You are on page 1of 9

Numerical Shape Optimization

in Industrial Glass Blowing


J. A. W. M. Groot
Department of Mathematics
and Computer Science,
Eindhoven University of Technology,
PO Box 513,
Eindhoven 5600, The Netherlands
e-mail: j.a.w.m.groot@tue.nl

C. G. Giannopapa
Department of Mathematics
and Computer Science,
Eindhoven University of Technology,
PO Box 513,
Eindhoven 5600, The Netherlands
e-mail: c.g.giannopapa@tue.nl

R. M. M. Mattheij
Department of Mathematics
and Computer Science,
Eindhoven University of Technology,
PO Box 513,
Eindhoven 5600, The Netherlands
e-mail: r.m.m.mattheij@tue.nl

Industrial glass blowing is an essential stage of manufacturing hollow glass containers,


e.g., bottles, jars. A glass preform is brought into a mold and inflated with compressed
air until it reaches the mold shape. A simulation model for blowing glass containers
based on finite element methods, which adopts a level set method to track the glassair
interfaces, has previously been developed [Giannopapa and Groot, 2007, A Computer
Simulation Model for the BlowBlow Forming Process of Glass Containers, Paper No.
PVP2007-26408, pp. 7986; Giannopapa, C. G., 2008, Development of a Computer
Simulation Model for Blowing Glass Containers, ASME J. Manuf. Sci. Eng., 130(4), p.
041003; Giannopapa and Groot, 2011, Modeling the BlowBlow Forming Process in
Glass Container Manufacturing: A Comparison Between Computations and
Experiments, ASME J. Fluids Eng., 133(2), p. 021103]. A considerable challenge in
glass blowing is the inverse problem: to determine an optimal preform from the desired
container shape. In previous work of the authors [Groot et al., 2009, Numerical Optimisation of Blowing Glass Parison Shapes, ASME Paper No. PVP2009-77946; Groot
et al., 2011, Development of a Numerical Optimization Method for Blowing Glass Parison Shapes, ASME J. Manuf. Sci. Eng., 133(1), p. 011010] a numerical method was
introduced for optimizing the shape of the preform. The optimization method described
the shape of the preform by parametric curves, e.g., Bezier-curves or splines, and
employed a modified LevenbergMarquardt algorithm to find the optimal positions of the
control points of the curves. A combined finite difference and Broyden method was used
to compute the Jacobian of the residual with respect to changes in the positions of the
control points. The objective of this paper is to perform an error analysis of the optimization method previously introduced and to improve its accuracy and performance. The
improved optimization method is applied to modeled containers of industrial relevance,
which shows its usefulness for practical applications.
[DOI: 10.1115/1.4028066]
Keywords: shape optimization, inverse problems, glass blowing, level set methods,
numerical simulation

Introduction

Glass blowing is a manufacturing process for the production of


hollow, glass containers, such as bottles and jars. In a glass blow
process a molten material is brought into a mold and inflated with
pressurized air to force it in the mold shape. First a so-called glass
parison or preform is constructed, usually by either a press stage
or a blow stage. The parison is inverted by means of a robotic arm
and two mold halves are closed around the parison just below the
neck. In the mold the parison is first left to sag due to gravity for a
short period. Then pressurized air is blown in the mold by means
of a blow head to force the glass in a mold shape. Finally, the
mold shape is left in the mold to cool down before it is ejected
from the mold. Figure 1 illustrates the process.
Glass blow processes take place at high production rates. At the
same time quality factors of the products, such as smoothness,
strength, weight, and cooling conditions, should be optimal. To
optimize and control the process, thorough knowledge is required.
Unfortunately, measurements are often complicated, considering
that glass blowing happens fast, in closed constructions and under
complicated circumstances, such as high temperatures. Furthermore, trial-and-error experiments with glass blowing equipment
are usually expensive and time consuming. Numerical process
simulation offers a good alternative.
Contributed by the Pressure Vessel and Piping Division of ASME for publication
in the JOURNAL OF PRESSURE VESSEL TECHNOLOGY. Manuscript received May 16, 2013;
final manuscript received July 20, 2014; published online September 4, 2014. Assoc.
Editor: Haofeng Chen.

Journal of Pressure Vessel Technology

Over the last few decades, mathematical models for process


simulation of glass blowing have become increasingly important
in understanding, controlling, and optimizing the process. A representative numerical simulation should give as output the containers final shape and wall thickness as well as the stress and
thermal deformations the glass and equipment undergo during the
process. The growing interest of glass industry for process simulation has been a motivation for a fair number of manuscripts on
this subject [111].

