 2, DIRECT ITERATION
C
IMPLICIT REAL*8 (A-H, 0-Z), INTEGER** (I-N)
DIMENSION RZ(2,1), URZ(2,1), B(4,8), EPS(l), TEPS(l)
DIMENSION QQ(1),PP(8,8),S2(2),W2(2)
DATA S2/-0.57735026918963DO,0.57735026918963DO/,
+ W2/2*1.0DO/
C
DO 10 I - 1, 8
QQ(I) - 0.
DO 10 J = 1, 8
PP(I,J) = 0.
10 CONTINUE
C
C CARRY OUT ONE POINT INTEGRATION
C
S = 0.
T = 0.
CALL STRMTXfRZ.B.WDXJ.SJ.IPLNAX.NEL)
WDXJ = WDXJ * 4
CALL VSPLON(QQ,PP,B,URZ,EPS,WDXJ,IDREC)
C
C REGULAR INTEGRATION
C
DO 100 I - 1, 2
S - S2(I)
DO 50 J - 1, 2
T - S2(J)
CALL STRMTXfRZ.B.WDXJ.S.T.IPLNAX.NEL)
WDXJ - WDXJ * W2(I)*W2(J)
CALL VSPLST(QQ,PP,B,URZ,TEPS,WDXJ,IDREC)
50 CONTINUE
100 CONTINUE
RETURN
END
C
SUBROUTINE FLWSTS(YS,FIP,STRAN,STRRT)
C
C USER SUPPLIED SUBROUTINE TO DESCRIBE THE MATERIAL
C FLOW STRESS.
C THIS SUBROUTINE SHOWS THE EXAMPLE OF
C YS - 10. * (STRAIN RATE) ** 0.1
Appendix The FEM Code, SPID 349
C
IMPLICIT INTEGER*4 (I.J.K.L.M.N), REAL*8 (A-H.O-Z)
COMMON /RIGD/ RTOL, ALPH, DIAT
C
IF(STRRT .LT. ALPH) GO TO 100
YS = 10. * STRRT ** 0.1
FIP = STRRT ** (-0.9)
RETURN
C
100 YO = 10. * ALPH ** 0.1
FIP - YO / ALPH
YS = FIP * STRRT
RETURN
END
c
SUBROUTINE FRCBDY(RZ,URZ,NBCD,TEPS,EPS,QQ,PP,IPLNAX)
C
C APPLY FRICTION BOUNDARY CONDITION
C
IMPLICIT INTEGERM (I-N), REAL*8 (A-H, 0-Z)
COMMON /DIES/ FRCFAC
COMMON /INOT/ INPT, MSSG, IUNIT, IUNI2
DIMENSION RZ(2,1),URZ(2,1),NBCD(2,1),EPS(5),QQ(1},PP(8,1),
+ ER(2,2),FR(2),XY(2,2),VXY(2,2)
C
DO 100 N = 1,4
11 = N + 1
12 = N
IF(N .EQ. 4) II = 1
IF(NBCD(2,I1) .NE. 3 .OR. NBCD(2,I2) .NE. 3) GO TO 100
CALL FLWSTS(FLOW,DUM,TEPS,EPS(5))
XY(1,1) - RZ(1,I1)
XY(2,1) = RZ(2,I1)
XY(1,2) - RZ(1,I2)
XY(2,2) = RZ(2,I2)
VXY(l.l) - URZ(1,I1)
VXY(2,1) = URZ(2,I1)
VXY{1,2) - URZ(1,I2)
VXY(2,2) - URZ(2,I2)
CALL FRCINT(XY,VXY,FLOW,FR.ER,FRCFAC,IPLNAX)
Jl = II * 2 - 1
J2 - 12 * 2 - 1
QQ(J1) - QQ(J1) + FR(1)
QQ(J2) = QQ(J2) + FR(2)
PP(J1,J1) = PP(J1,J1) + ER(1,1)
PP(J2,J2) - PP(J2,J2) + ER(2,2)
PP(J1,J2) - PP(J1,J2) + ER(1,2)
PP(J2,J1) = PP(J2,J1) + ER(2,1)
100 CONTINUE
RETURN
END
c
SUBROUTINE FRCINT(RZ,URZ,FLOW,FR.ER,FRCFAC,IPLNAX)
C
350 Metal Forming and the Finite-Element Method
C CHAPTER 7.3
C EQUATIONS (7.14) AND (7.15)
C INTEGRATION METHOD : SIMPSON'S FORMULA
C THIS ROUTINE CALCULATES THE FRICTION MATRIX
C USED FOR BOTH TYPES OF ITERATION SCHEME
C
IMPLICIT INTEGERM (I-N), REAL*8 (A-H, 0-Z)
COMMON /INOT/ INPT,MSSG,IUNIT,IUNI2
COMMON /ITRC/ ITYP.ICONV
DIMENSION RZ(2,1),URZ(1),QQ(1),PP(8,1)
DIMENSION SLIV(2),ER(2,2),FR(2)
DATA PI/3.1415926535898DO/
DATA UA/0.0005DO/
C
C INITIALIZE FR AND ER ARRAY
C
DO 10 I « 1, 2
FR(I) = 0.
DO 10 0 = 1, 2
ER(I,J) - 0.
10 CONTINUE
MINT = 5
FAC = DSQRT((RZ(1,2)-RZ(1,1))**2 + (RZ(2,2
FK FLOW * FRCFAC / SQRT(3.)
