Sei sulla pagina 1di 29

Journal of

ELSEVIER

Journal of Materials Processing Technology 47 (1995} 201-229

Materials Processing Technology

An analysis of the steady-state wire drawing of strain-hardening materials


U.S. Dixit, P.M. Dixit*
Department o/" Mechanical En~ineerinyL Italian Institute q[" Teclmologo', Kanpur 208 016. India

(Received May 24, 1993)

Industrial Summary
A comprehensive investigation of the steady-state wire drawing process has been done to study the effects of various process variables on important drawing parameters and deformation, the process variables considered being the reduction ratio, the die semi-angle, the coefficient of friction and the back tension, whilst the drawing parameters studied are the diepressure, the drawing stress and the separation force. The deformation is represented by contours of equivalent strain and of equivalent strain-rate. The quantitative effects of strain hardening on the drawing parameters and qualitative effects on the deformation are studied also. A comparison of the drawing parameters is made for three materials (copper, aluminium and steel).

Notation
A
f

F2

tr, t:

[K]
K l li lo
I1 I1

area of the d o m a i n average coefficient of friction global right-hand side vector drawing force unit vectors along the r and z directions global coefficient matrix factor in the strain-hardening relationship taper length of the die length of the inlet zone length of the exit zone multiplication factor for the calculation of inlet and exit zone length exponent in the strain-hardening relationship unit normal vector

*Corresponding author. 0924-0136/95/$09.50 ~O 1995 Elsevier Science S.A. All rights reserved SSDI 0 9 2 4 - 0 1 3 6 ( 9 5 ) 0 1 3 2 0 - Z

202 U.S. Dixit, P.M. Dixit / Journal o f Materials Processing Technology 47 (1995) 201-229

P P P Pf Pb r r% r, O, z

R1
RE
S

S
Sij , S

g
t t

ti

tr, tz tn ts U1 U2
Vr~ Yz
Fn Fs

Wp, Wr, C(

6~j

{A}
V

r.F
?
8

/:'ij, g
~;kk

i}j, ~' P
(Tij~ ~7o ~Tf
O'xb O'xf

hydrostatic pressure power drawing force in Eq. (49) power due to plastic work friction power power due to back tension reduction ratio (fractional) percentage reduction cylindrical coordinates initial radius of the wire final radius of the wire coordinate along the taper length of the die separation force deviatoric part of the stress tensor equivalent deviatoric stress time traction vector components of the traction vector components of the traction vector along the r & z directions die- pressure interfacial shear stress inlet velocity of the wire drawing speed velocity vector components of the velocity vector along the r & z directions velocity normal to interface velocity tangential to interface W z weight functions die semi-angle Kronecker's delta global vector of primary variables vector differential operator, del parts of the domain boundary equivalent (plastic) strain equivalent strain-rate strain-rate tensor trace of the strain-rate tensor deviatoric strain-rate tensor density stress tensor yield stress of material at zero plastic strain flow stress of material (Fig. 3) back tension average drawing stress

U.S. Dixit. P.M. Dixit / Journal of Materials Processblg Technology 47 (1995) 201 229 203

Gy
It Ho

()

d()

variable yield stress of material proportionality constant coefficient of friction in Eq. (49) constant used for non-dimensionalization of/~ non-dimensional quantity differential value of( )

1. Introduction

Since the appearance of the first paper on metal flow in wire drawing in 1886, by Smith [1], many investigators have studied the process. Wistreich [2] and Majors [-3] have obtained the die-pressure experimentally using the split-die technique and by measuring the hoop-strain respectively. Avitzur [-4] has described an experimental approach for determining the drawing force as a function of reduction, cone angle and friction. Cook and Wistreich [5] have described a method for measuring the diepressure by the photo-elastic method. The visioplasticity method introduced by Thomsen et al. [6] combines the results of experiment and analysis. Hoffman and Sachs [-7] proposed the slab method for the wire drawing process. Siebel [8] introduced a theory of wire drawing in which he assumed that the effects of homogeneous deformation, friction, and non-useful deformation were additive, giving an equation for drawing force. Avitzur [9] has proposed an extra term to account for redundant power in the drawing stress expression of Sachs [7]. Shield [10] has shown that when the von Mises criterion is used for an axisymmettic problem, the general equation is of elliptic type, the slip lines becoming real only under a special condition that reduces the problem to that of a plane-strain type. Avitzur [11,12] has applied the upper-bound theorem to the problem of wire drawing, dividing the wire into three zones, in each of which the velocity field was assumed to be continuous: at the interface, however, the tangential component of velocity was discontinuous. Avitzur [13] has applied this technique to a strain-hardening material also, where he considered a linear strain-hardening coefficient, using the upper-bound method, he also analyzed the central-burst defect in an extruded or drawn product

[14].
The early applications of the finite-element method (FEM) to metal forming were based on the plastic stress-strain matrix developed from the Prandtl Reuss equations. Iwata et al. [15] made an elastoplastic analysis of hydrostatic extrusion using FEM, the analysis being performed for the non-steady state in both plane-strain and axisymmetric extrusion. Lee et al. [16] also applied the FEM to the non-steady extrusion process, calculating the residual stresses for plane-strain and axisymmetric problems after the external loads were removed. From the results of non-steady state extrusion, results for the steady state were speculated. Lee et al. [17] undertook stress and deformation analysis of steady-state plane-strain extrusion with friction-less curved dies, using the elastic-plastic finite-element method, whilst Shah and Kobayashi [ 18] analyzed axisymmetric extrusion through friction-less conical dies by the rigid-plastic finite-element method, the technique involving the construction of

204 u.s. Dixit, P.M. Dixit / Journal ~?fMaterials Processing Technology 47 (1995) 201 229 flow lines from velocities and the integration of strain-rates numerically along flow lines to determine the strain distribution. Zienkiewicz et al. [19] have presented the flow-formulation approach in forming and extrusion, investigating two techniques, viz. the pressure-velocity formulation with Lagrangian constraints and the penalty-function approach. Tayal and Natarajan [20], Gunasekara et al. [21], Balaji et al. [22] carried out deformation analysis of the extrusion process by FEM. Bianchi and Sheppard [23] compared the viscoplastic finite-element method with, slip-fine field and upper-bound solutions, showing that the FEM gives better results. Some analysis of the extrusion process has been done by modifying standard packages [24,25]. Wire-drawing bibliography is not so rich as that of extrusion. Chen et al. [26] obtained the steady-state deformation characteristics in extrusion and drawing as functions of material properties, die work interface friction, die-angle and reduction. They observed that although it appears that the differences between extrusion and wire drawing are merely in geometrical quantities and hydrostatic stress components (extrusion being essentially a process of compression, whilst drawing is a process of tension), the finite-element results obtained in extrusion cannot be extrapolated to obtain results in drawing by taking into account the geometrical conditions and the concepts of pushing in extrusion and pulling in drawing. Chevalier [27] studied the influence of geometrical parameters and the friction condition on the quality of the final wire using finite-element simulation, an elasto-plastic model being used in the analysis. From the literature survey, it is evident that not much work has been done in the area of application of FEM to the wire drawing process considering the strainhardening effect.
1.1. Modelling o[" the drawing process