Fig. 1 Schematic drawing of glass blowing

C 2014 by ASME
Copyright V

DECEMBER 2014, Vol. 136 / 061301-1

Downloaded From: http://pressurevesseltech.asmedigitalcollection.asme.org/ on 11/13/2015 Terms of Use: http://www.asme.org/about-asme/terms-of-use

A key issue in glass blowing is the wall thickness distribution


of the final container. In practice often a container with a certain
wall thickness distribution is desired and the corresponding shape
of the preform is sought for in order to produce a container with
exactly this thickness distribution. Therefore, it is interesting to
consider the inverse problem, that is, to determine the shape of the
preform, given the desired thickness distribution of the final container. This class of inverse problems is quite challenging, as in
general coupled, highly nonlinear physical systems and complicated geometries are involved. Typically, numerical optimization
methods are employed to find solutions to inverse problems.
Over the last few decades, numerical optimization has become
increasingly popular in glass blowing. Early attempts to optimize
the container wall thickness distribution date from the early nineties and were first applied to the blow molding of plastic containers. In Ref. [12] a neural network approach to find the preform
thickness, the mold geometry and a representative rheological
parameter corresponding to the optimal wall thickness distribution
was presented. In the same year a combined NewtonRaphson
and profiled optimization routine to predict the parison thickness
distribution required for a specified wall thickness distribution
was presented [13]. Few years after an attempt was made to optimize the wall thickness distribution in stretch blow molding of
plastic bottles [14]. The authors used an optimization method
based on a method of feasible directions that attempts to find the
optimal thickness of the preform. Several other optimization
methods have been considered in literature to optimize the wall
thickness distribution in stretch blow molding [15] and extrusion
blow molding [1618] of polymers in three dimensions.
An engineering approach to find the optimal parison shape for
glass blowing was presented in Refs. [19,20]. The authors
combined a computer aided design model, a 3D thermomechanical finite element model with adaptive mesh techniques and an
optimization technique based on the LevenbergMarquardt
method. The algorithm attempted to optimize the geometry of the
mold for the preform blow stage, given the wall thickness distribution at the end of the final blow stage. Optimization methods
have also been used to estimate the heat transfer coefficient or the
initial temperature distribution in glass blowing [21,22].
A numerical optimization method to find an optimal preform
from the desired container shape has been previously developed
[23,24]. The method uses a modified LevenbergMarquardt algorithm combined with a secant method. The 2D axial-symmetrical
simulation model presented in Refs. [1,25] is applied to compute
the shape of the glass container. In Ref. [23] an initial guess of the
preform shape was constructed by means of an analytical approximation of the optimal shape. The optimization method was
applied to the blowing of a simple ellipsoidal glass container.
The objective of this paper is to perform an error analysis of
the optimization method previously introduced and to improve its
accuracy performance. The improved optimization method is
applied to a modeled container of industrial relevance.

Problem Formulation

The following problems are considered:


The forward problem. Find the shape of the container, given the
shape of the preform. The forward problem represents the physical process of blowing the preform into the mold shape.
The inverse problem. Find the shape of the preform, given the
shape of the container. Usually, a container with a certain shape is
desired, but in order to form this container first the preform needs
to be known.
In this section a mathematical formulation of both problems is
given. Let C1,0, C2,0, Ci, and Cm be the inner and outer preform
surfaces, the inner container surface, and the mold surface, respectively, as depicted in Fig. 2 for a 2D axial-symmetrical jar. The
dotted lines represent the unknown surfaces. Since the mold
surface is fixed, also one of the preform surfaces is fixed, thus
mapping one surface onto another.
061301-2 / Vol. 136, DECEMBER 2014

Fig. 2 Glass blowing problem description

The forward problem is formulated as follows: find the location


of inner glass surfaces C1 at t t*, such that C2(t*) Cm and
C2(t) 6 Cm for t < t*, given C1(0) C1,0 and C2(0) C2,0. In the
mathematical model used for the forward problem, the glass melt
is modeled as an incompressible, Newtonian fluid. Since viscous
forces dominate, the flow of glass can be described by a Stokes
flow problem

r  lru  rp qg 0
(1a)
ru0
Because of the strong temperature dependency of the glass
viscosity, the flow problem is coupled to an energy problem
involving the convectiondiffusion equation


@T
u  rT  r  kc rT 0
(1b)
qcp
@t
Glass and air are distinguished by the sign of a level set function h. The level set function is a solution of the evolution
equation
@h
u  rh 0
@t

(1c)

So, the glassair interfaces at time t are implicitly given by


h(x, t) 0 [2629].
Together Eq. (1) forms a fully coupled system of partial differential equations that describes the glass flow during a blow process. Boundary conditions include:

an inflow pressure at the mold entrance


free-stress conditions for air and no-slip conditions for glass
on the mold wall
2D axial symmetry on the axis

Further details and analysis regarding the problem formulation


of the glass blow simulation model are given in Refs. [1,2,25,30].
The inverse problem is to find the location of the initial glass
surface C2,0 C2(0), given the surface C2 at time t t*, such that
C2(t*) Cm and C2(t) 6 Cm for t < t*, and given C1(t*) Ci and
C1(0) C1,0.

Simulation Model

A Galerkin finite element method is employed for the discretization of the forward problem. The 2D axial-symmetrical finite
element model has been implemented in COMSOL 3.5 with MATLAB.
For the spatial discretization of the flow problem Mini-elements
are used [31]. For the spatial discretization of the level set and
energy problem linear Lagrange elements are used. For the temporal discretization an implicit differential-algebraic solver scheme
has been used [32], which employs a variable-order variable-stepsize backward difference formula method. The discretized system
Transactions of the ASME

