You are on page 1of 37

Accepted Manuscript

Title: Fourier series based finite element analysis of tube


hydroforming - generalized plane strain model

Authors: Yabo Guan, Farhang Pourboghrat

PII: S0924-0136(07)00634-6
DOI: doi:10.1016/j.jmatprotec.2007.06.044
Reference: PROTEC 11049

To appear in: Journal of Materials Processing Technology

Received date: 21-11-2006


Revised date: 3-6-2007
Accepted date: 15-6-2007

Please cite this article as: Y. Guan, F. Pourboghrat, Fourier series based finite element
analysis of tube hydroforming - generalized plane strain model, Journal of Materials
Processing Technology (2007), doi:10.1016/j.jmatprotec.2007.06.044

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
1

FOURIER SERIES BASED FINITE ELEMENT ANALYSIS OF TUBE


HYDROFORMING -
GENERALIZED PLANE STRAIN MODEL

p t
Yabo Guan1, Farhang Pourboghrat2

cri
1
Department of Neurosurgery, Medical College of Wisconsin,
Milwaukee, WI, 53226, USA

us
2
Department of Mechanical Engineering, Michigan State University,
East Lansing, MI, 48824-1226, USA

an
dM
pte

Corresponding author:
ce

Yabo Guan
Department of Neurosurgery
Medical College of Wisconsin
9200 West Wisconsin Avenue
Ac

Milwaukee, WI 53226, USA


Tel: 414-384-2000-ext 41387
Fax: 414-483-4393
E-mail: yguan8805@yahoo.com

Page 1 of 36
2

Abstract – In previous paper (Fourier series based finite element analysis of tube hydroforming --
an axisymmetric model. Engineering Computations. 23(7): 697-728, 2006), an axisymmetric
analysis of tube hydroforming process was discussed. In the present paper, a generalized plane
strain implicit formulation of the cross sectional expansion of an extruded aluminum tube with
pressurized fluid to fill a hydroforming die is presented. The cross-section of the tube is modeled

t
with thin straight and circular segments with constant thickness, and Fourier series are used to

p
approximate nodal displacements. The material of the tube is assumed to obey a rate-independent,

cri
elastoplastic model that takes into account work hardening and normal anisotropy. At the tube-die
interface, frictional stress is assumed, based on Coulomb friction, to be proportional to the contact
pressure whenever relative sliding occurs. The kinematics relationships are derived based on thin

us
shell theory, and the equilibrium equation is derived based on virtual work principle. The axial feed
is implemented by imposing either a compressive force or strain in the tube length direction.

an
Frictional boundary condition is introduced into the formulation in the form of a penalty function,
which imposes the constraints directly into the tangent stiffness matrix. The Newton-Raphson
iterative method is used to incrementally solve the resulting nonlinear equations. Two examples of
dM
tube hydroforming problems are solved and numerical predictions of the deformed shape,
hydroforming pressure, and deformation strains are compared with experimental and ABAQUS
results.
pte

Keywords: Aluminum; Tube Hydroforming; Finite Element; Contact Analysis; Fourier series;
Plane Strain
ce
Ac

Page 2 of 36
* Manuscript

1. Introduction

Tube hydroforming is receiving the greatest attention, especially in the auto industry,
because existing multi-piece, stamped/welded assemblies in auto body and frame
structures could be potentially replaced with less expensive hydroformed parts that are
lighter, stronger and more precise. Well known hydroformed automotive applications
include exhaust manifolds, radiator enclosure, dash assemblies, frame rail, and engine
cradles etc. [1, 2].

t
Prior to tube hydroforming, pre-bending and stretching operations take place to shape

p
the blank tube to fit into the hydroforming die. The analysis of this pre-forming is
necessary in order to accurately predict the formability of the tube during the

cri
hydroforming process. Wu and Yu [3] simulated the multi-operation tube hydroforming
of an automotive structural part with explicit LS-Dyna3D commercial code. Using the
explicit finite element code LS-Dyna3D, Srinivasan et al. [4] provided additional

us
correlation of experimental and simulation results for tube hydroforming, and Liu et al.
[5] provided analytical and experimental examination of tube hydroforming limits. Kaya
et al. [6] performed plane strain analysis of crushing and expansions of tube cross-
sections using the two-dimensional implicit finite element code DEFORM 2D. Kim et al.
an
[7] developed a rigid-plastic finite element method for the analysis of tube hydroforming
process. Hwang and Altan [8, 9] evaluated the quality of the tubes formed by
hydroforming and crushing in a square die and rectangular die respectively. A two
dimensional model for the bend-stretch-pressure forming process was developed by
dM
Corona [10]. Other numerical analyses of tube hydroforming performed recently can be
found in Refs. [11-28].
Tube hydroforming is the process whereby a closed-section hollow part with varying
cross sections is formed by applying internal fluid pressure and axial compressive loads
to force a tubular metal blank to conform into the shape of a given die cavity. Although
finite element method has been used widely for simulating the process, unfortunately it
pte

becomes very costly when three dimensional model of working piece is created and used
for analysis directly. Therefore, cost-efficient two dimensional finite element method
capable of simulating various cross section shapes is desirable. In our previous study
[29], an axisymmetric tube hydroforming finite element analysis program was developed.
Fourier series interpolation functions, which reduce the size of the global stiffness matrix
ce

and the number of variables considerably, were employed for approximating the
displacements. Some simplifications were assumed. The principal geometrical
assumption is that the representative meridian of the tube is initially straight. This
Ac

assumption however could be relaxed by using a curved, instead of a straight segment to


represent the initial geometry of the tube. All segments making up the meridian are
assumed to be relatively thin and of constant thickness. The deformation of the tube is
assumed not to vary along its cross-section, hence, the analysis could be considered to be
axisymmetric. The axisymmetric hydroforming program (AXHD) written based on this
formulation is very efficient in predicting the deformations for the free-forming stage of
tube hydroforming under simultaneous action of internal pressure and displacement
stroke. Failure model (FLC) based on shear instability was also incorporated into the
code to predict the onset of fracture for the steel tube. The hoop and axial strains

Page 3 of 36
predicted with AXHD code compared excellently with those from ABAQUS using plane
stress axisymmetric (SAX1) and four-node shell (S4R) elements.
In the present study, the principal geometrical assumptions of the model are that the
cross-section of the round extruded tube can be modeled with circular segments, and all
segments are relatively thin and of constant thickness. The axial feed is implemented by
imposing either a compressive force or strain in the tube length direction. The material of
the tube is assumed to be elastoplastic and to obey a plasticity model that takes into
account rate-independent, work hardening and normal anisotropy. The boundary friction

t
condition is introduced into the formulation in the form of a penalty function, which

p
imposes the constraint directly into the tangent stiffness matrix. The Newton-Raphson
algorithm is used to solve the nonlinear equations. The PSHD (Plane Strain

cri
Hydroforming) program has been written based on the above formulations.
The layout of present paper is as following: In Section 2 the thin shell model is
described. The kinematics assumptions and principal strain formulations are discussed in

us
Section 3. The constitutive model and contact algorithm are described in Sections 4 and 5
respectively. Section 6 describes the equilibrium equation formulations based on the
virtual work principle (VWP) and the application of Newton-Raphson iterative method to
solve the resulting non-linear equations. Finally, in Section 7 two examples are provided
an
in support of the section-analysis finite element model, where numerical predictions of
the deformed shape, hydroforming pressure, and deformation strains are compared with
experimental measurements and the nonlinear finite element code ABAQUS.
dM
2. Thin Shell Theory

The deformation of the mid-surface of an element will be considered based on thin


shell theory. Figure 1 shows the shell mid-surface at the reference time ot and current
pte

time t as it bends and stretches.