DH 2. / (NINT - 1)
S -1. - DH
CON 2. / PI * FK
WD DH / 3. * FAC * 0.5 * CON
C
DO 300 N - 1, NINT
S = S + DH
HI = 0.5 * (1. - S)
H2 = 0.5 * (I. + S)
WDXJ = WD
IF(IPLNAX .NE. 1) GO TO 90
RR = H1*RZ(1,1) + H2*RZ(1,2)
WDXJ - RR * WDXJ
90 CONTINUE
C
IF(N .EQ. 1 .OR. N .EQ. NINT) GO TO 100
NMOD = N - N/2*2
IF(NMOD .EQ, 0) WDXJ = WDXJ * 4
IF(NMOD .EQ. 1) WDXJ = WDXJ * 2
100 CONTINUE
C
US * H1*URZ(1) + H2*URZ(3)
AT = DATAN(US/UA)
IF(ITYP .EQ. 2) GO TO 200
US2 = US * US
USA - US2 + UA*UA
CT1 = AT*WDXJ
CT2 - UA/USA*WDXJ
GO TO 250
C
Appendix The FEM Code, SPID 351
c
SUBROUTINE NFORCE(QQ,FRZ,LM)
C
C ADD NODAL POINT FORCE
C
IMPLICIT REAL*8 (A-H.O-Z), INTEGERM (I-N)
DIMENSION QQ{1),FRZ(1),LM(1)
C
DO 100 I = 1, 8
N - LM(I)
FRZ(N) - FRZ(N) - QQ(I)
100 CONTINUE
RETURN
END
c
SUBROUTINE NONLIN
C
C THIS ROUTINE CONTROLS THE ITERATIONS
C
IMPLICIT INTEGERM (I-N), REAL*8 (A-H, 0-Z)
COMMON /INOT/ INPT.MSSG,IUNIT.IUNI2
COMMON /MSTR/ NUMNP.NUMEL.IPLNAX
COMMON /TSTP/ NINI,NCUR,NSEND,NITR,DTMAX
COMMON /ITRC/ ITYP, ICONV
COMMON /CNEQ/ NEQ,MBAND
COMMON /RVA1/ RZ(2,100), URZ(2,100), FRZ(2,100)
DIMENSION UNORM(2),ENORM(2),FNORM(2)
COMMON A(5000),B(200)
C
RTOL •= 0.00001
IF(ITYP .EQ. 2) RTOL - 0.0005
ACOEF - 1.
NSTEL - NEQ * MBAND
IF(NSTEL.LE.5000 .AND. NEQ.LE.200) GO TO 10
HRITE(MSSG,1010)
STOP
C
10 CONTINUE
DO 30 N - 1 , 2
UNORM(N) = 0.
ENORM(N) - 0.
FNORM(N) - 0.
30 CONTINUE
C
ITRMAX - 20
IF(ITYP .EQ. 2) ITRMAX - 200
DO 200 N = 1, ITRMAX
NITR = N
CALL STIFF(B,A,NEQ,MBAND,ITYP)
IDREC = 1
CALL NORM(FRZ,B,FDUM,DFN,NEQ,IDREC)
IF(ITYP .EQ. 2) DFN = 0.
CALL BANSOL(B,A,NEQ,MBAND)
IDREC - ITYP
354 Metal Forming and the Finite-Element Method
CALL NORM(URZ,B,UC,EC,NEQ,IDREC)
IF(ITYP .EQ. 1) WRITE(MSSG,1030) N
IF(ITYP .EQ. 2) WRITE(MSSG,1050) N
WRITE(HSSG,1070) UC.EC.DFN
IF(N .EQ. 1) GO TO 130
IF(EC .LT. RTOL .AND. DFN .LT. RTOL) GO TO 300
IF(ITYP .EQ. 2) GO TO 130
IF(EC .LT. ENORH(2)) GO TO 100
C
C ADJUST THE ACOEF
C
ACOEF = ACOEF * 0.7
GO TO 130
100 CONTINUE
IF(ENORM(1) .GT. ENORM(2) .AND. ENORM(Z) .GT. EC)
+ ACOEF - ACOEF * 1.3
IF(ACOEF .GT. 1.) ACOEF = 1.0
C
C VELOCITY UPDATE
C
130 CONTINUE
NB - 0
DO 150 I - 1, NUMNP
DO 150 J - 1,2
NB - NB + 1
IF(ITYP .EQ. 1) URZ(J.I) - URZ(J.I) + ACOEF * B(NB)
IF(ITYP .EQ. 2) URZ(J,I) = B(NB)
150 CONTINUE
C
170 CONTINUE
UNORM(l) UNORM(2)
ENORM(l) ENORH(2)
FNORH(l) FNORM(2)
UNORM(2) UC
ENORM(2) EC
FNORM(2) DFN
200 CONTINUE
C
C SET FLAG
C
ICONV - 2
RETURN
C
300 CONTINUE
C CONVERGED CASE
C SET FLAG
C
ICONV - 1
C
RETURN
1010 FORHAT(/,' YOU NEED MORE SPACE IN THE BLANK COMMON ')
1030 FORMAT(/,' N-R ITERATION NO. ',I5,/)
1050 FORMAT(/,' DRT ITERATION NO. ', 15,/)
1070 FORMAT( ' VELOCITY NORM = '.F15.7,/,
Appendix The FEM Code, SPID 355
DO 150 I = 1,3
STS(I.N) = 2./3. * EFSTS * (EPS(I.N)-EM) / AL + DIAT * EM * 3.