Downloaded From: http://pressurevesseltech.asmedigitalcollection.asme.org/ on 11/13/2015 Terms of Use: http://www.asme.org/about-asme/terms-of-use

of equations is solved using the BiCGstab method with a geometric multigrid method as a preconditioner [33]. A simulation stops
if either the mold wall is covered with glass or an upper limit of
the process time is reached.
A level set method is used to capture the glassair interfaces.
This approach allows for the computation of two-fluid flow solutions with moving interfaces while using a fixed finite element
mesh. Moreover, the level set function naturally deals with topological changes. A triangulated fast marching method is used as a
re-initialization algorithm to maintain the level set function as a
signed distance function [26,34]. Further details about the level
set method and fast marching algorithm used can be found in
Refs. [1,25].
Air is replaced by a fictitious fluid with a much larger viscosity
in the simulations, so that the flow of air can also be described by
the Stokes flow equations. The viscosity of the fictitious fluid
should still be much smaller than the viscosity of glass, so that the
pressure drop in the air domain is negligible compared to the pressure drop in the glass domain [1,35]. For further details the reader
is referred to Refs. [1,2,25,30].
The accuracy of the glass blow simulation model has been
successfully validated by the authors by means of verification of
conservation of volume [1,25] and comparison of the container
thickness with experimental data provided by industry [2].

Shape Optimization Strategy

The inverse problem can be formulated as an optimization


problem: find the shape of the preform, such that the difference
between the inner surfaces of the blow molded container and the
designed container is minimal. Figure 3 illustrates the optimization problem. In the optimization problem, Ci is known as in the
inverse problem, but C1(t*)  C1(t*;C0) is considered a function of
the unknown preform surface C0, not necessarily equal to Ci. If
inner preform surface C1,0 is known, then outer preform surface
C2,0 should be such that the difference between C1(t*;C2,0) and Ci
is minimal, where t* is such that C2(t*) Cm and C2(t) 6 Cm for
t < t*. If the inverse problem has a solution C2,0, then this difference is zero, i.e., C1(t*;C2,0) Ci.
4.1 Parametrization of the Unknown Preform Surface.
Subject to optimization is the location of the outer preform surface
C2,0. The optimization algorithm presented in Refs. [23,24]

Fig. 4 Parametrization of the unknown preform surface by a


cubic spline with six control points

describes the axial-symmetrical surface by a parametric curve,


e.g., a spline or Bezier curve. The optimization method adjusts the
control points in radial direction; the azimuthal coordinates are
fixed (see Fig. 4). Furthermore, the topmost control point is fixed
on the mold wall. Thus, for m variable control points P1,, Pm
the vector p comprising the parameters to be optimized, in terms
of 2D axial-symmetrical spherical coordinates, is defined by
Ref. [23]
p RP1 ; ; RPm T

(2)

Figure 4 illustrates the parameter directions for a cubic spline


with m 5. The control point P0 is fixed.
4.2 Objective Function. The optimization method minimizes
a scalar objective function that represents the difference between
the surfaces C1(t*) and Ci. A suitable choice of the objective function is the average distance from Ci to C1(t*) with respect to the
L2-norm
1
Up :
2

Ci

h2i;;p dC

1
0



h2i;;p xi sx0i sds

(3)

where xi: [0, 1] ! Ci is the parametrization of Ci. The subscript p


indicates that the level set function is evaluated for parameter vector p. Note that the level set function h1,*(Ci) is kept as the signed
distance function from Ci to C1(t*) by means of re-initialization.
The integral in Eq. (3) is computed by applying a composite
nG-point Gaussian quadrature rule
Up
_

Fig. 3 Difference between designed and computed container

Journal of Pressure Vessel Technology

nG
nint X




1X
xk x0 sjk h2i;;p xsjk
2 j k

(4)

Here fxk gk1;;nG are the weights of the quadrature rule,


fsjk gk1;;nG are the points of the quadrature rule in the jth subinterval of [0, 1] for j 1,, nint, and nint is the number of subintervals. For convenience the objective function is presented as
DECEMBER 2014, Vol. 136 / 061301-3

Downloaded From: http://pressurevesseltech.asmedigitalcollection.asme.org/ on 11/13/2015 Terms of Use: http://www.asme.org/about-asme/terms-of-use

U
_

1 T
r r
2

(5)

where the residual vector r(p) comprises the nres nintnG


elements
q


(6)
ri p : rj1nG 1k p xk jx0 sjk jhi;;p xsjk
The components of the residual vector represent
the distances
p
from Ci to C1(t*) in points x(sjk) with weights xk jx0 sjk j.

Constraints

Geometric constraints are imposed on the parameters to ensure


that the preform satisfies reasonable physical conditions. They
ensure that the control points of the outer surface are located
within an area that is contained within the mold at a reasonable
distance from the mold boundary and the inner surface. The constraints are given by
cl p > 0;

l 1; nc

(7)

where nc is the number of constraints. The constraints are incorporated in the optimization method by introducing a weighted objective function, given by
~
Up
Up fp

nc
X
wl
;
c
l p
l

(9)

with non-negative weights wl [19,3638].