Pourboghrat et al. [30] derived the principal curvatures and stretches of a shell of
revolution undergoing axisymmetric deformation using both total and updated
Lagrangian formulations. Below, the principal curvature and stretch of a thin shell
ce

undergoing generalized plane-strain deformation will be derived. The generalized plane


strain assumption implies that strain in the third principal direction (in this case the tube
length direction) could be specified as a constant value. When this constant value is
chosen to be zero, conventional plane strain assumption will result.
Ac

2.1 Principal Curvature and Stretch (updated Lagrangian)

After bending and stretching, the principal mid-surface curvature, k1, of a shell
element at the current configuration, t (=ot+Δt), could be calculated from the known
information about the element at the reference configuration (time ot, see Figure 1), as
follows:

Page 4 of 36
r = R + u Aˆ + w Nˆ (1)
~ ~ ~ ~

where R is the reference configuration at time ot and u and w are incremental


~

displacements defined in Figure 1. In Eq. (1), the unit tangent vector  and the unit
~

principal normal vector N̂ to the mid-surface of the reference configuration are defined
~

t
as

p
∂R
Aˆ = ~
=R (2)

cri
~ ∂S ~ ,S

Aˆ Nˆ

us
Nˆ = or, Aˆ = −
~ ,S ~ ,S
(3)
~ K1 ~ K1

~⎝
an
where K1 is the centerline curvature at the reference configuration. To calculate centerline
curvature at the current configuration, k1, the unit tangent vector aˆ ⎛⎜ = a a ⎞⎟ and the unit
~ ~ ⎠
dM
principal normal vector of the mid-surface of the shell nˆ ⎛⎜ = n n ⎞⎟ must be known. Using
~⎝ ~ ~ ⎠

Eq. (1), the tangent vector a could be calculated as


~

∂r
a= ~
=r = R + u , S Aˆ + u Aˆ + w, S Nˆ + w Nˆ (4)
∂S
pte

~ ~ ,S ~ ,s ~ ~ ,S ~ ~ ,S

By substituting from Eqs. (2) and (3) into (4), and after re-arranging, the following
expression results:
∂r
a= ~
=r = (1 + u , S − K 1 w) Aˆ + ( w, S + K 1u ) Nˆ = c Aˆ + d Nˆ (5)
ce

~ ∂S ~ ,S ~ ~ ~ ~

The principal incremental stretch of the mid-surface in the radial direction calculated
from the magnitude of the base vector a in Eq. (5) is
Ac

[
λ 1 = a = a⋅ a = c 2 + d 2 = (1 + u ,S − K 1 w) + (w,S + K 1u ) ]
1
2 2 2
(6)
~ ~ ~

The current length of the mid-surface of the shell in the radial direction, ds, is calculated
from the reference length, dS, and λ1 as follows:
ds = λ1dS (7)

Page 5 of 36
The unit principal normal vector of the surface of the current shell, n̂ , is
~

− d Aˆ + c Nˆ
nˆ = ~ ~
(8)
~ λ1

which, from Eqs. (5), (6) and (8), shows that nˆ⋅ aˆ = 0 . The current principal curvature of
~ ~
the shell, k1, could now be found as

t
1

p
k1 = − â⋅ n̂ = − r ⋅ n̂ = − a⋅ n̂ (9)
~ ~ ,s ~ ,s ~ ,s λ21 ~ ~ ,S

cri
where a is given by Eq. (4) and nˆ can be derived from Eq. (8):
~ ~ ,S

d nˆ λ1 ⋅ (d, S Aˆ + c, S Nˆ − K1c Aˆ − K1d Nˆ ) − λ1, S ⋅ (− d Aˆ + c Nˆ )


nˆ = ~
= ~ ~ ~ ~ ~ ~
(10)

us
~ ,S dS λ12

In Eq. (10), λ1,S is assumed to vanish within an element and the above expression
simplifies as

n̂ = −
~ ,S
an
(d ,S + K1c ) Â~ + (− c ,S + K1d ) N̂~
λ1
(11)
dM
By substituting from Eqs. (5), (6) and (11) into Eq. (9), the current centerline curvature of
the shell, k1, can be found:
cd ,S − dc ,S + K1λ21
k1 = (12)
λ31
pte

3. Kinematics of the Circular Segment

Using the updated Lagrangian formulation, exact expressions for membrane strains,
normal vector rotation, and principal curvatures of plane strain shell element were
ce

derived in Section 2 (Shell Element Model). By using in these expressions the values of
displacements and curvatures of the shell at previous time increment (i.e., t = ot ), one
would recover the incremental values of strains. However, by using in these expressions
Ac

the values of displacements and curvatures of the shell at the initial time (i.e., t = 0 ) one
would recover the total values of strains. In this paper, the difference between the total
strains at time t (current) and o t (previous) is used to calculate the incremental strains,
i.e., Δε = ε (Y ,t ) − ε (Y , ot ) , where Y corresponds to the location of a material point on
the cross section of the tube.

Page 6 of 36
3.1 Kinematics Assumptions

The hydroforming of a tube, shown in Figure 2a, proceeds by calculating the


deformation of its cross section (YZ-plane), while applying a compressive force or strain
in the length direction (X-axis) to simulate the axial feed. The cross section of the tube,
Figure 2b, could be a closed or open section comprised of straight and circular segments.
Only circular segment kinematics will be discussed in the present paper for the round
extrusion tube hydroforming application. Detail of kinematics formulation for straight

t
segment can found in a previous paper [29].

p
All segments making up the cross-section of the tube are assumed to be thin and of

cri
constant thickness. To meet the thin shell assumption, the radius of the circular segment
should be greater than eight times the thickness. Based on thin shell theory described in
Section 2, deformation (i.e., strain, rotation) and curvature expressions will be derived
only for circular segments as functions of displacements ( u, w ), and their derivatives

us
( u , s , u , ss , w, s , w, ss ). Details of these derivations are given in appendix A.

3.2 Principal Strains


an
For the circular segment shown in Figure 3, the local coordinates of the segment are
dM
defined by the angle φ and the through-thickness axis z. The initial geometry of the
segment is specified by: 1) the coordinates of the center of the arc, Yc and Z c , 2) the
angle φ 0 , which locates the line where φ = 0 , 3) the centerline radius R, 4) the angle
spanned by the arc Δφ , and 5) the thickness of the segment t. As in the straight segment,
the mid-surface is at z=0 and is indicated by the dashed line.
pte

The rotation of a through-thickness line, β , is derived in appendix A (Eq. A7) to be:


u − w,φ
β= (13)
R
ce

The true axial strain at any point along the tube cross section was given by
ε x = ε 0x + κ z ξ − κ Y η (14)
Ac

where ξ and η are defined as


ξ = Z c + ( R + w ) sin φ + u cos φ + z(sin φ + β cos φ ) (15a)

η = Yc + ( R + w) cos φ − u sin φ + z (cosφ − β sin φ ) (15b)

where φ = φ 0 + φ . The engineering tangential strain is given by

Page 7 of 36
eφ = eφ0 + κ φ z (16)

where membrane portion of the strain, eφ0 , and the local curvature κφ are derived in
appendix A (Eqs. A3 and A12) to be:
2 2
⎛ u + w ⎞ 1 ⎛ u − w,φ ⎞ 1 ⎛ w + u,φ ⎞
eφ = ⎜⎜ ,φ
0
⎟⎟ + ⎜⎜ ⎟⎟ + ⎜⎜ ⎟⎟ (17)
⎝ R ⎠ 2⎝ R ⎠ 2⎝ R ⎠

t
and

p
uφ − w,φφ
R

cri
κφ = (18)
u − w,φ
R 1− ( )2
R

us
The true strain is then calculated from εφ = ln(1 + eφ ) .