150 CONTINUE
STS(4,N) = EFSTS * EPS(4,N) / AL / 3.
STS(5,N) = EFSTS
200 CONTINUE
C
C UPDATE TOTAL EFFECTIVE STRAIN
C
DO 300 N = 1, NUMEL
TEPS(N) = TEPS(N) + EPS(5,N) * DTMAX
300 CONTINUE
RETURN
END
c
SUBROUTINE PRTINP
C
C THIS SUBROUTINE PRINTS THE INPUT DATA
C
IMPLICIT REAL*8 (A-H, 0-Z), INTEGERM (I-N)
CHARACTER TITLE*70
COMMON /TITL/ TITLE
COMMON /TSTP/ NINI,NCUR,NSEND,NITR,DTMAX
COMMON /RVA1/ RZ(2,100), URZ(2,100), FRZ(2,100)
COMMON /RVA2/ EPS(5,100), STS(5.100), TEPS(IOO)
COMMON /INVR/ NOD(4,100), NBCD(2,100)
COMMON /DIES/ FRCFAC
COMMON /RIGD/ RTOL, ALPH, DIAT
COMMON /MSTR/ NUMNP, NUMEL, IPLNAX
COMMON /INOT/ INPT, MSSG, IUNIT, IUNI2
C
C INPUT SUMMARY
C
WRITE(IUNIT,1010) TITLE
WRITE(IUNIT,1020)
WRITE(IUNIT,1030) NINI.NSEND,DTMAX
WRITE(IUNIT,1050) ALPH,DIAT
WRITE(IUNIT,1070) IPLNAX
WRITE(IUNIT,1110) FRCFAC
WRITE(IUNIT,1130) NUMNP
WRITE(IUNIT,1150)
WRITE(IUNIT,1180) (N,(RZ(I,N), 1=1,2), N = 1, NUMNP)
C
C PRINT NODE VELOCITY
C
WRITE(IUNIT,1220)
WRITE(IUNIT,1180) (N,(URZ(I,N), 1=1,2), N = 1, NUMNP)
C
C ELEMENT INFORMATION
C
WRITE(IUNIT,1270) NUMEL
WRITE(IUNIT,1330)
WRITE(IUNIT,1350) (N, (NOD(I,N),1=1,4), N=l,NUMEL)
c
Appendix The FEM Code, SPID 357
C BOUNDARY CONDITION
C
WRITE(IUNIT,1400)
WRITE(IUNIT,1430) (N, (NBCD(I,N),1-1,2), N=1,NUMNP)
C
C WRITE STRAIN DISTRIBUTION AT INPUT STAGE
C
WRITE(IUNIT,1500)
WRITE(IUNIT,1550) (N,TEPS(N), N=1,NUMEL)
RETURN
C
1010 FORHAT(1H1,///,5X,'OUTPUT OF S P I D ',//,
+ 5X.A,///)
1020 FORMAT(5X,'INITIAL INPUT SUMMARY'///)
1030 FORMATC INITIAL STEP NUMBER 15,/,
+ ' FINAL STEP NUMBER 15,/,
+ ' STEP SIZE IN TIME UNIT F10.5)
1050 FORMATC LIMITING STRAIN RATE F15.7,/,
+ ' PENALTY CONSTANT F15.7)
1070 FORMATC DEFORMATION CODE 15,/,
+ ' IF - 1, AXISYMMETRIC ',/
+ ' = 2, PLANE STRAIN ')
1110 FORMATC FRICTION FACTOR = ', F15.7,/)
1130 FORMAT{///' NUMBER OF NODAL POINTS = ', 15,/)
1150 FORMAT(//,' NODE COORDINATES ',//,
+ ' NODE NO X-COORD Y-COORD',/)
1180 FORMAT(5X,I5,5X,2F15.7)
1220 FORMAT(///,' NODE VELOCITY ',//,
+ ' NODE NO X-VELOCITY Y-VELOCITY'/)
1270 FORMAT(//, NUMBER OF ELEMENTS = ', 15,/)
1330 FORMAT(//, ELEMENT CONNECTIVITY ',//,
+ ELE NO. I J K L ',/)
1350 FORMAT(5I7
1400 FORMAT(//, BOUNDARY CONDITION CODE ',//,
+ NODE NO Xl-CODE X2-CODE ',/)
1430 FORMAT(3I7)
1500 FORMAT(///,' STRAIN DISTRIBUTION AT INPUT STAGE ',//,
+ ' NODE NO. STRAIN ',/)
1550 FORMAT(I5,5X,F15.7)
END
c
SUBROUTINE PRTSOL
C
C THIS SUBROUTINE PRINT THE SOLUTION RESULTS
C
IMPLICIT REAL*8 (A-H.O-Z), INTEGERM (I-N)
CHARACTER TITLE*70
COMMON /TITL/ TITLE
COMMON /INOT/ INPT,MSSG,IUNIT,IUNI2
COMMON /TSTP/ NINI,NCUR,NSEND,NITR,DTMAX
COMMON /MSTR/ NUMNP,NUMEL,IPLNAX
COMMON /RVA1/ RZ(Z.IOO), URZ(2,100), FRZ(2,100)
COMMON /RVA2/ EPS(5,100), STS(5,100), TEPS(IOO)
COMMON /INVR/ NOD(4,100), NBCD(2,100)
358 Metal Forming and the Finite-Element Method
C
C PRINT NODE COORDINATES
C
WRITE(IUNIT,1010) TITLE, NCUR