In the present study, only the steady-state part of the process is considered, hence an Eulerian formulation is used. The process is considered to be axisymmetric, a conical die shape only being considered. The material is assumed to be rigid plastic strain hardening and yielding according to the von Mises criterion. The elastic effects at the entry and exit are neglected, as these are small. The effects of temperature and strain rate (viscoplasticity effects) on the yield strength of the material are ignored in this work, the inclusion of these effects rendering the analysis quite complex, whilst the temperature rise in the presence of lubricants and at low speeds is quite low. At high speeds, whatever increase the strain rates may produce in the yield strength, most of this increase is compensated for by a decrease in the yield strength due to the temperature rise. The effect of die deformation on various design parameters and product quality has not been reported so far. It is believed that the expressions for the die deformation, which will be consistent with the present finite-element formulation, can be obtained only in an iterative way, i.e., the interracial pressure found in the first iteration by assuming the die to be rigid can be used to find the geometry of the deformed die by the finite-element method, this geometry of the deformed die then being used in the second iteration, the iterations being continued until the change in the die

U.S. Dixit, P.M. Dixit / Journal 01 Materials Processing Technology 47 (1995) 201-229 205

deformation between two successive iterations becomes negligible. After the deformation, a conical die surface will deform to some other surface. Since conical dies only are considered, it was decided not to include this effect in the present study. The plastic behaviour of the materials is represented by a relationship between the deviatoric part of the stress tensor and the rate of deformation (i.e. the symmetrical part of the velocity gradient) tensor, also called as strain-rate tensor. Since the constitutive relationship used is applicable only to the plastic deformation zone, the domain should consist of the portion of wire bounded by the die and the plastic boundaries. However, the plastic boundaries are not known a priori: initially therefore, the domain also includes a reasonable length of the wire in the inlet and exit regions, the plastic boundaries being determined later using a suitable criterion on the strain-rate invariant. At the die wire interface, the friction is modelled by Coulomb's law, subject to the constraint that the local shear stress cannot exceed the yield shear stress. Although it is true that the coefficient of friction varies along the line of contact because of its dependence on factors such as the interracial relative velocity, the die-pressure, the yield strength, etc., it was decided to use a constant coefficient of friction only, the main difficulty in using a variable coefficient of friction being the lack of a wellestablished relationship specifying this variation. Continuity and momentum equations of the metal flow in the domain are converted into non-linear algebraic equations using the Galerkin finite-element technique. A mixed pressure-velocity formulation is used, the resulting equations being solved by iteration using the Householder [28] method, to find the nodal velocities and pressure. To update the value of yield stress in each iteration, flow lines are constructed from the velocity field and the integration of strain-rates along the flow lines is carried out using Simpson's rule to determine the strain distribution. From the nodal velocities and strain distribution along the flow lines, the strain rates and stresses are found first at the Gauss points and then extrapolated to find the secondary quantities such as the die-pressure, the drawing and separation forces, the contours of strains and strain-rates and the plastic deformation zone. The validity of the model is tested by comparing the results for the drawing and separation forces with available experimental results.

2. Problem formulation
2.1. Material behavior

In flow formulation, the measure of deformation is the rate of strain tensor, which may be expressed mathematically as g = ~(Vv + (Vv)T). (1)

In order to express the stress as a function of strain rate in a convenient form, the stress and strain-rate tensors are divided into two parts:
tTij = -- p 6 U + Sij ,

(2)

206 U.S. Dixit. P.M. Dixit / Journal of Materials Processing Technology 47 (1995) 201-229 where p = --~ t r a is the hydrostatic part, S is the deviatoric part and 6 u is Kronecker's delta. Similarly,
eij = ~kkfi~ + gu,

(3)

where g'is the deviatoric part of the strain-rate tensor. In plastic deformation, since there is no change in volume, the hydrostatic part of stress is not related to the deformation. Another consequence of volume constancy is that the hydrostatic part of the strain rate is zero and its deviatoric part is the strain-rate tensor itself. For a rigid-plastic material, the deviatoric parts of the stress and strain-rate tensors are related by
Sij = 2#~ u,

(4)

where 2/t is the proportionality constant. Further, making the definition

(5)

and defining the second invariant of strain rate (also known as the equivalent strain rate) as
~

~, = ~ ,
2

(6)

then it follows that

g = 3/~'. For metals which yield according to von Mises criterion,


g = %.

(7)

(8)

Then,/~ for such metals is /~ = O'y/(3~), (9)

where Cry is the yield strength of the metal. In general, a s is a function of strain, temperature and strain rate. However, as stated in Section 1, the temperature and visco-plastic effects can be neglected, and hence it is a function of strain only. For a strain-hardening metal, if it is assumed that there is isotropic hardening and that the components of strain-increments bear a constant ratio to one another, then the yield stress can be expressed in terms of the equivalent plastic strain. The most common approximation for metals relating equivalent plastic strain to yield stress is given by O'y
=

Cro(1 + K g ) " ,

(10)

where ao is the yield stress of the metal at zero plastic strain and K and n are metal-dependent coefficients determined from experiment. The equivalent plastic strain is obtained by the time integration of equivalent strain-rate as given below:
t

(11)
0

u.s. Dixit, P.M. Dixit / Journal of Materials Processing Technology 47 (1995) 201-229 207
Since the material is assumed to be rigid-plastic, the elastic part of the strain is identically zero. Therefore, henceforth the adjective plastic is dropped when referring to the equivalent strain.

2.2. Governing equations and boundary conditions


When the deformation is axi-symmetric, it is convenient to use cylindrical polar co-ordinates. In polar co-ordinates, the governing equations for the steady, constant-volume flow can be written as ~ + ~00 + ~= = 0,

(12) (13)

{ (?v= c= p .\v r =vr + ,,= ~v.'~~ z ]/

{10(ro',,)~ 4- ~ \ r.
CZ f

=0.

(14)

The boundary conditions for the domain of the present problem (Fig. 1) are as follows: (1) Entry and exit boundaries tAB & EF): The control volume is so selected that its entry and exit boundaries are sufficiently distant from both sides of the deformation zone. Since the drawing velocity U2 is specified, the z-component of the velocity at every point of EF is known the r-component of this velocity obviously being zero at both EF and AB. The z-component of the velocity at AB (U~) can then be found from the continuity equation (v~-A)AB = (v:A)Ev. However, AEv/AAn = (1 -- r), where r is the reduction ratio. Hence,
(Vz)AB
=

(15)

(1 -

r)(v=)Ev.