5.1 Algorithm. A modified LevenbergMarquardt method
is used to solve the nonlinear least squares problem. The
LevenbergMarquardt method is modified in the sense that it
accounts for the geometric constraints on the preform. In the ith
iteration of the modified LevenbergMarquardt method, the following system of equations is solved:

i
(10)
Ji TJi ki I Hf si Ji Tri rp fi
where J is the nres  m Jacobian matrix of r, Hf is the m  m
Hessian matrix of f, and rpf is the gradient of f, all with respect
to p. The solution s(i) is a parameter increment, which is used to
adjust the control points
pi1 pi si

5.2 Computing Derivatives. The Jacobian matrix J is not


readily available, but has to be approximated each iteration. A forward difference approximation of J requires m solutions of the
forward problem with different parameter vectors, which can cost
a considerable amount of computational time. An alternative is a
secant method, which usually employs a Broyden update of the
Jacobian [3941].
Derivative-free optimization is another possibility. Derivativefree optimization algorithms do not require approximations of the
derivatives of the residual vector, but usually need more iterations. Derivative-free optimization can be recommended if m is
large, which is usually not the case for the applications in this
paper.
5.2.1 Finite Differences. The Jacobian J of a residual vector r
can be computed by means of the forward differences formula
Jpe

(13)

where e is a unit basis vector in Rm . The forward difference formula requires the computation of m additional function values in
p de for each basis vector e.
The error in the finite difference approximation directly influences the error of the optimization algorithm. Expression (13) suggests that the step size d is chosen as small as possible. However,
also the error in computing the residual r should be taken into
account. Let this error be denoted by eF. Then the sum of errors is
bounded by
2eF 1
Md
2
d

r
p
4
eF  eF
d
M

(15)

To improve the order of accuracy, a central difference formula


can be employed to compute the Jacobian matrix

(11)

The parameter k in Eq. (10) denotes the non-negative Levenberg


Marquardt parameter. It acts as a regularization parameter in case
the Jacobian matrix J(i) is nearly singular and interpolates between
a steepest descent and a GaussNewton iteration. If
~ i , the LevenbergMarquardt parameter k(i) is
~ i1 > Up
Up
increased and Eq. (10) is solved again.
The algorithm stops if
n
o

 i1 < g or si < g
(12)
max si ; Up
ki

(14)

where the error in computing the residual is bounded by eF and M


is an upperbound for Hrj pk ; j 1; ; nres . The upperbound for
the sum of errors is minimal for

Jpe

(i)

rp de  rp
Od
d

(8)

Here function f is the sum of the weighted penalty functions


fp

for some error tolerance g, which is yet to be determined. The


search step s(i) in the second stopping criterium is scaled with k(i)
to avoid that the algorithm stopswhen the LevenbergMarquardt
 pi1 is not small. If Up
 i1
parameter k(i) increases while U
i
< g the algorithm is proceeded until also s < g, in which
case the algorithm decides that convergence has been reached.

 
rp de  rp  de
O d2
2d

(16)

A drawback is that the central difference formula is almost twice


as expensive, since 2 m additional function values are required
instead of m function values for the forward difference formula.
The sum of errors is bounded by
eF 1 2
Ld
d 6

(17)

where L is an upperbound for the third order derivatives of the


residual vector. The upperbound for the total error is minimal for

where
~
Up

Up
:
;
jCi j

jCi j :

nG
nint X
X
j

xl jx0 sjl jxsjl

061301-4 / Vol. 136, DECEMBER 2014


d

6
eF
L

1=3

1=3

 eF

(18)

Transactions of the ASME

Downloaded From: http://pressurevesseltech.asmedigitalcollection.asme.org/ on 11/13/2015 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Since the error in the central difference formula is O(d2), the error
2=3
in the solution of the inverse problem is at least OeF , which is
not a considerable improvement considering the fact that the
method is twice as expensive. Still the central difference formula
can be used when the optimal accuracy is reached to gain a few
extra digits in the last few iterations.
5.2.2 Broyden Update. In Refs. [23,24] a secant method using
a Broyden update of the Jacobian matrix is applied [3941]. The
initial Jacobian matrix is approximated by a forward difference
method. A forward difference method is also used if no convergence is reached after several iterations with the secant method.
The Broyden update of the Jacobian matrix J(p(i1)) for iteration i 1 of the optimization method is
Ai1 Ai

i i

Measurement errors eM. Data provided by manufacturers


contains measurements errors. Since this is often the desired accuracy, it is assumed that the numerical accuracy of the optimization
method is of the order of magnitude O(eM). In practice, the measurement error can be smaller if the numerical accuracy is limited
by the computational effort.
Interpolation errors eI. The interpolation of the unknown preform surface by a parametric curve gives an interpolation error.
The error in the forward problem eF is not mentioned in this
list, as it is the result of the model error eC and the measurement
error eM. Neglecting the influence of rounding errors, the total
error of the optimization method is the sum of the contributions of
all errors
e eC eT eM eI

A s s T
si Tsi

(19)

where yi ri1  ri . Assume that the Jacobian matrix is Lipschitz continuous with Lipschitz constant c
i1


Jp
 Jpi  c si

(20)

In the following a quantitative analysis of the errors is performed. The analysis is based on the following assumptions:

In Ref. [41] the following upperbound for the error of the Broyden
update is given:
i1

3
A
 Jpi1  Ai  Jpi c si
2

(21)

Suppose J(p(0)) is approximated by forward p


difference
formula

(13). Then the initial error is O(d), where d  eF . Since the forward problem is not sensitive to changes in the parameters, c is
small. Assuming that r is twice differentiable, it holds that


Hj p  M
(22)
c
max
m

The model error eC is the sum of model simplifications and


discretization of the forward problem. The contribution of
model simplifications can be omitted, as it is irrelevant for
the choice of the error tolerance.
The time step is small compared to the mesh size with respect
to the order of accuracy of the discretization schemes. Then
the model error eC is O(h2) for linear finite elements, where h
is the typical mesh size. Re-initialization of the level set function preserves this property.
Given eM, the mesh size h is chosen such that h2 & eM.
A cubic spline is used for the interpolation of the unknown
interface, hence the interpolation error is O(n4), where n is
the largest distance between the control points. The number
of parameters is chosen such that n4 . h2. Thus, if is the
length of the axial symmetrical cross section of the surface,
then n  CI =m  1 and
1

m & CI h2 1

p2R ; j1;;n

where
i Hj is the Hessian matrix of rj. The order of magnitude of
s is not smaller than O(d), so the error of the Broyden update
is at least O(d) and can have larger order of magnitude far away
from the optimum.