WRITE(IUNIT,1020)
WRITE(IUNIT,1040) (N,(RZ(I,N), 1=1,2), N -1, NUHNP)
C
C PRINT NODE VELOCITY, NODAL FORCE
C
WRITE(IUNIT,1080)
WRITE(IUNIT,1100) (N,(URZ(I,N), 1-1,2),
cc + (FRZ(I,N), 1=1,2), N = 1, NUMNP)
C STRAIN RATE, STRESS, TOTAL EFFECTIVE STRAIN
C
WRITE(IUNIT,1130)
WRITE(IUNIT,1180) (N,(EPS(I,N),1-1,5), N - 1, NUMEL)
WRITE(IUNIT,1230)
WRITE(IUNIT,1180) (N,(STS(I,N),1=1,5), N = 1, NUMEL)
WRITE(IUNIT,1330)
WRITE(IUNIT,1360) (N, TEPS(N), N = 1, NUMEL)
RETURN
C
1010 FORMAT(1H1,///,5X,'OUTPUT OF S P I D',//,5X,A,//,
+ 10X, SOLUTION AT STEP NUMBER - ',I5,///)
1020 FORMAT( /, NODE COORDINATES',//,
+ NODE NO X-COORD Y-COORD',/)
1040 FORMAT(5X,I5,5X,2F15.7)
1080 FORMAT(///, NODAL VELOCITY AND FORCE',//,
+ NODE NO X-VELOCITY Y-VELOCITY',
+ ' X-FORCE Y-FORCE ',//)
1100 FORMAT(3X,I5,3X,4F15.7)
1130 FORMAT(///,' STRAIN RATE COMPONENTS ',//,
+ ELE. NO. Ell E22 E33'
+ E12 EBAR',//)
1180 FORMAT(I5,5F15.7)
1230 FORMAT(//, STRESS COMPONENTS',//,
+ ELE. NO. Sll S22 S33'
+ S12 SBAR',//)
1330 FORMAT(///,' TOTAL EFFECTIVE STRAIN ',//,
+ ' ELE. NO. EFFECTIVE STRAIN '//)
1360 FORMAT(5X,I5,5X,F15.7)
END
c
SUBROUTINE RSTFIL
C
C GENERATE RESTART FILE
C
IMPLICIT REAL*8 (A-H, 0-Z), INTEGERM (I-N)
CHARACTER TITLE*70
COMMON /TITL/ TITLE
COMMON /TSTP/ NINI,NCUR,NSEND,NITR,DTMAX
COMMON /RVA1/ RZ(2,100), URZ(2,100), FRZ(2,100)
COMMON /RVA2/ EPS(5,100), STS(5,100), TEPS(IOO)
Appendix The FEM Code, SPID 359
COMMON /INVR/ NOD(4,100), NBCD(2,100)
COMMON /DIES/ FRCFAC
COMMON /RIGD/ RTOL, ALPH, DIAT
COMMON /MSTR/ NUMNP, NUMEL, IPLNAX
COMMON /INOT/ INPT, MSSG, IUNIT, IUNI2
C
NN - NCUR + 1
OPEN(IUNI2,FILE='SPID.RSr,STATUS*'UNKNOWN',FORM='FORMATTED')
WRITE(IUNI2,1010) TITLE
WRITE(IUNI2,1040) NCUR.NN.DTMAX
WRITE(IUNI2,1060) ALPH.DIAT
HRITE(IUNI2,1080) IPLNAX
WRITE(IUNI2,1060) FRCFAC
WRITE(IUNI2,1080) NUMNP
WRITE(IUNI2,1120) (N,(RZ(I,N),1=1,2),N=l,NUMNP)
WRITE(IUNI2,1080) NUMEL
WRITE(IUNI2,1080) (N,(NOD(I.N),1-1,4), N=l,NUMEL)
WRITE(IUN12,1160) (N,(NBCD(I,N),1-1,2),N=1,NUMNP)
WRITE(IUNI2,1120) (N,(URZ(I,N),I-=1,2),N-1,NUMNP)
WRITE(IUNI2,1200) (N,TEPS(N), N-l,NUMEL)
CLOSE(IUNI2)
RETURN
C
1010 FORMAT(1X,A)
1040 FORMAT(2I10,F20.7)
1060 FORMAT(3F20.10)
1080 FORMAT(517)
1120 FORMAT(I5,2F20.10)
1160 FORMAT(3I7)
1200 FORMAT(I7,F20.10)
END
c
SUBROUTINE STIFF(B,A,NEQ,MBAND,ITYP)
C
C STIFFNESS MATRIX GENERATION
C ITYP - 1, NEWTON-RAPHSON ITERATION
C ITYP - 2, DIRECT ITERATION
C
IMPLICIT REAL*8 (A-H, 0-Z), INTEGER*4 (I-N)
COMMON /INOT/ INPT,MSSG,IUNIT.IUNI2
COMMON /RVA1/ RZ(2,100), URZ(2,100), FRZ(2,100)
COMMON /RVA2/ EPS(5,100), STS(5,100), TEPS(IOO)
COMMON /INVR/ NOD(4,100), NBCD(2,100)
COMMON /DIES/ FRCFAC
COMMON /MSTR/ NUMNP, NUMEL, IPLNAX
DIMENSION A(NEQ,1),B(1)
DIMENSION RZE(2,4), URZE(2,4), NBCDE(2,4),PP(8,8), QQ(8),
+ LM(8)
C
C INITIALIZE LOAD VECTOR, STIFFNESS MATRIX, AND
C NODAL POINT FORCE ARRAY
C
DO 20 N = 1, NEQ
B(N) = 0.