(16)

/> r~ [ vr=O, t z =0

/5 vo_Vr = O , t z : 0

Vz =UI RI Vr = 0

E]v.1 2
Vr = 0 IR 2
D

]-i

_1 Vr = O , t z = 0 -t

"L o

Z~I z
(Axis of symmetry}

/t-

Fig. 1. T h e d o m a i n a n d the b o u n d a r y c o n d i t i o n s .

208 U.S. Dixit, P.M. Dixit / Journal of Materials Processing Technology 47 (1995) 201 229 Thus the boundary conditions on these boundaries are v~=(1-r)U2, v~= U2, vr=0 vr=0 on AB, on EF. (17)

(2) The top f r e e surfaces (BC & D E ) and the axis o f s y m m e t r y (AF): The following boundary condition exists on these surfaces:
tz

0,

vr = 0.

(18)

(3) The Die wire interface (CD): The component of velocity in the direction normal to the die wire interface at any point on the interface is zero, i.e. v,=0 or vr+v~tan~=0, (19)

where ~ is the die semi-angle. The second boundary condition is given by Coulomb's friction relationship

Itsl =.fit. I

(20)

wherefis the coefficient of friction, ts is the tangential component of the stress vector and t. is the normal component of the stress vector. Further, the shear stress on the interface is subject to the constraint

Itsl ~< O'y/N/3

(21)

The finite-element formulation presented here does not require any pressure boundary conditions to be satisfied. However, there may be a spurious pressure distribution in the solution, i.e. the pressure values may be determined only up to an additive constant. This latter constant can be determined from the condition that, at the inlet boundary, the pressure values should be equal to one third of the back tension (zero, if there is no back tension). Since there may be slight difference in pressure values at different nodes, the average of these values is taken.
2.3. Non-dimensionalization

The non-dimensionalization of various physical quantities is undertaken using the following relationships:
= r/R2, = p/(ao/3); Z = z/R2; ~r = Ur/U2,
Uz ~- u z / U 2 ;

(22)
(23)

/~

where
(7 o

I~o - 3(Uz/R2)"

(24)

U.S. Dixit, P.M. Dixit / Journal o/ Materials Processing Technology 47 (1995) 201 229

209

The non-dimensionalization of continuity and momentum equations are obtained by substituting Eqs. (22) (24) into Eqs. (12)-(14):

~rr + L00 + ~= = 0,
t'o \c,~+~':~)a~ + 8e 0 =0,

(25) (26)

I'o

~Vr~+V:~)--\

.Z

=0.

(27)

Since Reynolds number (pU2R2/I~o) is found to be typically of the order of 10 v for all of the cases considered, inertial terms are ignored.

2.4. Mixed.finite-element formulation


Many workers have used the Galerkin weighted-residual method coupled with direct penalization for finite-element formulation of the governing equations of metal forming. However, in the penalty formulation, pressure values converge only for a range of penalty number and there is no systematic method for determining this range. For this reason, the mixed pressure-velocity formulation has been used here. To avoid the problem of ill-conditioning, the Householder method for solving the system of simultaneous equation has been used. This method does not require pivoting, and is a very effective method for the solution of an ill-conditioned set of equations [28].

2.4.1. Galerkin (weak)Jormulation


Let g:, /:r, /5 be the functions that satisfy all the essential boundary conditions exactly. Then v:, vr, p constitute a weak solution if the following integral equation is satisfied:

(/,',r+5,00+L=)wp+\ i ( i
A
7-

+ ?5

.: Wr (28)

+ \ ~

-c5 /

. 2rtfd~d,q = O, "j

where wp, Wr and w= are the weight functions that satisfy the homogeneous versions of the boundary conditions and A represents the area of the domain. Integrating the second and third parts of Eq. (28)

fl,2~fdFdY.+ f122r~FdFdS-f I32=FdFdS-f142~d?de=O,


A A F= F~

(29)

210 U.S. Dixit, P.M. Dixit / Journal t~f Materials Processing Technology 47 (1995) 201 229 where
I, = - w~[-~-rr + ~00 + ~..-=3, /2 = -- p E ~ . ( w ) + ~oo(W) + ~.=(w)] (30)

+ 2fi[/;.,,~,,(w) + ~oo~oo(W) + k=~==(w) + 2k,=~,=(w)], 13 = t-=w=, I4 = rrw,,

(31) (32)

and Fr and F= are respectively those parts of the boundary where the traction components t-r and ~ are specified. The terms {,~(w), ~oo(W), -~=(w) and L.,(w) are the components of the tensor ~(w) = (Vw + (Vw)T), where
w

(33)

= w , i , + w=i=.

(34)

2.4.2. Finite-element equations The finite-element formulation of the Galerkin integral (29) and the non-dimensional version of the boundary conditions (17)-(20) are similar to that of a flow problem of a non-Newtonian incompressible fluid, the details of which are described in any standard text, viz [29]. In the present work, 9-noded rectangular elements are used to descretise the domain (Fig. 2), with bi-quadratic approximation for the velocity components and bi-linear approximation for the pressure. Since the term ~oo~.oo(W)rof the integrand of Eq. (29) contains a 1/f term, 3 x 10 Gauss points are used in the Gauss-Legendre integration scheme for the evaluation of the elemental coefficient matrix. The assembly of the elemental coefficient matrices and the right-hand side vectors into the global matrix and vector is done by transferring the elements corresponding to a local degree of freedom in each elemental matrix/vector to positions of the corresponding global degrees of freedom in the global matrix/vector. The essential boundary conditions, except for those on the die-wire interface, are applied in the

55 Elements 2 5 3 Velocity nodes 7 2 Pressure nodes Nodes for velocity 8L pressure Nodes only for velocity

~, "U////~////~//~
e/

Fig. 2. The finite-elementmesh and a typical element ~e'.

U.S. Dixit, P.M. Dixit / Journal of Materials Processing Technology 47 (1995) 201 229

211

usual way. The conditions on the interface (Eqs. (19)-(20)) are applied by performing certain row operations on the global coefficient matrix and the global right-hand side vector, the details of which are given in the thesis of Dixit [30]. Letting the global equation after application of all the boundary conditions be [K]{A} = {F}, (35)

where [K] is the global coefficient matrix, {F} is the global right-hand side vector and {A} is the global vector of primary unknowns (i.e. velocity components and pressure).