3.2 Constraints
an
Since the cross-section is made up of several independent segments, with their own
local coordinate systems and variables, it is necessary to enforce compatibility of
dM
deformations at junctions of two or more segments. Two constraint equations are used to
ensure compatibility of displacements and one equation to ensure compatibility of
rotation between two segments. Therefore, at a junction where M segments come
together, 3(M-1) constraint equations need to be enforced. At a junction between two
straight segments having orientation angles θ1 and θ 2 and components u1 , w1 and u2 , w2 ,
the two displacement compatibility conditions can be written in terms of the displacement
pte

components in the Y and Z directions as follows:


u 2 cos θ 2 + w2 sin θ 2 − u1 cos θ1 − w1 sin θ1 = 0 (19)
u 2 sin θ 2 − w2 cos θ 2 − u1 sin θ1 + w1 cos θ1 = 0 (20)
ce

The rotation constraint requires that the angle between segments at a junction remain
unchanged. For the current example, if the rotations of the two members at the junction
are β1 and β 2 , compatibility condition is
Ac

β 2 − β1 = 0 (21)
4. Constitutive Equation

The elastic-plastic, rate-independent constitutive model implemented in the


generalized plane strain tube hydroforming analysis code assumes isotropic hardening
and is based on Pourboghrat et al. [30]. The uniaxial stress-plastic strain curve of the
material is assumed to have the following power-law form:

Page 8 of 36
σ = K (ε + ε o )n (22)

where σ is the effective stress and ε is the effective plastic strain. Parameters K, n and
ε 0 are material constants that are calculated by curve fitting Eq. (22) to stress-strain data
from a uniaxial tensile test. The elastic strain increment is related to the stress increment
through the equations of linear, isotropic elasticity with Young’s modulus E and
Poisson’s ratio ν. The yield function given below allows for anisotropic yielding of the
material:

t
σ x2 + σ s2 + R (σ x − σ s )
2

p
φ= −σ 2 = 0 (23)
1+ R

cri
where R is the normal anisotropy parameter and x and s indicate axial and hoop direction,
respectively.
.

us
During the loading, Hooke's law is used to calculate stress below the elastic limit; i.e.,
σ ≤ σ y , where σ y is the initial yield stress of the tube material obtained from a uniaxial
tensile test. Beyond the elastic limit; i.e., σ > σ y , the co-rotational time derivative of

an
stress (Jaumann stress rate) is calculated, for a given strain rate, from an elastic-plastic
constitutive equation:


⎡ L:P P:L
σ = ⎢L − ~ ~
( ) ( ) ⎤⎥ : D
~ ~
dM
(24)
~ ⎢~ h+ P:L:P ⎥ ~
⎣⎢ ~ ~ ~
⎦⎥

e p
Here σ and D( = D + D ) are the Jaumann rate of stress and strain rate tensors,
~ ~ ~ ~

respectively, σ is the stress tensor, h( = ∂σ ∂ε ) is the plastic hardening parameter, L is


pte

~ ~

the fourth order elastic tensor and P⎛⎜ = ∂φ ∂ σ ∂φ ∂ σ ⎞⎟ , where


~⎝ ~ ~ ⎠

∂φ ∂ σ = ∂φ ∂ σ : ∂φ ∂ σ , is the second order tensor representing the unit normal to


~ ~ ~
ce

the flow potential surface. The effective plastic strain rate, associated with Eq. (24), is
calculated from the following expression:
.

P:L:D
ε = ~ ~ ~
(25)
Ac

h + P:L:P
~ ~ ~

The fourth order elastic tensor L( = Lijkl ) used in this work is the standard tensor for the
~
isotropic elasticity, which has only two independent components.

Page 9 of 36
5. Contact Algorithm

The tube hydroforming simulation requires modeling the frictional contact between
the tube and the die. The contact analysis is complex because it requires accurate
tracking of the motion of multiple bodies, and the motion due to the interaction of these
bodies after the contact. The numerical objectives are to detect the motion of the bodies,
apply a constraint to avoid penetration, and finally apply appropriate boundary
conditions to simulate the frictional contact behavior. Each of these objectives will be

t
separately described next.

p
cri
5.1 Contact Detection

Many contact search algorithms including methods for global search and local search

us
using sheet (mesh) normal or tool normal were proposed in the sheet metal forming
simulation literature [31-39]. In the present paper, a global search method using tube
surface normal was employed.

an
To detect contact between the tube and the die, evenly spaced contact nodes are
initially defined along the tube cross section (e.g., at so , s1 = so + Δs , etc.). During the
contact analysis, the displacement of each contact node is checked for surface
dM
penetration, by determining whether it has crossed into the die. For this purpose, the
calculation of the tube surface normal is required, since it is used to determine which
segment on the die is closest to a potential contact node on the tube cross section. For
example, as shown in Figure 4, the closest segment on the die (i.e., Bi −1Bi or Bi Bi +1 ) to
the contact node ( Ak ) on the tube can be determined using the following cross-product
pte

algorithm:
uuuuur uuuuuuur
If: (A k Bi × nˆ ) ⋅ (A k Bi +1 × nˆ ) < 0 (26)

then, Bi Bi +1 will be the die segment associated with the contact node Ak .
ce

5.2 Projection Algorithm


Ac

A nodal position produced by the trial solution may penetrate the die. By using the
cross-product algorithm, the closest segment on the die corresponding to the contact node
can be found. The nodal coordinates are then modified by a projection scheme such that
the contact node just touches the die surface. There are two ways to bring the penetrated
contact node back to the die surface. As shown in Figure 5, PQ is assumed to be the die
segment associated with the penetrated contact node A, point B is the intersection point
between the normal vector and PQ, and O is the original location of the contact node.
Based on the following vector equation, the coordinate of point B could be calculated:

Page 10 of 36
OP + PB = OA + AB (27a)

PQ uuur
PB = η1 , AB = −η2 nˆ (27b)
PQ

where η1 , η 2 are scalar parameters. Once η1 , η 2 are solved for from Eqs. (27a) and (27b),
the coordinate of point B could be determined.

p t
5.3 Implementation of Contact Constraints

cri
A contact node projected on the die surface at time t + Δ t , is constrained to move in
the tangent direction defined by the trial solution, Δu * . The constraint on the
%
displacement vector δu = (δu , δw) , for contacting nodes is then:

us
%
δu ⋅ nˆ = 0 (28)
%

5.4 Separation of a Node in Contact


an
After a node on the tube comes into contact with the die surface, it is possible for it to
dM
separate from the die surface in a subsequent iteration or deformation increment.
Mathematically, a node should be separated from the die when the calculated reaction
force at the node becomes tensile, as it would imply that the node on the tube is being
pulled by a tensile force to keep it in contact with the die. In contrast, when the reaction
force on the node is compressive, the node continues to stay in contact with the die.
When contact occurs, a reaction force associated with the contact node balances the
pte

internal stress of the element sharing this node. When separation occurs, this reaction
force behaves as a residual force (as the force on a free node should be zero). This
requires that the internal stresses in the deformable body be redistributed.
ce

6. Equilibrium Equation

The equilibrium equation is satisfied using the virtual work principle (VWP). In
Ac

contrast to traditional finite element method, nodal displacements in this formulation are
approximated with Fourier series, which makes the implementation of contact constraints
and boundary conditions more challenging. The boundary friction condition is introduced
into the formulation in the form of penalty functions, which imposes those constraints
directly into the tangent stiffness matrix. The Newton-Raphson algorithm is then used to
solve the nonlinear equilibrium equations. In the following sections the VWP will be
discussed for bending, pressure loading and frictional contact modeling.