360 Metal Forming and the Finite-Element Method
DO 20 I = 1, MBAND
A(N,I) = 0.
20 CONTINUE
DO 50 N - 1, NUHNP
DO 50 I - 1, 2
50 FRZ(I.N) - 0.
C
DO 200 N - 1, NUMEL
C
C CHANGE RZ, URZ, AND NBCD FROM GLOBAL ARRANGEMENT TO ELEMENTAL
C ARRANGEMENT
C
DO 100 I - 1, 4
12 - 1 * 2
II = 1 2 - 1
NE = NOD(I,N)
RZE(1,I) - RZ(1,NE)
RZE(2,I) - RZ(2,NE)
URZE(l.I) - URZ(l.NE)
URZE(2,I) - URZ(2,NE)
NBCDE(l.I) = NBCD(l.NE)
NBCDE(2,I) - NBCD(2,NE)
LM(I2) - NOD(I,N)*2
LH(I1) - LM(I2) - 1
100 CONTINUE
C
CALL ELSHLF(PP,QQ,RZE,URZE,EPS{1,N),TEPS(N),IPLNAX,ITYP,N)
IF(ITYP .EQ. 1) CALL NFORCE(QQ,FRZ,LM)
IF(FRCFAC.NE. 0.)
+ CALL FRCBDY(RZE,URZE,NBCDE,TEPS(N),EPS(1,N),QQ,PP,IPLNAX)
CALL ADDBAN(B,A,NEQ,LM,QQ,PP)
200 CONTINUE
C
C APPLY DISPLACEMENT BOUNDARY CONDITION
C
CALL DISBDY(URZ,NBCD,B,A,NEQ,MBAND,ITYP)
RETURN
END
c
SUBROUTINE STRMTX(RZ,B,WDXJ,S,T,IPLNAX,NEL)
C
C CHAPTER 6.4
C EQUATIONS (6.25), (6.27) AND (6.35)
C EVALUATE STRAIN RATE MATRIX OF QUADRILATERAL ELEMENT
C
C B(4,8) : STRAIN RATE MATRIX
C RZ(2,4) : NODE COORDINATES
C (S,T) : NATURAL COORDINATE
C
IMPLICIT REAL*8 (A-H, 0-Z), INTEGERM (I-N)
COMMON /INOT/ INPT,MSSG,IUNIT,IUNI2
DIMENSION RZ(2,1), B(4,l)
C
R12 = RZ(1,1) - RZ(1,2)
Appendix The FEM Code, SPID 361
C
R - Q1*RZ(1,1) + Q2*RZ(1,2) + Q3*RZ(1,3) + Q4*RZ(1,4)
B(3,l) - Ql / R
B(3,3) - QZ / R
8(3,5} - Q3 / R
8(3,7) - Q4 / R
WDXJ - WDXJ * R
C
40 CONTINUE
B(4,l) - Yl
B(4,3) •= Y2
8(4,5) = Y3
B(4,7) - Y4
B(4,2) - XI
B(4,4) - X2
B(4,6) = X3
B(4,8) - X4
RETURN
C
1010 FORMAT(/,' SORRY, NEGATIVE JACOBIAN DETECTED AT ELEMENT NO.
+ 15)
1030 FORMAT(' DXJ.S.T = \3F15.7)
END
c
DO 60 I = 1, 4
EPS(I) - EPS(I) + B(I,J) * URZ(J)
60 CONTINUE
EB2 = (EPS(1)**2 + EPS(2)**2 + EPS(3)**2) * D(l) +
+ EPS(4)**2 * 0(4)
EPS(5) - DSQRT(EB2)
C
C EVALUATE VOLUMETRIC CONTRIBUTION Of STIFFNESS MATRIX
C
DO 80 I - 1, 8
IF(IDREC .EQ. 1)
+ QQ(I) = QQ(I) - DIAT * WDXJ * XVOL * XX(I)
TEN - DIAT * WDXJ * XX(I)
DO 80 J = I, 8
PP(I,J) = PP(I,J) + TEM * XX(J)
PP(J,I) = PP(I,J)
80 CONTINUE
RETURN
END
c
SUBROUTINE VSPLSTtQQ.PP.B.URZ.TEPS.WDXJ.IDREC)
C
C CHAPTERS 6.5
C EQUATIONS (6.43), (6.44) AND (6.46)
C FOUR POINTS INTEGRATION OF VOLUME STRAIN RATE
C
C PP = ELEMENTAL STIFFNESS MATRIX
C QQ = ELEMENTAL LOAD VECTOR
C B = STRAIN RATE MATRIX
C
IMPLICIT REAL*8 (A-H, 0-Z), INTEGER*4 (I-N)
COMMON /TSTP/ NINI,NCUR,NSEND,NITR,DTMAX
COMMON /RIGD/ RTOL,ALPH,DIAT
DIMENSION PP(8,8), QQ(8), 8(4,8), URZ(l)
DIMENSION D(6),FDV(8),E(4),XX(8)
DATA 0/3*0.6666666666666700, 3*0.333333333333333DO/
C
C ELIMINATE DIALATATIONAL COMPONENT FROM STRAIN RATE MATRIX
C
DO 20 I = 1, 8
XX(I) = (B(1,I) + B(2,I) + B(3,I)) / 3.