2.4.3. Formulation for strain hardening The equivalent strain at a point is obtained by the time integration of the equivalent strain rate along the particle path, as given by Eq. (11). The first step in the determination of the equivalent strain field is the construction of particle paths, or 'flow lines' as they are called. The slope of a flow line is given by
dr
-

~
(36)

dz

Vz '

Given a point (ii, 6) on a flow line, the coordinates (ii+ 1, ?i+1) of the adjacent point are found by using the relationship
r i + l ~ -I- (Zi+l -- zi) - ~ ,
Vz

(37)

where (iz+ 1 - f~) is chosen sufficiently small so that the path between the two points may be approximated by a straight line. In this manner, flow lines of various points on the inlet boundary can be found. Now, along a flow line, dt ds
Us

dz
Uz

dr
Dr

(38)

Substitution of this expression for dt in Eq. (11) gives the following expression for the equivalent strain:

,)t~ z

(39t

The equivalent strain is zero at the first plastic boundary. However, this boundary is not known a priori, therefore, the equivalent strain is taken zero on the inlet boundary. This is acceptable, since the strains it give rises to at the first plastic boundary are quite small. Then integrating expression (39) by Simpson's scheme, equivalent strains are found at various points along the flow lines. Nine points in each element are selected and the equivalent-strain values at these points are interpolated to obtain the equivalent strains at the Gauss points. These values are then substituted into Eq. (10) to update the values of cry for further iteration.

212 U.S. Dixit, P.M. Dixit / Journal q[" Materials" Processing Technology 47 (1995) 201 229

2.4.4. Implementation
The lengths li and lo of the inlet and exit zones are usually taken as
Ii =

nR>

Io = nR2.

(40)

The value of the multiplying factor 'n' has to be selected such that uniform conditions prevail at the beginning of the inlet zone and the end of the exit zone. After the conducting of several numerical experiments, a value of three was selected for this analysis. The strategy followed for obtaining the converged solution with and without strain hardening is as follows. First a converged solution (nodal velocities and pressures) without strain hardening is obtained. For this part of the analysis, an initial guess for is taken as 10. This solution is then used as the initial solution for the analysis with strain-hardening and the yield strength is updated by evaluating the strains from Eq. (39) using the immediately preceding solution in each successive iteration. A 55-element mesh is selected (Fig. 2). Because of discontinuity at the corner points, smaller size elements are chosen near to the corner points. Similarly, elements adjacent to the center line are also of smaller size. The high value of ~ in the inlet and exit zones renders the coefficient matrix ill-conditioned. To counter this problem, ~i is restricted to a threshold maximum, which is 100 in the present analysis. The convergence criteria used for primary variables is as follows. All values of velocities of less than 10 4 and all values of pressures of less than 10-3 are considered to be zero and not checked for convergence. A percentage difference of less than 10-2 for velocities and of 0.1 for pressure, between the solutions of two successive iterations, is considered to mark the convergence of the solution.

2.5. Evaluation of the secondary quantities


Once the solution of the problem is obtained in the form of nodal velocities and pressures, the secondary quantities, viz. the drawing force, the die-pressure, the separation force, the deformation zone, etc., are calculated as explained below. (i) Drawin9 force (F2) The drawing force can be calculated by integration of the drawing stress along the exit plastic boundary. However, since it is difficult to determine the plastic boundary accurately, the drawing force is calculated from the relationship
F2 = P / U 2 ,

(41)

where the total power P consists of the following three parts. (a) Internal power dissipation (Pp). The power dissipated due to plastic deformation is given by

Pp =

2r~f
A

Sij

~'ijr dr dz.

(42)

u.s. Dixit. P.M. Dixit / Journal ofi Materials Processing Technology 47 (1995) 201 229 213

Substitution of Eqs. (4), (6) and (9) into this equation leads to

Pp = 2r~ f a y ~'r dr dz.

(43)

(b) Friction power dissipation at the die wire interface (Pf). The power dissipated due to friction is given by
l

Pr = 2~ | tsvsrds.
Id

(44)

where vs is the resultant velocity along the interface and I is the taper length of the die. Since ts is a very small quantity, it is calculated from the relationship Itsl = f It, i. The method of calculating t, is explained later. (c) Power required to overcome back tension (if any) (Pb). The power required to overcome the back tension O'xbis given by
R1 P b = 27[ |
i]

O'xb U 1 r

dr,

(45)

where U1 is the velocity at the inlet. The average drawing stress (~xf) is found by dividing the drawing force by the exit cross-sectional area of the wire. (ii) Die pressure (t,) While calculating t,, first the stresses ~i~ are evaluated at 2 x 2 Gauss points and then these are extrapolated to various points on the die wire interface, after which t, is calculated from the expression t, = t.fi, where ti = aiinj. (47) (46)

and fi is the unit outward normal to the interface. (iii) Separation force (S) The separation force is the resultant of the vertical component of the stress vector on the upper half of the die. Since the stress vector is independent of 0, the integration along 0 direction leads to the product of the vertical component and the projected length 2r. Thus the separation force is given by
l

S = | 2(t, cos ~ - ts sin ~) r ds.


,d

(48)

214 u.s. Dixit, P.M. Dixit / Journal of Materials Processing Technology 47 (1995) 201 229 (iv) Contours The contours of ~ and ~"are plotted to exhibit the deformation pattern, the contours of~ being plotted using the data at various points of the flow lines, whilst for ~', plotting is done using the values evaluated at 2 x 2 Gauss points of the elements from the converged solution. Since the values of ~ in the inlet and exit regions are quite small, a sufficiently small value of ~ can be used as a cut-off to demarcate the plastic zone from the rest of the domain, the cut-off used here being 3% of the maximum value of ~, over the domain.

3. Results and discussion The finite-element modelling of the physical problem of wire drawing developed in the previous section has been applied to a number of cases involving three metals and various non-dimensional process variables, the metals considered being aluminium, steel and copper, having material properties as given in Table 1.
3.1. Validation

The results of the finite-element analysis have been compared with Wistreich's results [2] for copper (Fig. 3), good agreement being observed in all the cases. Wistreich used the following equation to relate the separation force S and the drawing force P: IL=tan -~ , (49)

where/~ is the coefficient of friction. Table 2 compares the separation force obtained by the FEM analysis with the separation force obtained from Wistreich's equation where the drawing force P has been obtained from the FEM analysis. It is observed that there is good agreement between the two values, except for combinations of low cone-angle and low reduction and of high cone-angle and high reduction. Fig. 4 compares the results of the FEM analysis with the experimental results of Chen et al. [26] for steel, there being fairly good agreement between the two. It is to be

Table 1 Material properties Material Aluminium Steel Copper Density (kg/m3) 2800 7870 8600 a0 (MPa) 50.30 584.68 70.30 n 0.26 0.06 0.49 I/K 0.050 0.262 x 10 4 0.022 Reference [31] [26] I-31]