Page 11 of 36
6.1 Tube Bending Problem

The principle of virtual work for tube bending could be expressed as following:

∑ ∫ (σ δε )
I J
i
x
i
x + σ sii δε sii dAi + ∑ λ j δC j = Tδε x0 (29)
Ai
i =1 j =1

t
where the integral on the left hand side represents the virtual internal work ( δ WInt ), the

p
term on the right hand side represents the virtual external work ( δ WExt = T δε x0 ) due to

cri
axial feed, index I corresponds to the total number of segments defining the cross-section,
index J corresponds to the total number of constraint equations, λ j are Lagrange

multipliers, C j are the constraint equations (from Eqs. (19-21)), T is the applied tension

us
(compression) due to axial feed, dA i = dsi dz i is the area of a straight segment, and
dA i = R i dφ i dz i is the area of a circular segment, x and s indicate axial and hoop direction,
respectively.
an
To solve Eq. (29), the displacement components wi , u i for each circular segment are
approximated using the following Fourier series expressions:
dM
n π φi N i
Ni
n π φi
i

w = α + ∑ α cos
i i
+ ∑ β n sin
i
(30a)
Δφ Δφ i
0 i n
n =1 n =1

n π φi N i
Ni
n π φi
i

u = γ + ∑ γ cos
i i
+ ∑ δ n sin
i
(30b)
pte

Δφ Δφ i
0 i n
n =1 n =1

After substituting from Eqs. (30a-b) into the principle of virtual work, Eq. (29), a
nonlinear expression of the following form will result:
(
f c, dκ Y , dκ Z , dT = 0 ) (31a)
ce

where
{
c = α 0i , α ni , β ni , γ 0i , γ ni , δ ni , δε x0 , λ j } (31b)
Ac

Equation (31a) should be incrementally solved for the unknown vector, c , given input
%
values for incremental bending curvatures, dκ Y , dκ Z , and incremental axial force, dT ,
applied to the ends of the tube (along X-direction) for the purpose of axial feeding. Since
Eq. (31a) is highly nonlinear, it is numerically solved using the Newton-Raphson method.
The final form of the Newton-Raphson iterative method used for solving c looks as
%
following:

10

Page 12 of 36
[K ][d c]= [R] = F
≈ ~ ~
Ext
~
− FInt
~
(32)

where K ( = ∂ 2WInt ∂c ∂c ) is the second variation of the virtual internal work with
≈ % %
respect to c , d c is the incremental c , and R is the residual. FExt ( = ∂WExt ∂c ) is the
% ~ % ~ ~ %
variation of the virtual external work with respect to c , and FInt ( = ∂WInt ∂c ) is the
% ~ %
∂c

t
variation of the virtual internal work with respect to c . The nodal force F ( = FInt ~ )
∂u

p
% ~ ~
~

∂u

cri
~
can be calculated from FInt and .
~ ∂c
~

us
6.2 Pressure Loading Model

Pressure loading is modeled as an external force to expand the tube. The virtual
external work done by a pressure p applied to the inside of the tube is equal to
I
an I
∂ (Δu i )
δW P
Ext = ∑ ∫ p nˆ δ (Δ u )dA i
~
i i
p = ∑∫ p nˆ i
% δ c~ dAp
∂c
i
(33)
dM
i =1 Ai i =1 Ai
p p ~

where δ (Δ u i ) is the incremental virtual displacement vector having two components


~

δ (Δw ) and δ (Δu i ) , and n̂ i is the unit outward normal to the tube surface A ip .
i

The principle of virtual work for bending and pressure loading of a tube then becomes
pte

I J I

∑ ∫ (σ xi δε xi + σ si δε si
i =1 Ai
i i
) dAi + ∑ λ j δ C j = T δε xo +
j =1
∑ ∫ p nˆ
i =1 Ai
i δ (Δu i ) dAip
%
(34)
p

The variation of the virtual external work due to internal pressure loading is:
ce

I ∂ ( Δu i )
F
%
P
Ext =p ∑∫ i =1 Ai
nˆ i

∂c
% dAip (35)
~
Ac

Due to the follower forces effect [40], the load stiffness matrix is
⎛ ∂ (nˆ i ) ∂ (Δu i ) i ⎞
i ∂ ( Δu )
I 2

∑∫ ⎜ % ⎟ dAp
P
K = p % + ˆ
n i
≈ Ext
i =1 Ai ⎜ ∂c ∂c ∂2 c ⎟
p ⎝ ~ ~ ~ ⎠
I
∂ (nˆ i ) ∂ (Δu i ) i
= p ∑∫
i =1 Ai ∂c
% dAp
∂c
(36)
p ~ ~

11

Page 13 of 36
Equations (35) and (36) will appear on the right- and left-hand sides of the Newton-
Raphson expression, Eq. (32), as follows
⎡ p ⎤ ⎡ ⎤ ⎡ ⎤
⎢⎣ K≈ + K≈Ext ⎥⎦ ⎣ d c~ ⎦ = ⎣ R~ ⎦ = F~Ext + F~Ext − F~Int
p
(37)

Once the above formulation was implemented into the numerical analysis code, it was
p
found that the load stiffness matrix K Ext has no or little effect on the solution, which

t
drove us to further simplify the formulation. To that end, by using the following

p
approximation, nˆ iδ (Δ u i ) ≈ δ (Δwi ) , the whole formulation was greatly simplified. The
~

details of the proof could be found in appendix B. Since ∂ 2 (Δwi ) ∂ 2 c = 0 , there is also

cri
~
p
no contribution to the stiffness matrix as a result of K Ext , appearing on the left hand side

of the Newton-Raphson expression, Eq. (37). When both methods were implemented into

us
the numerical code, the results turned out to be almost identical.

6.3 Frictional Contact Modeling


an
The most challenging task when developing a numerical code for metal forming
processes is to model frictional contact. To model the tooling-workpiece frictional
dM
contact correctly, the following two conditions were continuously monitored during the
equilibrium iteration:
(1) Penetration of the contact nodes into the die, and
(2) Nodal contact forces becoming tensile at the contact boundary (separation).
pte

Once the penetration of the contact nodes into the die has been detected, the
penetrated nodes were returned to the die surface and constrained to stay on the die
surface for the remainder of the equilibrium iterations. The nodes, which were returned to
the die surface, were constrained to move only tangent to the die surface and only
condition 2 stated above could cause the contacting node to be separated from the die
ce

surface. Figure 6 shows the schematic of a typical contact check during the Newton-
Raphson equilibrium iteration. The external work done by the frictional contact is added
to the virtual work principle Eq. (34) as following:
I J I I

∑ ∫ (σ δε )
Ac

i
i
x
i
x + σ sii δε sii dAi + ∑ λ jδ C j = T δε x0 + ∑ Pnˆiδ (Δu i )dApi + ∑ ∫ τ i δ (Δu i )dAτi (38)
i =1
A
j =1 i =1 % i =1 Ai %
τ

where τ i is the traction on the surface of the tube and δ (Δu i ) is the virtual incremental
%
displacement of the contacting nodes.
In order to improve convergence, a special algorithm was introduced. For each trial
set of contacting and non-contacting nodes, equilibrium iteration was performed. After
equilibrium was satisfied, the nodes were reexamined for non-penetration condition.

12

Page 14 of 36
Releasing or projecting certain nodes then updated the contact set and another
equilibrium iteration was initiated.
During contact iteration, the trial displacements were first updated according to the
Newton-Raphson procedure and the non-penetration contact condition was then applied
to these trial values by projecting the contact nodes to the die surface along the normal
vector. The modified trial solutions were then used for Newton-Raphson iteration. Within
this force equilibrium iteration, the internal force was calculated. The signs of the sheet
normal force at contact nodes were checked so that the nodes having non-compressive

t
(tensile) force were released.

p
7. Numerical Results

cri
The PSHD (Plane Strain Hydroforming) program has been written based on the above
formulations and examples of generalized plane strain tube hydroforming with an
aluminum tube were solved and verified with ABAQUS finite element code, and when

us
available, compared with experimental results. The results of these simulations will be
presented next.