20 CONTINUE
DO 40 I = 1, 8
DO 40 J = 1,3
B(J,I) = B(J,I) - XX(I)
40 CONTINUE
C
C CALCULATE STRAIN RATE
C
DO 60 J = 1, 4
E(J) = 0.
DO 60 I = 1, 8
E(J) = E(J) + B(J,I) * URZ(I)
60 CONTINUE
364 Metal Forming and the Finite-Element Method
EFSR2 - D(1)*E(1)*E(1) + D(2)*E(2)*E(2) + D(3)*E(3)*E(3) +
+ D(4)*E(4)*E(4)
IF(NITR.EQ.l .AND. NCUR.EQ.NINI .AND. IDREC.EQ.2)
+ EFSR2 - (ALPH*100.) ** 2
ALPH2 - ALPH **2
IF(EFSR2 .LT. ALPH2) EFSR2 - ALPH2
EFSR - DSQRT(EFSR2)
CALL FLWSTS(EFSTS, STRAT, TEPS, EFSR)
C
C CALCULATE FIRST DERIVATE OF EFSR **2
C
DO 80 I - 1, 8
FDV(I) - 0.
DO 80 J - 1 , 4
FDV(I) - FDV(I) + D(J)*E(J)*B(J,I)
80 CONTINUE
C
C ADD POINT CONTRIBUTION TO STIFNESS MATRIX
C
Fl - EFSTS / EFSR * WDXJ
IF(IDREC .EQ. 2) GO TO 200
F2 - STRAT / EFSR2 * WDXJ - Fl / EFSR2
DO 120 I - 1 , 8
QQ(I) - QQ(I) - FDV(I) * Fl
DO 110 0 - 1 , 8
TEM - 0.
DO 100 K - 1, 4
TEM - TEM + D(K)*B(K,I)*B(K,J)
100 CONTINUE
PP(I,J) - PP(I,J) + TEM*F1
IF(EFSR2 .LT. ALPH2) GO TO 105
PP(I,J) - PP(I,J) + FDV(I)*FDV(J)*F2
105 PP(J,I) - PP(I,J)
110 CONTINUE
120 CONTINUE
RETURN
C
200 CONTINUE
DO 300 1 - 1 , 8
DO 280 J - 1,8
TEM - 0.
DO 250 K = 1,4
TEM - TEM + D(K)*B(K,I)*B(K,J)
250 CONTINUE
PP(I,J) = PP(I.J) + TEM*F1
PP(J,I) - PP(I,J)
280 CONTINUE
300 CONTINUE
RETURN
END
3. Output File
The output file, SPID.OUT, contains the results of 20 solution steps. The
20th step (40% reduction in height) solution is printed here as an example
of the output file.
OUTPUT OF S P I D
SIMPLE UPSETTING, M = 0.5
SOLUTION AT STEP NUMBER 20
NODE COORDINATES
NODE NO X-COORD Y-COORD
1 0.0000000 0.6000000
2 0.2539333 0.6000000
3 0.5183240 0.6000000
4 0.7739461 0.6000000
5 1.1417572 0.6000000
6 0.0000000 0.3924766
7 0.2869400 0.3802143
8 0.5691376 0.4023207
9 0.8917992 0.4077725
10 1.2311768 0.5210101
11 0.0000000 0.2249196
12 0.3335209 0.2248486
13 0.6637572 0.2373209
14 1.0026870 0.2919173
15 1.2807303 0.3639948
16 0.0000000 0.1097141
17 0.3752439 0.0951748
18 0.7403535 0.1229732
19 1.0537857 0.1484054
20 1.3246827 0.1858872
21 0.0000000 0.0000000
22 0.3950873 0.0000000
23 0.7622312 0.0000000
24 1.0712774 0.0000000
25 1.3397564 0.0000000
Appendix The FEM Code, SPID 367
NODAL VELOCITY AND FORCE
NODE NO X-VELOCITY Y-VELOCITY X-FORCE Y-FORCE
STRESS COMPONENTS
ELE. NO. Sll S22 S33 S12 SBAR
1 0.1662418
2 0.1901851
3 0.3206123
4 0.5611510
5 0.4400065
6 0.4483966
7 0.5676523
8 0.4973681
9 0.7078742
10 0.6981489
11 0.6199824
12 0.4176614
13 0.8754011
14 0.8086128
15 0.6054953
16 0.4199022
~Z
Appendix The FEM Code, SPID 369
FIG. A.2 FEM grid distortions of simple compression predicted by SPID: (a) undeformed;
(b) 40% reduction in height.
/
Figure A.2 shows (a) the initial FEM mesh and (b) the deformed FEM
mesh at 40% reduction in he/ght predicted by SPID.