U.S. Dixit, P.M. Dixit / Journal of Materials Processing Technology 47 (1995) 201-229 215
lO Metal : Copper o o -=40% o
o

b~
t:~ O80

Experimenf01 [2] FEM 0n01ysJs (f = 0.05) o ._2.. o

2
0.60

0.40 10%"o" ~

u -~ 0.20 ~c

0.%

2'.0

4!0

6!0

8.0

,0'.0

,20

,4.0

,6.0

Die semiengle, a(deg) Fig. 3. Comparison of the relative drawing stress obtained from FEM analysis and from Wistreich's

experiments. Table 2 Comparison of the separation force obtained from FEM analysis and from Wistreich's equation [2] (~ = 0.03) Reduction (%) Die semi-angle (deg)
S/{R 2 20-0)

S/(R22ao)
(Wistreich)

(FEM)

Error(w.r.t. S by Wistreich) (%) 27.7 15.5 11.4 24.7 13.0 7.4 12.2 18.8

10 10 10 20 20 20 40 40 40

3 6 14 3 6 14 4 8 14

4.636 3.449 2.840 11.180 6.405 4.446 26.230 19.130 4.090

6.41 4,08 2,55 14.84 7.36 4.80 29.88 16,10 5.30

22.8

o b s e r v e d that the present F E M analysis gives a linear relationship between the d r a w i n g stress a n d R~/R2, as p r e d i c t e d also by the F E M analysis of Ref. [26]. H a v i n g c o m p a r e d the results of the present F E M analysis with e x p e r i m e n t a l l y available results, it was ascertained that the present m o d e l w o u l d be able to predict the m a g n i t u d e s of various d r a w i n g p a r a m e t e r s with r e a s o n a b l e engineering a c c u r a c y over m o s t of the range of process variables.

3.2. Parametric stud), 3.2.1. Effect of reduction ratio and die semi-angle
Fig. 5 shows the v a r i a t i o n of d r a w i n g stress with die semi-angle at different reductions. W h e n the die-angle is very small, the length of c o n t a c t between the wire

216

U.S. Dixit, P.M. Dixit / Journal o[ Materials Processing Technology 47 (1995) 201 229
bo

Metal : steel
o o 1.2 Unlubricated]~ . , r__l /

f = 0.10

+ +
- -

Lubricated
FEM

J" r_xp . . . . . .

to, I~,b]

Analysis

1.0 ~" = 8 * 0.8 "X3 0.6 0 04

=0

g
Z OC

0.2

I
I.I (R~ / R 2 ) 1.2 13

Fig. 4.

Comparison of drawing stress obtained from FEM analysis and from the experimentsof Chen et al.

16

Metal : steel
f = 003

o3

L2

r=45

0.8 1:3 c <3 o9 E 73 5 z 0.0


Die semi-angle,

04

~-"E-%

(deg)

Fig. 5.

Variation of drawing stress with die-angle at different reductions.

and die is high, which increases the net friction force, resulting in high drawing stress. When the die-angle is too large, then again the drawing stress is high because of the large distortion. The die-angle at which the drawing stress takes a minimum value is called the optimum die-angle in the literature. It is observed that for a given friction coefficient, the optimum die-angle increases with increasing reduction. Fig. 6 shows the effect of reduction on die-pressure distribution, for a fixed coefficient of friction and die-angle. As the percentage reduction decreases, the die-pressure increases. It is seen also that, in general, the die-pressure is not uniform along the taper length of die, although at low reduction it is generally uniform over the middle region. At higher reduction it has two peaks, one near to the entry and the other at the exit. Since the shear-stress in the present model is given by the die-pressure multiplied by the coefficient of friction, the shear-stress distribution pattern will be similar.

U.S. Dixit, P.M. Dixit ,/Journal o[ Materials Processing Technology 47 (1995) 201-229 217

Therefore, the assumption of constant shear-stress does not seem to agree with the real phenomenon, especially at high reduction. Fig. 7 shows the distribution of die-pressure along the taper length for different die-angles, keeping the reduction and the friction coefficient constant. At small die-angles, the change in the die-pressure due to die-angle is marginal, but at greater die-angles, the die-pressure increase is quite high. Further, the die-pressure/shearstress distribution pattern also changes with the die-angle. Fig. 8 shows the variation of separation force with die-angle at different reductions. For each reduction, the separation force decreases uniformly with increasing die-angle because of two effects, namely decrease of the contact length over which the diepressure acts and decrease in the resolved component of the normal stress. It is observed that the separation force increases with increasing reduction despite a

380
bo

330 2.80

o_ 250 c 1.80 o 1.50 c


o8O;o

o~o

o.Io

o.6o~':4~'o~o '
%

1.00

Fraction of taper length,

s/l

Fig. 6. Distribution of die-pressure at different reductions.

28

"~

~o

~ 23

Mel01 : st881

r=25%'f=O03

o '~[o
~ 08

~"

Z 0

01.2

I 0.4

0'~6

0.8

1.0

Fraction of taper length,s/l


Fig. 7. Distribution of die-pressure at different die-angles.

218 U.S. Dixit, P.M. Dixit / Journal q[ Materials Processing Technology 47 (1995) 201-229

decrease in the die-pressure at a fixed die-angle: this happens because of the increase in the contact length.

3.2.2. Effect o[ the coefficient of friction


Fig. 9 shows the variation of drawing stress with die-angle for different coefficients of friction. It is seen that the optimum die-angle increases with increasing coefficient of friction. Further, the variation of drawing stress with coefficient of friction is more significant at smaller die-angles in comparison to that at greater die-angles. This suggests that if the coefficient of friction is not known accurately in a process, the die should be designed according to maximum expected coefficient of friction. As an example, consider a process in which flies in the range of 0-0.1 and the die is to be

~3o
O3
Metol : steel

d
o 24

f = 0.03

g
o
m

12

r= 45 %

g
6 E
13

(3

5%

o
Z

0 Die semi-ongle, a (deg)

Fig. 8. Variation of separation force with die-angle at different reductions.

Ic~

18
~ Metol : steel

1.6 0.10 r = 25%

1.2

o L 13 o

0.8 0.6
f=O

04. 0.2

g
Z

0.0

~.o

~'.o

6'.o

~'.o ,o.o'

12.o' ,~.o

~o'.o 18.o

Die semi-ongle, a (deg)

Fig. 9. Variation of drawing stress with die-angle at different coefficients of friction.