5.3 Error Tolerance. In the optimization method there are


also other sources of errors than the numerical solution of the forward problem and approximation of the Jacobian matrix. In order
to estimate the total error of the optimization method, all sources
of errors should be taken into account. Finally, the error tolerances
g in Eq. (12) should be of the same order as the total error. The
following errors are considered.
Model errors eC The computation of the residual r is subject to
errors. The error between the computed residual and the exact
residual is represented by the model error eC.
Truncation errors eT. The error in the approximation of the
Jacobian by finite differences with the true residual is referred to
as the truncation error eT. Also the error in the Jacobian update
can be viewed as a truncation error and is denoted as eT. The composite quadrature rule in Eq. (4) also induces a truncation error.
The order of magnitude of the truncation error of the composite
nG-point Gaussian quadrature rule with nint subintervals of [0, 1]
is O1=nint 2nG . Clearly, not many interpolation points are
required for the composite quadrature rule to match the accuracy
of the numerical differentiation method that approximates the
1=4
Jacobian. For example, if nG 2, then nint & OeT is sufficient.
Thus the truncation error in Eq. (4) is negligible.
Rounding errors eR. Performing arithmetic operations in
machine precision results in a rounding error eR. In general,
rounding errors are negligible compared to model and truncation
errors.
Journal of Pressure Vessel Technology

(23)

(24)

for some constant CI that is independent of h and n. Since the


fourth order derivative of the surface along the cross section
is typically small, it is expected that CI 1. Consequently,
only few control points are required to meet the required
accuracy.
The forward difference method combined with the Broyden
method is used for the approximation of the Jacobian. The
truncation error eT is O(d). Furthermore, the optimal choice
for the forward difference scheme is used, i.e., eF CF h2
Ohp ; p > 2, so that the minimal upperbound (15) for d is
r
CF
d2
h
(25)
M
Unfortunately, constants M and CF are difficult to estimate
without any additional computational effort, so the precise
magnitude is unknown. Constant M can be estimated by
approximating the diagonal elements of Hrj , e.g., by means
of the BroydenFletcherGoldfarbShanno (BFGS) method.
Constant CF is independent of the mesh size, but it does
depend on the mesh quality, the geometry and the problem. It
can be estimated by Richardson extrapolation with different
mesh sizes. On the other hand, if the problem is well scaled,
it can be assumed that CF/M  1 [42], and the choice d 2 h
is fairly close to optimal.

Based on the foregoing assumptions, the errors are estimated by


p
eF  eC  h2 & eM eI ; eT  eF  h
(26)
As a result
e Oh

(27)

DECEMBER 2014, Vol. 136 / 061301-5

Downloaded From: http://pressurevesseltech.asmedigitalcollection.asme.org/ on 11/13/2015 Terms of Use: http://www.asme.org/about-asme/terms-of-use

that is, the error of the optimization method is dominated by the


mesh size. The error tolerance is chosen in the order of magnitude
of the total error, g 2 h.
From the error analysis it can also be concluded that the Broyden update does in general not provide sufficient accuracy for
minimization. Therefore, as a rule forward difference formula
(13) is used to approximate the Jacobian matrix. The Jacobian can
be updated by a Broyden update, provided:
(1) the order of magnitude of ksi k is not larger than O(h),
since in this case by Eq. (21) the Broyden update may be
insufficiently accurate
(2) the order of magnitude of ksi k is not smaller than O(h),
which could mean that the accuracy of the approximate
Jacobian matrix is not sufficient for convergence or the
parameter vector is near the optimum, hence a more accurate approximation is required
 pi1 is larger than O(h), since
(3) the order of magnitude U
optimal accuracy is required near of the optimum
(4) k(i) is not large compared to kJT pi1 Jpi1 k, since in
that case s(i) does not approximate a Newton direction,
which is a requirement for a Broyden update
5.4 Initial Guess. Convergence of iterative optimization
methods is best near the optimum, whereas far away from the
optimum they are often not reliable. This is particularly the case
for nonlinear problems such as in glass blowing. Therefore, finding a suitable initial guess is essential to the success of the optimization method.
In Ref. [23] an initial guess of the preform shape was constructed by means of an analytical approximation of the optimal
shape. The analytical approximation is based on the assumption
that the polar velocity component is zero: uu 0. By means of
this assumption and conservation of volume, the following
approximation of the spherical radius R2,0 of the outer preform

Fig. 5 Typical structured mesh for glass bottle

061301-6 / Vol. 136, DECEMBER 2014

surface as a function of the polar angle u is obtained in terms of


spherical radii Rm, Ri, and R1,0 corresponding to surfaces Cm, Ci,
and C1,0, respectively [23]

13
R2;0 u R3m u  R3i u R31;0 u ; 0  u  u0 (28)
Here u 0 corresponds to the symmetry axis and u u0 corresponds to the mold entrance.
The next step in the construction of the initial guess is the
choice of the angles of the control points. A suitable choice is an
equidistant distribution of the control points along the curve. Consider only the case in which the control points are on the parametric curve, e.g., a natural spline. The distance between the control
points is expressed in terms of the length of the curve
u0 q
R2 u2 R02 u2 du
(29)

The angles of the control points are determined by solving the following equations by an adaptive Simpson rule:
uj q

R2 u2 R02 u2 du
(30)
m
uj1
where uj is the angle of Pj, j 1,,m.