The computations were performed on an IBM PC/AT with Professional
FORTRAN compiler of Ryarj/McFarland Corp. and also on VAX 750 with
VMS operating system. Tar>le A.I summarizes the CPU time requirements
to run this simulation on/VAX-750 and IBM PC/AT.
basic principles and concepts in, Flat-tool forging, 293, 294, 295
73-74 Flow formulation, 4-5
basis of, 4 Flow rule, 63-66, 245-246
boundary conditions, 86-87, plasticity and, 63-66
117-121 porous metals, 245-246
computer-aided, 93-94 Flow stress, 28-30, 66
construction of model, 3-4 Folding, 153
direct iteration method, 121-122 Forging. See also Axisymmetric
discretization of problem, 90 forging; Closed-die forging
element strain-rate matrix, 101-107 classification of, 151
elemental stiffness equation, disk, 311, 312
108-110 flashless, 170, 171, 172
geometry updating, 123-125 flat-tool, 293, 294, 295
history of, 90 forward extrusion, 13
in-plane deformation processes, hot-die disk, 237, 238, 239
192-194 isothermal, 237, 238, 239. See also
introduction to, 3-4 Axisymmetric isothermal forging
Lagrange multiplier method, 110 open-die, 15
linear matrix solver, 115-117 orbital, 16
linearization, 92 process design objective, 35
metal forming and, 5-6 process design steps, 35, 36
nodal point velocities, 91 radial, 314
notation, 101 titanium alloy Ti6242, 234-237
numerical integrations, 111-114 Forward extrusion forging, 13
penalty function method, 110 Friction, 10, 30-33.
plane-strain problems, 131-133 at tool-workpiece interface,
porous metals, 246-249 163-165, 166
procedures, 90-94 Friction coefficient, 206
rectangular element family, 97-100, Friction hill, 138-139
105-106 Friction shear stress, 32
rezoning, 126-129 Frictional stress, 119-120
rigid region, treatment of, 87
ring compression, 160-163 Gaussian elimination, 116-117
sheet-metal forming of general Gaussian quadrature formula,
shapes, 209-210 111-113
stiffness equations, 92 Gear blank forging, 172
three-dimensional brick element, Generalized stress, 66
100-101, 106-107 Geometry updating, 123-125
three-dimensional problems,
276-278 H cross sections, preforms for, 309
time-increment, 123-125 Hasek method, 50
triangular element family, 95-97, Heading of cylindrical bars, 153-159
104-105, 113 Heat transfer
variational approach, 83-84, 86 in porous metals, 259-262
Flange drawing, 194-195, 197 in thermo-viscoplastic analysis,
Flange-hub shapes, axisymmetric 225-227
forging of, 253-256, 257 Hexahedral element, 276
Flashless forging, 170, 171, 172 Hill's general method, 3, 78-83, 88
Flat-face die, 36, 38 Hobbing, 14
Flat-tool compression, 314 Hooke's law, 63
374 Index
Hot compression, of steel cylinder, Metal forming and finite-element
231-234 method, 5-6
Hot-die disk forging, 237, 238, 239 Metal-forming processes, 8-24. See
Hot extrusion, 20 also specific processes
of rods and shapes, 36-39 backward extrusion forging, 14
Hot forming, 28-29, 34-35 brake-bending, 21
preform design, 315-318 classification of, 11-12
Hot nosing, 237-240 closed-die forging, 9, 12, 13
Hot pressing, under plane-strain closed-die forging, without flash, 12
compression, 262-266 coining operation, 13
Huber-Mises criterion, 59 deep drawing, 23
Hydrodynamic conditions, 31 deformation mechanics, 10. See
also Plasticity
Impression, 35-36, 37 description of, 11-12
Infinitesimal deformation theory, drawing, 20
55-56 extrusion process, 19, 20
Infinitesimal plastic strain, 63-64 forward extrusion forging, 13
In-plane deformation processes, friction, 10
192-195 hobbing, 14
Integration points ,111 ironing, 21
Ironing, 21 material variables, 9-10
Ironing press, 46-47 nosing, 15
Isoparametric elements, 96, 98, 99 open-die forging, 15
Isothermal forging, 237, 238, 239. See orbital forging, 16
also Axisymmetric isothermal product properties, 10
forging radial forging of shaft, 16
ring rolling, 18
Jauman derivative, 324 roll bending, 22
roll forming, 22
rolling, 17
Kirchhoff stress, definition of, 324
rotary tube piercing, 18
rubber-diaphragm hydroforming, 24
Lagrange multiplier method, 110 rubber-pad forming, 24
Lagrangian description, 55-57 shear forming from plate, 19
Lagrangian family, 98 sheet-metal, 11, 12
Lagrangian strain, 208 spinning, 23
Levy-Mises equations, 66 systems approach in, 8-10
Limiting drawing ratio, 205 tooling and equipment, 10
Lubrication, basic types of, 30-31 upsetting with flat-heading tool, 17
Lubricity, 32-33 Metal powders. See Porous metals
Methods of analysis, 73-88
Maximum plastic work principle, equations for, 73
64-65 finite-element method, 83-87, 88.