U.S. Dixit, P.M. Dixit / Journal of Materials Processing Technology 47 (1995) 201-229 219 designed by choosing the optimum value of ct according to Fig. 9. Iffis chosen as 0.1, then c~is 8 . On the other hand, iffis assumed to be zero, then ~ is 2 . Now suppossing that the actual value of f is 0.03, then the corresponding value of the optimum (non-dimensional) drawing stress (axf/ao) is 0.65. Corresponding to the actual f, in the first design (~ = 8 ) axf/ao is 0.675 whilst in the second design (~ = 2 ) O'xf/O" o = 0.825. Thus, the deviation from the optimum value of axf/ao corresponding to the actual value of f is much less (3.9%) in the first design than in the second (26.9%). It is observed that friction has very little influence on the die-pressure distribution pattern for the range of process variables selected here. Consequently, the separation force also does not vary greatly with change in the friction coefficient. 3.2.3. Effect of back tension As expected, the drawing stress increases with increasing back tension. The pattern of die-pressure distribution does not change greatly with back tension but its value decreases with increasing back tension, on account of which the separation force also decreases. 3.3. Effect of strain hardening
Figs. 10 and 11 show the distribution of die-pressure without and with strainhardening, for the two sets of process variables, these sets being selected so as to give extreme values to the drawing parameters over the assumed range of the process variables. It is observed that the pattern of die-pressure distribution does not change greatly with strain-hardening. However, its value is greater with strain-hardening than without strain-hardening, along the whole taper length of the die. Table 3 shows that whilst the percentage error in the estimation of maximum die-pressure is almost same for the two sets, the error in the estimation of separation force and drawing force is greater for set 2, this being so because the strain levels are greater for set 2.

3.CI

-----

Without strain hardening With strain hardening

Metal: steel

bo
c 25,~

- -

2C e~ -5 -5 E o c E q~

j" ./ IC

E o Z

0.~

0.0

01.2

01.4 Fraction

I 0.6 length,

~
s/I

I 0.8

1.0

of t a p e r

Fig. 10. Distribution of die-pressure with and without strain-hardening(r = 5%, ~ = 2:', / = 0.03).

220 U.S. Dixit, P.M. Dixit / Journal of Materials Processing Technology 47 (1995) 201-229
2.0
Metal bo 1.5 : steel

a.
nO

1.0

\Without strain hardening With strain hardening

"5 ._~ 0.5


-

--

E -p
Z O0

0!2

'

014

0[6

'

018

'

I.O

Fraction of taper l e n g t h , s / l Fig. 1 h Distribution of die-pressure with and without strain-hardening (r = 45%, ~ = 16' / ' = 0.1).

Table 3 Quantitative effects of strain-hardening on the drawing parameters Parameter Set Parameter value without Drawing stress Set 1 Set 2 Set 1 Set 2 Set 1 Set 2 0.1252 0.8730 1.8911 2.1990 1.51 1.20 with 0.1924 1.5146 2.8830 3.7056 2.41 1.88 Error(w.r.t. strain -hardening value) (%) 34.93 42.36 34.41 40.66 36.93 36.17

~rxf/ao
Separation force S/(R2 2o0) Max. die-pressure

t,/ao

Set 1: r = 5%, ~ = 2 , f = 0.03 Set 2: r = 45%, z~= 16"~,f= 0.10.

3.4. Comparison of the drawing parameters,for three materials


Fig. 12 shows the variation of drawing stress with die-angle. Since the yield strength at zero plastic strain (~o) is different for each material, the non-dimensionalization factor is also different. At all angles, the non-dimensionalized drawing stress is greatest for copper and least for aluminium, as copper is the most strain-hardening amongst the three materials (Table 1). The aluminium selected here is slightly less strainhardening than the steel. It is observed that although the optimum die-angle is generally the same for all of the three materials, the profile of the curve changes in moving from a low strain-hardening material to a high strain-hardening material. Fig. 13 shows the die-pressure distribution for the three materials. Since the copper is the most strain-hardening material, the corresponding die-pressure is largest. Further, the rise in die-pressure near to the exit is also very large for the copper. Fig. 14 shows

U.S. Dixit, P.M. Dixit / Journal o/Materials Processing Technology 47 (1995) 201 229

221

17
- -

Steel
Copper
Aluminium i ""

i / / ~' // i

15

.... ----

r:25%,f:O05 r~ c 11 2 "~ 13

I.I

c _o 0.9 c k5 i c o 7 O7 O. 2.0 ' 4I0 6.0 ' 8.0 ' I0.0 ' 120 I 14.0 '

,do

18.o

Die semi-angle, a (deg)


F i g . 12. Variation of drawing stress with die-angle for three different metals.

6.0
6o

.....
-----

5.0

Steel Copper Aluminium

r= 2 5 % , f = 0 . 0 3 , d-=6

,'",

4.0 cL 3.0 o c o
c

/ g/ /

, I

/ /

i
, i

2.0 ~

......

g
Z

I lO( 0

0.2

"~,'~'-- "--I--" - - ; - - ' - - [ - -

0.4

0.6

"~"

0.8

"~

IlO

Fraction of taper l e n g t h , s / I
F i g 13. Distribution of die-pressure for three different metals.

the variation of separation force with die-angle, the non-dimensionalized separation force being greatest for the copper. Except at low angles, the non-dimensionalized separation force is same for the steel and aluminium. However, the absolute value of separation force will be greater for the steel than for the aluminium.

3.5. The study qf contours


To study the deformation pattern and its variation with process variables, the contours of non-dimensional equivalent strain-rate and equivalent strain are considered for four sets of process variables, these four sets being so chosen that comparison of the contours of the first set with any of the successive sets illustrates the effect of one of the three process variables.

222

U.S. Dixit, P.M. Dixit / Journal of Materials Processing Technology 47 (19953 201 229

The equivalent strain-rate distributions for the four sets of process variables are shown in Figs. 15-18. It is observed that the shape of the plastic boundaries is not spherical as assumed in Avitzur's upper-bound analysis, but is dependent on the process variables. The plastic deformation zone contains some portion of the wire before the entry and after the exit also. Further, this zone expands in moving away from the centre of the wire to the die-wire interface. The strain rates increase as die-wire contact is approached, greater strain-rate gradients being observed around the first contact point. Near to the exit, the strain rates start to decrease because of unloading. The manner in which they vary between the two plastic boundaries depends on the process variables.

cO 2 O

o
g R
16 ~, \

.... -----

Steel Copper Aluminium


0.03

r = 25 %,f=

12

= o

4
o =

'

'

~, ' ,'o'

,'2'

,~'

,'~

18

Die semi(Ingle, ~ (deg) Fig. 14. Variation of separation force with die-angle for three different metals.

Metal

steel

0.43
|

JJ,
0.03).

Fig. 15. Equivalent strain-rate contours (r = 5 % , :~ = Z ' , f =

U.S. Dixit, P.M. Dixit / Journal of Materials Processing Technology 47 (1995) 201 229 223
Metal: steel
0.47

'L
Rigid

Rigid

't
Fig. 16. E q u i v a l e n t strain-rate c o n t o u r s (r = 5%, c~ = 2 ' , f = 0.1).