7.1 Hydroforming of a Round Aluminum Tube into a Square Die


an
The following example was chosen in order to illustrate the capability of the new
formulation to model the hydroforming of a round aluminum tube into a square die, and
dM
to also study:
(a) Effect of die-tube friction coefficient on predicted strains,
(b) Sensitivity of the contact solution to the number of segments used, and
(c) Effect of pre-bending on tube hydroforming.
pte

For case (a), the predicted shape of the hydroformed tube for an applied internal pressure,
and deformation strains were verified by direct comparison with experimental data. For
case (b), the shape of the hydroformed tube was verified against ABAQUS simulation
results. Verification of case (c) was not possible since no experimental data was available
and we were unable to get converged solution with ABAQUS.
ce

7.1.1 Tube Hydroforming Experiment


Ac

An 8.0” (203 mm) long aluminum 6061-T4 tube with an outer radius of 1.0” (25.4
mm) and a thickness of 0.049” (1.24 mm) was hydroformed into a square die with a side
length of 2.0” (50.8 mm) using a maximum internal pressure of 3040 psi (21 Mpa). The
mechanical properties of the aluminum 6061-T4 tube obtained from a uniaxial tensile test
are shown in Table 1.
Figure 7a shows the hydroformed tube and its cross section. The hoop strain
distribution of the hydroformed tube at the maximum pressure of 3040 psi was

13

Page 15 of 36
experimentally measured using circular grids etched on the straight tube. Table 2 shows
measured hoop strains as a function of the angular position along the cross section of the
hydroformed tube, as shown in Figure 7b. It can be seen, from Table 2, that the maximum
hoop strain is measured to be between 7-11%, which occurs in the 0-10 degree zone.
Near the boundary between the curved and flat portion of the tube (i.e., 10-20 degree
zone), hoop strains are about 7%. The tube was axially compressed a total of 0.1181” (3
mm) from each end during the hydroforming process, which resulted in a measured axial
strain of -3%.

p t
7.1.2 Tube Hydroforming Simulation with PSHD Code

cri
The tube hydroforming process was simulated using the new formulation to illustrate
its capability to model frictional contact. Due to symmetry, only one quarter of the tube
was modeled, as shown in Figure 7b, to reduce computational time. The tube was

us
modeled using 8 circular segments, and for each segment 6 Fourier series terms were
used (i.e., N=6). The best number of terms to use in the simulation was determined based
on whether or not a converged solution was obtained. It was found that N=4 to 6 (total of
an
9 to 13 coefficients) would result in numerically stable solutions. Divergence of solution
more commonly occurred when larger values of N were used. Sometimes using larger
values of N caused numerical instability during the contact search algorithm. To obtain
accurate contact solution, 20 evenly distributed contact nodes were used for each segment.
dM
The number of Gauss integration points used along the length of the segment was 12,
while that used through the thickness was 3.
The tube hydroforming simulation was carried out incrementally. That is, the total
pressure was divided into several hundred steps, and at each step 10 psi of pressure was
applied to the tube until the total pressure of 3040 psi was reached. The size of the
incremental pressure loading was decided based on whether or not a converged solution
pte

was obtained. It was found that at early stages of the deformation it is better to take very
small pressure loading increments, but larger increments was taken once the tube became
fully plastic. Figure 7b shows the predicted deformed shape of the tube at 2000 psi and
3040 psi. The friction coefficient used in this simulation was 0.1. It can be seen that at
3040 psi, the predicted cross section of the tube slightly underestimates the actual shape
ce

of the tube.
The convergence difficulties occurring in this example using the implicit code
indicated the sensitivity of the mesh normal direction method. It is difficult to define the
Ac

tube normal accurately at each node during each increment while the accurate die normal
can be determined from the die geometry. Although it was shown that mesh normal
contact search algorithm was cost efficient and robust in the dynamic explicit FE code for
sheet forming simulation [39], it was recommended to use tool normal instead in order to
obtain better simulation accuracy [41].

7.1.3 Effect of Die Friction Coefficient

14

Page 16 of 36
The predicted hoop strain variation as a function of three different friction
coefficients are shown in Table 2, and compared with experimental data. In this work the
actual friction coefficient was not measured. Instead, the sensitivity of predicted hoop
strain as a function of friction coefficient, as it increased from 0.0 to 0.3, was studied with
the PSHD program. It was found that by increasing the friction coefficient the strain
magnitude decreases only where the tube contacts the die, and there is little or no effect
on the hoop strain where the tube is freely expanding. At μ = 0 the maximum hoop
strain occurs in the 0-10 degree zone and the magnitude of the strain matches the

t
experimental one.

p
7.1.4 Sensitivity of the Contact Solution to the Number of Segments Used

cri
A fictitious tube hydroforming problem was devised to numerically study the
sensitivity of the contact solution to the number of segments used to model the tube. In

us
this problem, a larger square die with a side length of 3.0” (76.2 mm) was used so that the
tube with a radius of 1.375” (35 mm) and a thickness of 0.1378” (3.5 mm) had to first
expand before making contact with the die. Since no experimental data was available, for
verification purposes this tube hydroforming example was also simulated with the
an
commercial code ABAQUS. In this numerical example, the aluminum tube (Table 1) was
hydroformed with a maximum pressure of 7000 psi. Beyond this pressure, the code could
not converge to a solution due to tensile instability. The goal of this exercise was to
dM
determine the minimum number of circular elements required in order to predict a
comparable deformed tube shape as ABAQUS.
Due to the symmetry of the die, only one half of the tube was modeled for the
hydroforming simulation, as shown in Figure 8a. The ABAQUS model used 26 8-node
shell elements with reduced integration (S8R5) for a more efficient simulation. In all
ABAQUS simulations the number of elements to be used was decided by trial and error.
pte

That is, tube hydroforming simulation was performed with a few elements and the
number of elements was then increased until no additional improvement (changes) in the
simulation results occurred. Figure 8a shows the predicted shape of the hydroformed
tube with 1 and 4 circular segments, using the current formulation. In the simulation, 4
Fourier series terms were used to approximate displacements (i.e., N=4), 8 Gauss
ce

integration points were used along the length of the element, and 3 integration points
were used through the thickness of the tube. A friction coefficient of 0.1 was also used
for the simulation. Figure 8a also shows the predicted shape of the hydroformed tube by
ABAQUS. When using only 1 segment, although very CPU-efficient, the PSHD model
Ac

was “stiff” and had difficulty capturing the expansion and the true size of the contact area
between the tube and the die. However, by using 4 segments the model, similar to
ABAQUS, was able to capture the deformation of the tube. Using more than 4 segments
only slightly changed the result, however, the CPU time increased dramatically.
Figure 8b shows predicted hoop strain distribution as a function of the internal
pressure by the new formulation. It could be seen that predicted strains are constant
throughout the tube as long as no contact occurs between the tube and the die. However,
as soon as a finite size contact region develops, e.g., at 7000 psi, strain distribution

15

Page 17 of 36
changes from being constant. In fact, strains are largest around the contact region, e.g.,
Z = 1.5 .