Fig. 6

Designed bottle with optimal wall thickness distribution

Transactions of the ASME

Downloaded From: http://pressurevesseltech.asmedigitalcollection.asme.org/ on 11/13/2015 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Fig. 7 Initial guess of control points

Fig. 8

Results

This section shows numerical results for optimizing the preform


shape for blowing an axial-symmetrical glass bottle with desired
wall thickness. A beer bottle with optimal wall thickness distribution is designed and one is interested in blowing this bottle. The
dimensional height of the bottle is 20.1 cm and the dimensional
radius is 1.25 cm at the neck to 2.89 cm at the widest part (see
Fig. 5). Figure 6 shows the designed bottle. Two cases are considered: blowing with sagging for 0.5 s and blowing without sagging.
The bottle is produced by blowing air in the mold with an inlet
pressure of 10 kPa.
Let the origin of the spherical coordinate system be given at
xO (0, zO), with zO 5.58 cm (see Fig. 7). A cubic spline with
six control points is used for the parametrization of the outer
preform surface. An initial guess R2,0 for the iterative optimization
method with respect to origin xO is determined by Eq. (28).
Figure 7 shows the initial guess of the control points. The positions of the control points are subject to thickness constraints
RPj  R0;1 uj > 0:2;

j 1; ; m

(31)

Furthermore, the control points should be located within an area


within the mold. Therefore, they should satisfy the domain
constraints
rm jz0 rPj > 0:5;
zm jr0 rPj > 0:5;

j 1; ; m  1

(32)

j 2; ; m

(33)

Figure 8 illustrates the constrained domain.


Figure 9 shows the blown bottles corresponding to the initial
guess computed with and without sagging. Figure 9(a) shows the
Journal of Pressure Vessel Technology

Constrained domain of control points

mold shape computed with sagging for 0.5 s and Fig. 9(b) shows
the mold shape computed without sagging. Obviously, the bottle
computed without sagging is much closer to the designed bottle in
Fig. 6 than the bottle computed with sagging. This could be
expected as during sagging the mass flows in negative axial direction. Consequently, the mass flow in polar direction during sagging in clockwise polar direction is significant. On the other hand,
in the initial guess of the preform shape the mass flow in polar
direction is completely omitted. In contrast the approximation
without sagging is quite satisfactory and suitable to be used as an
initial guess.
The optimization method is used to find the optimal positions
of the control points of the outer preform surface as to minimize
the difference between the inner surfaces of the computed container without sagging and the designed container in the sense of
Eq. (3). Consider a finite element mesh with 40 elements along
the mold opening by 272 elements along the symmetry axis and a
mesh distribution as in Fig. 5. The LevenbergMarquardt method
with combined forward difference method and secant method is
used to find the optimal preform. The typical mesh size for this
example is h 0.059. Figure 10 shows the convergence results.
From the second iteration the Broyden method is used to update
the Jacobian matrix. However, in the sixth iteration no convergence is reached, so the Jacobian matrix is approximated by finite
differences. In the seventh iteration another Broyden update is
performed, but from the eighth iteration onward only forward difference approximations are used, since the parameter vector is
close to the optimum. Only 59 function evaluations are required
until convergence. The central processing unit (CPU) time for one
simulation is approximately 5.4  103 (1.5 h) and the total CPU
time for the optimization procedure is approximately 3.2  105.
The order of magnitude of the error of the optimization method
DECEMBER 2014, Vol. 136 / 061301-7

Downloaded From: http://pressurevesseltech.asmedigitalcollection.asme.org/ on 11/13/2015 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Fig. 11 Optimal preform and mold shape for beer bottle. (a)
Optimal preform and (b) optimal mold shape.
Fig. 9 Mold shapes for initial guess. (a) Computed with sagging and (b) computed without sagging.

result is encouraging and gives confidence that the method can be


used in practice.