Metal: steel

v ~

Rigid
.OZ

Rigid

Fig. 17. E q u i v a l e n t strain-rate c o n t o u r s (r = 5%. ~ = 10:~,f= 0.C::~.

Metal:steel

.062

Fig. 18. E q u i v a l e n t strain-rate c o n t o u r s (r = 45%, ~ = 2 : ' , f = 0.03).

Comparison of Fig. 15 with Fig. 16 shows that friction does not affect the strainrate distribution appreciably, except that there is a slight increase in the value of maximum strain-rate. Comparison of Fig. 15 with Fig. 17 shows that as the die-angle is increased keeping other process variables fixed, the plastic region near to the axis of

224

U.S. Dixit, P.M. Dixit / Journal o f Materials Processing Technology 47 (1995) 201 229

symmetry narrows down and the strain rates near to the die surface become very high compared to those near to the axis of symmetry. It is observed that as the die-angle is increased, at a particular angle the plastic zone disappears. In this case, near to the axis of symmetry the zone of the undeformed material and the zone of the fullydeformed product have a common boundary. Since the zone of fully-deformed material moves faster than the zone of undeformed material, they must separate and form a central burst. The central-bursting defect, which is observed in the case of large die-angles and small reductions, has been analyzed by Avitzur [14] using upperbound analysis, the derivation being based on the energy required for the flow of metal with a central-burst being lesser than that required for sound flow, when a central-burst occurs. Using the present FEM analysis, the central-burst defect can be studied by studying the strain-rate contours. However, the study of the central-burst defects as well as other of defects is left for future work. Comparison of Fig. 15 with Fig. 18 shows that increasing the reduction reduces the values of strain-rate and makes the strain-rate more uniform across the cross-section of the wire. Figs. 19 and 15 compare the strain-rate contours of a non-hardening material with that of a hardening material. Leaving aside minor details, the strain-rate distribution pattern is the same for both. However, the maximum strain-rate is greater for a non-hardening material. Figs. 20 and 21 compare the equivalent strain-rate contours of a non-hardening material with those of a hardening material for the other set of extreme process variables. Here also the same trend is observed, except that the value of maximum strain-rate is less for a non-hardening material. Figs. 22-25 show the distribution of equivalent strain for the four sets of process variables. In each set, over a cross-section, the equivalent strain is greatest at the die wire interface and least at the axis of symmetry, throughout the deformation zone. The value of the maximum equivalent-strain increases only marginally with the value off(Fig. 23) but significantly with the values of~ and r (Figs. 24 and 25). It is observed that increasing f o r ~ makes the deformation more non-uniform over the cross-section,
Metal : steel

0.7

Rigid -/

Rigid

Fig. 19. Equivalent strain-rate contours of non-hardening material (r - 5%, :t = 2 ' , f = 0.03).

U.S. Dixit, P.M. Dixit / Journal t~/Materials Processin~ Technology 47 (1995) 201 229 225
Metal : steel
1.7

i1(( (/
Fig. 20. E q u i v a l e n t strain-rate contours of non-hardening material (r = 4 5 % , ~ = 16 , . / = 0.11.
Metal : steel 3.3

Fig. 21. E q u i v a l e n t strain-rate contours (r = 4 5 % , ~ = 16 , / =

0.1).

Metal: steel

Contour level d i f f e r e n c e = 0 . 0 0 5

\, )

Fig. 22. E q u i v a l e n t strain contours 0" = 5 % , ~ = 2 , 1 = 0.03).

more pronouncedly near to the die wire interface in the exit region (Figs. 23 and 24). On the other hand, Fig. 25 shows that increasing reduction has the effect of improving the homogeneity of the product over the cross-section of the wire. Comparison of Fig. 26 with Fig. 22 and Fig. 27 with Fig. 28 shows that strain-hardening does not have an appreciable effect on the strain distribution pattern.

226

U.S, Dixit, P.M. Dixit / Journal o[ Materials Processing Technology 47 (1995) 201 229
Metol :

steel

Contour level difference = 0 . 0 0 5

~ ~, !

0.085

Fig. 23. E q u i v a l e n t s t r a i n c o n t o u r s (r = 5 % , ~ = 2 , f = 0.1).

Metol: steel

Contour level d i f f e r e n c e = 0 . 0 2

'd
Fig. 24. E q u i v a l e n t s t r a i n c o n t o u r s (r = 5 % , ~ = 1 0 , f = 0.03).

4. Conclusions

The steady-state wire drawing problem has been solved using the mixed pressure-velocity finite-element formulation. The resulting non-linear algebraic equations are solved iteratively by the Householder method. The material is considered to be rigid-plastic strain-hardening and a constant coefficient of friction is used at the die-wire interface. The model is shown to predict the drawing force and the separation force with reasonable accuracy.

U.S. Dixit, P.M. Dixit / Journal of Materials Processing Technology 47 (1995) 201 229 227
Metal : s t e e l .Ol
Contour level difference = 0.01

Fig. 25. Equivalent strain contours (r = 45%, ~ = 2 ,

f=

0.03).

M e t o l : steel

Contour

level d i f f e r e n c e = 0 . 0 0 5
e

Fig. 26. Equivalent strain contours of non-hardening material (r = 5%, ~ = 2', f = 0.03).

Metal : Steel

Contour level diff =0.05

'

au

0.1 Fig. 27. Equivalent strain contours of non-hardening material (r = 45%, ~ = 1 6 , f = 0.1).

The analysis suggests that the die-angle should be chosen according to maximum expected coefficient of friction if it is not known accurately. The results indicate that the die-pressure distribution and hence the shear-stress distribution, in general, varies along the taper length of die and at no time becomes uniform for any range of process variables. Thus the assumption of constant friction-factor, made often in the study of the drawing process, does not seem to be justified. The coefficient of friction has very little influence on the die-pressure distribution. It is observed that although the die-pressure decreases with increasing reduction, the separation force increases. Similarly, an increase in die-angle increases the die-pressure but decreases the separation force.

228 U.S. Dixit, P.M. Dixit / Journal o[ Materials Processing Technology 47 (1995) 201 229
Metol: Steel Contour level diff : 0.05

.80

.0,%.1 .15

Fig. 28. Equivalent strain contours (r = 45%, : = 16 , f - 0.1).