7.1.5 Effect of Pre-Bending on Tube Hydroforming

As mentioned in the Introduction section, it is common in tube hydroforming industry


to pre-bend the tube before hydroforming it. To study the effect of pre-bending, the

t
previous tube hydroforming example (7.1.4) was used with 4 segments, but this time the

p
tube was first bent to a maximum curvature of κ y = −0.04 / in before hydroforming it
with a maximum pressure of 7000 psi. Similar to previous examples, this combined

cri
bending/hydroforming problem was solved incrementally. That is, by applying a small
curvature at each step, the incremental tube-bending problem was solved until the total
curvature of κ y = −0.04 / in was reached. Then, by applying a small pressure the

us
incremental tube hydroforming process was solved, until the maximum pressure of 7000
psi was reached. As in the previous example (7.1.4), once pressure exceeded 7000 psi,
the code could not converge to a solution due to tensile instability.

an
Figure 9a shows the deformed shape of the tube after bending and hydroforming.
Compared to Figure 8a, the tube-die contact is asymmetric with respect to Z = 0 . Figure
9b shows the hoop strain distribution in the tube. Initially, strain distribution corresponds
to that of a bent tube, i.e., positive strains where 0 ≤ Z ≤ 1.5 , and negative strains
dM
where −1.5 ≤ Z < 0 . As the tube is further pressurized, the strain distribution continues
to grow, due to the superposition of a positive hoop strain. A comparison between
Figures 8b and 9b at the maximum pressure of 7000 psi shows that the maximum strain is
larger when the tube is bent first and then hydroformed, i.e., 26% vs. 22%. However, the
minimum strain is lower in the bent/hydroformed tube, i.e., 12% vs. 16%.
pte

8. Conclusions

A generalized plane strain assumption allows the user to specify a compressive load
or axial strain to each end of the tube to simulate axial feeding. Based on this formulation,
a tube hydroforming code (PSHD) was written and several examples of this process were
ce

investigated. Only a few circular segments and 4-6 Fourier series terms to approximate
displacement were required to model the cross section of the tube and accurately predict
the final deformed shape and strain distribution of the hydroformed tube. Numerical
codes such as the one described in this paper are useful engineering tools for quick and
Ac

efficient simulation of tube stretching, tube bending and tube hydroforming at early
stages of the process design.

Acknowledgements

The authors wish to thank the National Science Foundation for the partial support of
this project through the grant DMI 0084992 in conjunction with Alcoa through the
GOALI program.

16

Page 18 of 36
Appendix A

According to the total Lagrangian formulation discussed in Section 2 (Shell Element

Model), we derived the membrane strain, rotation of normal vector and the current

principal centerline curvature used in Section 3 (Kinematics Assumptions).

t
1. Membrane strain

p
For circular arc segment, the undeformed shell is non-flat and has an initial curvature K1,

cri
then rearranging Eq.(6) and approximating the square root , we obtain

1
λ1 ≈ 1 + [u ,2S + 2u , S + K 12 w 2 − 2 K 1 w − 2 K 1 wu , S + w,2S + K 12 u 2 + 2 K 1uw, S ] ,

us
2

then membrane strain is

e≈
1 2
2
1 an
( u ,S + w,2S ) + K 12 ( u 2 + w 2 ) + u ,S + K 1 ( uw,S − wu ,S − w )
2
(A1)

Consider the circular arc segment shown in Figure 3. Since Rdφ = dS , then
dM

∂ (⋅) 1 ∂ (⋅) ∂ (⋅)


= = K1 , Eq. (A1) becomes:
∂S R ∂φ ∂φ

1 2 2 1
pte

e≈ K 1 (u ,φ + w,2φ ) + K 12 (u 2 + w 2 ) + K 1u ,φ + K 1 ( K 1uw,φ − K 1 wu ,φ − w)
2 2

1 2 2
= K 1 [(u ,φ + w,2φ ) + (u 2 + w 2 ) + 2(uw,φ − wu ,φ )] + K 1 (u ,φ − w)
2
ce

2 2
⎛ u ,φ − w ⎞ 1 ⎛ u + w,φ ⎞ 1 ⎛ w − u ,φ ⎞
= ⎜⎜ ⎟+ ⎜
⎟ ⎜
⎟ + ⎜

⎟ (A2)
⎝ R ⎠ 2⎝ R ⎠ 2 ⎜⎝ R ⎟

Ac

As shown in Figure 3, if the positive w direction is in the opposite direction of the normal

to the mid-surface vector, n̂ , then Eq. (A2) becomes:


~

2 2
⎛ u ,φ + w ⎞ 1 ⎛ u − w,φ ⎞ 1 ⎛u + w⎞
e = ⎜⎜ ⎟⎟ + ⎜⎜ ⎟⎟ + ⎜⎜ ,φ ⎟ (A3)
⎝ R ⎠ 2⎝ R ⎠ 2 ⎝ R ⎟⎠

17

Page 19 of 36
which is used in Section 2 as Eq.(17) and Eq.(4.3) in Brush and Almroth [42].

Based on the following assumptions, we could derive the normal vector rotation and

current principal centerline curvature for both straight segment and circular arc segment.

Assumption 1: The rotation angle β is small,

p t
Assumption 2: Bending deformation is dominant and stretching is negligible, namely,

cri
ds
u , s ≈ 0, u , SS ≈ 0 or λ1 = ≈ 1.
dS

us
2. Rotation of normal vector

We assumed the angle between the current normal vector n̂ and the S (arc length) is
~

an
α, the angle between the current normal vector n̂ and normal vector N̂ at the reference
~ ~
dM
is β.

π
Then, we have β + = α , and nˆ ⋅ A
ˆ = cosα = -sinβ (A4)
2 ~ ~

Substituting Eq. (8) in Section 2 into Eq. (A4), we obtain:


pte

d w, S + K1u w, S + K1u
sin β = = = (A5)
λ1 λ1 (1 + u, S − K1w) 2 + ( w, S + K1u ) 2
ce

For circular arc segment ( K 1 ≠ 0) , according to assumptions 1 and 2, Eq. (A5) becomes:

1
β = K 1u + w, S Since dS = Rdφ and K 1 = , we obtain
Ac

u + w,φ
β = (A6)
R

For the coordinates shown in Figure 3, we have

18

Page 20 of 36
u − w,φ
β = (A7)
R

which is used in Section 2 as Eq. (13) and Eq. (4.4) in Brush and Almroth [42].

It should be noted that the rotation of a through-thickness line, β, was considered and
contributed to the hoop strain, however, there is no transverse strain considered in the
formulation.

tp
3. The current principal centerline curvature

cri
For circular arc segment, using Eq. (A5) and finite rotation assumption, we have

us
u − w,φ
sinβ = (A8)
R
an
According to Flugge [43], we have the change in curvature:
dM

β&
κφ = (A9)
R

By differentiating Eq. (A8), we obtain:


pte

dβ u − w,φφ
β& = = β ,φ = φ (A10)
dφ R cos β

where
ce

u − w,φ
cos β = 1 − sin 2 β = 1 − ( )2 (A11)
R
Ac

Substituting from Eqs. (A10), (A11) into Eq. (A9), we have

uφ − w,φφ
κφ = R (A12)
u − w,φ
R 1− ( )2
R

19

Page 21 of 36
Which is used in Section 3 as Eq. 18.

pt
cri
us
an
dM
pte
ce
Ac

20

Page 22 of 36
Appendix B

Here we will prove the equivalence of the two formulations in the pressure modeling, i.e.:

I ∂ ( Δu i ) I ∂ ( Δwi )
F
%
P
Ext =p ∑∫
i =1 Ai
nˆ i

∂c
% dAip ≈ − p ∑∫
i =1 Ai ∂c
dAip (B1)
~ ~

t
In Eq. (B1), the displacement vector Δu i has two components, Δu i and Δwi . For a point

p
%
at the mid-surface, i.e., z=0, of a circular segment, Eqs. (15a-b) will be:

cri
ξ i = Z c + ( R + wi ) sin φ + u i cos φ (B2.a)

us
η i = Yc + ( R + wi ) cos φ − u i sin φ (B2.b)

where φ = φ 0 + φ .