Fig. 10 Convergence of objective function

(27) can be verified from the convergence results. With the current
error tolerance, it can be observed that convergence is slowing
down in the last few iterations. If the tolerance is set stricter then
O(h), the optimization method will converge extremely slow or
not at all. This phenomenon has been observed for different mesh
sizes and confirms the theoretical error estimates.
Figure 11 shows the result of the optimization method.
Figure 11(a) shows the optimal shape of the preform and
Fig. 11(b) shows the resulting container. The resulting container
in is in good agreement with the designed container in Fig. 6.
The mean distance between the inner container surfaces is
7.83  102 cm and the maximum distance is 1.78  101 cm. The
061301-8 / Vol. 136, DECEMBER 2014

Conclusions

An optimization method was previously introduced in


Ref. [24]. The optimization method approximates the unknown
preform surfaces by parametric curves and uses a modified
LevenbergMarquardt algorithm to find the optimal positions of
the control points of the curves. In Ref. [23] the inverse problem
was reformulated, the optimization method was improved and an
analytical approximation of the solution of the inverse problem
was derived that was used as an initial guess for the iterative optimization algorithm. In Ref. [23], the optimization method was
applied to the blowing of a simple ellipsoidal glass container.
In this paper an error analysis was applied to determine the
optimal error tolerance for the stopping criterium and to establish
an efficient strategy to combine the finite difference method with
the secant method to compute the Jacobian matrix of the residual
vector. The improved optimization method is applied to the blowing of a 2D-axial-symmetrical beer bottle, which shows its efficiency for practical applications. In addition, it shows that the
initial guess is useful for real applications, although the analytical
approximation still requires improvement to account for sagging.

References
[1] Giannopapa, C. G., 2008, Development of a Computer Simulation Model for
Blowing Glass Containers, ASME J. Manuf. Sci. Eng., 130(4), p. 041003.

Transactions of the ASME

Downloaded From: http://pressurevesseltech.asmedigitalcollection.asme.org/ on 11/13/2015 Terms of Use: http://www.asme.org/about-asme/terms-of-use

[2] Giannopapa, C. G., and Groot, J. A. W. M., 2011, Modeling the BlowBlow
Forming Process in Glass Container Manufacturing: A Comparison Between
Computations and Experiments, ASME J. Fluids Eng., 133(2), p. 021103.
[3] Cesar de Sa, J. M. A., 1986, Numerical Modelling of Incompressible Problems
in Glass Forming and Rubber Technology, Ph.D. thesis, University College of
Swansea, Swansea, UK.
[4] Cesar de Sa, J. M. A., 1986, Numerical Modelling of Glass Forming Processes, Eng. Comput., 3(4), pp. 266275.
[5] Cormeau, A., Cormeau, I., and Roose, J., 1984, Numerical Simulation of
Glass-Blowing, Numerical Analysis of Forming Processes, J. F. T. Pittman,
O. C. Zienkiewicz, R. D. Wood, and J. M. Alexander, eds. Wiley, New York,
pp. 219237.
[6] Williams, J. H., Owen, D. R. J., and Cesar de Sa, J. M. A., 1986, The Numerical Modelling of Glass Forming Processes, Collected Papers of the XIV International Congress on Glass, H. C. Bharwaj, ed., Indian Ceramic Society, New
Delhi, India, pp. 138145.
[7] DeLorenzi, H. G., and Nied, H. F., 1987, Blow Molding and Thermoforming
of Plastics: Finite Element Modelling, Comput. Struct., 26(12), pp.
197206.
[8] Warby, M. K., and Whiteman, J. R., 1988, Finite Element Model of Viscoelastic Membrane Deformation, Comput. Methods Appl. Mech. Eng., 68(1),
pp. 3354.
[9] Chung, K., 1989, Finite Element Simulation of PET Stretch/Blow-Molding
Process, J. Mater. Shaping Tech., 7(4), pp. 229239.
[10] Cesar de Sa, J., Natal, R., Silva, C., and Cardoso, R. P., 1999, A Computational Model for Glass Container Forming Processes, Europe Conference on
Computational Mechanics Solids, Structures and Coupled Problems in Engineering, Munich, Germany, August 31September 3.
[11] Hyre, M., 2002, Numerical Simulation of Glass Forming And Conditioning,
J. Am. Ceram. Soc., 85(5), pp. 10471056.
[12] Diraddo, R. W., and Garcia-Rejon, A., 1993, Profile Optimization for the Prediction of Initial Parison Dimensions From Final Blow Moulded Part Specifications, Comput. Chem. Eng., 17(8), pp. 751764.
[13] Diraddo, R. W., and Garcia-Rejon, A., 1993, On-Line Prediction of Final Part
Dimensions in Blow Molding: A Neural Network Computing Approach,
Polym. Eng. Sci., 33(11), pp. 653664.
[14] Lee, D. K., and Soh, S. K., 1996, Prediction of Optimal Preform
Thickness Distribution in Blow Molding, Polym. Eng. Sci., 36(11),
pp. 15131520.
[15] Thibault, F., Malo, A., Lanctot, B., and Diraddo, R., 2007, Preform Shape and
Operating Condition Optimization for the Stretch Blow Molding Process,
Polym. Eng. Sci., 47(3), pp. 289301.
[16] Gauvin, C., Thibault, F., and Laroche, D., 2003, Optimization of Blow Molded
Part Performance Through Process Simulation, Polym. Eng. Sci., 43(7),
pp. 14071414.
[17] Hsu, Y. L., Liu, T. C., Thibault, F., and Lanctot, B., 2004, Design Optimization of the Blow Moulding Process Using Fuzzy Optimization Algorithm,
Proc. Inst. Mech. Eng., Part B, 218(2), pp. 197212.
[18] Yu, J.-C., Chen, X.-X., Hung, T.-R., and Thibault, F., 2004, Optimization
of Extrusion Blow Molding Processes Using Soft Computing and Taguchis
Method, J. Intell. Manuf., 15(5), pp. 625634.
[19] Lochegnies, D., Moreau, P., and Guilbaut, R., 2005, A Reverse Engineering
Approach to the Design of the Blank Mould for the Glass Blow and Blow
Process, Glass Technol., 46(2), pp. 116120. Available at: http://www.ingentaconnect.com/content/sgt/gt/2005/00000046/00000002/art00017
[20] Moreau, P., Marechal, C., and Lochegnies, D., 2001, Optimum Parison Shape
for Glass Blowing, XIXth International Congress on Glass, Society of Glass
Technology, Edinburgh, UK, July 16, pp. 548549.