Maxwell-Heuber-Mises criterion, 59 See also Finite-element method
Mechanics of deformation, 10. See Hill's general method, 3, 78-83, 88
also Plasticity for non-steady-state processes,
Metal flow, 27 175-176
in non-steady-state upset forging, 27 for steady-state processes, 174-176
in rolling, 42 upper-bound method, 3, 74-78, 88
in steady-state extrusion, 27 von Mises criterion, 59
Index 375
Modeling, process, 1-3 stress, 54-58
Multipass bar drawing and extrusion, virtual work-rate principle, 62-63
183-186 viscoplasticity, 70-72
yield criterion, 58-61
Nakajima method, 50 Plastometers, 29
Necking, 47 Plate
Newton-Raphson method, 92, 93, 121 bending of, 141-147, 327-331
Nominal stress, 324 rolling of, 41-45
Nonquadratic yield criterion, 217-220 Porous metals, 244-270
Non-steady-state flow, 27 axisymmetric forging of flange-hub
Nosing, 15 shapes, 253-256, 257
hot, 237-240 axisymmetric forging of pulley
shell, 301-305 blank, 256-259
compaction, 266-270
Open-die forging, 15 discretization, 247-248
Orbital forging, 16 finite-element modeling, 246-249
flow rules, 245-246
Parent element, 97 fully dense materials, 248
Penalty function method, 110 heat transfer in, 259-262
Piola-Kirchhoff stress, 208, 324 hot pressing under plane-strain
Piola-Kirchhoff stress tensor, 57 condition, 262-266
Plane plastic flow, 131. See also numerical procedures, 246-249
Plane-strain problems simple compression, 249-253
Plane-strain compression, hot pressing updating relative density, 248
under, 262-266 volume integration, 248-249
Plane-strain problems, 131-149 yield criterion, 245-246
closed-die forging with flash, Powder compaction, 266-270
133-136 Powder forming, 244. See also Porous
finite-element formulation, 131-133 metals
plate bending, 141-147 Prandtl-Reuss equations, 65-66
sheet rolling, 137-141 Preform design, 298-320
side pressing, 148-149 axially symmetric forging, 309-315
Plane-strain rolling, 305-309 backward tracing, 298-301,
Plastic anisotropy, 190-191 301-305, 305-308, 311-312,
Plastic strain-rate, 63-64 316-317, 318-320. See also
Plasticity, 54-70 Backward tracing
effective strain, 66-68 definition of, 298
effective stress, 66-68 H cross sections, 309
equilibrium equations, 61-62 hot forming, 315-318
extremum principles, 68-70, 71-72 method for design, 298-301
flow rule, 63-66 plane-strain rolling, 305-309
infinitesimal deformation theory, shell nosing, 301-305
55-56 Pressing, hot, 262-266
Lagrangian description, 55—57 Process modeling, 1-3
maximum plastic work principle, Product properties, 10
64-65 Pseudoconcentrations, 334
plastic potential, 63-66 Pulley blank, axisymmetric forging of,
strain, 54-58 256-259
strain-hardening, 66-68 Punch loads/displacements, 144-145,
strain-rate, 54-58 146
376 Index
Punch-stretching, 201-206 Shear stress criterion, 59
Sheet rolling, 137-141
Radial forging, 314 Sheet-metal forming, 47-52, 189-220
of shaft, 16 axisymmetric out-of-plane
Rectangular block compression, deformation, 195-201
278-281 axisymmetric punch-stretching,
Rectangular element family, 97-100 201-206
strain-rate matrix, 105-106 bore expanding, 194-195, 196
Reliability of simulation, 333-334 classification of processes, 11, 12
Rezoning, 126-129 deep-drawing processes, 201-206
Rigid-plastic formulation, 327-334 flange drawing, 194-195, 197
disadvantages of, 330 general shapes, 206-210
solid formulation vs., 327-334 in-plane deformation processes,
Ring compression, 159-165, 331-332, 192-195
333 nonquadratic yield criterion,
axisymmetric isothermal forging, 217-220
159-163 plastic anisotropy, 190-191
friction and, 163-165 square-cup drawing, 210-217
solid formulation, 331-332, 333 Shell nosing, 301-305
Ring rolling, 18 Side pressing, 148-149
Ring test, 33 Simple compression, 249-253,
Rods, 36-39, 45-47 364-369
drawing of, 45-47 of cylinder, 364-369
hot extrusion of, 36-39 of porous metals, 249-253
Roll bending, 22 Simple plastic incremental
Roll forming, 22 deformation. See SPID
Rolling, 17 Simpson's formula, 114
analysis of spread in, 289-292 Simulation of forming processes,
metal flow in, 42 332-334
plane-strain, 305-309 Slab method, 1
plate, 17, 43-44 Slip-line field method, 1-3, 138
ring, 18 Small-strain solid formulation,
of shapes, 44-45 321-323
sheet, 17, 137-141 Smooth entry dies, 38
strip, 42-43, 45 Solid formulation, 4-5, 321-335
Rotary tube piercing, 18 large deformation: incremental
Rubber-diaphragm hydroforming, 24 form, 326-327
Rubber-pad forming, 24 large deformation: rate form,
323-326
Shaft, radial forging of, 16 plate bending, 327-334
Shapes rigid-plastic (flow) solutions vs.,
drawing of, 45-47 327-334
flange-hub, axisymmetric forging of, ring compression, 331-332, 333
253-256, 257 small-strain, 321-323
hot extrusion of, 36-39 three-dimensional analysis with, 334
rolling of, 44-45 SPID (simple plastic incremental
sheet-metal forming of, 206-210 deformation), 338-369
Shear forming from plate, 19 description of major variables,
Shear strength of deforming material, 342-343
32 example solution, 364-369
Index 377