an
The unit outward normal to the surface A ip , n̂ i , is then defined as

∂ξ i ∂η i
dM
(− , )
∂φ ∂φ
nˆ i = (B3)
∂ξ i ∂η i
(− , )
∂φ ∂φ

where
pte

∂ξ i ∂wi ∂u i
− =− sin φ − ( R + w ) cos φ −
i
cos φ + u i sin φ (B4.a)
∂φ ∂φ ∂φ
ce

∂η i ∂wi ∂u i
= cos φ − ( R + wi ) sin φ − sin φ − u i cos φ (B4.b)
∂φ ∂φ ∂φ
Ac

1/ 2
∂ξ i ∂η i ⎧ ∂wi ∂u i 2 ⎫
(− , ) = ⎨( − u ) + [( R + w ) +
i 2 i
] ⎬ ≈ R + wi (B4.c)
∂φ ∂φ ⎩ ∂φ ∂φ ⎭

We also have,

∂ (Δu i ) ⎛ ∂ (Δη i ) ∂ (Δξ i ) ⎞


% =⎜ , ⎟ (B5)
∂c ⎜ ∂c ∂c ⎟
~ ⎝ ~ ~ ⎠

21

Page 23 of 36
where

Δξ i = Δwi sin φ + Δu i cos φ (B6.a)

Δη i = Δwi cos φ − Δu i sin φ (B6.b)

Combining (B3), (B4.a-c), (B5) and (B6.a-b), we have

t
⎛ ∂ (Δu i ) ∂u i ∂ (Δwi ) ∂wi ∂ (Δu i ) ⎞

p
∂ (Δu i ) 1 ∂ (Δwi )
nˆ i % =− ⎜ ( R + wi ) + ui + − ⎟
∂c R + wi ⎜ ∂c ∂c ∂φ ∂ c ∂φ ∂ c ⎟
⎝ ⎠

cri
~ ~ ~ ~ ~

∂ (Δwi )
≈−

us
(B7)
∂c
~

∂ 2 (Δw i )
Thus, we have proved that (B1) is correct. Since
an ∂2 c
~
= 0 , there will be no change in

the stiffness matrix and pressure will only show up on the right hand side of the Newton-
dM
Raphson expression, Eq. (37).
pte
ce
Ac

22

Page 24 of 36
References

[1] G. Morphy, Hydroforming high strength steel tube for automotive structural
applications using expansion, SAE International Congress & Exposition (1997).

[2] F. Dohmann, Ch. Hartl, Hydroforming: a method to manufacture light-weight parts, J.


Mater. Process. Tech. 60 (1996) 669-676.

t
[3] L. Wu, Y. Yu, Computer Simulations of Forming Automotive Structural Parts by

p
Hydroforming Process, 3rd International Conference: Numerical Simulation of 3D Sheet
Metal Forming Processes – Verification of Simulations with Experiments, Dearborn,

cri
Michigan (1996) 324-329.

[4] T.M. Srinivasan, J.R. Shaw, K. Thompson, Tubular hydroforming: correlation of


experimental and simulation results, SAE Technical Paper # 980448 (1998) 131-137.

us
[5] S.D. Liu, D. Mueleman, K. Thompson, Analytical and experimental examination of
tubular hydroforming limits, SAE Technical Paper # 980449 (1998) 139-150.

an
[6] S. Kaya, I. Gorospe, T. Altan, Preforming and expansion of an aluminum alloy in
tube hydroforming – comparison of FEA predictions with existing experimental data,
Transactions of NAMRI/SME, 30 (2002) 119-126.
dM
[7] J. Kim, B.S. Kang, H.H. Choi, Numerical analysis and design for tube hydroforming
process by rigid-plastic finite element method, Transactions of NAMRI/SME, 30 (2002)
127-134.

[8] Y.M. Hwang, T. Altan, Process fusion: tube hydroforming and crushing in a square
die, Proceedings of the Institution of Mechanical Engineers – Part B – Journal of
pte

Engineering Manufacture, 218 (2) (2004) 169-174.

[9] Y.M. Hwang, T. Altan, Finite element analysis of tube hydroforming processes in a
rectangular die, Finite Element in Analysis and Design, 39 (11) (2003) 1071-1082.
ce

[10] E. Corona, A simple analysis for bend-stretch forming of aluminum extrusions, Int. J.
Mech. Sci., 46 (2004) 433-448.
Ac

[11] A. Aydemir, J.H.P. de Vree, W.A.M. Brekelmans, M.G.D. Geers, W.H. Sillekens, R.
Werkhoven, An adaptive simulation approach designed for tube hydroforming processes,
J. Mater. Process. Tech., 159(3) (2005) 303-310.

[12] Y. Aue-U-Lan, G. Ngaile, T. Altan, Optimizing tube hydroforming using process


simulation and experimental verification, J. Mater. Process. Tech., 146 (1) (2004) 137-
143.

23

Page 25 of 36
[13] A. Kulkarni, P. Biswas, R. Narasimhan, , A.A. Luo, R.K. Mishra, T.B. Stoughton,
A.K. Sachdev, An experimental and numerical study of necking initiation in aluminium
alloy tubes during hydroforming, International Journal of Mechanical Sciences, 46(12)
(2004) 1727-1746.

[14] T. Papelnjak, Numerical analyses of tube hydroforming by high internal pressure,


Strojniski Vestnik-Journal of Mechanical Engineering, 50(1) (2004) 31-43.

[15] T. Hama, M. Asakawa, A. Makinouchi, Investigation of factors which cause

t
breakage during the hydroforming of an automotive part, J. Mater. Process. Tech., 150

p
(1-2) (2004) 10-17

cri
[16] T. Hama, M. Asakawa, H. Fukiharu, A. Makinouchi, Simulation of hammering
hydroforming by static explicit FEM, Iron and Steel Institute of Japan, 44(1) (2004) 123-
128.

us
[17] T. Hama, M. Asakawa, S. Fuchizawa, A. Makinouchi, Analysis of hydrostatic tube
bulging with cylindrical die using static explicit FEM, Materials Transactions, 44(5)
(2003) 940-945.
an
[18] F.C. Lin, C.T. Kwan, Application of abductive network and FEM to predict an
acceptable product on T-shape tube hydroforming process, Computers and Structures, 82
(15-16) (2004) 1189-1200.
dM

[19] M. Imaninejad, G. Subhash, A. Loukus, Influence of end-conditions during tube


hydroforming of aluminum extrusions, Int. J. Mech. Sci., 46(8) (2004) 1195-1212.

[20] P. Ray, B.J. Mac Donald, Determination of the optimal load path for tube
hydroforming processes using a fuzzy load control algorithm and finite element analysis,
pte

Finite Elements in Analysis and Design, 41(2) (2004) 173-192.

[21] K.I. Johnson, B.N. Nguyen, R.W. Davies, G.J. Grant, M.A. Khaleel, Numerical
process control method for circular-tube hydroforming prediction, Int. J. Plasticity, 20(6)
(2004) 1111-1137.
ce

[22] Q.C. Hsu, “Theoretical and experimental study on the hydroforming of bifurcation
tube”, J. Mater. Process. Tech., 142(2) (2003) 367-373.
Ac

[23] K.J. Fann, P.Y. Hsiao, Optimization of loading conditions for tube hydroforming, J.
Mater. Process. Tech., 140 Sp. Iss (2003) 520-524.