Journal of Pressure Vessel Technology

[21] Choi, J., Ha, D., Kim, J., and Grandhi, R. V., 2004, Inverse Design of Glass
Forming Process Simulation Using an Optimization Technique and Distributed
Computing, J. Mater. Process. Technol., 148(3), pp. 342352.
[22] Moreau, P., Lochegnies, D., and Oudin, J., 1998, An Inverse Method for Prediction of the Prescribed Temperature Distribution in the Creep Forming Process, Proc. Inst. Mech. Eng., Part C, 212(1), pp. 711.
[23] Groot, J. A. W. M., Giannopapa, C. G., and Mattheij, R. M. M., 2009,
Numerical Optimisation of Blowing Glass Parison Shapes, ASME Paper No.
PVP2009-77946.
[24] Groot, J. A. W. M., Giannopapa, C. G., and Mattheij, R. M. M., 2011,
Development of a Numerical Optimisation Method for Blowing Glass Parison
Shapes, ASME J. Manuf. Sci. Eng., 133(1), p. 011010.
[25] Giannopapa, C. G., and Groot, J. A. W. M., 2007, A Computer Simulation
Model for the BlowBlow Forming Process of Glass Containers, ASME Paper
No. PVP2007-26408.
[26] Sethian, J. A., 1999, Level Set Methods and Fast Marching Methods, Cambridge University, New York.
[27] Sussman, M., Smereka, P., and Osher, S., 1994, A Level Set Approach for
Computing Solutions to Incompressible Two-Phase Flow, J. Comp. Phys,
114(1), pp. 146159.
[28] Adalsteinsson, D., and Sethian, J. A., 1995, A Fast Level Set Method for Propagating Interfaces, J. Comp. Phys, 118(2), pp. 269277.
[29] Chang, Y. C., Hou, T. Y., Merriman, B., and Osher, S., 1996, A Level Set Formulation of Eulerian Interface Capturing Methods for Incompressible Fluid
Flows, J. Comp. Phys., 124(2), pp. 449464.
[30] Groot, J. A. W. M., Mattheij, R. M. M., and Laevsky, K., 2011, Mathematical
Modelling of Glass Forming Processes, Mathematical Models in the Manufacturing of Glass (Lecture Notes in Mathematics), A. Fasano, ed., Vol. 2010,
Springer, Berlin, pp. 156.
[31] Bathe, K., 1996, Finite Element Procedures, Prentice Hall, Englewood Cliffs,
NJ.
[32] Hindmarsh, A. C., Brown, P. N., Grant, K. E., Lee, S. L., Serban, R., Shumaker,
D. E., and Woodward, C. S., 2005, SUNDIALS: Suite of Nonlinear and Differential/Algebraic Equation Solvers, ACM Trans. Math. Software, 31(3),
pp. 363396.
[33] van der Vorst, H. A., 1992, Bi-CGSTAB: A Fast and Smoothly Converging
Variant of Bi-CG for the Solution of Nonsymmetric Linear Systems, SIAM
J. Sci. Stat. Comput., 13(2), pp. 631644.
[34] Kimmel, R., and Sethian, J. A., 1998, Computing Geodesic Paths on Manifolds, Proc. Natl. Acad. Sci. USA, 95(15), pp. 84318435.
[35] Haagh, G. A. A. V., 1998, Simulation of Gas-Assisted Injection Moulding,
PhD thesis, Eindhoven University of Technology, Eindhoven, Netherlands.
[36] Carroll, C. W., 1961, The Created Response Surface Technique for Optimizing
Nonlinear Restrained Systems, Oper. Res., 9(2), pp. 169184.
[37] Schnur, D. S., and Zabaras, N., 1992, An Inverse Method for Determining
Elastic Material Properties and a Material Interface, Int. J. Numer. Methods
Eng., 33(10), pp. 20392057.
[38] Gelin, J. C., and Ghouati, O., 1994, An Inverse Method for Determining
Viscoplastic Properties of Aluminium Alloys, J. Mater. Process. Technol.,
45(14), pp. 435440.
[39] Broyden, C. G., 1965, A Class of Methods for Solving Nonlinear Simultaneous Equations, Math. Comput., 19(92), pp. 577593.
[40] More, J. J., and Trangenstein, J. A., 1976, On the Global Convergence of
Broydens Method, Math. Comput., 30(135), pp. 523540.
[41] Dennis, J. E., and Schnabel, R. B., 1996, Numerical Methods for Unconstrained
Optimization and Nonlinear Equations, SIAM, Philadelphia, PA.
[42] Nocedal, J., and Wright, S. J., 1999, Numerical Optimization, Springer,
New York.

DECEMBER 2014, Vol. 136 / 061301-9

Downloaded From: http://pressurevesseltech.asmedigitalcollection.asme.org/ on 11/13/2015 Terms of Use: http://www.asme.org/about-asme/terms-of-use

You might also like