It is s h o w n that, for the selected steel, the i n c r e a s e in d i e - p r e s s u r e , s e p a r a t i o n force a n d d r a w i n g stress d u e to s t r a i n - h a r d e n i n g is a b o u t 3 5 % to 4 2 % . By c o n s i d e r i n g t h r e e m a t e r i a l s of different h a r d e n i n g characteristics, it is seen t h a t the d r a w i n g p a r a m e t e r s are f u n c t i o n s of the m a t e r i a l also. It is o b s e r v e d t h a t an i n c r e a s e in the r e d u c t i o n o r a d e c r e a s e in the die s e m i - a n g l e o r coefficient of friction tends to m a k e the p r o d u c t m o r e h o m o g e n e o u s . At v e r y low r e d u c t i o n a n d large die-angle, the plastic z o n e n e a r t h e axis of s y m m e t r y n a r r o w s d o w n a n d m a y e v e n t u a l l y d i s a p p e a r , c a u s i n g the c e n t r a l b u r s t i n g defect to occur. S t r a i n - h a r d e n i n g seems to h a v e v e r y little effect o n the d i s t r i b u t i o n of e q u i v a l e n t strain o r o f e q u i v a l e n t strain-rate.

References
[1] O. Smith, Flow of metals in the drawing process, J. Franklin Inst., 122 (1886) 321 346; 123 (1887) 232. [2] J.G. Wistreich, Investigation of the mechanics of wire drawing, Proc. Inst. Mech. En.qrs., 169 (19551 654 665. [3] H. Majors, Studies in cold-drawing; Part 3: Determination of friction coefficient, Trans. ASME, 78 (1956) 79 87. [4] B. Avitzur, Study of flow through conical converging dies, in: A.L. Hoffmanner (Ed.), Metal Formin~t: Interrelation Between Theory and Practice, Plenum Press, New York, 1971, pp. 1 46. [5] P.M. Cook and J.G. Wistreich, Measurement of die-pressure in wire drawing by photo-elastic methods, J. Appl. Phys., 3 (1952) 1 7. [6] E.G. Thomsen, C.T. Yang and T.B. Bierbower, An experimental investigation of the mechanics of plastic deformation of metals, Univ. Cal([i Publ. Eng., 5 (1954) 89 144. [7] O. Hoffman and G. Sachs, Introduction to the Theory qfPlasticity[)~r Engineers, Ch. 16, McGraw-HilL New York, 1953. [8] E. Siebel, Derderzeitige Stand der Erkenntnisse fiber die Mechanischen Vorgange beim Drahtziehen, Stahl Und Eisen, 66-67 (1947) 171 180. [9] B. Avitzur, Metal Forming: Process and Analysis, Chs. 8 9, McGraw-Hill, New York, 1968. [10] R.T. Shield, On the plastic flow of metals under conditions of axial symmetry, Proc. R. Soc. Lnnd., A233(1956) 267 287. [11] B. Avitzur, Analysis of wire drawing and extrusion through conical dies of small cone angle, Trans. ASME J. En.q. Ind., 85 (1963) 89-96. [12] B. Avilzur, Analysis of wire Drawing and extrusion through conical dies of large cone angle, Trans. A S M E J. Eng. Ind., 86 (1964) 305 316. [13] B. Avitzur, Strain-hardening and strain-rate effects in plastic flow through conical converging dies, Trans. A S M E J. En.q. Ind., 89 (1967) 556 562. [14] B. Avitzur, Analysis of central bursting defects in extrusion and wire drawing, Trans. ASME J. Eng. Ind., 90 (1968} 79 91.

U.S. Dixit, P.M. Dixit / Journal o f Material.; Processing Technology 47 (1995) 201 229 229

[15] K. lwata, K. Osakada and S. Fujino, Analysis of hydrostatic extrusion by the finite element method, Trans. A S M E J. En 9. Ind., 94 (1972) 697 703. [16] C.H. Lee, H. lwasaki and S. Kobayashi, Calculation of residual stresses in plastic deformation processes, Trans. A S M E J. Eng. Ind., 95 (1973) 283 291. [17] E.H. Lee, R.L. Mallet and W.H. Yang, Stress and deformation analysis of the metal extrusion process, SUDAM No. 76-2, Stanford University, June 1976. [18] S.N. Shah and S. Kobayashi, A theory on metal flow in axisymmetric piercing and extrusion, J. Prod. Engr., 1 (1977)73. [19] O.C. Zienkiewicz, P.C. Jain and E. Onate, Flow of solids during forming and extrusion: Some aspects of numerical solutions, Int. J. Solids Structs., 14 (1978) 15 ~38. [20] A.K. Tayal and R. Natrajan, Extrusion of rate-sensitive materials using a viscoplastic constitutive equation and the finite element method, Int. J. Mech. Sci., 23 (1981) 89-98. [21] J.S. Gunasekara, H.L. Gegel, J.C. Malas, S.M. Doraivelu and J.T. Morgan, Computer aided process modelling of hot forging and extrusion of aluminium alloys, Ann. CIRP, 31 (1982) 131 135. [22] A.P. Balaji, T. Sundararajan and G.K. Lal, Viscoplastic deformation analysis and extrusion die design by FEM, Trans. A S M E J. Appl. Mech., 58 (1991) 644 650. [23] J.H. Bianchi and T. Sheppard, A comparison ofa viscoplastic finite-element model with slip-line field and upper-bound solutions for non-hardening material subjected to plane strain and axisymmetric extrusion, Int. J. Mech. Sci., 29 (1987) 61 81. [24] T. Reinikainen, K. Andersson, S. Kivivuori and A.S. Korhonen, Finite-element analysis of copper extrusion processes, J. Mat. Proc. Tech., 34 (1992) 101 108. [25] D. Bhattacharya, P.J. Richards and A.A. Somashekar, Modelling of metal extrusion using the PHOENICS package, J. Mat. proc. Tech., 35 (1992} 93 111. [26] C.C. Chen, S.I. Oh and S. Kobayashi, Ductile fracture in axisymmetric extrusion and drawing: Part 1: Deformation mechanics of extrusion and drawing, Part 2: workability in extrusion and drawing, Trans. A S M E J. Eng. Ind., 101 (1979) 23 44. [27] L. Chevalier, Prediction of defects in metal forming: Application to wire drawing, J. Mat. Proc. Tech., 32 (t992) 145 153. [28] K.J. Bathe, Finite Element Procedures in Englineerinq Analysis, Prentice-Hall, New Delhi, 1982, pp. 475 478. [29] J.N. Reddy, An Introduction to the Finite Element Method, McGraw-Hill, New York, 1984, pp. 280 293. [30] U.S. Dixit, Finite element analysis of wire drawing of strain-hardening materials, M. Tech. Thesis, liT Kanpur, 1993, pp. 38 40. [31] F.A.R. A1-Salehi, T.C. Firbank and P.R. Lancaster, An experimental determination of the roll pressure distributions in cold rolling, lnt. J. Mech. Sci., 15 (1973) 693 710.

Potrebbero piacerti anche