[24] J. Kim, B.S. Kang, Analytical approach to bursting failure prediction in tube
hydroforming based on plastic instability, Advances in Engineering Plasticity and its
Applications, Parts 1 and 2 – Key Engineering Materials, 274-276 (2004) 601-606.

24

Page 26 of 36
[25] J. Kim, B.S. Kang, S.M. Hwang, H.J. Park, Numerical prediction of bursting failure
in tube hydroforming by the FEM considering plastic anisotropy, J. Mater. Process. Tech.,
153-54 Special Issue Part 1 (2004) 544-549.

[26] J. Kim, S.J. Kang, B.S. Kang, A prediction of bursting failure in tube hydroforming
processes based on ductile fracture criterion, International Journal of Advanced
Manufacturing Technology, 22(5-6) (2003) 357-362.

[27] L.H. Lang, S.J. Yuan, X.S. Wang, Z.R. Wang, Z. Fu, J. Danckert, K.B. Nielsen, A

t
study on numerical simulation of hydroforming of aluminum alloy tube, J. Mater. Process.

p
Tech., 146 (3) (2004) 377-388.

cri
[28] J.W. Yoon, K. Chung, F. Pourboghrat, F. Barlat, Design optimization of extruded
preform for hydroforming processes based on ideal forming design theory, Int. J Mech.
Sci. 48 (2006) 1416-1428.

us
[29] Y. Guan, F. Pourboghrat, W.R. Yu, Fourier series-based finite element section
analysis of tube hydroforming - an axisymmetric model, Engineering Computations, 23(7)
(2006) 697-728.
an
[30] F. Pourboghrat, M.E. Karabin, R.C. Becker, K. Chung, A Hybrid Membrane/Shell
Method for Calculating Springback of Anisotropic Sheet Metals Undergoing
Axisymmetric Loading, Int. J. Plasticity, 16 (2000) 677-700.
dM

[31] J.O. Hallquist, G.L. Goudreau D.J. Benson, Sliding interfaces with contact-impact in
large-scale Lagrangian computations, Comput. Methods Appl. Mech. Eng. 51 (1985)
107-137.

[32] T. Belytschko, J.I. Lin, A Three-Dimensional Impact-Penetration Algorithm With


pte

Erosion, Comp. Struct. 25 (1987) 95-104.

[33] E Nakamachi, A finite element simulation of the sheet metal forming process, Int. J.
Numer. Meth. Engng. 25 (1988) 283-292.
ce

[34] Z.H. Zhong, L. Nilsson, A contact searching algorithm for general contact problems,
Comp. Struct. 33 (1989) 197-209.

[35] Y.T. Keum, E Nakamachi, R.H. Wagoner, J.K. Lee, Compatible description of tool
Ac

surfaces and FEM meshes for analysing sheet forming operations, Int. J. Numer. Meth.
Engng. 30 (1990) 1471–1502.

[36] D.J. Benson, J.O. Hallquist, A single surface contact algorithm for the post-buckling
analysis of shell structures, Comput. Methods Appl. Mech. Eng. 78 (1990) 141-163.

[37] T. Belytschko, M.O. Neal, Contact-impact by the pinball algorithm with penalty and
Lagrangian methods, Int. J. Numer. Meth. Engng. 31 (1991) 547-572.

25

Page 27 of 36
[38] M. Oldenburg , L. Nilsson, The position code algorithm for contact searching, Int. J.
Numer. Meth. Engng. 37 (1994) 359 – 386.

[39] S.P. Wang, E. Nakamachi, The inside-outside contact search algorithm for finite
element analysis, Int. J. Numer. Meth. Engng. 40 (1997) 3665 – 3685.

[40] H.D. Hibbitt, Some follower forces and load stiffness, Int. J. Num. Mech. Engng., 14
(1979) 937-941.

p t
[41] T. Hama, M. Takamura, C. Teodosiu, A. Makinouchi, H. Takuda, Effect of Tool

cri
Modeling Accuracy on Sheet Metal Forming Simulation. Key Engineering Materials
Vols. 340-341 (2007) 743-748.

[42] D.O. Brush, B.O. Almroth, Buckling of Bars, Plates, and Shells, McGraw-Hill, 1975.

us
[43] W. Flugge, Stress in Shells, 2nd Edition. Springer-Verlag. New York, 1973, pp. 362.

an
dM
pte
ce
Ac

26

Page 28 of 36
Figure

n a
z ds time t

∆w ∆u

t
Z

p
e2 w
N u

cri
A
dS time o
t
θ
e1

us
R
~
r
~

an X
dM
Figure 1. Shell mid-surface at reference time ot and current time t.
pte
ce
Ac

Page 29 of 36
Tube

X Y

p t
(a) Arc
Segment

cri
Bending X Straight
Y Segment
Surface

us
Z
an (b)
Figure 2. (a) The assumed global coordinate system for the tube, (b) The rectangular
dM
cross section of the tube is defined with 4 straight and 4 circular arc segments.

Y Yc X
pte

Zc
φο ∆φ
z φ R
ce

ξ u
w t
Ac

β η
Z

Figure 3. Geometry of a circular segment.

Page 30 of 36
n

Ak
Tube Segment

p t
Bi-1
Bi

cri
Die Segment
Bi+1

us
Figure 4. The associated die segment with contact node.
an
dM
n

η2 Tube Segment
pte

P η1 Q

B Die Segment
ce
Ac

Figure 5. The projection Algorithm.

Page 31 of 36
Tube Tooling Surface

p t
Node on
The Tube

cri
P0 P1 P1 P1

us
Symmetry Line

(a) (b)
an (c) (d)
dM
Figure 6. A schematic of tool-workpiece contact check; (a) shows the tube and the
tooling, (b) shows initial penetration of some of the contact nodes as internal pressure
increases from P0 to P1, (c) shows how those nodes are returned to the tooling surface,
and finally (d) shows how the equilibrium shape is reached after several iterations.
pte
ce
Ac

Page 32 of 36
p t
(a)

cri
us
Actual
cross section

an
dM
pte
ce

(b)
Ac

Figure 7. A 6111-T4 aluminum alloy tube was hydroformed with a maximum pressure of
3040 psi into a square die, (a) deformed tube and its cross section, and (b) predicted
intermediate tube shapes compared with the actual tube cross section.

Page 33 of 36
p t
cri
us
an
(a)
dM
pte
ce
Ac

(b)
Figure 8. A tube expanded into a symmetric square die,
(a) deformed tube shape, (b) strain distributions.

Page 34 of 36
p t
cri
us
an
(a)
dM
pte
ce
Ac

(b)
Figure 9. A tube bent and then expanded into a symmetric square die,
(a) deformed tube shape, (b) strain distributions.

Page 35 of 36
Table

Table 1. Material properties of aluminum 6061-T4 tube

Young’s Yield
Material Poisson K-value
Modulus Stress R-value N-value ε0
Type Ratio (psi)
(psi) (psi)
Aluminum
1.03E+7 0.33 18,730 0.82 69,183 0.2646 0.0
6061-T4

p t
Table 2. Predicted and measured strains at the maximum pressure of 3040 psi

cri
Axial
Hoop Strains
Strains
Friction 0~10 10~20 20~30 30~45
Coefficient degree degree degree degree

us
μ=0.0 6.5%~12.1% 6.7%~6.8% 6.9%~7.2% 6.7%~7.0% -3.25%
μ=0.1 6.4%~6.8% 6.8%~7.1% 6.8%~7.1% 6.8%~7.3% -3.34%
μ=0.3 6.3%~6.7% 6.7%~7.0% 6.7%~7.0% 6.7%~7.3% -3.25%
Experiment 7-11% 6-7%
an 6-7.5% 6-8% -3%
dM
pte
ce
Ac

Page 36 of 36

You might also like